Вы находитесь на странице: 1из 24

Combustion Theory and Modelling

ISSN: 1364-7830 (Print) 1741-3559 (Online) Journal homepage: https://www.tandfonline.com/loi/tctm20

The modelling of premixed laminar combustion in


a closed vessel

Khizer Saeed & C R Stone

To cite this article: Khizer Saeed & C R Stone (2004) The modelling of premixed laminar
combustion in a closed vessel, Combustion Theory and Modelling, 8:4, 721-743, DOI:
10.1088/1364-7830/8/4/004

To link to this article: https://doi.org/10.1088/1364-7830/8/4/004

Published online: 20 Aug 2006.

Submit your article to this journal

Article views: 116

View related articles

Citing articles: 21 View citing articles

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=tctm20
INSTITUTE OF PHYSICS PUBLISHING COMBUSTION THEORY AND MODELLING
Combust. Theory Modelling 8 (2004) 721–743 PII: S1364-7830(04)69216-3

The modelling of premixed laminar combustion in a


closed vessel
Khizer Saeed1 and C R Stone2,3
1 Department of Mechanical Engineering, Aligarh Muslim University, Aligarh, 202002, India
2 Department of Engineering Sciences, Oxford University, Parks Road, Oxford, OX1 3PJ, UK

E-mail: khizersaeed@yahoomail.com and richard.stone@eng.ox.ac.uk

Received 19 September 2003, in final form 25 June 2004


Published 5 August 2004
Online at stacks.iop.org/CTM/8/721
doi:10.1088/1364-7830/8/4/004

Abstract
Premixed laminar combustion in closed vessels has been widely used for the
determination of laminar burning velocities. A novel multiple burnt gas zone
model has been developed to describe the different aspects of premixed laminar
combustion in a closed spherical vessel. The mixture is divided into burnt and
unburnt gases with the flame front as a discontinuity. Unburnt gases are assumed
to be in a frozen chemical state. However, burnt gases are divided into a number
of burnt gas zones. For the methane–air mixture the model is used to determine
the temperature distribution within the burnt gas, the relationship between the
pressure rise and mass fraction burnt, the variation of the combustion products
with temperature and pressure, the spatial distribution of the gas density, the
influence of elevated initial temperature, pressure and molecular structure. This
computation allows for the variation in heat capacity of the constituents, and
solves the equilibrium combustion equation for the ten major species (N2 , O2 ,
H2 , CO, CO2 , H2 O, O, H, NO, OH). This eliminates some of the simplifications
made by Bradley and Mitcheson (1976 Combust. Flame 26 201–17) or Takeno
and Iijima (1979 7th Int. Colloq. on Gas Dynamics of Explosions and Reactive
Systems (Göttingen, Germany) pp 20–4), such as the constant specific heat for
the burnt and unburnt gases, incorporation of the flame front thickness, use of
the Rankine–Hugoniot relation for finding the burnt gas state.

1. Introduction

The laminar burning velocity has been widely determined from pressure measurements in
closed vessels [1–18]. The advantage of measuring the burning velocity using the closed vessel
method over other methods is that, from a single test, burning velocities can be calculated over
3 Author to whom any correspondence should be addressed.

1364-7830/04/040721+23$30.00 © 2004 IOP Publishing Ltd Printed in the UK 721


722 K Saeed and C R Stone

a wide range of temperatures and pressures. However, two conditions must be fulfilled for
the correct determination of the burning velocities from this method. First, the pressure rise in
the vessel due to combustion must be recorded correctly and, second, the equation used for the
determination of the burning velocities should be valid throughout the range of pressure rise.
Most burning velocity equations for the closed vessel have been found to be valid only in the
initial stages of combustion (generally taken as P  1.1Pi ) [1]. The two most widely quoted
results are probably those of Lewis and von Elbe [2] and of Metghalchi and Keck [3]. Lewis
and von Elbe [2] assumed that the mass fraction burnt is proportional to the fractional pressure
rise. Metghalchi and Keck [3] divided the closed vessel into two zones, burnt and unburnt, but
assumed a uniform temperature and concentration in each zone. They calculated the burning
velocities using the conservation of energy and volume equations. It has been shown recently
by Stone et al [5] that the differences between the results of Metghalchi and Keck [3] and the
assumption of Lewis and von Elbe [4] are not significant for methane–air mixtures. However,
Hopkinson in 1906 established that there is a temperature gradient in the burnt gas due to the
consecutive nature of burning inside the closed vessel. O’Donnovan and Rallis [9] proposed
its incorporation into an equation of the burning velocity for the closed vessel. However,
contradictory results have been found when the temperature gradient is modelled (Bradley
and Mitcheson [19], Ryan and Lestz [16], Metghalchi and Keck [7] and Hill and Hung [7]).
In view of this, a more novel and more rigorous multizone computational analysis has been
undertaken in the present study to analyse the combustion process inside the closed vessel. In
addition to predicting the combustion behaviour, the present multiple burnt gas zones model
has been successfully used for the determination of the laminar burning velocity [1].

2. Literature review

Computational studies, which model the combustion phenomenon inside a closed constant
volume vessel are those by Bradley and Mitcheson [19] and Takeno and Iijima [20]. However,
both aim to predict the rate of pressure rises as opposed to using the models for deducing the
burning velocity from experimental pressure records.
Bradley and Mitcheson [19] were probably the first investigators to model flame
propagation inside a constant volume spherical vessel. The approach taken by them was
that of the conservation of the volume, which is made up of three volumes—unburnt, reacting
and burnt gas. The fuel–air mixture inside the vessel is divided into n equal mass intervals
dmu , and the flame propagation is seen as the consumption of these unburnt mass intervals. For
the nth unburnt mass interval dmu,n , it first moves into the reaction zone with a temperature
Tu,n−1 , burns at the constant pressure Pn−1 , and a proportion of the gas attains the ideal
equilibrium temperature Tf,n , for these conditions. This reaction increases the pressure and in
the model this is assumed to follow the constant pressure combustion and to be isentropic. The
pressure throughout the vessel becomes Pn and the uniform unburnt gas temperatures becomes
Tu,n . At this stage the radius of unburnt gas is rb,n . Using the volume conservation equation,
the total volume of the burnt gas at any instant of time is calculated by summing up the n
intervals. Equilibrium temperatures of the individual mass decrements are computed from the
JANAF thermochemical data. The variation of the burnt gas temperature showed an asymptotic
behaviour at the instant of peak pressure. A difference of nearly 500 K is obtained between
the first burnt mass element and the last, on the completion of the combustion process inside
the vessel. Bradley and Mitcheson [19] concluded that modelling the burnt gas temperature
distribution with a multizone model led to a slightly higher final pressure over a single burnt
gas zone model. This work of Bradley and Mitcheson [19] used constant specific heats for
the calculation of the burnt and unburnt gas properties and flame front thickness. This study
Modelling of premixed laminar combustion 723

