Download as pdf or txt
Download as pdf or txt
You are on page 1of 47

pubs.acs.

org/EF Review

Flammability Limits: A Comprehensive Review of Theory,


Experiments, and Estimation Methods
Andrés Z. Mendiburu,* João A. Carvalho Jr., and Yiguang Ju
Cite This: https://doi.org/10.1021/acs.energyfuels.2c03598 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: Flammability limits play an important role in combustion research,


industrial applications, and fire safety. This article provides a comprehensive review of
Downloaded via UNIV ESTADUAL PAULISTA on March 14, 2023 at 02:56:37 (UTC).

recent developments in the fundamental understanding of flammability limits and their


experimental determination as well as estimation methods for pure fuels and fuel mixtures.
The article begins with a discussion of the importance and challenges of determining
flammability limits. It then presents the theoretical, computational, and experimental
methods available to understand the mechanism of flammability limits and to quantify
them. The experimental setups using cylindrical and spherical vessels to determine the
flammability limits are discussed. The effects of buoyancy, thermal radiation, and flame
stretch are examined. The relationship between the fundamental flammability limits and
the extinction limits of stretched flames via strain and radiation is presented. The effects of
initial temperature, pressure, mixtures of different fuels, and diluents are examined, and
available estimation methods are presented. Finally, the flammability limits of renewable
and alternative fuels are addressed and strategies for estimating the flammability limits of
these fuels are presented.

1. INTRODUCTION temperature and pressure conditions. Similarly, the UFL is


The purpose of this review article is to present the state of the art defined as the richest composition of the fuel−air mixture above
in the study of flammability limits of premixed flammable gas which a flame fails to propagate under certain temperature and
mixtures. Flammability limits are the concentration limits pressure conditions. Therefore, the FLs are important for the
beyond which a flame fails to propagate. The flammable application of safety measures in industry to prevent the
mixtures studied in this review are formed by homogeneous formation of flammable mixtures in locations where an ignition
premixing of a gaseous fuel and an oxidizer. In most cases, the source may be present. The formation of flammable mixtures
oxidant is air (actually, it is the O2 present in air), but other can occur in different ways and in different applications.
oxidants can also be used to determine the flammability limits of As aforementioned, the FLs of a mixture are temperature and
a mixture. pressure dependent. Thus, if one of these parameters is varied
In the presence of an ignition source, a planar and unstretched
while the other is held constant, the values of the FLs will also
flame can form and propagate freely in a fuel−oxidizer mixture if
the mixture is within the fundamental flammability limits change. It can be observed that an increase in the initial
(FLs);1,2 otherwise no flame propagation is possible. The temperature, while the initial pressure remains constant,
fundamental FLs are determined for certain standard initial decreases the value of the LFL and increases the value of the
conditions of temperature and pressure. The FLs of a fuel−air UFL. In other words, less fuel is needed on the lean side and less
mixture at certain initial conditions Ti and pi should depend only oxidizer is needed on the rich side to sustain flame propagation.
on the composition of the mixture and not on experimental On the other hand, increasing the initial pressure while
configurations. There are two fundamental FLs: a lower maintaining the initial temperature leads to an increase in the
flammability limit (LFL) for a lean fuel−air mixture and an UFL. Experimental observations have not clearly demonstrated
upper flammability limit (UFL) for a rich fuel−air mixture. The the effect of initial pressure on the LFL; in most cases it appears
FLs are generally expressed as fuel concentrations in the fuel−air
mixtures. The FLs measured in a laboratory experiment often
depend on the experimental configurations and flame Received: October 24, 2022
geometry.2 Therefore, the fundamental FLs must be appropri- Revised: February 6, 2023
ately extrapolated from the observed FLs to the planar and
unstretched flame conditions.
The LFL is defined as the leanest composition of the fuel−air
mixture below which a flame will not propagate under certain

