Download as pdf or txt
Download as pdf or txt
You are on page 1of 220

Prepared by

Dr. Waleed Shaaban Abd-Elsalam

1
2021
1
CONTENTS

Chapter 1
Introduction
Chapter 2
Thermochemistry of Combustion
Chapter 3
Introduction to Mass Transfer
Chapter 4
Chemical Kinetics
Chapter 5
Laminar Premixed Flame
Chapter 6
Laminar Diffusion Flame
Chapter 7
Droplet Evaporation and Burning
2
Chapter 1

INTRODUCTION

Combustion is the oldest technology of mankind; it has been


used for more than one million years. At present, about 90% of
our energy worldwide energy is provided by combustion.

1.1. COMBUSTION DEFINITION

Combustion is a rapid oxidation that generates heat and


frequently with light. The definition clearly implies the
importance of chemical reaction for combustion. The definition
also indicates that combustion is important as it transforms
chemical energy into heat.
3
1.2. IMPORTANCE OF STUDYING COMBUSTION

The fields of combustion are very broad and touch


directly or indirectly nearly all aspects of our life. The
electronic devices we use are generally powered by fossil-fuel-
fired power plants. The cars we drive use internal combustion
engines. The aircrafts we fly in use jet-fuel-powered turbine
engines. Most of the materials we use have been made through
some type of heating or melting process.
About 90% of energy used in the world comes through
combustion. Electricity generation is primarily provided by
fuel combustion. Transportation systems are based almost
entirely on combustion. Ground vehicles are exclusively
powered by burning fuels using gas turbines or spark ignition
engines while most trains are driven by diesel engines.
Combustion has been the foundation of worldwide
industrial development for the past 200 years. As shown in
Table 1.1, industries rely heavily on combustion process. The
major uses for combustion in industry are reported in Table
1.2. A major use for combustion is metal industries including
iron, steel, aluminum and other metal refining. Petroleum

4
refining and cement industries extensively use energy produced
by combustion. Many others industrial processes meet their
energy need by combustion include among others: boilers,
metal melting, metal heating, metal treating, glass melting,
glass treating, fluid heating, clay firing and drying. The
demand for energy in the industrial sector is expected to
continue to increase rapidly. Most of the energy (88%) is
produced by the combustion of fossil fuels like oil, natural gas,
and coal.

Combustion is also used to dispose waste materials.


Incineration (waste disposal by combustion) is an old method;
5
however, it is receiving an attention as the availability of
landfill sites is limited and it safely disposes toxic wastes.

6
However, combustion has some downsides related to
environmental pollution. The major pollutants generated due to
combustion include unburned hydrocarbons, nitrogen oxides,
carbon monoxide, sulfur oxides and particulate matter. Primary
pollution concerns relate to specific health hazards, smog, acid
rain, global warming and ozone depletion.

1.3. COMBUSTION MODES AND FLAME TYPES

Combustion can occur in two modes. The first mode is a


flame mode in which fuel oxidation occurs in a certain region.
The second mode of combustion is a non-flame mode where
fuel oxidation occurs nearly in all volume. This is the case of
autoignition that causes knock phenomenon in spark ignition
engine.
In combustion processes, oxidizer and fuel should be
mixed and burned. The flame can be classified based upon the
sequence of mixing and burning. The flame is called premixed
flame if the fuel and oxidizer are mixed first and burned later.
On the other hand, it is called non-premixed or diffusion flame
if mixing and burning occur simultaneously. Each of these
modes can be further subdivided based on whether the fluid
7
flow is laminar or turbulent. Table 1.2 gives combustion
system examples for each mode of flame.

Table 1.3 Examples of combustion system for flame modes


based on premixedness and fluid flow type

8
Chapter 2

THERMOCHEMISTRY OF COMBUSTION

In this chapter, some basic relations for ideal gases and their
mixtures are perused as they are an integral part of combustion studying.
Several thermodynamics topics related to combustion and reacting system
are also discussed. Chemical equilibrium is developed and applied to
combustion product mixtures.

2.1 REVIEW OF PROPERTIES RELATIONS


2.1.1 Extensive and Intensive properties

The extensive property is one that increases in proportion to the


size of the system such as volume (V), internal energy (U), enthalpy (H).
It is usually denoted with capital letters. On the other hand, the intensive
8
property is one that unchanged when the size of the system is increased
by adding identical systems, like specific enthalpy (h), specific internal
energy (u), pressure (P) and temperature (T). The intensive properties are
usually denoted with lower case letters except pressure and temperature.
Extensive properties are simply obtained from the corresponding
intensive properties by multiplying the property value per unit mass (or
mole) by the amount of mass, m, (or number of moles,n), i.e.,

V  mv  nv (2.1a)
U  mu  nu (2.1b)
H  mh  nh (2.1c)

Where the symbols with overbar, e.g., u and h are molar based intensive
properties.

2.1.2 Equation of State


An equation of state is the relationship among the pressure, P,
temperature, T, and volume V of a substance. For an ideal gas, the
equation has the forms:

PV  mRT (2.2a)
PV  NR T (2.2b)

Pv  RT (2.2c)
P  RT (2.2d)

Where the specific gas constant R is related the universal gas constant Rµ
(=8315 J/lmol-K) and the molecular weight MW by
R  R / MW (2.3)

01
The assumption of ideal gas appears appropriate for all combustion
systems considered as the temperatures associated with combustion
processes are high that results in low densities.

2.1.3 Calorific Equation of State


The calorific equations of state are the expressions relating internal
energy (or enthalpy) to temperature and volume (or pressure), i.e.,
u  u(T , v) (2.4a)
h  h(T , P) (2.4b)
The above equations can be written in differentiation form as:
u u
du  ( ) v dT  ( )T dv (2.5a)
T v
h u
dh  ( ) P dT  ( )T dP (2.5b)
T P
u
For ideal gas, the partial derivatives with respect to volume, ( )T , and
v
u
with respect to pressure, ( )T are zero and the calorific equations of
P
state may be written as:
T
u (T )  uref   cv dT (2.6a)
Tref

T
h(T )  href   c p dT (2.6b)
Tref

where cv and cp are the constant-volume and constant-pressure specific


heats, respectively.
u
cv  ( )v (2.7a)
T

00
u
cp  ( )P (2.7b)
T
In general, the more complex the molecule, the greater its molar
specific heat is. The specific heats cv and cp are functions of temperature.
The more complex molecules have also greater temperature dependence.

2.1.4 Ideal-Gas Mixtures

Constituents’ mass fractions and mole fractions are two useful


concepts to characterize the composition of a mixture of gases. The mass
fraction of species i, Yi, is defined as the ratio between the mass of
species i, mi, to the total mass of mixture, mmix:
mi
Yi  (2.8)
mmix
In the same way, The mole fraction of species i,  i , is the ratio between
the number of moles of species I, Ni, to the total number of moles of the
mixture, Nmix.
Ni
i  (2.9)
N mix
It is obvious, by definition, that the sum of all the constituent mass (or
mole) fractions must be unity, i.e.,

 Yi  1 (2.10a)
i

 i  1 (2.10b)
i

Mass fractions and mole fractions can be converted from to another using
the following expressions
i  Yi MWmix / MWi (2.11a)

01
Yi  i MWi / MWmix (2.11b)

Where MWmix is the mixture molecular weight and can be calculated as:

MWmix  1 /  Yi / MWi (2.12a)


i

MWmix    i MWi (2.12b)


i

For ideal gases, the mixture pressure is the sum of the constituent partial
pressures,
P   Pi (2.13)
i

The partial pressure may be related to the mixture composition and total
pressure as:
Pi  i P (2.14)

The different properties of mixtures can be determined as mass (or


mole) fraction weighted sums of the individual species properties. For
instance, specific internal energies of a mixture are calculated as
u mix    i ui (2.15a)
i

u mix   Yi ui (2.15b)
i

Entropy of a mixture is also calculated as a mass (or mole) fraction


weighted sum of constituents,
smix (T , P)    i si (T , Pi ) (2.16a)
i

smix (T , P)  Yi si (T , Pi ) (2.16b)


i

On the contrary of properties like enthalpy and internal energy of


ideal gases, the entropies of pure species depend on the species partial

02
pressures. The constituent specific entropies can be determined using
standard-state (Pref=Po=1 atm) values as:
Pi
si (T , Pi )  si (T , Pref )  R ln (2.17a)
Pref
Pi
si (T , Pi )  si (T , Pref )  R ln (2.17b)
Pref

2.1.5 Latent Heat of Vaporization


In many combustion processes, a liquid-vapor phase change is
important. For example, a liquid fuel droplet must first vaporize before it
can burn. Water vapor in combustion products can condense if they are
sufficiently cooled. Latent heat of vaporization (also known as the
enthalpy of vaporization) may be defined as the heat required to
completely vaporize a unit mass of liquid in a constant pressure process at
a given temperature, i.e.,

h fg (T , P)  hvapor (T , P)  hliquid (T , P) (2.18)

The latent heat of vaporization at a given saturation pressure and


temperature is frequently used with the simplified Clausius-Claperyon
equation to estimate saturation pressure variation with temperature:

dPsat h fg dTsat
 2
(2.19)
Psat R Tsat

The above equation, assuming hfg is constant, can be integrated from


point 1 (Psat,1,Tsat,1) to point 2 (Psat,2,Tsat,2) to permit, for example,
estimating Psat,2 from a knowledge of Psat,1,Tsat,1 and Tsat,2, i.e.,
03
Psat , 2 h fg 1 1
 exp[  (  )] (2.20)
Psat ,1 R Tsat , 2 Tsat ,1

2.2 FIRST LAW OF THERMODYNAMICS


2.2.1 First Law-closed system (Fixed mass)

Conservation of energy is the fundamental principle embodied in


the first law of thermodynamics. Considering a fixed mass presented in
Fig. 2.1a, energy conservation can be expressed for a finite change
between states, 1 and 2, as

1 Q2 1W2  E12 (2.21)


where

1 Q2 is the heat added to the system in going from state 1 to state 2,

1W2 is the work done by the system on the surroundings in going from
state 1 to state 2,
E1 2 is the change in total system energy (sum of internal, kinetic
and potential energies) in going from state 1 to state 2.

2.2.2 First Law-Open System (Control Volume)

For a control volume, fluid may flow across the boundaries, see
Fig. 2.1b. In the case of steady-state, steady-flow (SSSF), the first law
can be expressed as:

Q cv  Wcv  m
 eo  m
 ei  m
 ( Po vo  Pi vi ) (2.22)
where

04
Q cv is the rate of heat transferred across the control surface from the
surroundings to the control volume
W cv is the rate of all work done by the control volume, including shaft
work, but excluding flow work,
m eo is the rate of energy flowing out of the control volume,
m ei is the rate of energy flowing into of the control volume,
 ( Po vo  Pi vi ) is the rate of work associated with pressure forces where
m
fluid crosses the control surface, flow work
m is the rate of mass crossing the control volume,
e is the specific energy of the system.

Fig. 2.1 (a) Schematic of fixed-mass system with moving boundary above the piston.
(b) Control volume with fixed boundaries and steady state flow.

Considering that enthalpy is the sum of internal energy and flow work,

h  u  Pv (2.23)

and the total system energy is the sum of internal, kinetic and potential
energies,

05
1
E  m(u   2  gz ) (2.24)
2

The first law for control volume can have the following form:
1
Q cv  W cv  m
 [(ho  hi )  ( o2   i2 )  g ( z o  z i )]
2
(2.25)
or on a specific mass basis,
1
q cv  w cv  (ho  hi )  ( o2   i2 )  g ( z o  z i ) (2.26)
2

Where lower case letters are used to denote mass specific quantities.

06
2.3 REACTANT AND PRODUCT MIXTURES
2.3.1 Stoichiometry

The stoichiometric quantity of oxidizer is just that


amount needed to completely burn a quantity of fuel. A
mixture of fuel and oxidizer is said to be fuel-lean, or just lean if it
contains a quantity of oxidizer more than stoichiometric; while that
containing less than stoichiometric is said to be fuel-rich or rich
mixture. The stoichiometric oxidizer (or air) fuel ratio (mass) is
determined by writing simple atom balances, assuming that the fuel reacts
to form an ideal set of products. The stoichiometric relation for a
hydrocarbon fuel represented by CxHy can be expressed as

C x H y  a(O2  3.76 N 2 )  xCO2  ( y / 2) H 2O  3.76aN 2


(2.27)
where

a  x y/4 (2.28)

Considering that the composition of air is 21% O2 and 79% N2 (by


volume), i.e., that for each mole of O2 in air there are 3.76 moles of N2,
the stoichiometric air-fuel ratio can be found as

mair 4.76a MWair


( A / F ) stoic  ( ) (2.29)
m fuel 1 MW fuel

where MWair and MWfuel are the molecular weights of air and fuel,
respectively. Table 2.1 reports stoichiometric air-fuel ratios for methane
07
and solid carbon. It is also reported the oxygen –fuel ratio for combustion
of H2 in pure O2. It is noted that (O/F)ratio much greater than one
for all systems.
A/F 12-18 Spark ignition
A/F 18-70 Diesel Engine
A/F 100-200 Gas turbine

Table2.1 Some combustion properties of methane, hydrogen and solid carbon


for reactants at 298 K

Equivalence Ratio, Φ, is used to indicate quantitatively whether


a fuel-oxidizer mixture is rich, lean or stoichiometric. The equivalence
ratio is defined as

( A / F ) stoic ( F / A)
  (2.29)
(A/ F) ( F / A) stoic
The definition implies that
for fuel-rich mixtures,   1
for fuel-lean mixtures,   1 .
for a stoichiometric mixture, Φ = 1.

08
There are other different parameters used to define relative
Stoichiometry. Among others,
Percent stoichiometric is defined as

100%
%stoichiometric _ air 

(2.30)

Percent excess air, or

(1  )
%excess _ air  *100%
 (2.31)

Example 1.
Butane (C4H10) burns with air at an equivalence ratio of
0.75. Determine the number of moles of air required per mole of
fuel.

Solution
The following expression may be used to calculate the (A/F) stoich for a
hydrocarbon fuel of the form CxHy.

MWair
( A / F ) stoic  4.76a
MW fuel (aa)

Where a  x  y / 4  4  10 / 4  6.5

28.85
( A / F ) stoic  4.76 * 6.5 *  15.33
58.12
11
( A / F ) stoic 15.33
By definition ( A / F )    20.44
 0.75
MW fuel
( A / F ) molar  ( AlF )
MWair
58.12
 20.44 *  41.17 Mole air/mole fuel
28.85

Another solution
MW fuel ( AlF ) stoic MW fuel
( A / F ) molar  ( AlF ) 
MWair  MWair
Using the eqn. (aa), we can get

4.76a 4.76 * 6.5


( A / F ) molar    41.25
 0.75 Mole air/mole fuel

Example 2

Consider a fuel which is an equimolar mixture of propane (C3H8) and


natural gas (CH4). Write out the complete stoichiometric combustion
reaction for this fuel burning with air and determine the stoichiometric
fuel-ratio on a molar basis. Also, determine the molar air-fuel ratio for
combustion at an equivalence ratio, Φ, of 0.8.

Solution:
The complete stoichiometric combustion reaction for the described fuel
can be expressed as

10
C3 H 8  CH 4  7 * (O2  3.76 N 2 )  4CO2  6H 2 O  26.32 N 2

The stoichiometric fuel-ratio on a molar basis is determined as


Number of mole of fuel
( FlA) Stoic,molar 
Number of mole of air
11
( FlA) Stoic,molar   0.06
7 * (1  3.76)
The equivalence ratio is define as

( FlA) ( FlA) molar


 
( F / A) stoic ( F / A) molar, stoic
( FlA)molar   * ( F / A)molar, stoic  0.8 * .06  0.048

( AlF )molar  1/( F / A)molar  20.83

11
Enthalpy of Formation and Absolute Enthalpy

o
The enthalpy of formation, h f ,i of a particular species,

i, is defined as the heat of reaction per mole of product


formed isothermally from elements of their standard states.

The standard state is chosen as the most stable form of the element at
1 atm and 25 oC. For oxygen and Nitrogen the standard state is
gaseous O2 and N2.

The enthalpy of formation of elements in their standard states is


assigned a value zero.

The absolute enthalpy (Standardized enthalpy)


is the sum of the enthalpy of formation, hf, at reference
temperature, which is associated with chemical bonds, and the
sensible enthalpy change, Δhs, which is associated with temperature
change. The molar absolute enthalpy for species i can be written as

hi (T )  h (Tref )  hs ,i
o
f ,i

Where hs ,i  hi (T )  h fo,i


For standard state, temperature, Tref and pressure, Po, are
considered to be 25 oC and 1 atm, respectively.