through computer simulation showed some insight into the combustion phenomenon inside a
spherical bomb, due to the consecutive nature of combustion. However, the method of analysis
was not used to derive the burning velocity from experimental pressure records.
Takeno and Iijima [20] outlined a model of flame propagation in a closed vessel based on a
quasi-steady flame surface model. In the analysis, the thermodynamic aspects are completely
decoupled from the gas dynamic aspect. In the thermodynamic analysis, the state of the burnt
and unburnt gas temperature and peak pressure is determined. The state of the unburnt gas
is determined as a function of pressure. The state of the burnt gas element is determined, by
first finding its state just behind the flame front using the Rankine–Hugoniot relation. The
subsequent state of this element at any pressure is then determined by making the equilibrium
calculation in which the entropy is kept constant at a value just behind the flame front. The
peak value of the pressure is determined by numerically solving the equation for volume
conservation. The total volume at any instant of time is assumed to consist of the unburnt and
burnt gas. In the gas dynamic analysis, the overall mass conservation equation is solved in
conjunction with the mass continuity equation to obtain a time-dependent differential equation
for the pressure, unburnt gas and burnt gas particle positions. The results obtained by them
include an instantaneous spatial distribution of temperatures, density, velocity, flame trajectory
and particle path lines and a burnt gas temperature profile.

3. Multizone model and assumptions

The modelling here is for a rigid closed spherical vessel of one-dimensional geometry. The
vessel is filled with combustible mixtures. The mixture is ignited at time t = 0 at the centre
of the vessel by the two electrodes. As a result of ignition, a symmetric flame is established
instantaneously, and begins to propagate in the outward direction to reach the walls of the
vessel at the termination of the process. The following assumptions are made in the analysis
of the combustion inside the vessel.

1. The unburnt gas is initially at rest and has a uniform temperature, pressure and composition.
2. The flame front (the visible part of the reaction zone) and a thermal boundary layer (the
preheat zone) are of negligible thickness and the gas within the bomb consists of a burnt
fraction x and an unburnt gas fraction 1 − x.
3. The pressure is assumed to be uniform across the bomb.
4. When the vessel is divided into multiple zones, the zone that is currently being burnt will
contain burnt and unburnt gas, separated by the flame front. The burnt gas in each zone
will be at a different temperature and thus have a different composition, and the unburnt
mass fraction 1 − x is in local thermodynamic equilibrium.
5. The gases are assumed to behave semi-perfectly.
6. There is no heat transfer within the burnt gas volume.

Using the above assumptions a mathematical derivation of the first-order ordinary differential
equations for the pressure and the burnt and unburnt gas temperatures is carried out for the
single burnt gas zone model. The derivation is done by first dividing the mixture into single
burnt and unburnt zones separated by the flame front of zero thickness. The single burnt
zone is then extended to multiple zones, which enables the incorporation of the burnt gas
temperature gradient in the model. At any instant during combustion, the contents of the
closed vessel consist of the burnt and the unburnt gases separated by the flame front of zero
thickness. The equations for the conservation of volume and internal energy are given by
724 K Saeed and C R Stone

equations (1) and (2):


V
v= = xvb (P , Tb ) + (1 − x)vu (P , Su,o ), (1)
M
E
e= = xeb (P , Tb ) + (1 − x)eu (P , Su,o ). (2)
M
Where v is the specific volume, V is the volume of the vessel, M is the mass, E is the
internal energy, x is mass fraction burnt, vu and vb are unburnt and burnt gas specific volumes,
respectively, eu and eb are the unburnt and burnt gas internal specific energies, respectively,
P is the pressure, Su,o entropy of the unburnt gas and Tb is the burnt gas temperature.
The formulation of the equations is based on the approach adopted by Ferguson [21],
the above two equations are solved as a set of ordinary differential equations for the
rate of change of pressure and burnt and unburnt gas temperatures for a single burnt zone.
These differential equations have been extended to the multiple burnt gas zones for the
multizone analysis. The differential equations for the pressure and unburnt and burnt gas
temperatures are given by the following equations:
dP A+B
= , (3)
dt C+D
 
dTu −hĀ(Tu − Tw ) vu ∂ ln vu A+B
= + × , (4)
dt M(1 − x)Cp,u Cp,u ∂ ln T C+D
 
dTb vu ∂ ln vu Tb ∂ ln Tb vu ∂ ln vu ∂ ln vu A+B
= −(1 − x) × × ×
dt Tu ∂ ln Tu xvb ∂ ln vb Cp,u ∂ ln Tu ∂ ln Tu C+D
vu ∂ ln vu Tb ∂ ln Tb hĀ(T u − Tw ) Tb ∂ ln Tb
+ × × −
Tu ∂ ln T xvb ∂ ln vb mCp,u xvb ∂ ln vb
  
vb ∂ ln vb vu ∂ ln vu A+B Tb ∂ ln Tb dx
× x + (1 − x) − (vb − vu ) .
P ∂ ln P P ∂ ln P C+D xvb ∂ ln vb dt
(5)
Where,
vu ∂ ln vu (Tu − Tw )
A= × hĀ , (6)
Cp,u ∂ ln Tu Tu
 
vb dx
B = (vu − vb ) + (hb − hu ) , (7)
Cp,b Tb dt
   
vb2 ∂ ln vb 2 vb ∂ ln vb
C=x + , (8)
Cp,b Tb ∂ ln Tb P ∂ ln P
   
vu2 ∂ ln vu 2 vu ∂ ln vu
D = (1 − x) + . (9)
Cp,u Tu ∂ ln Tu P ∂ ln P

Where h is the heat transfer coefficient, Ā is the inside area of the spherical test vessel, hu is
the specific enthalpy of the burnt gas, hb is the specific enthalpy of the burnt gas, Cp,u is the
constant pressure specific heat of the unburnt gas, Cp,b is the constant pressure specific heat
of the burnt gas, M is the total mass of the mixture, t is the step, Tw is the wall temperature of
the test vessel, Tu is the unburnt gas temperature and Tb is the burnt gas temperature.
Modelling of premixed laminar combustion 725