© XXXX American Chemical Society https://doi.org/10.1021/acs.energyfuels.2c03598


A Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

to be very weak. The effects of temperature and pressure on FLs had formed there.19 Therefore, FLs are of great importance for
are discussed in detail in section 6. the adoption of safety measures in the application of alternative
Another important problem is the determination of the FLs of aviation fuels. Moreover, understanding the chemical kinetic
a mixture of two or more fuels with an oxidizer. For example, mechanisms of alternative aviation fuels20 is necessary for the
adding hydrogen to natural gas can improve flame stability3 and development of theoretical models for the prediction of FLs.
combustion efficiency.4 Therefore, the fuel mixture of hydrogen Despite the importance of FLs, few reviews have been
and natural gas could improve the combustion process in some published on this topic. As far as the authors are aware, the most
applications. However, due to safety concerns, it would be recent were published in 2001 by Ju et al.,2 in 2012 by Coronado
necessary to determine the FLs of this fuel mixture. The FLs of et al.,21 and in 2022 by Qi et al.22 In the case of the article by Ju et
fuel mixtures can be determined experimentally or estimated by al.,2 the focus was on the theoretical developments up to 2001,
the law of Le Chatelier.5 A comprehensive validation of this law while Coronado et al.21 focused on the experimental methods to
has recently been carried out, taking into account the available determine FLs for aerospace applications. In the recent article by
experimental data.5 This law will be discussed later in this article. Qi et al.22 the flammability limits of gases at elevated
The FLs also depend on the composition of the oxidant temperatures and pressures were reviewed. This article, on the
present in the fuel−oxidant mixture. As an example, consider the other hand, aims to be comprehensive by including discussions
difference between pure oxygen and air, which is the high on the theoretical developments, experimental procedures, and
concentration of nitrogen in the latter and its absence in the estimation methods that enable the understanding and
former. Likewise, air can be said to be further diluted by adding determination of flammability limits.
more nitrogen or another type of diluent such as carbon dioxide, Thus, the objective of this article is to review the state of the
water vapor, or helium, to name a few possibilities. These art in the study of flammability limits using a comprehensive
mixtures are studied in the literature as fuel−diluent−oxidant approach by addressing the following topics:
mixtures. Under certain temperature and pressure conditions, • the definition and existence of fundamental FLs
the addition of diluents results in an increase in the LFL and a • theoretical approaches to the determination of FLs
decrease in the UFL. • experimental methods for determining the FLs
Furthermore, the burning limits of a mixture can be • the available estimation methods to determine the FLs in
significantly changed by flame stretch.1 Therefore, it is air under reference conditions
important to understand the relationship between the burning
• the effects of initial temperature and pressure conditions
limits of a stretched flame and the fundamental FLs. In addition,
• the effect of adding diluents to the fuel and oxidant
for low temperature combustion of fuels such as n-alkanes,
mixture
alcohols, and ethers, the flammability limits of a cool flame can
differ significantly from those of a hot flame for the same • the estimation of FLs of fuel mixtures according to Le
mixture.6 A transition from cool flame to hot flame can change Chatelier’s law
the FLs of a hot flame. Therefore, it is also important to • the FLs of renewable and alternative fuels and future
understand the differences between the FLs of cool and hot perspectives
flames for the same mixtures. A review of the dynamics of cool
flames has recently been published.6 Therefore, only a brief 2. THE THEORY OF FLAMMABILITY LIMITS
discussion is provided in the Supporting Information. In an earlier review on FLs, Lovachev et al.23 adopted the
Some examples of the importance of the FLs for practical following definition for a flammable mixture: “We propose to
applications can be listed. For instance, in nuclear reactors, define a gaseous mixture as flammable per-se at a stated
mixtures of combustible gases rich in hydrogen can form and temperature and pressure if, and only if, it will propagate flame
come into contact with the hot walls of the reactor.7 Nuclear indefinitely, the unburnt portion of the mixture being
waste stored in underground tanks can also produce flammable maintained at that temperature and pressure”.
mixtures.8 In addition, the introduction of renewable fuels into From this definition, the FL can be defined as follows:23 “The
the energy matrix leads us to consider fuel mixtures involving flammability limits of a gaseous mixture, at a determined
gaseous fuels such as hydrogen and ammonia,9−11 and also temperature and pressure, are the leanest and richest
liquid fuels involving blends of biodiesel12 and ethanol,13,14 compositions for which the mixture can be considered as
among others. In either case, a flammable mixture could form flammable.” As Ju et al.1 have pointed out, the fundamental FLs
and FLs must be known for safe operation. of a mixture at a given pressure and temperature can only be
Some of the fuels mentioned are already present in industrial defined for the propagation limit of a planar unstretched flame,
processes, such as mixtures of carbon monoxide and hydrogen which is governed only by the chemistry of the fuel−oxidant
produced during the reforming of fossil fuels.15 Ammonia is used mixture and the thermal radiation. Therefore, a definition of the
in the polymer industry as a source of nitrogen and hydrogen fundamental FLs can be stated as follows: The fundamental
and at high temperatures in the nitration process for hardening flammability limits of a gaseous mixture, at a determined
steel.16 Due to the need to replace chlorofluorocarbons, various temperature and pressure, are the leanest and richest
gas mixtures are used in cooling processes. For example, composition for which a planar unstretched flame can propagate
chlorofluorocarbons have been replaced by hydrofluorocarbons, through the mixture. As will be seen later, the experimental
hydrofluoroethers, and mixtures of hydrofluorocarbons with determination of FLs does not consider planar flames in most
hydrocarbons.17,18 cases.
Further evidence of the importance of FLs for practical It is known that for a given mixture there are two extinction
applications is the tragic accident of Trans World Airlines Flight limits for a stretched premixed flame, namely the stretch
800, which crashed into the Atlantic Ocean on July 17, 1996. extinction limit at a high stretch rate, at which the characteristic
The cause was determined to be an explosion in the center wing burning time becomes longer than the flow residence time, and
fuel tank due to the ignition of a flammable fuel−air mixture that the radiation extinction limit at a low stretch rate, at which the
B https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