12
Example
Determine the absolute enthalpy for the given mixture at
temperature of 1000 K, on both a mole basis (kJ/kmol) and a mass
basis (kJ/kg)
Species CO CO2 H2O N2 NO
No. of moles 0.095 6 7 34 0.005

Solution
Using the mole fraction definition

xi  N i / N T
and using tables of thermodynamics properties of gases

Species CO CO2 H2O N2 NO total


No. of moles 0.095 6 7 34 0.005 47.1
xi 0.002 0.127 0.149 0.7219 0.0001 1
h fo,i (kJ/kmol) -110541 -393546 -241845 0 90297

hs ,i (1000) 21697 33425 25993 21468 22241

hs ,i  hi (T )  h fo,i

The absolute enthalpy of the mixture is calculated as

hmix   xi hi   xi [h fo,i  hs ,i (T )]


=0.002*(-110541+21697)+0.127*(-393546+33425)+
0.149*(-241854+25993)+0.7219*(0+21468)+
0.0001*(90297+22241)
=-78065 kJ/kmolmix

13
The molecular weight of mixture is required to determined hmix

MWmix   xi MWi
=0.002*28.01+0.127*44.011+.147*18.016+0.7219*28.013
+0.0001*30.006
=28.52 kmol/kg

Then,

hmix  hmix / MWmix


=-78065/28.52
=-2737 kJ/kgmix

14
Enthalpy of Combustion and Heating Values

The science of thermochemistry is concerned with the heat


changes associated with chemical reaction; in other words, it deals
essentially with the conversion of chemical energy into heat energy,
and vice versa.
Applying the steady flow form of the first law, the definition of
the enthalpy of reaction, or the enthalpy of combustion, ΔhR, (per
mass of mixture), may be mathematically expressed as

Figure 2.3 Steady-flow system used to determine enthalpy of combustion

  1 2
Qcv  Wcv  m [(ho  hi )  ( o   i2 )  g ( z o  z i )]
2

hR  qcv  hprod  hreac

15
The combustion process is assumed to be complete. A stoichiometric
mixture is considered where reactants enter and products exit, both
at standard state conditions (25 oC, 1 atm).

The heat of combustion, Δhc, (heat value), is numerically equal


to the enthalpy of reaction, but with opposite sign. The upper or
higher heating value, HHV, is the heat of combustion calculated
assuming that all of the water in the products has condensed to
liquid. In this case a greater amount of energy is liberated, hence the
designation "upper". The lower heating value, LHV, is
corresponding to the case where none of water is assumed to
condense.

Figure 2.4 Enthalpy of combustion for stoichiometric methane air mixture with
water in the vapor state.

16
Example
The lower heating value of vapor n-decane (C10H22) is 44,597
kJ/kg at T=298 K. The enthalpy of vaporization of n-decane is 276.8
kJ/kg of n-decane. The enthalpy of vaporization of water at 298 K is
2442.2 kJ/kg of water. Determine the lower heating value of liquid n-
decane in kJ/kg. Determine also the higher heating value of vapor n-
decane at 298 K.

Solution

LHVl  LHVg  h fg , n  decane


LHVl  44597  276.8
=44320.2 kJ/kgC10H22

The higher heating value of vapor n-decane at 298 K, HHVg, may be


calculated in this way

C10 H 22  15.5(O2  3.76 N 2 )  10CO2  11H 2 O  58.28N 2

MWH 2O y / 2
HHVg  LHVg  h fg , H 2O *
MWC x H y
18.016 * 22 / 2
HHVg  44597  2442.2 *
142.284
=47999 kJ/kg n-decane
17
Figure 2.5 Different heating values

Example
Determine the enthalpy of formation in kJ/kmol for methane,
CH4, given the lower heating value of 50,016 kJ/kg at 298 K.

Solution
For 1 mole of CH4, the combustion equation can be written as

CH 4  2(O2  3.76 N 2 )  CO2  2H 2 O  7.52 N 2


18
The LHV is expressed as

H R  MW fuel * N f * LHV  H prod  H reac


H Re ac  H prod  MW feul *1* LHV
1* h fo,CH 4  (1* 393546  2 * 241845  0)  16.043 *1* 50016

h fo,CH 4  74829 kJ/kmolCH4

21
Adiabatic Flame Temperature

There are two adiabatic flame temperatures to be defined: one


for constant-pressure combustion and one for constant-volume.
When a fuel-air mixture burns adiabatically at constant pressure, the
absolute enthalpy of the reactants at the initial state (say, T=298 K,
P=1 atm) and the absolute enthalpy of the products at the final state
(T=Tad, P=1 atm) are equal. Appling the first law of thermodynamics
for steady state, steady flow results in

hR  qcv  hprod  hreac


H reac (Ti , P)  H prod (Tad , P)

Or, equivalently, on a per-mass-of-mixture basis

hreac (Ti , P)  hprod (Tad , P)

The constant-pressure adiabatic flame temperature would be suitable


when dealing with a gas turbine and furnace. On the other side, the
analysis of an ideal Otto cycle may require the constant-volume
adiabatic flame temperature. For fixed mass, constant volume
process, the first law is reduced to

U reac (Ti , P)  U prod (Tad , Pf )

20
As most compilations or calculations of thermodynamic properties
provide values for enthalpy rather than internal energy, the above
equation can be rearranged to the following form:

H reac  H prod  V ( Pini  Pf )  0

Appling the ideal-gas law the PV term can be replaced

PiniV   N i Ru Tini  N reac Ru Tini


reac

PprodV  N R T
prod
i u ad  N prod RuTad

Thus,

H reac  H prod  Ru ( N reacTini  N prodTad )  0

An alternative form on a per-mass-of –mixture basis, can be


expressed as

Tini Tad
hreac  hprod  Ru (  )0
MWreac MWprod

where MW  m / N and h  H /m

21
Example
Determine the adiabatic flame temperature for constant-
pressure combustion of stoichiometric propane (C3H8)–air mixture
assuming reactants at 298 K, no dissociation of the products, and
constant specific heats evaluated at 1300 K. Also, determine the
adiabatic flame temperature for constant- volume combustion,
estimating the final pressure.

Solution

For 1 mole of CH4, the stoichiometric combustion equation can


be written as

C3 H8  5(O2  3.76 N 2 )  3CO2  4H 2O  18.8N 2

For adiabatic flame temperature at constant pressure, the first law of


thermodynamics is expressed as

H reac (Ti , P)  H prod (Tad , P)


Using tables of thermodynamics properties of gases and fuels
Species Enthalpy of formation Specific heat @
@ 298 K, kJ/kmol 1300 K, kJ/kmol.K
C3H8 -103,847 -
CO2 -393,546 56.984
H2O -241,845 45.027
O2 0 -
N2 0 34.113

22
H react   Ni hi 
react
 i f ,i  hs, i)
N (
react
h o

 1* (-103,847  0)  5 * (0  0)  18.8 * (90  0)

= - 103,847 kJ

Calculating Hprod considering hs ,i  c p,i * (Tad  Tref )


H prod  3 * [393,546  56.984 * (Tad  298)]  4 * [241,845
 45.027 * (Tad  298)]  18.8 * [0  34.113 * (Tad  298)]

H prod  992 .4Tad  2,443749


Equating Hreac to Hprod

992.4 * Tad  2,443,749  103,847

solving for Tad yields

Tad  2357 K

For adiabatic flame temperature at constant volume, the first law of


thermodynamics is expressed as

H reac  H prod  Ru ( N reacTini  N prodTad )  0

23
We will evaluate the properties at a higher temperature; say 1500 K,
as the constant-volume adiabatic flame temperature is expected to be
quite higher than the constant-pressure adiabatic flame temperature.

Using tables of thermodynamics properties of gases and fuels

Species Enthalpy of formation Specific heat @


@ 298 K, kJ/kmol 1500 K, kJ/kmol.K
C3H8 -103,847 -
CO2 -393,546 58.292
H2O -241,845 47.103
O2 0 -
N2 0 34.805

The enthalpy of reactants is still the same, thus

H reac  - 103,847 kJ
H prod  3 * [393,546  58.292 * (Tad  298)]  4 * [241,845
 47.103 * (Tad  298)]  18.8 * [0  34.805 * (Tad  298)]

H prod  1017.6 * Tad  2,451,269

Ru ( N reacTini  N prodTad )  8.315 * (24.8 * 298  25.8 * Tad )

 61451  214.5 * Tad


24
Reassembling the energy equation and solving for Tad yields

- 103,847 - (1017.6 * Tad  2,451,269)  (61451  214.5 * Tad )  0

Tad  2846 K
The final pressure, Pf, may be calculated as

Tad * N prod 2846 * 25.8


Pf  Pi  1*
Tini * N reac 298 * 24.8
Pf  9.935 atm

25
CHEMICAL EQUILIBRIUM

The ideal combustion products for burning a hydrocarbon fuel


with air are CO2, H2O, O2 and N2. Actually these species dissociate
yielding many other species, such as H2, OH, CO, H, O, NO and N,
this is particularly true at higher temperatures. Determination of the
mole fractions of all of the product species at a given temperature
and pressure is the subject of this part. The equilibrium-constant
approach is used, limiting our study to the ideal gases applications.

Second Law Considerations


For a fixed-volume, adiabatic reaction process (constant U, V
and m), the second law requires that the internal entropy change

dS  0
It implies that composition of a system will spontaneously shift
toward the point of the maximum entropy when approaching from
either side, since dS is positive. Once the maximum entropy is
reached, no further change in composition is allowed, since this
would require the system entropy to decrease in violation of the
second law of thermodynamics. Mathematically, the condition of
equilibrium can be expressed as.

26
(dS mix )U ,V ,m  0

Figure 2.7 Chemical equilibrium for fixed-mass isolated system

The entropy of the product mixture can be calculated by summing


the product species entropies, i.e.
n
S mix (T f , P)   N i S i (T f , Pi )
i 1
where Ni is the number of moles of species i in the mixture. The
individual species entropies are obtained from

27
Tf
dT Pi
S i  S (Tref )  c  Ru ln o
o
i p ,i
Tref
T P
where the ideal-gas behavior is assumed and Pi is the partial pressure
of the ith species.

Gibbs Function

Gibbs free energy, G, is a very useful thermodynamics


properties in establishing chemical equilibrium and calculating the
composition of a mixture at a given temperature, pressure and
stoichiometry.
Mathematically, Gibbs free energy may be defined in terms of
other thermodynamic properties as

G  H  TS
(dG)T , P,m  dH  SdT  TdS
(dG)T , P,m  TdS
(dG)T , P,m  0 since dS  0
This inequality states that the Gibbs function always decreases
for spontaneous, isothermal, isobaric change of a fixed-mass system
28
in the absence of all work except boundary (-PdV) work. Therefore,
the Gibbs function attains the minimum in equilibrium, in contrast to
the maximum in entropy. Hence, at equilibrium

(dG)T , P,m  0
For a mixture of ideal gases, the Gibbs function for the ith species is
given by

g i ,T  g io,T  RuT ln( Pi / P o )

Where g io,T is the Gibbs function of the pure species at the standard-
state pressure (i.e. Pi=Po) and Pi is the partial pressure. The standard
–state pressure, Po, by convention is taken to be 1 atm.
The Gibbs function for a mixture of ideal gases can be expressed as

Gmix   N i g i ,T   N i [ g io,T  Ru T ln( Pi / P o )]

where Ni is the number of mole of the species i.

For fixed temperature and pressure, the equilibrium condition


becomes

31
dGmix  0
Or

 dN [ gi
o
i ,T  RuT ln( Pi / Po )]   Ni d[ gio,T  RuT ln( Pi / Po )]  0

The second term in the above equation can be shown to be zero by

recognizing that d (ln Pi )  dPi / Pi , V  Ru TN i / Pi and

 dP  0 , since the sum of all changes in the partial pressure


i

must be zero because the total pressure is constant. Therefore,

dGmix  0   dN i [ g io,T  RuT ln( Pi / P o )] xx

For the general system, where

aA  bB  .....  eE  fF  ......

The change in the number of moles of each species is directly


proportional to the stoichiometric coefficient, i.e.

dN A  ka yy
dN B  kb
.
.
dN E  ke
dN F  kF
30
Substituting Eqn. yy into Eqn. xx and canceling the proportionality
Constant k, it is obtained

 a[ g Ao ,T  Ru T ln( PA / P o )]  b[ g Bo ,T  Ru T ln( PB / P o )]  ......


 e[ g Eo ,T  Ru T ln( PE / P o )]  f [ g Fo ,T  Ru T ln( PF / P o )]  0

The above equation can be rearranged and the log terms grouped
together to yield

(eg Eo ,T  fg Fo ,T  .........  ag Ao ,T  bg Bo ,T  ..........)


( PE / P o ) e ( PF / P o ) f .etc
  RuT ln
( PA / P o ) a ( PB / P o )b .etc

The term in parentheses on the left-hand-side is called the

standard-state Gibbs function change, GT , i.e.,


o

GTo  (eg Eo ,T  fg Fo ,T  ......  ag Ao ,T  bg Bo ,T  ......)


or, alternatively,

GTo  (eg of , E  fg of , F  ......  ag of , A  bg of , B  ......)T

where g of ,i is the Gibbs function of formation of species i which is


frequently employed in dealing with reacting system and is defined as

31
Where the  j are the stoichiometric coefficients of the elements required
to form one mole of the compound of interest. For example, the
coefficients are  O 2  1 / 2 and  C  1 for forming a mole of CO from O2
and C, respectively. The Gibbs functions of formation of the naturally
occurring elements are assigned values of zero at reference state.

The equilibrium constant, Kp is defined as

o e o f
( PE / P ) ( PF / P ) .etc
Kp  o a o b
( PA / P ) ( PB / P ) .etc

The equation of chemical equilibrium at constant temperature and


pressure is then given by

G   RuT ln K p
o
T
Or

K p  exp(G / RuT ) o
T ss

If GTo is positive, reactants will be favored since ln K p is negative,


which requires that Kp itself is less than unity. On the contrary, if

GTo is negative, the reaction tends to favor products. Physical

32
insight to this behavior can be obtained by recalling the definition of
G in terms of the enthalpy and entropy.

G  H  TS
o
T
o o

which can be substituted in Eqn. ss:

(  H o / RuT ) ( S o / RuT )
Kp  e .e
The last expression implies that the reaction tends to favor the

product, which means K p  1, when the enthalpy change, ΔHo, is


negative, i.e, the reaction is exothermic and energy of the system
becomes lower. Additionally, positive changes in entropy, which

indicate greater molecular chaos, leads to values of K p  1 .

Example
Consider the equilibrium reaction O2  2O in a closed vessel
contains 1 mole of O2 when there is no dissociation. Calculate the
mole fraction of O2 and O for the following conditions:
A. T=2500 K, P=1.0 atm
B. T=2500 K, P=3.0 atm

Solution
Two equations are required to find the mole fractions, χO2 and χO.
The first equation is due to an equilibrium expression recognizing
that a=1 and b=2.
Evaluating the standard-state Gibbs function change at T=2500 K

33
GTo  (bg of ,O  ag of ,O 2 )T  2500K
 (2 * 88203  1* 0)
=176206 kJ/kmol.

From the definition of Kp, we can have

( PO / P o ) 2
Kp 
( PO2 / P o )1
The equilibrium constant Kp can be rewritten in terms of mole
fractions.

( O )2
Kp  ( P / P o
)
(  O2 )1

Reassembling the chemical equation equilibrium

( O )2
( P / P o
)  exp(  G o
T / RuT )
(  O2 )1

 176206
 exp( )
8.315 * 2500
 O2
( P / P o )  0.9916
O 2
V1

34
The second equation can be obtained by equating the entire mole
fractions sum to unity

i
i 1

Or

O  O  1
2 V2

Solving Eqn.s V1 and V2, we can get χO2, and χO

T=2500 K T=2500 K
P=1 atm P=3 atm
χO2 0.3834 0.5671
χO 0.6166 0.4329

Example 2.8

Solution

35
36
or

37
38
41
Complex Systems

Actually, in most combustion systems, many species and


several simultaneous equilibrium reactions are important. If n
reactions are considered, there will be n reaction equilibrium
equations. Additionally, m element-conservation expression may be
obtained. Thus, we have a set of (n+m) equations for (n+m)
unknowns. A method of simultaneously solving nonlinear equations
such as generalized Newton-Raphson's method should be applied to
the system.

EQUILIBRIUM PRODUCT OF COMBUSTION

Full Equilibrium
The adiabatic flame temperature and the detailed composition
of the product of combustion can be obtained by solving
simultaneously the equations that represent the first law, complex
chemical equilibrium principles and atom conservation.
Major products of lean combustion are usually H2O, CO2, O2,
and N2; while for rich combustion, they are H2O, CO2, CO, H2 and
N2. It is interesting to indicate that the maximum flame temperature
occurs not at stoichiometric, but, rather, at slightly rich equivalence
ratio. This later is a consequence of the heat combustion, ΔHC, and
heat capacity of the products ( N prod .c p , prod ) . For equivalence ratios

between   1 and  (Tmax ) , the heat capacity decreases more rapidly

40
with  than ΔHC; while beyond  (Tmax ) , ΔHC fall more rapidly than
does heat capacity.

Water-Gas Equilibrium
As indicated above, H2O, CO and H2 are major products for
rich combustion. Under these conditions the so-called water-gas shift
reaction must be taken into account. The reaction is described as

CO  H 2 O  CO2  H 2
The water-gas equilibrium is central to steam reforming in the
petroleum industry and gasification processes.

Pressure Effects
Pressure has a significant effect on dissociation with reactions
having changes in total number of moles. Increasing pressure results
in a decrease in CO2 and H2O dissociation. The trend is consistent
with Le Chatelier principle that states that any system initially in
equilibrium state when subjected to change will shift in composition in
such a way to minimize the change. For example, an increase in
pressure upon CO2  CO  1 / 2O2 reaction, this should cause a
shift to the CO2 side, i.e., in the direction to produce fewer moles.
For equimolar reactions, pressure has no effect.