For the multizone analysis, the above equations are extended to N zones. Now, N + 3
equations need to be solved simultaneously. The extension of the single burnt gas zone to
multiple zones is done in the following manner:

dP A + B u + Bi
=  , (10)
dt Cn + D
  
dTu −hĀ(Tu − Tw ) vu ∂ ln vu A + Bu + Bn
= + ×  , (11)
dt M(1 − x)Cp,u Cp,u ∂ ln Tu Cn + D
  
dTbn vu ∂ ln vu Tbn ∂ ln Tbn vu ∂ ln vu ∂ ln vu A + B u + Bn
= −(1 − x) × × × 
dt Tu ∂ ln Tu xvbn ∂ ln vbn Cp,u ∂ ln Tu ∂ ln Tu Cn + D
vu ∂ ln vu Tbn ∂ ln Tbn hĀ(Tu − Tw ) Tbn ∂ ln Tbn
+ × × −
Tu ∂ ln T xvbn ∂ ln vbn MCp,u xvbn ∂ ln vbn
   
vbn ∂ ln vbn vu ∂ ln vu A + Bu + B n
× x + (1 − x) 
P ∂ ln P P ∂ ln P Cn + D
Tbn ∂ ln Tbn dx
− (vbn − vu ) , (12)
xvbn ∂ ln vbn dt

where (Bn ) = B1 + B2 + B3 + · · · + BN
vu ∂ ln vu (Tu − Tw )
A= × hA(1 − x 0.5 ) , (13)
Cp,u ∂ ln Tu Tu
dx
Bu = (vu − hu ) , (14)
dt
 
vbn dx
Bn = −vbn + hbn , (15)
Cp,bn Tbn dt
   
2
vbn ∂ ln vbn 2 vbn ∂ ln vbn
Cn = x + , (16)
Cp,bn Tbn ∂ ln Tbn P ∂ ln P
   
vu2 ∂ ln vu 2 vu ∂ ln vu
D = (1 − x) + , (17)
Cp,u Tu ∂ ln Tu P ∂ ln P

where n is the zone number, which varies from n = 1, 2, . . . , N number of total zones. To
solve these equations, a BOMB program was evolved based on the work of Raine et al [24] at
the Oxford Engine Group.

4. Burn rate laws

The mass inside the spherical vessel is divided into multiple zones. The zones are divided
so that they consist of equal masses or equal radii. In the equal mass zones model (hereafter
called the EQM model) every zone consists of the same mass, and in the equal radius zones
model (hereafter called the EQR model) every zone has the same increase in radius. The EQM
and EQR models are shown in figures 1 and 2. Each individual zone is then further divided
into a number of elemental shells equal to the number of steps/zones. For example, in each
zone in the 10 zones and 200 step model the 20 elemental shells are consecutively consumed
and the flame propagation is seen as the consecutive consumption of these elemental shells.
726 K Saeed and C R Stone

0.9
0.8

Mass Fraction Burnt (x)


0.7
EQM Model
0.6

0.5
0.4
0.3 EQRModel

0.2
0.1

0
0 20 40 60 80 100 120 140 160 180 200
No of Steps

Figure 1. Behaviour of the mass fraction burned with the number of steps for the EQM and EQR
models distribution.

(a) (b)

ri
ri+dri

EQR model EQM model


(Zone 2 Burning) (Zone 1 Burning)

Figure 2. Radial distribution of the multiple zones inside a spherical vessel. (a) Five zones EQR
model; (b) Five zones EQM model. Hatched portion indicates the elemental shell consumed at
one step.

The burn rate laws adopted in the present study to represent linear (EQM) model and
cubic (EQR) model as represented in figures 1 and 2 are given by the equations (18) and (19),
respectively:
xθ = xθ−1 + x EQM, (18)
xθ = {tx} 3
EQR, (19)
where θ is the step number, which varies from 1, 2, 3, . . . , t number of total steps. x(1/t)
is the elemental mass fraction burnt within the shell within the elemental time of t, xθ is the
mass fraction burnt at the step number θ, where x varies from 0 to 1 and x0 = 0.
At the start of the ignition, the consumption of the very first or the inner-most elemental
shell (θ = 1) inside the 1st or the inner-most zone takes place at the initial temperature of
Modelling of premixed laminar combustion 727

Tu,θ (= Ti ), a constant initial pressure Pu,θ (= Pi ) and an initial radius ri (figure 2). As a result,
a step increase in the temperature of the shell takes place: this value is Tu,θ +1 ; which leads to
the expansion of the shell from Rf,θ to Rf,θ +1 , and the movements of the boundaries of the
zone 1 and the remaining zones in the outward direction, where Rf,θ +1 is the radius of the shell
after the expansion of the mass fraction enclosed within the elemental shell . The expansion
of the shell compresses the remaining shells (or unburnt gases, when considered as a single
entity) and increases the pressure to Pu,θ +1 . Before the consumption of the next shell (θ + 1),
the pressure is assumed to become uniform throughout the vessel at Pu,θ +1 and the shell (θ +1)
is then consumed at pressure Pu,θ +1 , with the temperature of the unburnt gas being Tu,θ +1 . The
expansion of this shell takes place both in the outward and inward directions, which increases
the zones radii in the outward direction but adiabatically recompressing the burnt shell and
hence increasing the temperature inside the shell. When all the shells of zone 1 are completely
consumed, zone 1 then moves into the burnt gas region which is now considered as a single
entity, and instead of taking the adiabatic recompression of the individual shells, the adiabatic
recompression of the whole zone and the variation in its mean bulk temperature due to the
expansion of the shells inside the subsequent zones is modelled. This establishes a temperature
and density gradient inside the burnt gas region zones. There is no difference in the zone and
shells when the numbers of zones are taken to be same as the number of shells. However, we
have found that modelling the whole burnt gas zone instead of the individual shells leads to
satisfactory results.
At every step, the flame front radius and burnt gas volume is determined using the following
equations:
   1/3
Pu,θ −1 1/γu,θ −1
Rf,θ = R 1 − (1 − x) , (20)

   
Pu,θ −1 1/γu,θ −1
Vb,θ = V 1 − (1 − x) , (21)

where V is the volume of the vessel, R is the radius of the spherical vessel and Rf is the flame
radius; for θ = 1, Pu,θ −1 = Pi and Tu,θ −1 = Ti .
The volume of the individual zones can be traced using similar equations.
At the time of the consumption of the nth zone, i.e. when the elemental shell θ is within
the nth zone:
   
Pu,θ −1 1/γu,θ −1
Vn,θ = V 1 − (1 − x) + ωV , (22)
Pt
where Vn,θ is the volume of the zone n when shell θ inside it is consumed.
At every step the BOMB program gives as an output the unburnt gas volume, which is
divided according to the burn rate law to obtain V at every step. Where ω = {n∗ (t/N )−θ } is
the number of remaining elemental shells inside the nth (n = 1, 2, 3, . . . , N number of zones)
zone when the θth elemental shell is burning inside it.
At the time when the flame front has completely consumed the zone, i.e. when the zone
is within the burnt gas region:
 