Figure 1. Schematic representation of the flammability limits of the ethanol−air mixture. Experimental data from Zabetakis.24

heat loss by radiation becomes comparable to the chemical heat Some of the above assumptions need explanation. Note that
release from the flame.1 It should be noted that these extinction the goal of this formulation is to obtain a simple relationship
limits should not be confused with the fundamental flammability between the heat released by the reaction, the mixture
limits. It is interesting to point out that, in the review by Ju et al.,2 properties, the heat losses, and the laminar flame speed.
they used the general term “combustion limits” instead of the Therefore, the assumptions of a one-step reaction and a linear
classical flammability limits defined for premixed flames. The temperature profile are reasonable. Furthermore, it is not
fundamental flammability limits, as well as the associated necessary to assume a thermal radiation model to understand
extinction limits for premixed flames, are discussed in this review the relationship between the above parameters.
article. Hereafter, the fundamental flammability limits will be The assumption that cp is independent of temperature can be
referred to simply as flammability limits. The combustion limits justified by the definition of the mean of a function, i.e., cp =
with nonzero flame stretch will be referred to as the extinction (∫ TTbu ∑Yicp,i)/(Tb − Tu). In a mixture where one species is
d

limits. strongly present in excess, the cp value can be considered


FLs are usually expressed as the percent mole fraction of fuel independent of the composition in such cases. Note that the
in the gas mixture. FLs are then expressed as shown in eq 1: Lewis number for stoichiometric methane−air mixtures without
nF and with some degree of dilution by argon, nitrogen, and carbon
FL = · 100% dioxide is close to unity.29 The Lewis number for methane−air
nF + nox (1)
mixtures with a ϕ value between 0.7 and 1.5 at 100 kPa and 293
where nF and nox are the numbers of moles of fuel and oxidant, K was calculated by the authors to be close to 1.09. The
respectively. The FL could also be expressed in terms of the assumption Le = 1 is often used to simplify the mathematical
equivalence ratio (ϕ). However, this is not a common practice in formulation, since it implies that k/ρcp = D.26
fire safety regulations. For example, the FLs of the ethanol−air Applying the assumptions and introducing the heat loss term,
mixture are shown schematically in Figure 1, with experimental the following expression is obtained:
data taken from the work of Zabetakis.24
A first interpretation of the FLs can be achieved by an analysis dT k d2T 1 L
mS = hf,0i iWi +
of the energy, mass, and species conservation equations, dx cp dx 2 cp cp (3)
considering a freely propagating, one-dimensional, planar
flame with some simplifying assumptions: The mixture composition can be either lean or rich. In the first
(i) The heat capacities and thermal conductivities are case, complete combustion is assumed, while in the second case
independent of temperature and composition. incomplete combustion is considered. By using these two
(ii) The process is at steady state. limiting cases, a simplified relationship for FLs can be derived.
The summation term on the right-hand side of eq 3 represents
(iii) The Lewis number is unity: Le = k/ρcpD = 1.
the volumetric heat release rate, and for a complete combustion
(iv) The combustion reaction can be represented by a one- reaction can be expressed as ω̇ fWfΔhC. Therefore, the total
step kinetic reaction. combustion heat release rate is related to the fuel consumption
(v) The heat losses by radiation are represented by a single rate (ω̇ f), the fuel’s molecular weight (Wf), and the fuel’s heat of
term, L. combustion (ΔhC). The expression is substituted into eq 3, and
(vi) The temperature profile is linear. integration is performed by applying the boundary conditions
and variable substitution shown below:
The above assumptions are compatible with the Shvab−
Zeldovich formulation in one dimension,25−28 which is shown in at → −∞:
eq 2.
Ä ÉÑ T = Tu and dT /dx = 0 (4)
d ÅÅÅÅ Ñ
cp dT ÑÑÑÑ
d d
mS
dx ( cp dT + ) dx ÇÅ
ÅÅ D
dx ( ÖÑ
) at x = δ/2:
0
= hf, i f Wi (2) T = (Tb + Tu)/2 and dT /dx = (Tb Tu)/ (5)