41
Chapter 3

INTRODUCTION TO MASS TRANSFER

Understanding combustion requires a combined knowledge of


thermodynamics, heat and mass transfer, and chemical kinetics.
Mass transfer is a fundamental topic in chemical engineering which
is quite complex. This chapter is dedicated to provide a simplified
treatment of the fundamental rate laws and conservation principles.
Imagine opening a bottle of perfume and placing the opened
bottle in the center of a room. The process whereby the perfume
molecules are transported from a region of high concentration, near
the bottle, to a region of low concentration, far from the bottle, is a
subject of mass transfer. Mass may be transferred by molecular
processes which are relatively slow and operate on small spatial
scales. A second mode whereby mass is transported is turbulent
processes that depend upon fluid dynamic conditions.

35
Mass Transfer Rate Laws
Fick's Law of Diffusion Fick's law describes the rate
at which one species, A, diffuses through the other, B. It is
on a mass basis

dYA
  YA (m
m A
 )  DAB
 m
" "
A
"
B
dx (3.1)

where:

m "A is the mass flux or the mass flow per unit area of
species A (kg/m2s)

m B" is the mass flux of species B (kg/m2s)

 "A  m
YA (m  B" ) is the mass flow of species A associated with bulk
flow per unit area (kg/m2s)
dYA
DAB is the mass flow of species A associated with
dx
molecular diffusion per unit area (kg/m2s)

YA is the mass fraction of species A

ρ Density (kg/m3)

DAB is the binary diffusivity (m2/s)

The mass flux is defined as the mass flow of species A per unit area
perpendicular to the flow:

35
 "A  m
m A/ A (3.2)
The second term on the right-hand-side is representing the diffusion
 "A,diff , superimposed on the bulk flow
flux of species A, m

dYA
m "
A, diff   DAB
dx (3.3)

The expression demonstrates that the diffusion flux is proportional to


the gradient of mass fraction. It is obvious that species A travels from
the region of high concentration to a region of low concentration.
This is analogous to energy transferring in the direction from high
temperature to low. It is also worthwhile to note that the negative
sign causes the flux to be positive in the x-direction when the
concentration gradient is negative.
Applying the concepts of the kinetic theory of gases, the binary
diffusivity, DAB, is expressed as
1/ 2
2  k BT  T
D AB   
3  m A   2P (3.4)

where kB is Boltzmann's constant (1.380 6504×10−23 J·K-1), mA is the


mass of a single A molecule and σ is the diameter of both A and B
molecules.
From Eqn. (3.4), we may express

DAB  T 3/ 2
/P (3.5)

33
It is evident that that the diffusivity depends strongly on temperature
and inversely with pressure. The mass flux of species A, however,
depends on the product DAB , which has square-root temperature
dependence and is independent of pressure

DAB  T 1/ 2
(3.6)
In many simplified analyses of combustion processes, the weak
temperature dependence is neglected and DAB is treated as constant.

Species Conservation

dm A
dt
[m A A]x [m A A]x x

 A
m

x
Figure 3.1. Control volume for one dimensional analysis of conservation
of species A.

35
Referring to Fig. 3.1, species A flows into and out of the control volume
due to both of bulk flow and diffusion. Species A may be generated or
disappears due to chemical reaction. The net rate of change in the mass of
A within the control volume can be expressed as.

dmA,CV
 [m A A]x  [m A A]x  x  m AV
dt (3.7)

 A is the mass production rate of species A per unit volume


where m
(kgA/m3s) and A is the area of the control volume perpendicular to the

flow. Taken into consideration that mA,CV  YAmCV  YA VCV and,


Equation 3.7 can be rewritten as:

( YA )  dY 
Ax  AYA m   DAB A  
t  dx  x

 dYA 


AYA m  DAB
   m AAx
 dx  x  x (3.8)

Dividing through by AΔx and taking the limit as x  0 , Eqn. 3.8


becomes

( YA )   dYA 
  YA m   DAB   A
m
t x  dx 
(3.9)

 ( YA )
For the case of steady flow where
t , Eqn. (3.9) becomes
35
d  dYA 
m A  YA m A  DAB  0
dx  dx  (3.10)

SOME APPLICATIONS OF MASS TRANSFER

Stefan Problem

Consider the diffusion system shown 3.2. Liquid A is


maintained at a fixed height in a tube using some device. A stream of
gas mixture A-B flows at the top of the tube. When there is a
difference in concentration, a deriving force for mass transfer will
cause species A to diffuse from the liquid –vapor interface (region of
higher concentration) to the flowing gas stream (region of lower
concentration). Considering a case of steady state and assuming that
B is insoluble in liquid A, therefore, the net flux of species B in the
tube is zero and a stagnant layer of B is established in the tube.

35
Stream mixture A-B

x=L YA,∞ at x=L

m "A

YA,i at interface
x=0
Liquid A

Figure 3.2 Diffusion of vapor A through a stagnant column of gas B.

Mathematically, the mass flux is given by Eqn. 3.1

dYA
  YA (m
m "
A
 )  DAB
 m "
A
"
B
dx (3.1)

Since m B  0 , the above Eqn. can be reduced to


"

dYA
m  YAm  DAB
"
A
"
A
dx (3.12)
Rearranging and separating variable, we obtained

35
m A dYA
 dx 
DAB 1  YA (3.13)

Assuming the product ρDAB to be constant, Eqn. 3.13 can be


integrated to yield

m A
 x   ln[1  YA ]  C
DAB , (3.14)

where C is the integration constant. Using the boundary condition

YA ( x  0)  YA,i

The mass-fraction distribution in exponential form can be obtained

m A x
YA ( x)  1  (1  YA,i ) exp( )
DAB (3.15)

The mass flux of species A, m A , can be determined by putting

YA ( x  L)  YA, in Eqn. 3.15

56
DAB 1  YA , 
m A  ln[ ]
L 1  YA , i (3.16)

Liquid-Vapor Interface Boundary Conditions

In Stefan problem, it usually required to estimate the gas-phase


mass fraction of diffusing species at the liquid-vapor interface, YA,i.
This can be realized through the assumption of existing equilibrium
between the liquid and vapor phases of species. Considering, also,
ideal gases assumption, the partial pressure of species A on the gas
side of the interface may be taken equal to the saturation pressure
corresponding to the liquid temperature, then

PA,i  Psat (Tliq,i ) (3.17)

The mass fraction can be expressed in term of partial pressure as

Psat (Tliq,i ) MW A
Y A ,i  (3.18)
P MWmix ,i
where the molecular weight of the mixture also depends on Psat.

The evaporation process of species A requires energy transfer


from the gas to the liquid surface, Qg. Some of this energy goes into

56
heating the liquid, Ql, while the remainder causes the phase change.
The energy balance can be expressed as

Q g  Ql  m
 (hvap  hliq )  m
 h fg (3.19)

56
Example 3.1
Consider water in a 25 mm diameter test tube into dry air at 1 atm
and 298 K. The distance from the water-air interface to the top of the
tube is L=15 cm. The mass fraction of the water vapor at the water
air interface is 0.0235 and the binary diffusivity for water vapor in air
is 2.6*10-5 m2/s.

a. Determine the mass evaporation rate of the water.


b. Determine the water vapor mass fraction at x=L/2.
c. Determine the fraction of the water mass flow that is contributed
by flow and the fraction contributed by diffusion at x=L/2.

Solution

Eqn. 3.16 can be applied to determine the mass flux since the
configuration represents the Stefan

DAB 1  YA , 
m A  ln[ ]
L 1  YA , i
First it is required to determine the mean density of the mixture

P
mix , av 
( Ru / MWmix , av )T
The average molecular weight may be given by

55
1
MWmix ,av  ( M mix ,i  M mix , )
2
Or

1
MWmix,av  [1 /  Y j ,i / MW j ,i  1 /  Y j , / MW j , ]
2 j j

1
MWmix,av  [1 /( 0.0235 / 18  0.9765 / 28.85)  1(1 / 28.85)]
2

=28.71 kg/kmol.

Then the average mixture density is

101325
mix, av 
(8135 / 28.71) * 298
=1.2 kg/m3

The mass flux of water vapor is

1.2 * 2.6 *10 5 1 0


m H 2O  ln[ ]
0.15 1  0.0235

55
= 0.495 *10-5 kg/m2s

The mass evaporation rate is given by

 H 2O * A
 H 2O  m
m
 0.495 *10-5 * ( * 0.252 / 4)

=2.43*10-7 kg/s
The water vapor mass fraction is given by Eqn. 3.15

 A x
m
YH 2O ( x)  1  (1  YH 2O ,i ) exp( )
DH 2O  Air
0.495 *10-5 * 0.075
YH 2O ( L / 2)  1  (1  0.0235) exp( 5
)
1.2 * 2.6 *10

YH 2O (15 / 2)  0.01181

The fraction of the water mass flow that is contributed by bulk flow
and the fraction contributed by diffusion at x=L/2.
The fraction of the water mass flow that is contributed by bulk flow,

 H 2O,bulk is given by
m

 H 2O,bulk ( L / 2)  YA ( L / 2)(m
m  H" 2O ) A
53
 H 2O,bulk (0.15 / 2)  0.01181* 0.495 *10-5 *  (0.25)2 /4
m

=2.868*10-9 kg/s

The fraction contributed by diffusion at x=L/2.

dYA
m H 2O,diff   DAB A
dx
The gradient of mass fraction at position x can be obtain
by differentiation of the following expression

 A x
m
YH 2O ( x)  1  (1  YH 2O ,i ) exp( )
DH 2O  Air
then,

dYH 2O ( x)  A
(1  YH 2O,i )m  A x
m
 exp( )
dx DH O  Air
2
DH 2O  Air

dYH2O ( L / 2) (1  0.0235)0.495 *10-5 0.495 *10-5 * 0.075


 5
exp( 5
)
dx 1.2 * 2.6 *10 1.2 * 2.6 *10

55
= -0.1568 m-1

dYA
m H 2O,diff   DAB A
dx
 1.2 * 2.6 *105 * 0.1568 * ( * 0.252 ) / 4
=2.40*10-7 kg/s
Example 3.2
For the above example, consider that the interface vapor mass
fraction is unknown. Find the mass evaporation rate of the water
when the liquid water is at 20 oC, assuming equilibrium at the
interface.

Solution
At t=20 oC P=2.339 kPa, hfg=2454.1 kJ/kg

At t=100 P=101.35 kPa hfg=2257 kJ/kg


Considering that there is equilibrium at interface, the partial pressure of
water vapor equals the pressure corresponding to the interface liquid
temperature, then

PA,i  Psat (T liq,i )


Using steam table we can obtain

PH 2O,i  2.339 kPa


The mass fraction of water vapor, YH2O, can be calculated as
MWH 2O PH 2O MWH 2O
YH 2O  x H 2O *  *
MWmix P MWmix

55
The interface mixture molecular weight can be determine as
MWmix,i  x H 2O,i MWH 2O,i  xair,i MWair,i
or
PH 2O,i P
MWmix,i  MWH 2O,i  (1  H 2O,i ) MWair,i
P P
2.339 2.339
MWmix ,i  *18  (1  ) * 28.85
101,35 10135
=28.6 kg/kmol.
Thus,

2.339 18
YH 2O  *
101.35 28.6
=0.0145
Then, the solution can be realized as described in the
above example.

If steam tables aren't available, instead, the properties of boiling


point are only known. In this case, it is possible to use Clausius-
Clapeyrion equation to estimate the saturation pressure at the liquid
temperature as follows

Clausius-Clapeyrion equation for ideal gases is expressed as

dP h fg dT

P Ru / MW T 2
By integration

55
Psat h fg 1 1
 exp[  (  )]
P Ru / MW T Tboil

At reference pressure, the boiling point of water has the following


values
P=1 atm Tboil=373 K hfg=2257 kJ/kg
Thus, at T=293 K, Psat can be calculated as

Psat 2257 1 1
 exp[  (  )]
1 8.135 / 18 293 373
Psat  2.62 kPa

55
Droplet Evaporation

Studying droplet evaporation demonstrates the application of


mass-transfer concepts to a problem of practical interest. The present
treatment is a part of the droplet evaporation and combustion which
introduces the concept of an evaporation constant and droplet life
time.

Evaporating droplet
rS
r Y

(a)
Y
1.0
YS

Y
0 r
rS
(b)
Figure 3.3 Evaporation of a liquid droplet in a quiescent environment

The problem is described in figure 3.3 where a droplet has


spherical symmetric coordinate. The droplet radius at the liquid-

56
vapor interface is denoted rs. Very far from the droplet

surface (r  ) , the mass fraction of the droplet vapor is YA,∞.

Physically, droplet liquid vaporizes due to energy transferred


from surroundings, then, vapor diffuses from droplet surface to the
ambient gas. The estimation of instantaneous evaporation rate from
the droplet surface is a main aim. The problem is somewhat complex
when considering all conservation laws of mass, species and energy.
However, a simplified solution can be achieved when making the
following assumptions
1. The evaporation process is quasi-steady state, i.e., the process
might be considered in steady state at any instant in time.
2. The temperature of droplet is uniform and has some fixed
value below the boiling point.
3. The mass fraction of vapor at the droplet surface is determined
by liquid vapor equilibrium at the droplet temperature.
4. All thermo-physical properties are constant, in particular, the

product D .

Evaporation Rate
The above assumptions reduce the case to Stefan's problem
while the change in coordinate system should be taken into
account. Overall mass conservation is expressed as

 ( r )  cos tan t  4r 2 m


m   (3.20)

56
It is worthwhile to note that the mass flow rate is constant with
r while mass flux isn't. Species conservation is expressed as

dYA
m  YA (m  m )  DAB
"
A
"
A
"
B
dr (3.21)

Since m B  0 , the above Eqn. can be reduced to


"

dYA
m  YAm  DAB
"
A
"
A
dr (3.22)
Substituting Eqn. 3.20 into Eqn. 3.22, and rearranging to solve for

 ( m A ) , yields
evaporation rate, m

D AB dYA
m  4r 2

1  YA dr (3.23)

m dr dYA

4DAB r 2
1  YA
Integrating the 3.23 yields

m
  ln[1  YA ]  C
4DAB r (3.24)

where C is the integration constant. Using the boundary condition


56
YA (r  rs )  YA, s
The mass-fraction distribution in exponential form can be obtained

(1  YA,S ) exp[ m /( 4D AB r )]


YA (r )  1 
exp[ m /( 4D AB rS )] (3.25)

The evaporation rate can be obtained by rearranging Eqn. (3.25) and

substituting YA  YA, for r 

 (1  YA, ) 
m  4rs DAB ln  
 (1  YA, S 
)  (3.26)

The argument of the logarithm in the above equation can be used to


define a useful dimensionless number, namely, Transfer number, BY:

1  Y A, 
1  BY 
1  Y A, S (3.27)

or

Y A, S  Y A , 
BY 
1 YA, S (3.28)

The evaporation rate is, then, expressed as

55
  4rs DAB ln(1  BY )
m (3.29)

BY may be interpreted as a driving potential for mass


transfer.

Droplet Mass Conservation


A mass balance on an evaporating droplet states that the rate
at which the mass of the droplet decreases is equal to the rate at
which the liquid is vaporized. Thus,

dmd
 m
dt (3.30)
The droplet mass can be expressed in terms of droplet diameter,
D=2rS, and density, ρl,

md   l D / 6 3
(3.31)
Substituting Eqn. 3.29 and Eqn. 3.31 and performing the
differentiation yields

dD 4 D AB
 ln(1  BY )
dt l D (3.32)

Eqn. 3.32 is more commonly expressed in terms of D2 rather than D.


So, Eqn. 3.32 may be rearranged to have the form

55
dD 2 8DAB
 ln(1  BY )
dt l (3.33)

Equation 3.33 indicates that D2 varies linearly with t with the

slope  (8DAB /  l ) ln(1  BY ) , as shown in Fig 3.4. This slope is


defined to be the EVAPORATION CONSTANT, K.

K  (8DAB / l ) ln( 1  BY ) (3.34)

Thus, Eqn. 3.33 becomes

dD 2
 K
dt (3.35)
The droplet lifetime, td, can be determined by integrating Eqn. 3.35
which yields

td  D / K 2
o (3.36)
A more general expression for the variation of D with time (t) is

D (t )  D  Kt
2 2
o (3.37)

53
2
Equation 3.37 is known as the D LAW for droplet evaporation
which is also used to describe burning of fuel droplet.

D2
 8DAB
Slope   K  ln( 1  BY )
Do2 l

t
0
tS

Figure 3.4 The D2 low for droplet evaporation.

55
Example 3.3
Determine the influence of the ambient water vapor mole fraction on
the lifetimes of 50-μm diameter water droplets. The droplets are
evaporating in air at 1 atm. Assume the droplet temperature is 75 oC
and the mean air temperature is 200 oC. Use values of  H O,  0.1,
2

0.2, and 0.3.

Solution
Equation 3.36 can be used to estimate the droplet lifetime.

td  D / K 2
o (3.36)

This requires calculating the evaporation rate constant, K, by Eqn.


3.34.