R̄b,θ Tb,n,θ
Vb,i,θ = mb,n , (23)

where Vb,i,θ is the volume of the burnt zone i = n − 1, n − 2, . . ., at the time when step θ is
consumed inside the nth zone in equation (22); and Pθ , Vb,i , Rb,i , Tb,i and mb,i are the pressure,
burnt gas volume, gas constant, temperature and mass of the nth burnt gas zone, respectively.
728 K Saeed and C R Stone

The BOMB program uses the Toronto University IVODE [25] differential equation solver
for the solution of the ODE. The input and output of the BOMB program are given in the
appendix. Since the burnt gas in each zone will be at a different temperature, they will thus
have a different composition. At each temperature the composition of the burnt gases is solved
for the equilibrium of the following combustion products: CO, CO2 , H2 O, H2 , H, OH, O,
N2 , NO and O2 . The equilibrium combustion calculations are solved by the minimization of
the Gibbs free energy, using a routine written by Ferguson [21]. The gases are assumed to
behave as semi-perfect gases, and the internal energy, entropy and specific heat capacities are
calculated from polynomial functions that describe the specific heat capacity variation with
temperature. As the specific heat capacity variation with temperature has a ‘knee’ between 900
and 2000 K, a single polynomial is never likely to give a satisfactory description. Instead, two
polynomials can be used which give identical values of specific heat, at the transition between
the lower range and the higher range. The data used here are from Gordon and McBride [22],
who used a transition temperature of 1000 K. At every step, the BOMB program outputs the
x, P , Tu , vu , ri , Rf , Tb,i , vb , Vb,i , XCO , XCO2 , XH2 O , XH2 , XH , XOH , XO , XN2 , XNO and XO2
parameters as shown in the appendix.

5. Model results

A parametric study was made for the flame front propagation and zone size, final radial
temperature distribution across the vessel, the temperature and pressure variation inside the
vessel, the influence of the burnt gas temperature gradient on pressure, errors in the mass
fraction and burning velocity at the initial elevated temperature and pressure, and the effects
of equivalence ratio on pressure and temperature.

6. Final temperature distribution

At the instant of the end of combustion in a spherical vessel, the radial temperature distributions
for the EQR and EQM models are shown in figure 3. It is seen from the figures that with the
increase in number of zones the shape of the temperature profile becomes the same. It might be
thought that a very large number of zones would give the most accurate prediction. However,
only very slight differences were found in the burnt temperature distribution profile when the
value of the number of zones was increased from 20 zones to 50 zones. This is shown in
table 1. Therefore, a finite number of zones can give a sufficiently accurate analysis of the
phenomenon.
The effect of step size on the end pressure for the fifty-zone model is presented in table 2.
It can be seen from the table that with an increase of steps from 200 to 1000 a very small change
in the end pressure takes place. To get sufficiently accurate results and to reduce the run time
of the simulation, it was decided to use 10 zones and 200 steps for all further investigations.
In both the EQR and EQM models, temperatures are highest at the centre of the vessel
and decrease towards the wall of the vessel. The temperature difference obtained between the
first and the last zone for the 20-zone case is more than 450 K for both the EQM and EQR
models. With an increase in the number of zones, the temperature profile becomes smooth.
A smaller number of zones underpredict the final temperature distribution range in the vessel.
A comparison of a 20-zone model with a 2-zone model for the final temperature distribution
shows a difference of nearly 90 K between the first-burnt zones, and a difference of 110 K in the
last-burnt zones. A comparison of the temperature profile obtained in the present study (using
the model with 10 zones and 200 steps) with that obtained by Bradley and Mitcheson [19] and
Modelling of premixed laminar combustion 729

3000
2 Zones
2900 5 Zones
10 Zones
Final Temperature (K)

2800
20 Zones

2700

2600
EQR Model
2500

2400

2300
0 20 40 60 80 100
Percentage Of Vessel Radius

3000
2 Zones
2900 5 Zones
10 Zones
Final Temperature (K)

2800
20 Zones
2700

2600

2500 EQM Model

2400

2300
1 21 41 61 81
Percentage Radius Of Vessel

Figure 3. Variation of 10-zones final temperatures with the zones in (a) EQR and (b) EQM model
at the initial conditions of 1 bar and 298.15 K for stoichiometric methane–air mixture.

Table 1. Predicted end pressure (bar) for the different numbers of zones for a stoichiometric
methane–air mixture at initial conditions of 300 K and 1 bar.
Number of zones 1 2 3 10 20 50
EQM model 8.798 8.782 8.768 8.761 8.751 8.723
EQR model 8.800 8.779 8.769 8.761 8.750 8.729

Takeno and Iijima [20] is shown in figure 4. It can be seen from figure 4 that the shape of the
temperature profile with the change in percent of radius is the same as that obtained by Bradley
and Mitcheson [19] and Takeno and Iijima [20]. However, it was found that the temperature
profile obtained by these investigators gives a relatively higher temperature in comparison to
that obtained in the present study. This could be due to the errors associated with the use of
burn rate laws and the methods of analysis used by these authors.
In the present study, the EQM zone and EQR zones model have been used and were found
to give very close agreement for the final temperature in the first zone at the end of combustion
730 K Saeed and C R Stone

Table 2. Predicted end pressure with different numbers of steps with 50-zones, for a stoichiometric
methane–air mixtures at initial conditions of 300 K and 1 bar.
Number of End pressure (Bar) End pressure (Bar)
steps EQM model EQR model
100 8.419 8.355
200 8.729 8.723
500 8.691 8.655
1000 8.652 8.652

Table 3. Comparison of the end pressure zones for stoichiometric methane–air mixtures, with
initial conditions of 300 K and 1 bar.
Bradley and Takeno and
Mitcheson (1976) Iijima (1979) Present study
End pressure (Bar) 8.90 8.88 8.76

3000
Present Work
2900 Bradley & Mitcheson [19]
Takeno & Iijima [20]
2800
Final Temperature (K)

2700

2600

2500

2400

2300
0 0.2 0.4 0.6 0.8 1 1.2
Percentage Radius Of Vessel

Figure 4. Comparison of the temperature profiles at the end of combustion for a stoichiometric
methane–air mixture initially at 298 K and 1 bar with that of Bradley and Mitcheson [19] and
Takeno and Iijima [20] (initial conditions are 290 K and 1 atm).

with the temperature predicted by STANJAN [22]. Therefore, these models can be used for
the calculation of the burning velocities inside a closed vessel.