C https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

Using L̅ and f to represent the mean values of heat loss and for ξ → +∞:
fuel consumption rate, respectively, denoting the flame thickness d
by δ, and applying the variable substitution, then eq 3 can be = 0, =0
integrated, and the expression shown in eq 6 is obtained: d (11b)
The dimensionless variables and parameters are shown in eqs
1 ijjj L hC yz
zz 2 mS k 12−18. In those expressions, ω̇ A,b is species A’s consumption
j f Wf z (Tb Tu) + (Tb Tu)
2 jj cp 2cp zz 2 cp rate at the adiabatic flame temperature (Tb) and at the initial
k { mass fraction of fuel (YA,u) in kmol/(m3·s). The parameter kb is
=0 (6) the thermal conductivity at Tb, and Lb is the heat loss rate at Tb
where ṁ S is the mass flux rate. The expression shown in eq 6 is a per unit volume.
quadratic equation; therefore, a real solution is possible only if T Tu
the inequality shown in eq 7 is true. = dimensionless temperature
Tb Tu (12)
ij L hC yz
cp(Tb Tu)(mS)2 8k jjjj f Wf z
zz 0 ij cp SL yz
j cp 2cp zz = expjjj xzzz dimensionless coordinate
k { (7)
k k { (13)
The mass flux rate can be related to the laminar flame speed by YA
ṁ S = ρuSL. Rearranging eq 7, the expression shown in eq 8 is = dimensionless mass fraction of A
obtained. YA,u (14)

ij L hC yz hC A,bWAk b
1 8k jj zz = dimensionless heat release
SL j f Wf z
cp(Tb Tu) jj cp 2cp z z (Tb Tu)(cp SL)2
u k { (8)
(15)
The above expression provides an estimate of the minimum
k A WA
value for the laminar flame speed at which FLs are found. The = dimensionless reaction rate
key parameter is the balance between the heat losses due to kb A,bWA (16)
thermal radiation and the heat release rate. Note that both are
negative. This result is consistent with experimental observa- kL
= dimensionless heat loss
tions of a nonzero laminar flame speed at FLs. Thus, the k bL b (17)
condition for the existence of FLs is heat loss by thermal
radiation from the flame to the surroundings. Note that the Lb
K= ratio of heat loss to heat release
assumption of a linear temperature profile allowed the A,bWA hC (18)
integration of eq 3, which is not so easily performed when an
exponential profile is adopted. The results of more detailed The parameter K represents the ratio between the heat loss to
investigations are presented in the next section. Although eq 8 the surroundings at the adiabatic flame temperature (Lb) per
was obtained for a lean mixture, a similar expression would be unit volume and the heat release rate of the combustion reaction
obtained for a rich mixture, with the main difference in the value at the adiabatic flame temperature (ω̇ A,bWAΔhC) per unit
of ΔhC. Another theoretical relationship for the flammability volume. To understand the parameter λ, it is first necessary to
limit of a planar flame is given by Ju et al.2 in their review. Some notice that the flame thickness is equal to δ ∼ kb/cpṁ S. Then, λ is
more rigorous theoretical analyses have also been developed to the ratio between the heat release rate of the combustion
explain and predict FLs. Sections 2.1−2.3 discuss some reaction per unit area (ΔhCω̇ AWAkb/cpṁ S) and the rate of
important aspects of these theories. Due to the nature of this sensible enthalpy transport by convection per unit area cpṁ S(Tb
article, full formulations are not provided. − Tu). Considering that A is the fuel, Spalding30 showed that, for
2.1. Theoretical Developments with One-Step Chem- a type of reaction characterized by eqs 19 and 20, there is a
ical Kinetics. Spalding30 used a one-dimensional approach to critical value of K, above which there are no real values of λ.
develop a theory about the FLs. In his analysis, a one-step Thus, the critical values KC and λC define the FL.
reaction was considered in the form shown in eq 9. n
= (19)
A+B C (9)
l
o m
for >1
o
Starting from the one-dimensional energy and species =m
o
o 0 for 1
conservation equations,28 it is possible to model the problem n (20)
by the dimensionless equations in eqs 10 and 11, subject to the The solution obtained by Spalding shows the relationship
boundary conditions shown in eqs 11a and 11b. between λ, K, and θ as a function of n and m, which typically take
d2 values in the ranges 6−15 and 1−5, respectively.30 The simple
= ( +K ) results given in eqs 21 and 22 can be used to plot the graphs
d 2 2
(10)
depicted in Figure 2 by varying θ between 0.005 and 1.
d2 (n + 1)(n + 2)
= =
d 2 2
(11) 2 n+2 (21)