K  (8DAB / l ) ln( 1  BY ) (3.34)

First we will evaluate required properties

Tboil=100+273=373 K
P=101.35 kPa
hfg=2257 kJ/kg
MWH2O=18 kg/kmol.
DAB=2.2*10-5 m2/s @T=273 K

55
An integrated form of Clausius-Clapeyrion equation can be used to
estimate the saturation pressure at the liquid temperature

Psat h fg 1 1
 exp[  (  )]
P Ru / MW T Tboil
Psat 2257 1 1
 exp[  (  )]
1 8.135 / 18 (348 373
Psat=0.382 atm=38.72 kPa
Using steam table we may also find Psat=38.58 kPa at Tsat =75 oC.

The interface mixture molecular weight can be determined as

MWmix,S   H 2O,S MWH 2O   air ,S MWair


or
PH 2O,i PH 2O,i
MWmix,S  MWH O  (1  ) MWair
P 2
P

0.382 0.382
 *18  (1  ) * 28.85
1 1
=24.71 kg/kmol.

The mass fraction, YA,S, at the surface is given by

55
MWH 2O PH 2O MWH 2O
YH 2O   H 2O *  *
MWmix P MWmix

0.382 18
YH 2O  *
1 24.71
=0.278

The mass fraction, YA,∞, in the far surrounding can be also


calculated by the same expression. The solution will proceed only for
χH2O,∞=0.1, while all of results will be summarized in a table.

For χH2O,∞=0.1

18
YH 2O,  0.1*
(0.1*18  0.9 * 28.85)
=0.0648

It is the time to evaluate the transfer number, BY

Y A, S  Y A , 
BY 
1 YA, S

55
0.278  0.0648

1  0.278
=0.295

To evaluate the evaporation constant K, we need first to estimate the


product ρDAB. Since the value of DAB is tabulated at T=273 K, it is
required to be extrapolated to 473 K, using Eqn. 3.5.

DAB  T 3/ 2
/P (3.5)
Or

T2 3 / 2
DAB (T2 )  DAB (T1 )( )
T1
473 3 / 2 5
DAB (473)  2.2 *10 ( )
273
=5.017*10-5 m2/s
1
MWmix,av  [ MWmix, S  MWmix,  ]
2
1
MWmix,av  [24.71  0.1 *18  .9 * 28.85]
2
=26.24 kg/kmol.

56
The average density can be calculated by the ideal gas law

P
 av 
Ru / MWmix ,av )T
101325

(8135 / 26.24) * 473
=0.691 kg/m3

Thus,

K  (8DAB / l ) ln( 1  BY )
 (8 * 0.691* 5.017 *105 / 975) ln( 1  0.295)
=7.35*10-8 m2/s

Then, the droplet lifetime is

td  D / K 2
o

 (50 *106 )2 / 7.35 *108


=0.034 s

56
χH2O,∞ Y H2O,∞ BY K(m2/s) td (s)
0.1 0.0648 0.295 7.35*10-8 0.034
0.2 0.1349 0.1982 5.036*10-8 0.0496
0.3 0.211 0.0928 2.42*10-8 0.1033

56
Chapter 4

CHEMICAL KINETICS

38
Chapter 4

CHEMICAL KINETICS

Chemical kinetics is a specialized field of physical chemistry


that studies elementary reactions and their rates. Rates of chemical
reaction coherently control the rate of combustion processes and
determine pollutants formation and reduction. Moreover, ignition
and flame extinction are intimately related to chemical processes.
Thus, in this chapter the basic chemical kinetics concepts will be
explored since their understanding is essential to combustion study.

GLOBAL VERSUS ELEMENTARY REACTION


The overall reaction of a mole of fuel with (a) moles of an
oxidizer to form (b) moles of combustion products can be expressed
by the global combustion mechanism

F  aOx  b Pr (4.1)

Experimentally, the fuel consumption rate may be expressed as

d[ X F ]
 kG (T )[ X F ] [ X Ox ]
n m
dt (4.2)

Where

38
[Xi] is the molar concentration of species i, in kmol./m3 or gmol/cm3
kG is the global rate coefficient which is a strong function of
temperature

Equation 4.2 indicates that the reaction is nth order with respect to
the fuel, mth order with respect to the oxidizer and (n+m)th order
overall. For global reactions, n and m arise from curve-fitting
experimental data.
The global reaction is a simplified approach for solving some
problems. However, the approach tells nothing about what is actually
happening in the system from chemical point of view. In fact, for the
above reaction to be completed, many sequential processes can occur
involving many intermediate species.
For instance, to effect the global conversion of hydrogen and
oxygen to water

2H 2  O2  2H 2O (4.3)

The following elementary reactions, among others, are important:

H 2  O2  HO2  H (4.4)

H  O2  OH  O (4.5)

OH  H 2  H 2O  H (4.6)

H  O2  M  HO2  M (4.7)

38
During the uncompleted mechanism described above, many
radicals or free radicals (intermediate species) occur such as HO 2
(hydroperoxy), HO (hydroxyl), H and O. Free radicals are reactive
atoms or molecules that have unpaired electrons. The group of
elementary reactions necessary to express the details of an overall
reaction is called a reaction mechanism.

ELEMENTARY REACTION RATES

Bimolecular Reactions

In bimolecular reactions, two molecules collide and react to


form two different molecules. It may be expressed as

A B C  D (4.8)

The reaction rate is directly proportional to the concentrations


(kmol./m3) of the two reactant species

d [ A]
 kbimolec [ A][ B] (4.9)
dt
All elementary bimolecular reactions are overall second order, being
first with respect to each of the reacting species. The SI units for
kbimol. are m3/kmol-s.
The bimolecular rate coefficient can be expressed by empirical
Arrhenius form,

38
k( T )  A exp(  EA / RuT ) (4.12)

where A is termed pre-exponential factor or the frequency factor.


EA is the activation energy (J/kmol)

Other Elementary Reactions

Unimolecular reactions are that involve a single species


undergoing rearrangement to form one or two product species, i.e.

AB (Isomerization) (4.13)

A BC (Decomposition) (4.14)

A typical example is the dissociation of of O2, O2  O  O

Unimolecular reactions are first order at high pressure

d [ A]
 kuni [ A] (4.15)
dt

38
At low pressures, the reaction becomes second order where it also
depends on the concentration of some molecules, M, with which the
reacting species may collide.

d [ A]
 k[ A][ M ] (4.16)
dt

Termolecular reactions involve three reactant species and have


the following general form.

A B  M C  M (4.17)

A typical example of termolecular involved in combustion is


recombination reactions, such as, H  OH  M  H 2O  M
Termolecular reactions are third order and their rates can be
expressed as

d [ A]
 kter [ A][ B][ M ] (4.18)
dt
Where M may be any molecule and is normally referred as a third
body.

33
RATES OF REACTION FOR MULTISTEP
MECHANISMS

Net Production Rates

The reaction mechanism is defined as a sequence of


elementary reactions that leads from reactants to products.
The net rates of production or destruction for any species
participating in the mechanism can be determined if the rates of the
different elementary reactions are known. Let us consider the H2-O2
reaction mechanism that includes forward and reverse reactions as
shown by the  symbol.

k f1

H 2  O2  HO2  H (4.19)
k r1

kf2

H  O2  OH  O (4.20)
kr 2

kf3

OH  H 2  H 2O  H (4.21)
kr 3

kf4

H  O2  M  HO2  M (4.22)
kr 4

38
Where kfi and kri are the elementary foreword and reverse rate
coefficients, respectively, for the ith reaction. The net rate of
production of O2, for instance, is the sum of all the individual
elementary rates producing O2 minus the sum of all the rates
destroying O2, i.e.,

d [O2 ]
 k r1[ HO2 ][ H ]  k r 2 [OH ][O ]
dt
 kr 4[ HO2 ][ M ]  .....
 k f 1[ H 2 ][O2 ]  k f 2[ H ][O2 ]
 k f 4[ H ][O2 ][ M ]  ... (4.23)

Similar expressions can be written for each species participating in


the mechanism. This yields a system of first order differential
equations that describes the evolution of the chemical system starting
from given initial conditions.

Compact Notation

A compact notation has been developed to represent a reaction


mechanism and the individual species production rates because many
elementary steps and many species may be involved in. For the
mechanism, one may write

89
N N
 ji X j   ji X j
j 1 j 1 (4.24)

For i=1,,2,…..L

Where  ji and  ji are the stoichiometric coefficients on the

reactants and products sides of the chemical equations, respectively,


for jth species in the ith reaction. When considering the four reactions
4.19-4.22 that involve the eight species O2, H2, H2O, HO2, O, H, OH,
and M. Defining j and i as follow:

j Species i Reaction
1 O2 1 Eqn. (4.19)
2 H2 2 Eqn. (4.20)
3 H2O 3 Eqn. (4.21)
4 HO2 4 Eqn. (4.22)
5 O
6 H
7 OH
8 M

The stoichiometric coefficient matrices can be written considering i


as the raw index and j as the column index.

89
1 1 0 0 0 0 0 0
1 0 0 0 0 1 0 0
 ji   
0 1 0 0 0 0 1 0 (4.25a)
 
1 0 0 0 0 1 0 1

And

0 0 0 1 0 1 0 0
0 0 0 0 1 0 1 0
 ji   
0 0 1 0 0 1 0 0 (4.25b)
 
0 0 0 1 0 0 0 1

The net production rate for each species,  , can be compactly


expressed using the following relations:

L
 j   ji qi for j=1,2,…,N (4.26)
i 1
where

 ji  ( ji  ji ) (4.27)

and

89
N N
  
qi  k fi  [ X j ]  kri  [ X j ]
ji ji

(4.28)
j 1 j 1

Relation between Coefficients and Equilibrium Constant

Thermodynamic measurements or calculations are considered


more accurate and precise either for determination of rate
coefficients of elementary reactions or in solving chemical kinetics
problem. For a bimolecular reaction

kf

A B  C  D (4.29)
kr

The net production rate of species A is expressed as

d [ A]
 k f [ A][ B]  kr [C ][ D] (4.30)
dt
At equilibrium conditions the net time rates of species must be zero.
Thus, Eqn. 4.30 is reduced to

0  k f [ A][ B]  kr [C ][ D] (4.31)

By rearranging

88
[C ][ D] k f (t )
 (4.32)
[ A][ B] kr (t )

The equilibrium constant was defined in terms of partial pressure as

( PC / P o )c ( PD / P o ) d ....
Kp 
( PA / P o ) a ( PB / P o )b .... (4.33)

where the superscripts are the stoichiometric coefficients, i.e.,


 i  a,b,... and  i  c, d ,... Since the molar concentrations relate to
the mole fractions and partial pressures as follows,

[ X i ]   i P / RuT  Pi / RuT (4.34)

An equilibrium constant, Kc, can be defined based on molar


concentrations as follows,

 
[ X i ] i

[C ]c [ D]d .... prod


Kc   
[ A]a [ B]b .... [ X i ] i (4.35)

reac

It is obvious from Eqs. (4.32), (4.33) and (4.35) that

88
 
[ X i ] i

prod k f (T )
Kc  

[ X i ] i k r (T )
reac

 K p ( RuT / Po )    (4.36)

For a bimolecular reaction K c  K p since       0 . Equation

(4.36) can be used to determine a reverse reaction rate from the


knowledge of the forward rate and the equilibrium constant for the
reaction; or, conversely.

Example 4.1

In methane combustion, the following reaction pair is important:

CH 4  M  CH3  H  M
where the reverse reaction coefficient is given by

k r (m6 / kmol2  s)  2.82 *105 T exp[ 9835 / T ].


At 1500 K, the equilibrium constant Kp has a value of 0.003691 based
on a reference-state pressure of 1 atm (101325 Pa). Derive an
algebraic expression for the forward rate coefficient kf. Evaluate your
expression for a temperature of 1500 K. Give units.

88
Solution
An algebraic expression for the forward rate coefficient, kf, can be
obtained using the following relation,

k f (T )
 K p ( RuT / P o )   
k r (T )
Substitute with

K p exp( GTo / RuT )

k r (m6 / kmol 2  s)  2.82 *105T exp[ 9835 / T ].


      1
and rearrange, the forward rate coefficient, kf, can be expressed as

k f (T )  2.82 *105 T exp[ 9835 / T ] * K p ( RuT / P o ) 1


or

k f (T )  2.82 *105T exp[ 9835 / T ] *

exp( GTo / RuT )( RuT / P o ) 1

At 1500 K, the forward rate coefficient, kf, can be estimated as

k f (1500)  2.82 *105 *1500 * exp[ 9835 / 1500] *


0.003691* (8.135 *1500 / 101.325) 1
=18.44 m3/mol-s
88
Chain and Chain-Branching Reactions

Chain reactions are the most common type of chemical


reactions. They involve production of a radical species that
subsequently reacts to generate another radical. This radical, in turn,
reacts to produce yet another radical. This sequence of events keeps
on until some reaction breaks the chain in which a stable species is
created from two radicals. For a chain reaction that is globally
presented by

A2  B2  2 AB (4.37)

It involves the chain initiation reaction

k1
A2  M  A  A  M (4.38)
The chain propagating reactions involving free radicals A and B

k2
A  B2  AB  B (4.39)

k3
B  A2  AB  A (4.40)

The chain terminating reaction

k4
A  B  M  AB  M (4.41)

88
Chain-branching reactions involve formation of two

radical species due to a reaction that consumes only one radical. A


typical example of the chain-branching reaction is

O  H 2O  OH  OH (4.42)
Existence of a chain-branching step in a certain mechanism may
cause an explosive effect. In a system with chain branching, the
concentration of a radical species may be built up geometrically,
causing the rapid formation of the products. Chain-branching
reactions are responsible for a flame being self-propagating.

Steady-State Approximation
Many combustion mechanisms involve a highly reactive
intermediate radical for which the rate of formation is slow while the
reaction consuming is very fast. Under these conditions, the radical
concentration builds up for a short period, after that, the radical is
consumed as rapidly as it is created. Therefore, concentration of the
radical is quite small in comparison with those of reactants and
products. Analyses of such systems may be simplified by applying
the Steady-State Approximation.
A typical example of this system is zeldovich mechanism for
formation of nitric oxide, in which N atom is the reactive
intermediate of interest.

k1
O  N 2  NO  N (4.43)

83
k2
N  O2  NO  O (4.44)
Reaction 4.43 is rate limiting since it is slow whereas reaction 4.44 is
extremely fast. The net rate production of N radicals may be written as

d[ N ]
 k1[O][ N 2 ]  k 2 [ N ][O2 ] (4.45)
dt

After a very short transient period, d [ N ] / dt  0 since the two terms


on the right hand sides become equal. The steady concentration of N is,
then, given by

k1[O][ N 2 ]
[ N ]ss 
k 2 [O2 ] (4.46)

[N]ss may change as it rapidly readjusts according to Eqn. 4.46. The


time rate of change of [N]ss may be obtained by differentiating Eqn.
4.46, then

d [ N ]ss d  k1[O][ N 2 ] 
 
dt dt  k 2 [O2 ]  (4.47)

88
Example 4.2
As mentioned above, the Zeldovich mechanism is a famous
chain mechanism. It describes formation of nitric oxide from
atmospheric nitrogen at high temperature, i.e., thermal mechanism.

k1 f
O  N 2  NO  N (4.43)

k2 f
N  O2  NO  O (4.44)
For this mechanism, it was found that the second reaction is
much faster than the first. Therefore, the N-atom concentration may
be evaluated by applying the steady state approximation technique.
Moreover, O2 and O may be assumed in equilibrium since NO
formation is much slower than other reactions involving O2 and O.

O2  2O
The global mechanism is

kG
N 2  O2  2 NO
and NO formation rate is expressed as

d [ NO]
 kG [ N 2 ]m [O2 ]n
dt

999
Determine kG, m and n using the elementary rate coefficients, etc,
from the detailed mechanism.

Solution
From the elementary reactions, one may express

d [ NO]
 k1 f [ N 2 ][O]  k 2 f [ N ][O2 ]
dt
d[ N ]
 k1 f [ N 2 ][O]  k2 f [ N ][O2 ]
dt
where reverse reaction rates are assumed negligible

Applying steady state approximation requires that d [ N ] / dt  0,


thus

k1 f [O][ N 2 ]
[ N ]ss 
k 2 f [O2 ]
Substituting [N]ss into the above expression for d [ NO] / dt , yields

999
d [ NO]
 2k1 f [ N 2 ][O]
dt

[O] can be eliminated through the equilibrium approximation


 
 i
[ X ] i

prod o    
Kp  
( Ru T / P )
[ X i ] i

react

[O]2 RuT 2 1
 ( o )
[O2 ] P
or

 o 1 / 2
K pP
[O]  [O2 ] 
 RuT 

Therefore

1/ 2
d [ NO ]  K pP o

 2k1 f   [ N 2 ][ O2 ] 1 / 2
dt  RT 
 u 

999
Comparing the above expression of NO formation rate with that of
global reaction, one can infer

 K pP o 1 / 2
kG  2k1 f  
 Ru T 
 
m 1
n  1/ 2

998
SOME IMPORTANT CHEMICAL
MECHANISMS

In this part, some chemical mechanisms of major importance


to combustion and combustion-generated air pollution are presented
and discussed.