7. Zone radius and density

Figure 5 gives a greater physical insight to combustion inside the bomb with respect to how the
radius of each zone varies during the propagation of the flame front. As the flame front travels
through the combustion bomb, each zone burns in succession. It can be seen from figure 5 that
as the flame front travels through each zone an increase in the radius of that zone takes place.
This is due to the expansion of the products of combustion immediately behind the flame front.
And when the flame front has passed the zone, which is indicated by the maximum radius of a
Modelling of premixed laminar combustion 731

80

70

60

Zone Radius (mm)


50

40

30

20 Flame Front

10

0
0 0.2 0.4 0.6 0.8 1
Mass Fraction Burnt

Figure 5. Particle path a spherical vessel for a 10-zone model with the initial conditions of 1 bar,
298.15 K and stoichiometric methane–air mixture for EQM model.

0.0013
10
9*
9 7*
5*
0.0011 3*
8 2*
1*
7
0.0009
Density (g/cm3)

f 5

0.0007 4

3
0.0005
2

0.0003 1

0.0001
0 0.2 0.4 0.6 0.8 1
Mass Fraction Burnt

Figure 6. Density variation of the burnt gas with the mass fraction burnt for a stoichiometric
methane–air mixture with initial conditions of 1 bar and 298 K.
(This figure is in colour only in the electronic version)

particular zone, recompression of the zones takes place due to the expansion of the next zone.
Due to the recompression, a decrease in the radii and hence the volume of the burnt zones takes
place. This results in an increase in the temperature of the burnt gas zones, which establishes
the radial burnt gas temperature gradient. No change in the radius of zone 10 takes place as it
represents the wall of the vessel.
Figure 6 shows the density variation of the burnt gas zones for a spherical vessel test
run. The bold line indicates the variation in ‘flame front’ density (the density immediately
behind the flame front, ρf ). It can be seen in figure 6 that the flame front density increases
732 K Saeed and C R Stone

1.0E-01
3.5 bars
0.5 bars

1.0E-02 O2
OH
NO O2

Mole Fraction
OH NO
H
1.0E-03 O

H
O

1.0E-04

1.0E-05
0 1 2 3 4 5 6 7 8 9 10 11
Pressure (bars)

1.0E+00
N2

H2O

1.0E-01 CO2
Mole Fraction

0.5bars
CO 3.5bars
CO
1.0E-02
H2
H2

1.0E-03
0 1 2 3 4 5 6 7 8 9 10 11 12
Pressure (bars)

Figure 7. Variation of the combustion products composition with pressure and temperature for an
stoichiometric methane–air mixture with an initial temperature of 298 K for EQM model.

with a simultaneous increase in the pressure and temperature of the unburnt gas. Takeno and
Iijima [20] had applied this flame density in their model for the calculation of the burning
velocity. However, the density of the burnt gas is significantly different from the flame front
density due to the recompression of the burnt gas. This is seen in figure 5, which shows
how the recompression of the burnt gas zones causes an increase in the zone density. The
maximum change in density due to recompression is in zone 1, as it undergoes the maximum
recompression from 1 to 1∗ . Similarly, zone 2 undergoes a density variation from state 2 to 2∗ ,
zone 3 from 3 to 3∗ and so on. No density variation is observed in zone 10. Most investigators
have either used the average density [3–7, 9, 15] or the flame front density (8, 20) in their model
for the determination of the burning velocity. It can be seen from figure 6 that a density gradient
is established in the burnt gas region due to the subsequent expansion of the unburnt gas.

8. Combustion products

Figure 7 shows the variation of the burnt gas products with pressure. Pressure and temperature
are coupled in closed vessel combustion. It can be seen from the figure that the mole fractions
Modelling of premixed laminar combustion 733

3000
1*
2*
3*
4"
5*
2500 6*
8*
9*10*

5' 6' 7' 8' 9' 10'


2' 3' 4'
1'

2000
Temperature (K)

1500

1000

9 10
8
500 5 6 7
3 4
2
1

0
0 0.2 0.4 0.6 0.8 1
Mass FractIon Burnt

Figure 8. Temperature variation for the 10 zones with mass fraction burnt using the EQM model
for a stoichiometric methane–air mixture with initial conditions of 1 bar and 298.15 K.

of H2 O and CO2 decrease with the simultaneous increase in temperature and pressure, while
those of other species increase. This could be due to the different dissociation rates of the
species at varying temperature and pressure conditions.

9. Temperature and pressure variation inside the vessel

Figure 8 shows the variation of the temperature of the unburnt gas and burnt gas zones. As
combustion starts, the temperature in zone 1 increases from an initial temperature and pressure
of 298.15 K and 1 bar, and as a result zone 1 expands. This expansion compresses and hence
increases the temperature of the remaining unburnt gas. The flame front starts consuming
zone 2 at a temperature higher than the unburnt gas in zone 1. As the flame consumes zone 2,
its temperature jumps from the position 2 to 2 . The recompression of zone 1 and zone 2 to the
higher temperatures 1∗ and 2∗ , respectively, takes place due to the expansion of the subsequent
zones. The curve connecting points 1–10 indicates the adiabatic compression of the unburnt
gas. The curve connecting points 1 –10 indicates the variation of the flame temperature of the
various zones immediately after each has started to burn. From the curve joining 1 –10 it is
seen that the flame temperature increases with the increase in pressure. However, it can also
be seen that the slope of the curve joining 1 –10 is smaller then that of the curve joining 1–10,
indicating that the combustion temperature rise decreases with the increase in pressure. This is
734 K Saeed and C R Stone

–1

Figure 9. Pressure against specific volume histories across the vessel for the 5-zones equal radii
(EQR) model.

because of increased dissociation at higher temperatures. However, the slopes of the curves
joining 1 –1∗ , 2 –2∗ , 3 –3∗ etc are greater than those of the curve joining 1–11, indicating that
a greater temperature rise takes place in the zones due to recompression in the burnt gas region
than due to compression of the unburnt gas. It can be seen that for zone 10, the temperature rise
due to compression in the unburnt gas is about 80 K, while the rise due to the recompression
in the burnt gas region of the zone is about 450 K in the EQR model. However, the slopes
of all the zones are not the same. Curve 1 –1∗ has the highest slope and the curve 10 –10∗ ,
the lowest. Also, the differences 1∗ –2∗ > 2∗ –3∗ > 3∗ –4∗ > 4∗ –5∗ etc, indicate that the first
zone attains the highest temperature due to the largest recompression and the last zone is at
the lowest temperature as no recompression of it takes place.
The cause of the radial and temperature variation of the zones in the burnt gas region
(as obtained in figures 3 and 8 ) can also be explained with the help of the pressure versus
specific volume history for each zone. This is shown in figure 9. The first zone burns and
expands from a to b at an almost constant pressure. As a result of its expansion, zone 1 does
some work on the remaining unburnt gas. This compresses the unburnt gas between a and c.
Therefore, the combustion in zone 2 starts at a higher pressure i.e. at point c. When zone 2
expands between c and c , it does more work than zone 1. This results in the recompression
of zone 1 from b to c∗ , and that of the remaining unburnt gas from c to d. Therefore, zone 3
burns at a pressure higher than that of zone 2, and does more work than zone 2 when it is made
to expand from d to d  . Zone 3 recompresses zone 1 between c and d  , zone 2 between c and
c and the remaining unburnt gas between d and e. This process continues to repeat itself until
the last but one zone is burnt. When the very last element of gas is burnt, it is almost at the
maximum pressure reached so when it expands, the amount of work done is much higher than
the previously burnt zones. Since, no recompression of the last zone takes place, it results in
a net loss of energy. Hence, the process leaves it at a lower temperature once combustion has
occurred, in comparison to the first zone. The recompression work done on the first zone is
larger than the expansion work done by it, this results in the increase in its energy and leaves
it at the highest temperature.
Modelling of premixed laminar combustion 735