for ξ = 0: 2 n+ 2 m
K= (1 )
= 0, =1 (11a) (n + 1)(n + 2) (22)

D https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

where the mass fraction of fuel at the initial condition and the
mass flow rate for the adiabatic conditions are represented by Yf,u
and ṁ S,ad, respectively. Considering the heat loss term at the
flame temperature of the nonadiabatic flame (Lb), the limit
condition given in eq 28 is obtained.
2 2
1 mS,ad RcpTb
= 2L b
e Ea (28)
where Ea and e are the activation energy and Euler’s number,
respectively. The limit condition can also be written as shown in
eq 29, which represents the decrease in the mass burning rate
due to the decrease in flame temperature or the increase of heat
loss.
Figure 2. Results for λ and K as functions of θ according to eqs 21 and ij m yz
2
22. jj S zz = exp[ Ea /RTf ]
jj zz
k mS,ad { exp[ Ea /RTb] (29)
The critical values θC, λC, and KC are given by eqs 23−25. In
36
addition, the theory proposed by Spalding can be used to Williams suggested that a first approximation for the FL can
determine the critical value of the laminar flame speed (SL,C), be obtained by varying ṁ S,ad while holding constant the other
given in eq 26. parameters in eq 28 in order to establish the limit conditions.
n+2 m
Then, a critical value of exp[−Ea/RTb] can be identified due to
C = its proportionality with ṁ S,ad. The lower and upper flammability
n+3 m (23) limits would be represented by the critical values of exp[−Ea/
2 n+2 m
RTb].
KC = C It is worth noting that Hertzberg37−40 published four
(n + 1)(n + 2)(n + 3 m) (24)
interesting papers in the 1970s−1980s dealing with various
(n + 1)(n + 2) (n + 2)
phenomena related to FLs, such as natural convection,
C = C conductive−convective heat loss, radiative heat loss, and flow
2 (25)
gradients. Hertzberg found that the limiting velocity of the
1 k bL b laminar flame at the LFL is between 6 and 8 cm/s.37
SL,C = 2.2. The Relationship between Extinction Limits and
cp (Tb Tu)K C C (26) Fundamental FLs. More detailed analyses of premixed flames
Although this theory explains the relationship between heat subject to heat losses, flame curvature, different reaction orders,
release and heat loss at the FL, it is not directly applied to and transport properties can be found in early studies by Joulin
determine the fuel mole fraction at the FLs. However, the value and Clavin,41 Buckmaster,42 and Mitani43 in the 1970s.
of SL,C can be estimated as follows: At the LFL the properties of However, the relationship between flame extinction limits and
the mixture can be approximated by those of air (cp = 1004 J/kg· the fundamental flammability limits was still not clear.
K, ρ = 1.169 kg/m3, kb = 0.100 W/m·K31). As given by Beginning in the 1980s, research on combustion in micro-
Spalding,30 the exponent m of the dimensionless heat losses χ = gravity44 sparked renewed interest in the study of combustion
θm varies from 1 to 5. A larger value of m means that the heat limits. Analytical and experimental studies showed that, in
losses due to thermal radiation are more important than those addition to flame chemistry, flame extinction can be influenced
due to thermal conduction. Moreover, the exponent n of the by the coupling between flame stretch and thermal radia-
dimensionless reaction rate φ = αθn varies between 6 and 15. tion.1,45−47 To understand the coupling effect of flame stretch
Therefore, the values of m = 8 and n = 2 are within the intervals and radiation on flame extinction and FLs, Ju et al.48 applied an
recommended by Spalding. asymptotic analysis to include this effect. The results showed
In previous work, the authors determined the Tb value at that the stretch and radiation extinction limits are the two
initial conditions of 298 K and 101 kPa for 120 hydrocarbons at solutions of eq 30.
LFL,32 273 oxygenated compounds at LFL,33 115 hydrocarbons 2
1 2CLeF jij Ta zyz ji T zy
at UFL,34 and 208 oxygenated compounds at UFL.35 It is = jjj zzz expjjjj a zzzz
noticeable that the mean value of Tb at the LFL is 1548 K and ae (YO YF LeO/LeF ) k Tf { k Tf { (30)
that at the UFL is 1067 K, with standard deviations of 107 and
147 K, respectively. Therefore, the assumption of Tb = 1500 K where
for a mixture at the LFL is reasonable. Following Spalding, Lb = 2
WF( + 1)YF
0.4186 J/cm3·s, and assuming Tb = 1500 K, m = 8, and n = 2, eq C= 3
26 gives the value SL,C = 1.342 cm/s. 2 0vO LeF (31)
Another theoretical approach was developed by Williams,36 The stretch rate at the radiation extinction limit (ae) is related
who applied a one-dimensional formulation including a heat loss to the dimensionless concentrations of the fuel (Y̅ F) and the
term. In this theory, the heat loss rate is given by the expression oxidant (Y̅ O), the dimensionless flame temperature (T̅ f), the
shown in eq 27. dimensionless activation temperature (T̅ a), the Lewis number of
L the fuel (LeF), and the Lewis number of the oxidant (LeO). Also,
K= in the above equations, κ is a geometric coefficient of the
hCYf,umS,ad 2cp (27) stretched flame, κ0 is the frequency factor of the reaction, vO is
E https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