H2-O2 System
The hydrogen –oxygen system is important in its own right as,
for example, in rocket propulsion. It is also important as a subsystem
in the oxidation of hydrocarbons. The oxidation of hydrogen may be
outlined as follows.

The initiation reactions are

H2  M  H  H  M (very high temperature) (H.1)

H 2  O2  HO 2  H (other temperature) (H.2)

Chain-reaction steps involving O, H and OH radicals are

H  O2  O  OH (H.3)

O  H 2  H  OH (H.4)

998
H 2  OH  H 2O  H (H.5)

O  H 2O  OH  OH (H.6)

Chain terminating steps involving O, H and OH radicals are the


three body recombination reactions:

H  H  M  H2  M (H.7)

O  O  M  O2  M (H.8)

H  O  M  OH  M (H.9)

H  OH  M  H 2O  M (H.10)

The mechanism may be completed by including reactions involving


HO2, the hydroperoxy radicals and H2O2, hydrogen peroxide. When

H  O2  M  HO2  M (H.11)
becomes actives, then the following reactions and the reverse of H.2,
come into play:

HO2  H  OH  OH (H.12)

HO2  H  H 2O  O (H.13)

HO2  O  O2  OH (H.14)

998
and

HO2  HO2  H 2O  O (H.15)

HO2  H 2  H 2O2  H (H.16)


with

H 2O2  OH  H 2O  HO2 (H.17)

H 2O2  H  H 2O  OH (H.18)

H 2O2  H  HO2  H 2 (H.19)

H 2O2  M  OH  OH  M (H.20)

In modeling H2-O2 system about 40 reactions may be taken into


consideration involving eight species: H2, O2, H2O, OH, O, H, HO2
and H2O2, since the reverse reactions of all the above can be
important depending upon system conditions.
The mechanism described above can be used to explain
explosion characteristic of H2-O2 system burning in a spherical
vessel, shown in Fig. 4.1. The figure illustrates that there are distinct
explosion and non-explosion zones depending upon the initial
temperature and pressure of the charged mixture. The area below
the first limit line has non-explosion feature. This behavior may be
ascribed to destruction of radicals, produced in reaction H.2-H.6, on
the vessel walls in a heterogeneous reaction. A heterogeneous

998
reaction is the reaction of gaseous species at solid surfaces and is very
important in solid combustion and catalysis.
In a second region, where pressure is set between the first and
second limit, and temperature is greater than 400 oC, the mixture has
explosion feature. Under these conditions, the chain sequence H.3-
H.6 prevails over the radical destruction at the walls. Since
increasing pressure increases radical concentration linearly and
increases the reaction rate geometrically.
The mixture becomes again non-explosive while the pressure
goes above the second limit. This lack of explosion is a result of
competition for H atoms between the chain branching reaction H.3
and reaction H.11 which is an effective terminating step at low
temperature. Reaction H.11 is a chain terminating step because the
hydroperoxy radical, HO2, is relatively un-reactive at these
conditions, and consequently, it can diffuse to the wall to be
destroyed.

998
Figure 4.1. Explosion limits for a stoichiometric hydrogen-oxygen
mixture in a spherical vessel.

Surpassing the third limit, another explosive regime is attained.


At these conditions, reaction H.16 adds a chain-branching step which
launches the H2O2 chain sequence.

993
Carbon Monoxide Oxidation
Oxidation of carbon monoxide is extremely important either in
its own right or to the oxidation of hydrocarbon. Hydrocarbon
combustion simplistically can be characterized as a two-step process:
the first step involves the breakdown of the fuel to carbon monoxide,
with the second being the final oxidation of carbon monoxide to
carbon dioxide.
It is well known that CO oxidation is slow; however, presence
of small quantities of hydrogen containing species, H 2O or H2, can
have a tremendous effect on the oxidation rate. This is because the
CO oxidation step involving the hydroxyl radical is much faster than
the steps involving O2 and O.
Let water be the primary hydrogen-containing species, the
oxidation of carbon monoxide may be described by four steps.

CO  O2  CO2  O (CO.1)

O  H 2O  OH  OH (CO.2)

CO  OH  CO2  H (CO.3)

H  O2  OH  O (CO.4)

The reaction CO.1 doesn't contribute considerably to formation of


CO2, but rather serves as the initiator for the chain sequence. The
main step for CO oxidation is CO.3 reaction. It is also a chain
propagating step, generating H atom which, in turn, react with O2 to

998
form OH and O through reaction CO.4. These radicals feed back into
the oxidation step (CO.3) and the first chain branching step (CO.2).
The key reaction of the mechanism is CO  OH  CO2  H (CO.3)

When H2 is used as the primary hydrogen-containing species,


the following steps are involved.

O  H 2  OH  H (CO.5)

OH  H 2  H 2O  H (CO.6)

With hydrogen presence, the entire H2-O2 relation system (H.1-H.20)


should to be also considered to describe CO oxidation.

Nitric Oxide Formation

Nitric oxide is an important minor species in combustion due to


its contribution in air pollution. During combustion, nitric oxide may
be formed from air nitrogen through three chemical mechanisms:
thermal or Zeldovich mechanism, Fenimore or prompt mechanism
and N2O-intermediate mechanism. The thermal mechanism
dominates in high temperature combustion over a wide range of
equivalence ratio. The Fenimore mechanism is significant in rich
combustion while the N2O-intermediate mechanism plays an
important role in very lean, low temperature combustion processes.

999
The thermal or Zeldovich mechanism consists of two chain
reactions:

O  N 2  NO  N (N.1)

N  O2  NO  O (N.2)

Which can be extended by adding the reaction

N  OH  NO  H (N.3)

The N2O-intermediate mechanism is important at fuel-lean


(   0.8 ), low temperature conditions. The three steps of this
mechanism are

O  N 2  M  N 2O  M (N.4)

H  N 2O  NO  NH (N.5)

O  N 2O  NO  NO (N.6)
This mechanism becomes important in NO control strategies that
involve lean premixed combustion in gas turbines.

In Fenimore mechanism, some hydrocarbon radicals react with


molecular nitrogen to form amines or cyano compounds. The amines
and cyano compound are then converted to intermediate compounds

999
that finally form NO. For equivalence ratios less than about 1.2, the
Fenimore mechanism can be described as:

CH  N 2  HCN  N (N.7)

C  N2  CN  N (N.8)

HCN  O  NCO  H (N.9)

NCO  H  NH  CO (N.10)

NH  H  N  H 2 (N.11)

N  OH  NO  H (N.12)

For equivalence ratios richer than 1.2 the mechanism becomes much
more complex.
Hydrogen cyanide HCN

999
Chapter 5

LAMINAR PREMIXED FLAME

5.1. PHYSICAL DESCRIPTION

Laminar premixed flames have application in many


residential, commercial, and industrial devices and processes.
Laminar flame is normally found in conjunction with diffusion
flame. Typical examples include gas ranges and Bunsen burner.
Study of premixed laminar flames is important by itself; however, it
is more important for understanding turbulent flame.
A flame may be defined as a self-sustaining propagation
of a localized combustion zone at subsonic velocities.

113
The definition implies that the flame occupies a thin zone and
moves sub-sonically (deflagration wave). The flame is approximately
1 mm thick and moves at 0.5 m/s for stoichiometric hydrocarbon
mixtures at ambient air. Pressure drop through the flame is very
small, about 1 Pa, and temperature in the reaction zone is high (2200-
2600 K). It is also possible for combustion wave to propagate at
supersonic velocities that is called detonation.

Consider a one dimensional flow where the unburned mixture


enters the flame in the direction normal to the flame sheet. The
product leaves the flame zone with much higher temperature and
lower density, see Fig. 5.1. Continuity requires that

u S L A  u vu A  bvb A (5.1)

114
where the subscripts u and b refer to the unburned and burned
gases, respectively. SL is the burning velocity (flame velocity) defined
as the velocity of the flame relative to the unburned reactant mixture.

Figure 5.1 Laminar flame structure. Temperature and heat-release


rate profiles based on experiments by Friedman and Bruke.

As shown Fig. 5.1 a flame may be divided into two zones. A


preheat zone where little energy is released and a reaction zone
where the bulk of chemical energy is released. The earliest thin part
of the reaction zone is characterized by very fast chemistry including
destruction of fuel molecules and creation of many intermediate
species. Therefore, gradients of temperature and species
concentration are very great in this region. These gradients cause
transfer of heat and radical species to the preheat zone. Actually,
these steep gradients are the driving forces that make the flame to be
self-sustaining.
Hydrocarbon flames are characterized by visible radiation.
The fast reaction zone appears blue with an excess of air due to
115
radiation from excited CH radicals in the high temperature zone.
The zone appears blue-green when air is reduced to less than
stoichiometric due to radiation from exited C2. For still richer
mixture, the flame appears bright yellow to dull orange depending on
the flame temperature due to soot radiation.

Figure 5.2 (a) Bunsen burner schematic (b) Laminar flame speed
equals normal component of unburned gases, vu,n.

The classical device to generate a laminar premixed flame is


the Bunsen burner shown in Fig. 5.2. Gaseous fuel from the fuel
supply enters through an orifice into the mixing chamber, into which
air is entrained through adjustable openings from the outside. The
cross sectional area of the fuel orifice may be adjusted by moving the
needle through an adjustment screw into the orifice. Thereby the
velocity of the jet entering into the mixing chamber may be varied
and the entrainment of the air and the mixing can be optimized. The
mixing chamber must be long enough to generate a premixed gas
issuing from the Bunsen tube into the surroundings. The typical

116
Bunsen-burner flame is a dual flame: a fuel rich premixed inner
flame surrounded by a diffusion flame. The shape of the flame is
determined by the combined effects of the velocity profile and heat
losses to the tube wall. For the flame to remain stationary, the flame
speed must equal the speed of the normal component of unburned
gases at each location. Thus,

S L  vu sin  (5.2)

5.2. SIMPLIFIED ANALYSIS

A simplified approach given by Spalding is pursued. It is an


analysis that couples principles of heat transfer, mass transfer,
chemical kinetics, and thermodynamics to understand the factors
governing flame speed and thickness. The main objective is to have a
simple expression for the laminar flame speed.

Assumptions
1. One dimensional, constant area, steady flow.
2. Kinetic and potential energies, viscous shear work, and thermal
radiation are all neglected.
3. The small pressure drop across the flame is neglected, i.e., the
pressure is constant.
4. The Lewis number, Le, is unity. Where Lewis number is defined
as the ratio of thermal diffusivity to mass diffusivity, i.e.,

117
k 
Le  
D c p D (5.3)

This leads to k / c p  D that greatly simplifies the energy

equation.
5. Specific heats are assumed to be constant and equal for all
species.
6. The oxidation process is a single-step exothermic reaction.
7. The fuel is completely oxidized in the flame zone.

Conservation Laws
Conservations of mass, species and energy are applied to the
differential control volume illustrated in Fig. 5.3.

Figure 5.3 Control volume for flame analysis.


constant
Mass conservation

118
d (  x )
0 (5.4a)
dx
Or

    x  constant
m (5.4b)

Species Conservation

dm i
 m i (5.5)
dx
Applying Fick's law yields

dYi
d [m Yi  D ]
dx  m i (5.6)
dx
  is the mass production rate of species i per unit volume
where m
(kg/m3s).

The global stoichiometry for the simplified reaction is

1 kg fuel   kg oxidizer  (1  )kg products. (5.7)

Accordingly,

119
1 1
 F  m
m  Ox
    Pr
m 
  1 (5.8)

Applying Eqn. (5.6) for each species yields

Fuel

dYF
d [ D ]
dYF dx  m
m    F (5.9a)
dx dx
Oxidizer

dYOx
d [ D ]
dYOx dx  m
m    F (5.9b)
dx dx
Products

dYPr
d [ D ]
dYPr dx  - (1   )m
m    F (5.9c)
dx dx

Energy Conservation
Using the first law of thermodynamics with considering the
assumptions given above, the following expression may be obtained

121
dT d dT
m c p  i
 [( Dc p ) ]   h of ,i m
dx dx dx
(5.10a)
where:

dT
m c p is the rate of sensible enthalpy transported
dx
by convection per unit volume

d dT
[( Dc p ) ] is the rate of sensible enthalpy transported
dx dx
by diffusion per unit volume

  h f ,i m i
o
is the rate of enthalpy production by

chemical reaction per unit volume


The right hand side of Eqn. (5.10) may be expanded

  hof ,i m  F  hof ,Oxm


 i [hof , F m  F  hof ,Pr (1  )m
 F]

 F hc
 m
where hc is the heat of combustion of fuel. According to the

assumption 4, Dc p can be replaced with k. Eqn. (5.10) becomes

121
dT
d (k )
dT 1 dx m F hc
m   
dx c p dx cp (5.10b)

The laminar flame velocity is simply related to the mass flux by

   u S L
m (5.11)

Solution

To integrate Eqn. (5.10b), it is required to assume a temperature


profile that satisfies the following boundary conditions. The
boundary conditions far upstream of the flame are

T ( x  )  Tu (5.12a)

dT
( x  )  0
dx (5.12b)

and far downstream of the flame,

T ( x  )  Tb (5.12c)

dT
( x  )  0 (5.12d)
dx
122
For simplicity, a linear temperature profile over the flame thickness,
δ, is assumed as shown Fig. 5.4. Integrating Eqn. (5.10b), yields

Figure 5.4 Assumed temperature profile for laminar premixed flame.

dT / dx 0
k  dT   hc 
m [T ]TT TTb   dx    m F dx
u
cp dT / dx 0 c p 
(5.13)
By evaluating the limits, it becomes

123

 hc
m (Tb  Tu ) 
cp  m dx

F
(5.14)

The limit on the right hand side can be changed from space to

temperature, as m  has values only within the zone δ, i.e.,

dT Tb  Tu

dx  (5.15a)

or


dx  dT
Tb  Tu (5.15b)

After changing the variables,

 hc  T b

m (Tb  Tu )   m F dT
c p (Tb  Tu ) T u

(5.16)
Using the average reaction rate definition

T
1 u

m F   m F dT
(Tb  Tu ) T b
(5.17)

124
Equation (5.16) is simplified to

 hc
m (Tb  Tu )  m F
cp (5.18)

  and
The algebraic Equation (5.18) involves two unknown m
δ, so it is required another equation to have a complete solution. This
may be realized through a reasonable assumption that considers the
fuel consumption occurs only in the high-temperature region. That is,

m F is zero in the interval    x   / 2 . Now integration can be


done from x   to x   / 2 . At x   / 2 , we may note that

Tb  Tu
T
2 (5.19)

and

Tb  Tu
dT  dx
 (5.15b)

Applying the current limits on the Eqn. (5.13), we obtain

dT / dx(Tb Tu ) /  
k  dT   hc
m [T ]T Tu  m dx
T (Tb Tu ) / 2
  dx   F
cp dT / dx0 cp 

125
  / 2  k / c p  0
m (5.20)

Solving Eqns. (5.18) and (5.20) simultaneously yields

1/ 2
 k (h ) 
m   2 2 c
m F 
 c p (Tb  Tu )  (5.21)

and

  2k /(c p m ) (5.22)

Applying the definitions of flame speed,   / u ,


SL  m thermal

diffusivity,   k / u c p , and recognizing that

hc  (  1)c p (Tb  Tu ) , the final results are attained

1/ 2
 m F 
S L   2 (1   ) 
 u  (5.23)

126
1/ 2
  2 u 
  
 (1   )m F  (5.24a)

Or in terms of SL,

  2 / S L (5.24b)

Example 5.1
Estimate the laminar flame speed of stoichiometric propane-air
mixture using the simplified theory expression. Use the given global
one-step reaction to estimate the mean reaction rate. The adiabatic
flame temperature of propane may be taken 2260 K.

d [C3 H 8 ]  15098
 8.6 *1011 exp( )[C3 H 8 ]0.1[O2 ]1.65
dt T

Solution
The laminar flame speed is expressed as

1/ 2
  F 
m
S L   2 (1   ) 
  u 
 F , and the
Thus, it is essential to evaluate the rate of fuel reaction, m

thermal diffusivity, . An average temperature over the second half

127
of the flame thickness, where chemical reaction is assumed to occur,
is given by

1 1
T  [ (Tb  Tu )  Tb ]
2 2
1 1
 [ (2260  300)  2260]
2 2
=1170 K

where it is assumed that Tb  Tad and Tu  300K .


The mean concentrations that may be used in the reaction rate
equation are

1
YC H  (YC H ,u  YC H ,b )
3 8
2 3 8 3 8

1
 (0.06015  0)
2
=0.0301
and

1
YO  (YO ,u  YO ,b )
2
2 2 2

1
 (0.233(1  0.06015)  0)
2
=0.1095

128
Where A/F of a stoichiometric propane-air mixture is

1
  1  15.625 and mass fraction of O2 in air is 0.233.
YC H
3 8
,u

The average density in the reaction zone is calculated as

P 101325
 
( Ru / MW )T (8315 / 29) *1770
=0.1997 kg/m3

The molar reaction rate is given by

d [C3H8 ]  15098
 F   8.6 *1011 exp( )[C3H8 ]0.1[O2 ]1.65
dt T
0.1 1.65
 15098 1.75  YC H   YO 
 F  8.6 *10 exp(
11
)     3 8 2

T  MWCH   MW O 
 3 8 2

 15098
 8.6 *10 exp( 11
)(0.1997 )1.75 *
1770
0.1 1.65
 0.301  0.1095 
 44   32 
=-2.439 kmol/ m3s

The reaction rate on mass basis is

129
 F   F * MWF
m
 -2.439* 44
=107.3 kg/ m3s

To determine the thermal conductivity, the appropriate mean


temperature should be obtained over the entire flame thickness.