Figure 10. Variation of the pressure with the mass fraction burnt for the 10-zones model for the
stoichiometric methane–air mixture with the initial conditions of 1 bar and 298.15 K for the EQM
model.

Figure 10 shows the variation of the pressure with the mass fraction burnt. A comparison
of this variation with that of the Lewis and von Elbe [2] expression is also shown. It can be
seen from figure 10 that the pressure variation with the mass fraction obtained in this study
is non-linear, and this is contrary to the assumed linear variation of Lewis and von Elbe [3].
Takeno and Iijima [20] also obtained a linear variation of the mass fraction burnt compared
to the pressure rise. However, from the present study it can be seen that throughout the
process there is an overestimation of the pressure if a linear relationship is used, and the error
is a maximum during the central part of combustion. Figure 11 shows the percentage of error
in pressure with the varying mass fraction burnt. It can be seen from figure 11 that the error is
at a maximum between mass fractions of 0.1 and 0.2, burnt for all the equivalence ratios. The
maximum error is found to be for stoichiometric combustion (nearly 5%). It can be seen from
figure 11 that the error is a minimum for the equivalence ratio 0.6. However, the error for the
equivalence ratio 1.2 is greater than for 0.8, and the error for equivalence ratio 1.4 is greater
than for 0.6.

10. Influence of the burnt gas temperature gradient on pressure

In figure 3, it was seen that in the burnt gas a temperature gradient exists in a spherical vessel
combustion, a fact established by Hopkinson in 1906. In the present analysis, a temperature
difference of more than 450 K is found to exist at the end of combustion for the stoichiometric
methane–air mixture. Also, a considerable temperature difference is found to exist between
the predictions made by a single zone model and the multiple zones model. Figure 12 shows
the variation of the final temperature in the first- and last-burnt zones for the 10-zone model.
It is seen that the difference is around 500 K.These differences can become higher for the fuels
that undergo large mole changes during combustion or if they have higher adiabatic flame
temperatures.
736 K Saeed and C R Stone

Figure 11. The error in pressure when a linear relation is assumed between the pressure rise and
the mass fraction burned for the 10-zone EQM model.

Bradley and Mitcheson [19], Metghalchi and Keck [3], Ryan and Lestz [16] and Hill and
Hung [7] have reported contradictory results when the temperature gradient is modelled to study
its effect on the pressure and hence on burning velocity. With the present multizone modelling,
this effect can be easily studied and the effect of the burnt gas temperature gradient on the
pressure can be decoupled. For methane, the variation of end pressure with the number of
zones was shown in table 1. It can be seen from the table that with the increase in number
of zones from 1 to 30, a reduction in the final pressure takes place. Results obtained in
the present study show that the burnt gas temperature profile has definite effects on the end
pressure for methane, and the studies of this effect on the higher hydrocarbon fuel are shown
in figure 13. It can be seen from the figure that the percentage of maximum error in pressure
is methane > benzene > propane > isoctane, when a linear relation is assumed between the
pressure rise and the mass fraction that is burnt.
Modelling of premixed laminar combustion 737

Figure 12. Variation in the difference in the final temperatures between the first and the last zone
with the equivalence ratio in the 10-zones model with initial conditions of 298 K and 1 bar for
methane–air for the EQR model.

Figure 13. Effect of molecular structure on the error in pressure when a linear relation is assumed
between the pressure rise and the mass fraction burnt for the initial conditions 298 K and Pi = 1 bar
for a stoichiometric mixtures in the EQM model.

11. The effect of varying equivalence ratio

Variations of equivalence ratio have a profound effect on the properties of the burnt and
unburnt mixtures. Varying the equivalence ratio gives a further insight into combustion
inside the spherical vessel. Using the multizone model, figure 14(a) shows the effects of
the equivalence ratio on the end pressure. It is seen that the pressure curve peaks on the rich
side of stoichiometric in the range of the equivalence ratios 1 to 1.1. This can be related to
a maximum in the adiabatic flame temperature. Since, for most hydrocarbon fuels the flame
temperature peaks on the rich side of the mixture, as in the rich mixture system there is greater
oxygen utilization, and the mean specific heat of the products is lower, owing to the formation
of more diatomic molecules in comparison to the triatomic molecules in the fuel lean side.
738 K Saeed and C R Stone

(a)

(b)

Figure 14. Variation in the (a) end pressure and (b) final temperature with equivalence ratio for
the 10-zones EQR model at initial conditions of 298 K and Pi = 1 bar, methane–air.

This is demonstrated in figure 14(b), which shows that at φ = 1.2 attains a higher temperature
than that at φ = 0.8, although both are equidistant from being stoichiometric. The adiabatic
flame temperature for a methane–air mixture with an equivalence ratio of 1.2 will be higher
than when the equivalence ratio is 0.8. However, with the weaker mixture the ratio of heat
capacities (γ ) will be higher. Thus, when the inner zones are compressed by the combustion
of the subsequent zones, then with a weaker mixture the early zones will be compressed to a
higher temperature.
In other words, higher pressure and temperature conditions are observed on the fuel-rich
side of stoichiometric than on the fuel-lean side

12. Burning velocity

The multizone model has been implemented for the determination of the laminar burning
velocity from the spherical vessel pressure history [1, 26]. A 10-zone EQM model was
Modelling of premixed laminar combustion 739

–1

Figure 15. Laminar burning velocity as determined from the 10-zones EQM model and Lewis and
Von Elbe [2] for the stoichiometric methane–air mixture at the initial conditions of 1 bar and 298 K
for methane–air.
(This figure is in colour only in the electronic version)

selected, in which the burnt gas is divided into zones of equal mass. A post-processing
computer program BURNVEL was written to determine the burning velocity and other
properties of the mixture at every time step. In order to determine the burning velocities
and other properties of the fuel–air mixture, the BURNVEL program takes the output files of
the BOMB program and files from the experiments. A centrally ignited spherical test vessel of
16 cm diameter was used for obtaining the pressure time test data. A detailed description
of the experimental test facility used for obtaining the pressure time data is provided in
[1, 26]. The appendix gives the flow chart for the determination of the burning velocity using
BOMB and BURNVEL programs. The burning velocity is calculated using the following
expression:
   2  1/γu
dp dri ri Pi
Su = , (24)
dt dp rb P

where the first term on the right-hand side, dp/dt, is calculated from the experimental data
file. The second term, dri /dp, is calculated from the BOMB program. Where,

ri = x 1/3 R. (25)

It has been found in the present study that there is a non-linear relationship between the mass
fraction burnt and pressure rise, which contrasts with the linear relationship assumed by the
Lewis and von Elbe [2]. As seen in figure 11 the error in pressure due to this is up to about 6%
and the error is a maximum for stoichiometric conditions. Also, Clarke [6] had established
that if the end pressure as calculated from the STANJAN [23] is used in the Lewis and von
Elbe [2] method then the Metghalchi and Keck [3] single burnt gas zone method and Lewis and
Von Elbe [2] methods give similar burning velocities. Figure 15 shows the burning velocity, as
determined from the multizone model of 10 zones and 200 steps and from the Lewis and Von
Elbe method [2], for a stoichiometric methane–air mixture at the initial conditions of 298 K and
1 bar. It can be seen from the figure that the burning velocities from the Lewis and Von Elbe [2]
740 K Saeed and C R Stone

method are lower than those given by the present multizone model. The difference in the
burning velocities with the simultaneous increase of pressure and temperature is found to be
around 5–6%.
During the initial stages of combustion, the flame front has high values of flame stretch due
to the large curvature that is associated with a small spherical surface. However, in a spherical
vessel the radius of curvature increases and the flame stretch approaches zero asymptotically,
and the ideal case for the flame front is approached. The flame stretch effect is ignored in the
present study and the justification for this is provided by Stone et al [5].
In practice, the flame front (the visible part of the reaction zone and preheat zone) that
extends into the unburnt gases, is treated as a discontinuity across which the change from
unburnt to the burnt gas takes place, Matalon and Matkowsky [27]. Most investigators have
considered this discontinuity as a shell of negligible thickness, e.g. Hill and Hung [7] and Stone
et al [5]. However, some investigators have tried to apply a correction due to the flame front
thickness. Rallis and Garforth [11] argued that the mass of gas contained in the spherical flame
front, particularly at the early stages of combustion can account for a significant proportion
of the mass of burnt and unburnt gases in it. They supported this argument with the work of
Andrews and Bradley [28], who used the flame front thickness correction in the calculation
of the burning velocity. Andrews and Bradley [28] found that the thin flame equation when
used for the calculation of the burning velocity is in error by as much as 22% for a flame
radius of 25 mm at Pi = 1 atm and Ti = 300 K. They reported that Dixon-Lewis [29] and
Janisch [30] have measured the profile of a stoichiometric methane–air flame at 1 atm and
obtained a flame thickness of 0.75 mm and 1.1 mm, respectively. Garforth and Rallis [11],
however, showed that the correction for methane–air mixtures was not as large as suggested by
Andrews and Bradley [28]. They argued that the flame front thickness is inversely proportional
to the pressure. Thus, with the increase in pressure, the flame front thickness decreases further.
Metghalchi and Keck [4] calculated the correction due to the preheat layer ahead of the reaction
front and found it to be substantially less than 1%. The justification for treating the flame front
as a thin shell of negligible thickness is also discussed by Hill and Hung [7], who found that
the volume occupied by the flame front in the spherical vessel combustion is of the order of
0.1% of the chamber volume. Thus, the flame itself can be satisfactorily represented by a
thin discontinuity. More recent work by Göttgens et al [31] shows that for equivalence ratios
richer than 0.6 and for pressures above 1 bar, the preheat zone is always less than 0.6 mm
and its radius reduces rapidly with increasing equivalence ratio and pressure, and the reaction
zone is of similar thickness to that of the preheat zone. Also, the flame front is taken to be of
negligible thickness because results presented by Stone et al [5] using this assumption show
that the burning velocity data of methane–air mixtures plotted at 1 bar are in agreement with
measurements from two alternative methods.
The pressure distribution through a flame front is considered as uniform. Recently, Hill and
Hung [7] have observed that since all particle velocities are very much lower than the speed of
sound, and any pressure difference associated with particle acceleration is very small compared
to the pressure rise due to combustion. The assumption of spatial uniformity of pressure is
therefore reasonable for the laminar burning velocity and is not expected to introduce any
appreciable errors.
The assumption is made for both the burnt and the unburnt gas, that there is no heat loss or
gain from these regions. However, Rallis and Garforth [11] observed that transfer of heat may
occur during the combustion process by radiation from the burnt gas to the unburnt gas and to
the containing walls; and by conduction from the unburnt gas to the wall; and by conduction
along the spark electrodes. Experimental investigations have been carried out to quantify these
losses by numerous investigators such as Metghalchi and Keck [4], Rhodes and Keck [15] and
Modelling of premixed laminar combustion 741

Clarke et al [6]. Metghalchi and Keck [4] examined the effects of the heat losses to the walls
of the spherical vessel from the steadily compressing unburned gas, heat losses to the electrode
from the burned gases, and the radiative heat losses due from the burnt gases. They found that
corrections for these losses in no case exceeded 1% and in most cases and were substantially
less than 1%. Rhodes and Keck [15] and Clarke et al [6] inserted two dummy electrodes in the
spherical vessel and for the methane air mixtures at Pi = 1 atm and Ti = 298.14 K there was no
observable difference in the combustion with and without the dummy electrodes. Hill and Hung
[7] have examined the radiative heat losses from the burnt gas region of methane–air mixtures
at stoichiometric conditions and found that radiation emitted from the burnt gas was mostly
absorbed by the chamber walls. They, proposed that it is therefore, reasonable to assume that
the bulk of the unburnt gas is compressed adiabatically and the energy radiated from the unburnt
gas region will be of the order of 1% of the heat of combustion. Clarke [32] has examined the
adiabatic compression of the unburnt gas during the combustion process inside the spherical
vessel by evaluating the heat transfer to the vessel walls using a surface mounted, thin film
platinum resistance thermometer. From the temperature–time history the heat flux into the wall
was calculated. For the stoichiometric methane–air mixtures at Pi = 1 bar and Ti = 298.14 K,
he found the total errors in the pressure measurement due to the losses from the heat transfer
to be less than 0.01%. Clarke [32] concluded that the compression process of the unburnt
gases during the combustion inside the spherical vessel could therefore safely be treated as
adiabatic.