1 1
Tt  (Tb  Tu )  (2260  300)
2 2
=1280 K
P 101325
u  
( Ru / MW )Tu (8315 / 29) * 300
=1.18 kg/m3
Then

k (Tt ) 0.0809
 
u c p (Tt ) 1.18 *1186
=5.78*10-5 m2/s
Then, the laminar flame speed is

131
1/ 2
 (2 * 5.78 *10 5 )(1  15.625)( 107.3) 
SL   
 1 .18 
=0.418 m/s=41.8 cm/s

The thickness of the flame is

  2 / S L  2 * 5.78 *105 / 0.418


=0.27 mm
The experimental value of SL for this mixture is 38.9 cm/s. Thus, the
analytical expression appears to predict SL with a reasonable
accuracy.

131
5.3. FACTORS INFLUENCING FLAME
VELOCITY AND THICKNESS

5.3.1 Temperature
Using Eqns (5-23) and (5-24), the temperature dependencies of
flame speed, SL, and thickness, δ, may be inferred as

SL  T 0.375TuTbn / 2 exp(  EA / 2RuTb ) P(n2) / 2 (5.25)

  T 0.375Tbn / 2 exp(  EA / 2RuTb ) P n / 2 (5.26)

Where n is the overall reaction order, and T  0.5(Tu  Tb ) .


Equation 5.25 indicates that the laminar flame speed is a strong
function of temperature. It predicts the flame speed to increase by a
factor of 3.64 when the unburned gas temperature is increased from
300 K to 600 K., see table 5.1 case B. The table also illustrates the
influence of variation in burned gas temperature that may be due to
heat transfer or shifting from the equivalence ratio of maximum
temperature, case C. It is apparent that the flame speed decreases
while flame thickness significantly increases with reducing the
burned gas temperature.
Table 5.1 Estimate of effects of unburned and burned gas
temperatures on laminar flame speeds and thickness
Case A B C
Tu(K) 300 600 300
Tb(K) 2000 2300 1700
SL/SLA 1 3.64 0.46
δ/δA 1 0.65 1.95

132
An empirical correlation of speed of laminar flame for stoichiometric
methane-air flames at 1 atm is given by Andrews and Bradley.

S L (cm / s)  10  3.71*10 4[Tu ( K )]2 (5.27)

The correlation is presented in Fig. 5.5, along with data from other
investigators.

Figure 5.5 Effect of unburned gas temperature on laminar flame


speeds for stoichiometric methane-air flames at 1 atm.

133
5.3.2 Pressure
1/ 2
  F 
m
S L   2 (1   ) 
  u 

Experimental measurements indicate that pressure has a negative


effect on laminar flame velocity. Andrews and Bradley give the
following correlation for methane-air flame.

S L (cm / s)  43[ P(atm)]0.5 (5.28)

It fits their data for pressure greater than 5 atm, see Fig. 5.6.

134
Figure 5.6 Effect of pressure on laminar flame speed of
stoichiometric methane-air mixture for Tu=16-27 oC.

5.3.3 Equivalence Ratio


The main effect of equivalence ratio on flame speed is directly
related to its impact on flame temperature. Thus, the flame speed is
expected to exhibit a maximum at a slightly rich mixture and falls of
on either side, see Fig. 5.7, in similarity to flame temperature. On the
other side, flame thickness exhibits an inverse trend, i.e., it has a
minimum value at a slightly fuel rich mixture as shown in Fig. 5.8.

135
Figure 5.7 Influence of equivalence ratio on the laminar flame speed
of methane-air mixtures at atmospheric pressure.

136
Figure 5.8 Flame thickness and quenching distance for methane –air
mixture at atmospheric pressure.

5.3.4 Fuel Type


Figure 5.9 reports measured flame speed for C2-C6 paraffins
(single bonds), olefins (double bonds), and acetylenes (triple bonds).
As a general trend acetylenes have highest flame speeds while
parrafinns have the lowest ones. Hydrogen has the maximum flame
speed due to some reasons. Hydrogen has much higher values of
thermal diffusivity and mass diffusivity than hydrocarbon fuels.
Moreover, the reaction kinetics of H2 are comparatively very rapid
since the relatively slow CO  CO2 step which is a major aspect in
hydrocarbon combustion is absent. Table 5.2 reports laminar flame
speed for a number of pure fuels.

137
Figure 5.9 Relative flame speed for C1-C6 hydrocarbon fuels. The
reference flame speed is based on propane using the tube method.

Table 5.2 Laminar flame speeds for different pure fuels burning in air
(=1, P= 1 atm and Tu=room temperature)
Fuel Formula Laminar flame speeds, SL (cm/s)
Methane CH4 40
Ethane C2H6 43
Propane C3H8 44
Ethylene C2H4 67
Acetylene C2H2 136
Hydrogen H2 210

138
5.4. QUENCHING, FLAMMABILITY
AND IGNITION

Flame quenching and flame ignition are two important


transient processes that will be treated here. The attention will be
confined to examining limit behavior, i.e., conditions under which a
flame will either extinguish or not, or ignite or not.
Many techniques can be used to extinguish flames. Flames may
be extinguished while passing through narrow passageways. Addition
of diluents is other techniques. Adding water may extinguish the
flame due to thermal effect while adding chemical suppressants, such
as halogens, extinguishes the flames as a result of altering chemical
kinetics. Flames can be effectively extinguished by blowing the flame
away from the reactants.
Three concepts including quenching distance, flammability
limits, and minimum ignition energies are discussed.

5.4.1 Quenching by a Cold Wall


When a flame enters a relatively large passageway it will
propagate. However, reducing the passageway to a certain size will
cause the passing flame to extinguish. For a circular tube, the
quenching distance is the critical diameter where the
flame extinguishes rather than propagates. The
quenching distance is experimentally determined by observing
whether stabilized flame does or doesn't flashback for a particular
tube diameter when the reactant flow is rapidly shut off. For a high-

139
aspect-ratio rectangular- slot burners, the quenching distance is
taken as the slot width. Tubes based quenching distances are
somewhat larger (  30% ) than slit based ones.
Williams provides two criterions governing ignition and flame
extinction.
1. Ignition will only occur if enough energy is added to the gas to
heat a slab about as thick as a steadily propagating laminar
flame to the adiabatic flame temperature.
2. The rate of liberation of heat by chemical reactions inside the
slab must approximately balance the rate of heat loss from the
slab by thermal conduction.

Simplified Quenching Analysis


Williams' second criterion can be applied to a flame that starts
to go through a rectangular slot, as shown in Fig. 5.10. An energy
balance can be obtained by equating the rate of heat produced by
reaction to the heat lost by conduction to the walls

Q V  Q cond (5.29)

141
Fig. 5.10 Flame quenching in rectangular slot.

Where the volumetric heat release rate Q  is expressed as

Q   m Fhc (5.30)

The heat loss from the flame slab to the wall can be obtained using
Fourier's law as

 dT
Qcond  kA
dx in gas (5.31)
at wall
where thermal conductivity, k, and the temperature gradient are
evaluated in the gas at wall. The area A is 2L , where δ the
adiabatic laminar flame thickness, L is slot length (perpendicular to
the page) and the factor 2 accounts for the flame being in contact

141
with wall on each side. The temperature gradient is assumed to be
linear and is expressed as

dT Tb  Tw

dx d /b (5.32)

Where d is the quenching distance and b is an arbitrary constant


which is introduced to increase the gradient close to the actual one.
Now, the Eqn. 5.29 becomes

Tb  Tw
(m F hc )(dL)  k (2L) (5.33)
d /b
or

2kb(Tb  Tw )
d  2
 m F hc (5.34)

Let Tw  Tu and using the relations

1/ 2
 m F 
S L   2 (1   ) 
 u  (5.23)

142
1/ 2
  2 u 
  
 (1   )m F  (5.24a)

hc  (1  )c p (Tb  Tu ) (5.35)

Eqn. 5.34 becomes

d  2 b / S L (5.36a)
or

d  b (5.36a)

Quenching distances for different fuels are reported in Table 5.3.


Quenching distance is usually greater than the flame thickness.

143
Table 5.3 Quenching distances, flammability limits and minimum
ignition energies for various fuels

113
Example
Consider the design of a laminar-flow, adiabatic, flat flame
burner consisting of a square arrangement of a thin-walled tubes as
illustrated in the sketch. Fuel-air mixture flows through both the
tubes and the interstices between the tubes. It is desired to operate
the burner with stoichiometric methane-air mixture exiting the
burner tubes at 300 K and 4 atm. The quenching distance is
expressed as d  6 / S L for methane-air mixture. You may

consider that   T
1.5
/ P.

A. Determine the mixture mass flowrate per unit cross-sectional area


at the design conditions.
B. Estimate the maximum tube diameter allowed so that flashback
will be prevented.

114
Solution
A. For a flat flame to stabilize, the mean velocity of unburned
mixture must be equal the laminar flame speed at the design
temperature and pressure. Applying the correlation by Andrews and
Bradley for methane-air mixture, SL is given as

S L (300 K )  43 * P 0.5
=43*(4)-0.5
=0.215 cm/s
  , is expressed as
The mass flux, m

   m
m  / A  u S L

Assuming that the mixture is ideal gas

MWmix  CH MWCH  (1  CH )MWair


4 4 4

 0.095 * (16)  (1  0.095)28.85


=27.6 kg/kmol
and

P 5 *101325
u  
( Ru / MWmix )Tu (8315 / 27.6)(300)
=5.61 kg/m3
Thus, the mass flux is

   u S L  5.61* 0.215
m

115

  1.21 kg/s - m
m 2

B. The diameter of the tube may be taken reasonably less than the
quenching distance to make the burner operate without danger of
flashback.

The quenching distance is given by

d  6 / S L
Thermal diffusivity should be obtained at the mean temperature.

1 1
Tt  (Tb  Tu )  (2260  300)
2 2
=1280 K

P 101325
u  
( Ru / MW )Tu (8315 / 29) * 300
=1.18 kg/m3

Then

k (Tt ) 0.0809
 
u c p (Tt ) 1.18 *1186
=5.78*10-5 m2/s
116
The quenching distance at 1 atm, d1, is

d 1 6 * 5.78 *10 / 0.215 -5

d1= 1.61 mm

Using d  6 / S L and   T 1.5 / P , the quenching distance can


be expressed as

d   / S L  T 1.5 /( PS L )

The quenching distance at 4 atm, d2, is given as

T2 1.5 P1 S L,1
d 2  d1* ( ) ( )( )
T1 P2 S L,2
1 0.43
d 2  1.61 * (1) 1.5
( )( )
4 0.215
d2=0.805 mm

Considering 20% less as a factor of safety, the design diameter may


be taken as

d design  0.65mm

117
To verify that the flow is laminar in the tube of the design diameter,
it is needed to determine Reynolds's number using air properties.

P 101325 * 4
u  
( Ru / MW )Tu (8315 / 29) * 300
=4.72 kg/m3
then

u d design S L
Re 

4.72 * 0.00065 * 0.215

15.89 *106
=41.5

The value of Reynolds's number is well blow the transient value, i.e,
the flow is laminar.

118
5.4.2 Flammability Limits

Experimentally, it is found that a flame propagates only


through a mixture of equivalence ratio, Φ, laying within a certain
range that is called flammability limits. The leanest mixture that
allows steady flame propagation is called the lower limit while the
richest mixture represents the upper limit. Table 5.3 reports
flammability limits for a number of fuel-air mixtures at atmospheric
pressure. To measure the flammability limits, a standard vertical
tube of 1.22 m (4 ft) length with an inner diameter of 5 cm is used. If
a mixture filling the tube is capable of propagating a flame
indefinitely away from the ignition source, it is called flammable. By
adjusting the mixture strength, the flammability limits can be
attained.

Example
A full propane cylinder from a camp stove leaks its contents of
0.5 kg into a (3.7*4.3*2.5) room at 20 oC and 1 atm. After a long time,
the fuel gas and room air are well mixed. Is the mixture in the room
flammable?

Solution:
From table 5.3, propane-air mixtures are flammable for
0.51<Φ<2.83. To solve the problem, one should obtain the
equivalence ratio of the mixture filling the room.
Assuming ideal gas behavior, the partial pressure of propane is
determined as

119
mC H ( Ru / MWC H )T
PC H  3 8 3 8

3 8
Vroom
0.5 * (8135 / 44) * 293

3.7 * 4.3 * 2.5
=680 Pa

The propane mole fraction is

PC H 680
C H  3 8

3 8
P 101325
=0.0067
and

 air  (1  0067)  0.9933


The air-fuel ratio of the mixture in the room is

Yair MWair  air 28.85 * 0.9933


(A/ F)   
YC H MWC H  C H
3 8
44 * 0.0067
3 8 3 8

= 97.21
The equivalence ratio of the mixture in the room is

( AlF ) stoich 15.6


 
( A / F ) actual 97.21
=0.1604
It is obvious that the mixture is not flammable since the equivalence
ratio is less than the lower flammability limit (Φ=0.51).

121
5.4.3 Ignition
Spark ignition is the most common method used for ignition
since it is highly reliable and does not require a pre-existing flame.
Spark ignition is used in many devices including SI engines, gas
turbines; and different industrial, commercial, and residential
burners. This section is dedicated to discussion of minimum energy
concept.

Simplified Ignition Analysis


Consider a spherical volume of gas that represents the incipient
propagating flame generated by a spark ignition engine. By applying
William's second criterion, it is possible to define a critical gas
volume radius such that a flame will not propagate if the actual
radius is smaller than the critical one. The second step is to apply
William's first criterion. That is the minimum ignition energy to be
provided by a spark is the energy needed to heat the gases of the
critical volume to the flame temperature.
The critical radius, Rcrit, can be determined by writing an
energy balance equation in which the rate of heat librated by
reaction equals the rate of heat lost to the cold gases, a schematic of
the problem is shown in Fig. 5.11.

Q V  Q cond (5.37)


Or

121
dT
 F hc 4Rcrit
m 3
/3   k 4Rcrit
2
dr Rcrit
(5.37)

Figure 5.11 Critical volume of gas for spark ignition

The temperature gradient beyond the sphere is assumed as

dT (Tb  Tu )

dr Rcrit Rcrit (5.38)

Substituting Eqn. 5.38 into Eqn. 5.38 yields

3k (Tb  Tu )
2
Rcrit 
 m F hc (5.39)

122
Using the laminar flame speed and thickness expression, one can
obtain


Rcrit  6
SL (5.40)

Rcrit  ( 6 / 2)  1.224 (5.41)

The minimum ignition energy, Eign, is the energy required to heat the
gases of the critical volume to the flame temperature, Tb, then

Eign  mcrit c p (Tb  Tu ) (5.42)

where the mass critical sphere, mcrit, is b (4Rcrit


3
/ 3) . Using
the expression 5.40 for Rcrit, we have


Eign  61.6 b c p (Tb  Tu )( )3
SL (5.42)

Using the ideal gas law b can be replaced to yield the final result:

c p Tb  Tu  3
Eign  61.6 P( )( )( )
Rb Tb SL (5.43)

123
Pressure and Temperature Dependencies
2
EignP (5.44)

Figure 5.12 Effect of pressure on minimum ignition energy

Increasing the initial mixture temperature generally decreases ignition


energies as reported in Table 2.

124
Table 5.4 Influence of temperature on ignition energies

125
FLAME STABILIZATION

Flashback Lift of-Blow off

126
127
Chapter 6

LAMINAR DIFFUSION FLAME

6.1. JET FLAME PHYSICAL DESCRIPTION

The main features of the laminar jet flame are described in


Figure 6.1. As the fuel flows along the flame axis, it diffuses radially
outward, while the oxidizer diffuses radially inward. The flame
surface is defined to exist where the fuel and oxidizer meet in
stoichiometric proportions. The products generated at the flame
surface diffuse radially both inward and outward. The flame length,
Lf, is simply determined by the axial point where equivalence ratio is
unity.

Figure 6.1 Laminar diffusion flame structure.

828
The region where chemical reaction occurs is generally quite
narrow. The high-temperature reaction zone occurs in an annular
region until the flame tip is reached.
For hydrocarbon flames, soot is frequently present, giving the
flame its typical orange or yellow appearance. In furnaces and
boilers, a certain amount of soot may be beneficial as a means of
increasing radiant heat transfer. However, in engines, soot causes
additional heat loss which is undesirable. The emission of soot
particulates to the atmosphere is undesirable as it may contain
condensed polycyclic aromatic hydrocarbons (PAHs), which have the
potential to cause cancer. Thus, soot emissions from stacks and
tailpipes need to be control.

Figure 6.2 Soot formation and destruction zone in laminar jet flames.

Soot is formed on the fuel side of reaction zone and is


consumed when it follows into an oxidizing region. Soot formation
and destruction zones are illustrated in Fig. 6.2. Some soot may go
beyond the high-temperature zone without oxidation generating soot
wings. This soot breaks through the flame and is generally referred
to as smoke. In general, the soot formation proceeds in four steps:

829
1. Formation of precursor species: Acetylene (C2H2) may be
considered a precursor species that causes the formation of
larger ring structures, PAHs.
2. Particle inception: It involves the formation of small particles of
critical size from growth by chemical means and coagulation.
3. Surface growth and agglomeration.
4. Particles oxidation. At some point, soot particles must pass
through an oxidizing zone of the flame. If all of soot particles are
oxidized, the flame is termed nonsooting, while incomplete
oxidation yields a sooting flame.

831
6.2 FLAME LENGTHS

The following expressions are developed by Roper and verified


experimentally for different burner geometries including circular, square
and slot. The flow regimes (momentum-controlled, buoyancy-controlled
and transitional) are considered. Flame Froude number, Frf, is used to
establish the regime of the flow. Physically, Froude number is defined as
the initial jet momentum flow to the buoyant force experienced by the
flame, i.e.,

(Ve IYF , Stoic ) 2


Frf  (6.1)
aL f
The flow regime can be determined by the following criteria

Frf  1 Momentum-controlled

Frf  1 Mixed (transient) -controlled

Frf  1 Buoyancy-controlled

Where I is the ratio of the actual initial momentum flow from the slot to
that of uniform flow

I  J e, act / mF ve ,
If the flow is uniform, I=1, and for a fully developed, parabolic exit
velocity profile (for h>>b), I=1.5.
YF,Stoic is the stoichiometric fuel mass fraction.
Where a is the mean buoyant acceleration evaluated from
Tf
a  0.6 g (  1) where g is the gravitational acceleration.
T

838
6.2.1 Circular-Port
The expressions are valid for momentum-controlled or buoyancy-
controlled.
Theoretical Expression
0.67
QF (T / TF )  T 
L f ,thy 
4D ln( 1  1 / S )  T f  (6.2)

where

QF  R 2 ve is the initial volumetric flow rate, m3/s

T is the oxidizer stream temperature, K


TF is the fuel stream temperature, K

Tf is the mean flame temperature, K

S is the molar stoichiometric oxidizer-fuel ratio

D is the mean diffusion coefficient for oxidizer at T

Experimental Expression

QF (T / TF )
L f , exp t  1330
ln( 1  1 / S ) (6.3)

where all quantities are evaluated in SI units. Note that the burner
diameter does not explicitly appear in either of these expressions.

832
6.2.2 Square-Port

The expressions are valid for momentum-controlled or buoyancy-


controlled.
Theoretical Expression
0.67
QF (T / TF )  T 
L f ,thy   
16 D [inverf ((1  S ) )]  T f 
 0.5 2  (6.4)

Experimental Expression

QF (T / TF )
L f , exp t  1045
[inverf ((1  S )  0.5 )]2
(6.5)

where inverf is the inverse error function. Values of the error function erf
are tabulated in Table 6.1. Again all quantities are in SI units.

6.2.3 Slot Burner, Momentum Controlled


Theoretical Expression
2 0.33
b QF  T 
2
 Tf 
L f ,thy      (6.6)
hIDYF , stoic  TF   T 
Experimental Expression
2
b 2QF  T 
L f , exp t  8.6 *10 4
  (6.7)
hIDYF , stoic  TF 
where b is the slot width, h is the slot length and the function β is given by

833
1
 ,
4inverf [1 /(1  S )]

Table 6.1 Gaussian error function

The

Gaussian error function is defined as


The inverse error function can be calculated as

834
6.2.4 Slot Burner, Buoyancy Controlled

Theoretical Expression
2/9
 9 4QF4 T4  T f 
L f ,thy   2 4 4   (6.8)
 8D ah TF  T 
Experimental Expression

3
QF4 T4 
4
L f , exp t  2 *10  4 4  (6.9)
 ah TF 
Where a is the mean buoyant acceleration evaluated from

TF
a  0.6 g (  1) (6.10)
T
where g is the gravitational acceleration. TF is the mean flame
temperature and may be assumed 1500 K.

6.2.5 Slot Burner, Transient Controlled


For transient regime where both jet momentum and buoyancy are
important, the following expression is recommended

3  3 2 / 3 
4  L f ,B    L f ,M  
L f ,T  L f ,M    1  3.38    1
L   
9  f ,M    L f ,B   
 
835
(6.11)
where the subscripts M, B and T refer to momentum-controlled,
buoyancy-controlled and transient, respectively.

Example 6.1
It is desired to operate a square-port diffusion flame with a 50-
mm high flame in a laboratory. Determine the volumetric
flowrate required if the fuel is propane. Also, determine the heat
 hc ) of the flame. What flowrate is required if
release ( m
methane is substituted for propane?

Solution
The following correlation is used to calculate the flame length for
square port

QF (T / TF )
L f , exp t  1045
[inverf ((1  S )  0.5 )]2
(6.5)

For a certain flame length the corresponding fuel flow rate can be
determined using the above correlation.
The molar stoichiometric air-fuel ratio, S, is calculated as

S  x  y / 4 * 4.76
=(3+8/4)*4.76=23.8 kmol/kmol

Thus, using Table 6.1

836
Assuming that T  TF  300 K and solving eqn 6.5 for QF,
yields

or

The propane density, at P=1 atm, T=300K) may be estimated using ideal
gas law and heat of combustion can be taken from Appendix B. Then, the
heat released rate is

Repeating the problem for methane, it is found that

Thus,

and

837
6.3 FACTORS AFFECTING FLAME LENGTH

6.3.1 Flowrate and Geometry

As shown in Fig. 6.3, the flame length has a linear dependence on


flowrate for the circular port. In the case of slot burner, the dependence
on flowrate is somewhat greater than linear. It is also noted that as the
slot-burner ports become more narrow (h/d increasing), the flames
becomes significantly shorter for the same flowrate.

Figure 6.3 Effect of flowrate and geometry of flame length

838
6.3.2 Fuel Type

By definition, the molar stoichiometric ratio, S, varies with fuel


type,

 moles ambient fluid 


S  stoic (6.12)
 moles nozzle fluid 
For a generic pure hydrocarbon, CxHy, the stoichiometric ratio can
be expressed as

x y/4
S
O2 (6.13)

where  O 2 is the mole fraction of oxygen in the air.

Figure 6.4 Dependence of flame length on fuel type

839
Figure 6.4 shows the flame length for different fuels relative to
methane calculated using Eqn. 6.3, assuming equal flowrate for all fuels.
The flame length increases as H/C ratio of the fuel decrease. This is less
significant for higher hydrocarbons. It is also noted that the carbon
monoxide and hydrogen flames are much smaller than the hydrocarbon
flames.

6.3.3 Primary Aeration


In gas-burning appliances a part of air, typically 40-60% of the
stoichiometric, is premixed with fuel gas before it burns as a laminar
diffusion flame. The primary aeration makes the flames short and
prevents soot formation, resulting in the familiar blue flame. The
maximum ratio of primary air is limited by safety considerations. If too
much air is added, the premixed mixture may exceed the rich
flammability limit that may lead condition of flashback.

The influence of primary aeration on the methane flame length is


shown in Fig. 6.5. Flame length is reduced by 85-90% from the original
no-air added length. The factor S, as defined by Eqn. 6.12, depends on the
chemical compositions of both the nozzle-fluid stream and the
surrounding fluid. Considering the nozzle fluid as a mixture of the pure
fuel and primary air, S can be evaluated as

1  pri
S
 pri  (1 / S pure) (6.14)

841
where ψpri is the mole fraction of the stoichiometric requirement met by
the primary air, and Spure is the molar stoichiometric ratio associated with
the pure fuel.

Figure 6.5 Effect of primary aeration on laminar jet flame length

848
6.3.4 Oxygen Content of Oxidizer

The concentration of oxygen in oxidizer strongly affects the length


of flame as shown in Fig. 6.6. As reported in the figure, the flame is
greatly stretched with reduction in oxygen concentration below 21%. On
the other side, with pure-oxygen, flame length reduces to one-quarter its
value in air.

Figure 6.6 Effect of oxygen content in the oxidizer stream on flame length

842
6.3.5 Fuel Dilution with Inert gas

Diluting the fuel with an inert gas also reduces the flame length
and lessens soot formation down to the variation in the stoichiometric
ratio, S. For hydrocarbon fuels,

x y/4
S
 1 
   O2 (6.15)

 1   dil 

where χdil is the diluent mole fraction in the fuel stream.

843
Example 6.2
In a study of nitric oxide formation in laminar jet flames,
the propane fuel is diluted with N2 to suppress soot formation.
The nozzle fluid is 60 percent N2 by mass. The burner has a
circular port. The fuel, nitrogen and air are all at 298 K and 1
atm. Compare the flame lengths for the following two cases

 F  5.10 6 kg/s). What is the


with the undiluted base case ( m
physical significance of your results? Discuss.
A. The total flowrate of the diluted flow (C3H8+N2) is 5.10-6
kg/s.
B. The flowrate of the C3H8 in diluted flow is 5.10-6 kg/s.

Solution
To calculate the flame length, it is needed to determine the flame
temperature, Tf . As a first approximation the flame temperature may be
taken equal to the adiabatic flame temperature. The general
stoichiometric chemical balance is

C3 H8  N 2  5(O2  3.76 N 2 ) 
3CO2  4H 2O  (18.8   ) N 2
The adiabatic flame temperature can be determined by equating the
enthalpy of the reactant to the enthalpy of the products.

H reac (Ti , P)  H prod (Tad , P)

844
 103847  3 *[-393546 + 56.21* (Tad  289)] 
4 *[241845  43.87 * (Tad  289)])
 (18.8   ) * 33.71* (Tad  298

After rearranging, we can get

2335573  10046 *
Tad 
(977.9  33.71* ) (AA)

The parameter α can be expressed using the mass fraction, YN2

YN 2 MWC3 H 8

(1  YN 2 ) MWN 2

1. The base case, the undiluted fuel with  F  5.10 6 kg/s.


m

The flame length may be given by


0.67
QF (T / TF )  T 
Lf 
4D ln( 1  1 / S )  T f 

For Propane C3H8, the molar stoichiometric oxidizer-fuel ratio, S, can be


determined as

1
S  ( x  y / 4) /[( ) O2 ]  4.76 * ( x  y / 4) /(1   )
1   dil

845
1
S  4.76 * (3  * 8) /(1   )
4
=23.8
where   0

The density of propane at 300 K is

P 101325
f  
( Ru / MWF )T (8315 / 44) * 300
=1.787 kg/m3

The exit flow rate, QF , can be determined as

5 *106
m f
QF  
f 1.787
=2.798*10-6 m3/s

Assume DO2 air  D  2.1*10 5 m 2 /s

Adiabatic flame temperature can be obtained putting α=0 in Eqn. AA

Tad=Tf= 2388 K

846
Thus, the flame length is

2.798 *10 6 (300 / 300)  300 


0.67
Lf  5  
4 * 2.1*10 ln( 1  1 / 23.8)  2388 

=0.064 m =6.4 cm

2. The total flowrate of the diluted flow (C3H8+N2) is 5.10-6 kg/s.

The adiabatic flame should be calculated first

YN 2 MWC3H8 0.6 44
  *
(1  YN 2 ) MWN 2 (1  .6) 28
=2.36
Then,

2335573  10046 * 2.36


Tad 
(977.9  33.71* 2.36)
=2231 K

The molar stoichiometric oxidizer-fuel ratio, S, is

1
S  4.76 * (3  * 8) /(1  2.38)
4
=7.041

847
To calculate the exit flowrate, the mean density of mixture must be
determined

1
MWmix   32.77
0.6 / 28  0.4 / 44
101325
f 
(8315 / 32.77) * 300
=1.33 kg/m3

The exit flow rate, QF , can be determined as

5 *106
m f
QF  
f 1.33
=3.759*10-6 m3/s

Thus, the flame length is

3.759 *10 6 (300 / 300)  300 


0.67
Lf   
4 * 2.1*10 5 ln( 1  1 / 7.041)  2231 
=0.028 m=2.8 cm

848
3. The flowrate of the C3H8 in diluted flow is 5.10-6 kg/s.

In this case the exit flowrate is

m f / 0.6 5 *106 / 0.6


QF  
f 1.33
=6.265*10-6 m3/s

Thus, the flame length is

6.265 *10 6 (300 / 300)  300 


0.67
Lf   
4 * 2.1*10 5 ln( 1  1 / 7.041)  2231 
=0.047 m=4.7 cm

849
Example 6.3
Design a natural-gas burner for a commercial cooking range that has a
number of circular arranged in a circle. The circle diameter are
constrained to be 160 mm. The burner must deliver 2.2 kW at full load
and operate with 40% primary aeration. For stable operation, the loading
of an individual port should not exceed 10 W per mm2 of port area. Also,
the full-load flame height should not exceed 20 mm. Determined the
number and the diameter of the ports.

Solution

Step 1. Apply port loading constraint. The total port area is

And the constraint is

Thus,

At this point, it is possible to choose a value for N or D and calculate D or


N as first trial for the design.
Choosing N=36 yields D=2.79 mm.
Step 2. Determine flowrates. The design heat-release rate determines the
fuel flowrate:

851
The primary aeration determines the flowrate of air premixed with fuel:

The total volumetric flowrate is

where  is the air-fuel mixture density that can be determined applying


the ideal gas law using the mean molecular weight.

Where Z is the primary molar air-fuel ratio:

Thus,

858
and

Step 3. Check flame length constraint.


The flowrate per port is

The molar ambient to nozzle fluid stoichiometric ratio, S, is evaluated as

The flame length is calculated as

852
A flame length of 19.8 mm meets the requirement of L f  20mm

Step 4. Check practically of design. If the 36 ports are arranged equally


spaced on a 160 mm diameter circle, the spacing between ports is

This spacing appears reasonable.

The flame length, L f , for circular port burners may be given by


0.67
QF (T / TF )  T 
Lf 
4D ln( 1  1 / S )  T f  (6.1)

where

QF  R 2 ve is the initial volumetric flow rate, m3/s

T is the oxidizer stream temperature, K

TF is the fuel stream temperature, K

853
Tf is the mean flame temperature, K

S is the molar stoichiometric oxidizer-fuel ratio


D is the mean diffusion coefficient for oxidizer at Tf

Example 6.1
In a study of nitric oxide formation in laminar jet flames, the
propane fuel is diluted with N2 to suppress soot formation. The nozzle
fluid is 60 percent N2 by mass. The burner has a circular port. The
fuel, nitrogen and air are all at 298 K and 1 atm. Compare the flame
lengths for the following two cases with the undiluted base case

F
(m  5.10 6 kg/s). What is the physical significance of your
results? Discuss.
C. The total flowrate of the diluted flow (C3H8+N2) is 5.10-6 kg/s.
D. The flowrate of the C3H8 in diluted flow is 5.10-6 kg/s.

Solution
To calculate the flame length, it is needed to determine the flame
temperature, Tf . As a first approximation the flame temperature
may be taken equal to the adiabatic flame temperature. The general
stoichiometric chemical balance is
C3 H8  N 2  5(O2  3.76 N 2 )  3CO2  4H 2O  (18.8   ) N 2

The adiabatic flame temperature can be determined by equating the


enthalpy of the reactant to the enthalpy of the products.

854
H reac (Ti , P)  H prod (Tad , P)
-103847 kJ/kg

 103847  3 *[-393546 + 56.21* (Tad  289)] 


4 *[241845  43.87 * (Tad  289)]  (18.8   ) * 33.71* (Tad  298)

After rearranging, we can get

2335573  10046 * 
Tad 
(977.9  33.71*  )
(AA)

The parameter α can be expressed using the mass fraction, YN2

YN MWC H
 2 3 8

(1  YN ) MWN2 2

1. The base case, the undiluted fuel with  F  5.10 6 kg/s.


m

The flame length may be given by

0.67
QF (T / TF )  T 
Lf 
4D ln( 1  1 / S )  T f 
For Propane C3H8, the molar stoichiometric oxidizer-fuel ratio, S,
can be determined as

855
1
S  4.76 * (3  * 8) /(1   )
4
=23.8

where  0

The density of propane at 300 K is

P 101325
f  
( Ru / MWF )T (8315 / 44) * 300
=1.787 kg/m3

The exit flow rate, QF , can be determined as

5 *106
m f
QF  
f 1.787
=2.798*10-6 m3/s

Assume DO air  D  2.1*105 m2 /s


2

Adiabatic flame temperature can be obtained putting α=0 in Eqn. AA

Tad=Tf= 2388 K

856
Thus, the flame length is

2.798 *10 6 (300 / 300)


0.67
 300 
Lf  5  
4 * 2.1 *10 ln( 1  1 / 23.8)  2388 

=0.064 m =6.4 cm

4. The total flowrate of the diluted flow (C3H8+N2) is 5.10-6 kg/s.

The adiabatic flame should be calculated first

YN MWC H 0.6 44
 2 3 8
 *
(1  YN ) MWN2 2
(1  .6) 28
=2.36
Then,

2335573  10046 * 2.36


Tad 
(977.9  33.71* 2.36)
=2231 K

The molar stoichiometric oxidizer-fuel ratio, S, is

1
S  4.76 * (3  * 8) /(1  2.38)
4
=7.041

857
To calculate the exit flowrate, the mean density of mixture must be
determined

1
MWmix   32.77
0.6 / 28  0.4 / 44
101325
f 
(8315 / 32.77) * 300
=1.33 kg/m3

The exit flow rate, QF , can be determined as

5 *106
m f
QF  
f 1.33
=3.759*10-6 m3/s

Thus, the flame length is

3.759 *10 6 (300 / 300)


0.67
 300 
Lf  5  
4 * 2.1 *10 ln( 1  1 / 7.041)  2231 
=0.028 m=2.8 cm

5. The flowrate of the C3H8 in diluted flow is 5.10-6 kg/s.

In this case the exit flowrate is

858
m f / 0.6 5 *106 / 0.6
QF  
f 1.33
=6.265*10-6 m3/s

Thus, the flame length is

6.265 *10 6 (300 / 300)


0.67
 300 
Lf  5  
4 * 2.1*10 ln( 1  1 / 7.041)  2231 
=0.047 m=4.7 cm

Example 6.3
Design a natural-gas burner for a commercial cooking range that has a
number of circular arranged in a circle. The circle diameter is constrained
to be 160 mm. The burner must deliver 2.2 kW at full load and operate
with 40% primary aeration. For stable operation, the loading of an
individual port should not exceed 10 W per mm2 of port area. Also, the
full-load flame height should not exceed 20 mm. Determined the number
and the diameter of the ports.