13. Conclusions

A novel multizone model has been used to study the combustion process inside a closed vessel.
It is observed from the present study that:

1. A 10-zone and 200-step model gives an accurate description of combustion inside the
spherical vessel.
2. A temperature difference exists between the first- and the last-burnt gas zones. For methane
and air fuel mixture this difference is as large as 500 K.
3. Earlier studies of Bradley and Mitcheson [18] and Takeno and Iijima [19] over-predict the
end pressure and temperature distribution.
4. The relationship between the pressure rise with the mass fraction burnt is non-linear. The
errors associated with the linear assumption peak at 0.1–0.2 mass fractions burnt for any
equivalence ratio burnt. The maximum error is found to be at stoichiometric.
5. The error decreases with an increase in molecular structure.
6. The burnt gas temperature profile has an effect on the end pressure for methane and the
error due to it can be as large as 5%.
7. A multiple burnt gas zone model has been successfully implemented for the measurement
of laminar burning velocities.
8. At any temperature and pressure conditions, burning velocity measured from the Lewis
and von Elbe [2] method is found to be lower than burning velocities measured from the
multiple zones method.
9. The difference in burning velocities as measured from the spherical vessel using multiple
burnt gas zone model of the present study and Lewis and von Elbe [2] method is found to
be around 5–6%.
742 K Saeed and C R Stone

Appendix

Flow chart for the determination of the burning velocity

References

[1] Saeed K 2002 Laminar burning velocity measurements DPhil Thesis Oxford University
[2] Lewis B and von Elbe G 1934 Determination of the speed of flames and the temperature distribution in a spherical
bomb from time–pressure explosion records J. Chem. Phys. 2 283–90
[3] Metghalchi M and Keck J C 1980 Laminar burning velocity of propane–air mixtures at high temperature and
pressure Combust. Flame 38 143–54
[4] Metghalchi M and Keck J C 1982 Burning velocities of mixtures of air with methanol, isooctane, and indolene
at high pressure and temperature Combust. Flame 48 191–210
[5] Stone R, Clarke A and Beckwith P 1998 Correlations for the laminar burning velocity of methane/diluent/air
mixtures obtained in free-fall experiments Combust. Flame 114 546–55
[6] Clarke A, Stone C R and Beckwith P 1995 Correlations for the laminar burning velocity of methane/diluent/air
mixtures obtained in free-fall experiments J. Inst. Energy 68 130–6
Modelling of premixed laminar combustion 743

[7] Hill P G and Hung J 1988 Laminar burning velocities of stoichiometric mixtures of methane with propane and
ethane additives Combust. Sci. Technol. 60 7–30
[8] Iijima T and Takeno T 1986 Effects of temperature and pressure on burning velocity Combust. Flame 65 35–43
[9] O’Donovan K H and Rallis C J 1959 A modified analysis for the determination of the burning velocity of a gas
mixture in a spherical constant volume combustion vessel Combust. Flame 3 201–14
[10] Garforth A M and Rallis C J 1978 Laminar burning velocity of stoichiometric methane–air: pressure and
temperature dependence Combust. Flame 31 53–68
[11] Rallis C J and Garforth A M 1980 The determination of laminar burning velocity Prog. Energy Combust. Sci. 6
303–29
[12] Eschenbach R C and Agnew J T 1958 Use of the bomb technique for measuring burning velocity Combust.
Flame 2 273–85
[13] Babkin V S and Kozachenko Y G 1966 Equations for determining normal flame velocity in a constant-volume
spherical bomb Fiz. Goreniya Vzryva 3 268–75
[14] Babkin V S and Kozachenko Y G 1969 Analysis of equations for determining the normal burning velociti by
the constant volume bomb method Fiz. Goreniya Vzryva 5 84–93
[15] Rhodes D B and Keck J C 1985 Laminar burning speed measurements of indolene-air-diluent mixtures at high
pressures and temperatures SAE paper 850047
[16] Ryan T W and Lestz S S 1980 The laminar burning velocity of iso-octane, n-heptane, methanol, methane and
propane at elevated temperature and pressures in the presence of a diluent SAE paper 800103
[17] Lavoie GA 1979 Correlations of combustion data for S.I engine calculation-laminar flame speed, quenching
distance and global reaction rates SAE paper 780229 pp 1015–33
[18] Milton B E and Keck J C 1984 Laminar burning velocities in stoichiometric hydrogen and hydrogen–hydrocarbon
gas mixtures Combust. Flame 58 13–22
[19] Bradley D and Mitcheson A 1976 Mathematical solutions for explosions in spherical vessels Combust. Flame
26 201–17
[20] Takeno T and Iijima T 1979 Theoretical study of non-steady flame propagation in closed vessels 7th Int. Colloq.
on Gas Dynamics of Explosions and Reactive System (Göttingen, Germany) pp 20–4
[21] Ferguson C R Internal Combustion Engines (New York: Wiley)
[22] Gordon S and McBride B 1971 Computer program for calculation of complex chemical equilibrium composition,
rocket performance, indirect and reflected shocks and chapman–jouquiit detonation Report NASA SP-273
[23] STANJAN 1987 Chemical Equilibrium Solver Stanford University
[24] Raine R R, Stone C R and Gould J 1995 Modelling of NO formation in spark ignition engines with a multizone
burned gas Combust. Flame 102 241–55
[25] IVODE 1987 Differential Equation Solver University of Toronto
[26] Saeed K and Stone C R 2004 Determination of the methanol air laminar burning velocity in a constant volume
vessel using multizone model Combust. Flame at press
[27] Matalon M and Matkowsky B J 1982 Flames as gasdynamic discontinuities J. Fluid Mech. 124 239–59
[28] Andrews G E and Bradley D 1972 The burning velocity of methane–air mixtures Combust. Flame 19 275–88
[29] Dixon-Lewis G and Wilson M J G 1951 A method for the measurement of the temperature distribution in the
inner cone of the Bunsen flame Trans. Faraday Soc. 47 1106–14
[30] Janisch G 1971 Chem. Ing. Technol. 43 561
[31] Göttgens J, Mauss F and Peters N 1992 Analytical approximation of burning velocities and flame thickness of
lean hydrogen, methane, ethylene, ethane, acetylene and propane flames 24th Symp. (Int.) on Combustion
(Pittsburg, PA: The Combustion Institute) pp 129–35
[32] Clarke A 1994 Laminar burning velocity measurements for methane air diluent mixtures obtained in free fall
experiments DPhil Thesis Oxford University

Вам также может понравиться