Solution

Step 1. Apply port loading constraint. The total port area is

And the constraint is

859
Thus,

At this point, it is possible to choose a value for N or D and calculate D or


N as first trial for the design.
Choosing N=36 yields D=2.79 mm.
Step 2. Determine flowrates. The design heat-release rate determines the
fuel flowrate:

The primary aeration determines the flowrate of air premixed with fuel:

The total volumetric flowrate is

where  is the air-fuel mixture density that can be determined applying


the ideal gas law using the mean molecular weight.

861
Where Z is the primary molar air-fuel ratio:

Thus,

and

Step 3. Check flame length constraint.


The flowrate per port is

The molar ambient to nozzle fluid stoichiometric ratio, S, is evaluated as

868
The flame length is calculated as

A flame length of 19.8 mm meets the requirement of L f  20mm

Step 4. Check practically of design. If the 36 ports are arranged equally


spaced on a 160 mm diameter circle, the spacing between ports is

This spacing appears reasonable.

862
Chapter 7

DROPLET EVAPORATION AND BURNING

Burning of droplet is a nonpremixed combustion system.


Knowledge of evaporation rates and droplet life time is important in
the design and operation of practical devices. Those include diesel,
rocket and gas turbines engines, as well as, oil fired boilers, furnaces,
and process heaters.

Simple Model of Burning Droplet


The following assumptions have been made to simplify the
model analysis; however, the model still preserves the essential
physics and agrees reasonably with experimental results:
1. The burning droplet is surrounded by a spherically symmetric
flame and exists in a quiescent, infinite medium.
2. The burning process is quasi-steady.
3. The fuel is a single-component liquid with zero solubility for gases.
4. The pressure is uniform and constant.
5. The gas phase consists of only three species: fuel vapor, oxidizer
and combustion products. The gas-phase region is divided into
two zones. An inner zone between the droplet surface and the
flame sheet contains only fuel vapor and products. An outer zone
around the flame sheet consists of oxidizer and products. Thus,
binary diffusion prevails.

361
6. Fuel and oxidizer react in stoichiometric proportions at the flame
sheet. Chemical kinetics are assumed to be infinity fast, resulting
in an infinitesimal thin flame sheet.
7. The Lewis number, Le   / D  k g /( c pg D) , is unity.

8. Radiation heat transfer is neglected.


9. The gas-phase thermal conductivity, specific heat and the product
D are all constants.
10. The liquid fuel droplet is the only condensed phase.

The basic model is shown in Fig. 7.1.

Figure 7.1 Temperature and species profiles for simple droplet burning model

361
For given an initial droplet size and conditions far from the
droplet, i.e., the temperature, T∞, and the oxidizer mass fraction,
Yox,∞, the model analysis aims to yield expressions that allow us to
evaluate the following five parameters:
1. The droplet mass burning rate, m F .
2. The flame temperature, Tf.
3. The flame radius, rf.
4. The fuel vapor mass fraction at the droplet surface, YF , s .

5. The droplet surface temperature, Ts.

Thus, five equations are needed to solve for the five unknowns. We
can derive five equations through
1. an energy balance at the droplet surface
2. an energy balance at the flame sheet
3. the oxidizer distribution in the outer region
4. the fuel vapor distribution in the inner region
5. phase equilibrium at the liquid-vapor interface expressed, for
example, by the clausius-Clapeyron relationship.

7.1.1 Mass Conservation


Overall mass conservation in the gas phase is described as

(7.1)

361
7.1.2 Species Conservation
Inner region Applying Fick’s law in the inner region where the fuel
vapor is the important diffusing species, we get

(7.2)
where the subscripts A and B denote fuel and products, respectively:

(7.3)

(7.4)
And the  -operator in spherical coordinates is (..)  d (..) / dr as
the only variations are in the r-direction. Thus, the Fick’s law is

(7.5)
The first-order ordinary differential equation must satisfy two
boundary conditions: at the droplet surface, liquid-vapor equilibrium
prevails, so

(7.6)
and at the flame, the fuel disappears, so

(7.7)

366
The general solution of Eqn. 7.5 is

(7.8)

where Z F  1 / 4D
Applying the droplet surface condition (Eqn. 7.6) to evaluate C1, we
obtain

(7.9)
Applying the flame boundary condition, a relation involving
the three unknowns YF,s, m F , and rf:

(7.10)
The combustion products mass fraction can be expressed as

(7.11)

Outer Region

In this region, the important diffusing species is the oxidizer, which is


transported radially inward to the flame. At the flame, the oxidizer and
fuel combine to stoichiometric proportions according to

(7.12)

361
Where ν is the stoichiometric (mass) ratio and includes any
nonreacting gases that may be a part of the oxidizer stream. This
relationship is illustrated in Fig. 7. 2. The vector mass fluxes in Fick’s
law are thus

(7.13)

(7.14)

(7.15)

Figure 7.2 Mass flow relationships at flame sheet. Note that the net mass
F
flow in both the inner and outer regions is equal to the fuel flowrate, m

361
with boundary conditions

(7.16)

(7.17)
Integrating Eqn. 7.15 yeilds

(7.18)
Applying the flame condition to eliminate C1 yields

(7.19)
Applying the condition at r   (Eqn. 7.17), It is obtained an
 F , and the flame radius,
algebraic relation between the burning rate, m
rf, i.e.,

(7.20)
The product mass distribution is

(7.21)

361
7.1.3 Energy Conservation
As the chemical reaction is confined to occur at the flame sheet, the
reaction rate is zero in the inner and outer zones. The energy equation
is then given by

(7.22)

Putting ZT  c pg / 4k g the energy equation becomes

(7.23)

Note that when Lewis is unity Z F  ZT , i.e, c pg / k g  D .


The boundary conditions applied to Eqn. 7.23 are
For inner region

(7.24)

(7.25)
For outer region

(7.26)

(7.27)

Of the three temperatures, only T∞ is considered to be known; Ts and


Tf are two of the five unknowns in the present problem.

311
7.1.4 Temperature distributions
The general solution of the Eqn. 7.23 is

(7.28)

Applying Enq.s 7.24 and 7.25, the temperature distribution in the

inner zone is obtained for

(7.28)

In the outer region, application of boundary conditions 7.26 and 2.27


yields for

(7.29)

313
7.1.5 Energy Balance at droplet surface

Figure 7.3 Surface energy balance at droplet liquid-vapor interface (top) and
at flame sheet (bottom)

The conduction heat transfer rates and the enthalpy fluxes at the
surface of the evaporating droplet are shown in Fig. 7.3.

(7.30)
or

(7.31)
311
The heat conducted into the droplet interior, Q i  l , can be handled in
several ways. One common approach is to model the droplet as
consisting of two zones: an interior region existing uniformly at its
initial temperature, T0; and a thin surface layer at the surface
temperature, Ts. For this so-called “onion-skin” model,

(7.32)
is the energy required to heat up from To to Ts the fuel that is
vaporized. For convenience, we define

(7.33)
So, for the onion model

(7.34)

Another common treatment of Q i  l is to assume that the droplet


behaves as a lumped parameter, i.e., it has a uniform temperature, with
a transient heat up period. For the lumped parameter,

(7.35)
and

311
(7.36)
where md is the mass of the droplet. Implementation of the lumped-
parameter model requires solving energy and mass conservation
equations for the droplet as a whole in order to obtain dTs/dt.
A third approach and the simplest, is to assume that the droplet
rapidly heats up to steady temperature, Ts. This, in effect, says the
thermal inertia of the droplet is negligible. With this assumption of
negligible thermal inertia,

(7.37)
The conduction heat transfer from the gas phase, Q i  g , can be
evaluated by applying Fourier’s law and using the temperature
distribution in the inner region to obtain the temperature gradient, i.e.,

(7.38)
where

(7.39)

Evaluating the heat-transfer rate at r=rs, Eqn.7.38 becomes, after


rearranging and substituting the definition of ZT,

311
(7.40)

 F , Tf, Ts and rf.


Equation 40 contains four unknowns: m

7.1.6 Energy Balance at Flame Sheet

Figure 7.3 Surface energy balance at droplet liquid-vapor


interface (top) and at flame sheet (bottom)

Referring to Fig. 7.3, the various energy fluxes at the flame sheet are
reported. Since flame temperature is the highest flame in the system,

311
heat is conducted both toward the droplet Q f  i and away to infinity,

Q f   . The chemical energy released at the flame is taken into


account by using absolute enthalpy fluxes for fuel, oxidizer and
products. A surface energy balance at the flame sheet can be written
as:

(7.41)

The enthalpies are defined as:

(7.42)

(7.43)

(7.44)

(7.45)

The mass flowrates of fuel, oxidizer, and products are related by


stoichiometry, Eqn. 7.12. Note that although products exist in the inner
region, there is no net flow of products between the droplet surface and
the flame; so, all the products flow outward away from the flame. Thus,
Eqn. 7.41 becomes

316
(7.46)

Substituting Eqns. 7.42-7.45 into the above yields

(7.47)
Considering that cpg is constant and Δhc is independent of temperature;
thus, we choose the flame temperature as a reference state to simplify
Eqn. 7.47.

(7.48)

Applying Fourier’s law at the two sides of the flame sheet, we can replace

Q f  i and Q f   , thus

(7.49)

The temperature distributions in inner (Eqn. 2.28) and outer regions



(Eqn. 2.29) can be differentiated to the temperature gradient at r  r f

and at r  r f , respectively. Performing these substitutions and
rearranging, the flame sheet energy balance is finally expressed as

311
(7.50)

Equation 7.50 is a nonlinear algebraic relation involving the same four

unknowns that appears in Eqn. 7.40.

7.1.7 Liquid-Vapor Equilibrium So far there are four


equations and 5 unknowns. The fifth equation may be obtained through
the Clausius-Clapereyron equation considering the equilibrium between
the liquid and vapor phases at the fuel droplet surface. At the liquid-vapor
interface, the partial pressure of the fuel vapor is given by

(7.51)

Where A and B are constants obtained from the Clausius-Clapeyron


equation as:

 h fg 
A  exp  
 RTboil 
(7.52)

(7.53)

311
The fuel partial pressure can be related to the fuel mole fraction and mass
fraction as follows:

(7.54)
and

(7.55)
Substituting Eqns. 7.51 and 7.54 into Eqn. 7.55 yields

(7.56)

Thus, the mathematical description of the simplified droplet burning


model is completed.

311
7.1.8 Summary and Solution
Table 7.1 summarizes the five equations that must be solved for the
 F , r f , T f , Ts and YF , s . The system of nonlinear
five unknowns: m

equations can be reduced to a useful level by simultaneously solving


 F , r f , and T f treating Ts as a known parameter for the time
Eqns. for m

being. Following this procedure, the burning rate is

(7.57)

or in terms of transfer number, Bo,q, is defined as

(hc / )  c pg (T  Ts )
Bo, q 
qi l  h fg (7.58)

4k g rs
m F  ln( 1  Bo,q ) (7.59)
c pg

The flame temperature is

qi l  h fg
Tf  [Bo, q  1]  Ts
c pg (1   ) (7.60)

and the flame radius is

311
ln[ 1  Bo,q ]
r f  rs
ln[(  1) / ] (7.61)

The fuel mass fraction at the droplet surface is

Bo,q  1 /
YF ,s 
Bo,q  1 (7.62)

The droplet surface temperature, Ts, is

  YF , s PMW pr 
Ts  B  
 A(YF ,s MWF  Y F ,s MW pr  MWF

(7.63)

Where:
kg is the thermal conductivity
rs is the fuel droplet radius
cpg is the specific heat
qi-l is the heat conducted into the droplet interior
hfg is the latent heat of vaporization
ν is the stoichiometric mass ratio and includes any non-
reacting gases that may be a part of the oxidizer
stream.
P is the surrounding pressure
MW molecular weight
Δhc is the heat of combustion of fuel

313
A  exp[ h fg / RTboil ]

B  h fg / R

R is the universal gas constant


Tboil is the boiling temperature of fuel.

An iteration technique can be used to evaluate the


above parameters. A value for Ts is assumed close to boiling
F,
temperature, then, Eqns. 7.58-7.62 are used to evaluate m
Tf, rf, YF,s. Equation 7.63 is used to improve the value of Ts.
Eqns. 7.58-7.62 can be re-evaluated and the process repeated
until convergence is obtained.

The problem can be greatly simplified by assuming


that the fuel droplet is at the boiling temperature, i.e.,

Ts  Tboil . With this assumption, m F , Tf and rf can be


directly estimated using Eqns. 7.58-7.61 without iteration
and YF,s is unity as Ts  Tboil .

311
7.2 BURNING RATE CONSTANT AND DROPLET
LIFETIME
The droplet radius history can be obtained by writing a mass
balance at the droplet surface.
mass of the droplet decrease = rate at which the liquid is vaporized

dmd
 m F (7.64)
dt
md  lD3 / 6 (7.65)

dD 2
 K (7.66)
dt

where K is the burning rate constant and is expressed as

8k g
K ln( 1  Bo,q )
l c pg (7.67)

The burning rate constant is truly constant only after a steady-state


surface temperature is reached, since, then, Bo,q becomes constant.
Assuming the transient heat-up period is neglected compared
to the droplet lifetime, a D2 law can be obtained for burning droplet.

D 2 (t )  Do2  Kt (7.68)

The droplet lifetime is obtained by letting D (t )  0 ,


2

td  Do2 / K (7.69)

311
The D2 law is a good representation of experimental results after the
heat-up transient as shown in Fig. 7.4.

Figure 7.4 Experimental data for burning droplets illustrating D2 law behavior
after initial transient.

311
The properties cpg, kg, and ρl that appear in Eqn. 7.67 can be determined
as the following:

(7.68)

(7.69)

(7.70)
where

(7.71)

Example
Estimate the mass burning rate of a 1 mm diameter n-hexane droplet
burning in air at atmospheric pressure. Assume no heat is conducted
into the interior of the liquid droplet and that the droplet
temperature is equal to the boiling point. The ambient air
temperature is 298 K. Also, calculate the ratio of the flame radius to
droplet radius and flame temperature; and estimate the droplet
lifetime.

Solution
Using table B.1., the properties of n-hexane, C6H14, can be obtained
MW=86 kg/kmol Δhc=48696 kJ/kg
Tboil=69 oC=342 K hfg=335 kJ/kg
311
ρl=659 kg/m3 Tad=2273 K

For simplicity we will take the thermal properties of gases as that of


air at the average temperature.

Tav  0.5(Ts  T f )  0.5(Ts  Tad )


 0.5(342  2273)
=1307 K
Using table C.1, the thermal properties of air can obtained at 1300 K
cpg=1.189 kJ/kg K kg=0.082 W/m.K

The stoichiometric mass ratio, ν, for C6H14 is

  4.76 * (6  14 / 4) * MWC H / MWair 3 8

 4.76 * (6  14 / 4) * 86 / 28.85
=1 5.17
Thus, the transfer number can be calculated

(hc / )  c pg (T  Ts )
Bo,q 
qi l  h fg
(48696 / 15.17)  1.189(298  342)

0  335
=9.426

The mass burning rate of the droplet is

316
4k g rs
m F  ln( 1  Bo,q )
c pg
4 * 0.082 * 0.0005
 ln( 1  9.426)
1189
=2.876*10-7 kg/s

The dimensionless flame radius, r f / rs , can be calculated as

rf ln[ 1  Bo, q ] ln[ 1  9.426]


 
rs ln[(  1) / ] ln[( 15.17  1) / 15.17]
=36.72

The flame temperature can be evaluated by

qi l  h fg
Tf  [Bo, q  1]  Ts
c pg (1   )
0  335
 [15.17 * 9.426  1]  342
1.189(1  15.17)
=2816 K

To evaluate the droplet lifetime, we calculate the evaporation rate


constant, K

311
8k g
K ln( 1  Bo,q )
l c pg
8 * 0.082
 ln( 1  9.426)
659 *1189
=1.963*10-6 m2/s

Thus, the droplet lifetime, td, is

td  Do2 / K  (0.001) 2 /(1.963 *106 )


=0.509 s

311

You might also like