the stoichiometric coefficient of the oxidant, ρ is the density, and reduces to eq 30. At low fuel concentration (YF = 0.029) below
YF∞ is the initial mass fraction of the fuel. As such, the solution of the flammability limit, the flame can exist in a narrow range of
eq 30 gives the extinction limits of a stretched flame as a function stretch rate bounded by a radiation extinction limit and a
of the fuel concentration at different Lewis numbers of fuel or stretched extinction limit. The results of eqs 32 and 33 for flame
oxidizer.48 location as a function of flame stretch at LeF = 0.9 at different fuel
As can be seen in Figure 3, there are two extinction branches concentrations (YF) are shown in Figure 4. Note that Figure 4a
for each fuel concentration. The upper one shows the stretch has a different scale than the other three diagrams in Figure 4.

Figure 3. Effect of Lewis number of fuel on the flammability limit.


Reproduced with permission from ref 48. Copyright 2000 Elsevier.

extinction limit, and the lower one is the radiation extinction


limit. As the fuel concentration decreases, these two branches
merge, defining the burn limit of a stretched premixed flame.
Note that this burn limit is different from the fundamental FL. It
can also be seen that, as the Lewis number of the fuel decreases
(increases), the stretched flame extinction limit, the radiation
extinction limit, and the burn limit of a stretched flame are Figure 4. Flame location as a function of flame stretch for LeF = 0.9 at
expanded (narrowed). Therefore, for a stretched flame the flame different fuel concentrations (YF) predicted by the analytical solution in
burn limit will be affected by the Lewis number of the mixture, eqs 32 and 33.
flame stretch, and radiation, all of which are coupled together.
However, in the formulations in eqs 30 and 31 the relationship It is seen that, at a normalized fuel mass fraction, YF = 0.029,
between the fundamental flammability limit and the burn limits below the standard flammability limit, there is a stretched flame
of stretched flames was still unknown. For example, it was island bounded by a radiation extinction limit at a low stretch
unclear whether the burn limit of a stretched flame shown in rate and a stretch extinction limit at a high stretch rate [Figure
Figure 3 would extend to the fundamental flammability limit. To 4a]. Outside these two limits there is neither a flame with a lower
answer this question, computational1 and analytical49,50 efforts stretch rate nor a flame with a higher stretch rate than these two
were made to understand the flame extinction limits of a limits. This result is consistent with Figure 3.
stretched flame subjected to radiation at near zero stretch rate. However, as the fuel concentration is increased close to the
The analytical relationship50 between the normalized mass standard flammability limit, YF = 0.0295, a new weak flame island
burning rate (m), normalized flame location (ηf), flame stretch appears at very low stretch rates [Figure 4b]. This weakly
(a), normalized Lewis number deviation from unity (l), and stretched flame is located very far from the stagnation point of
flame temperature change (pf), due to stretch and heat losses the counterflow flame. Therefore, both the weakly stretched
before (H−) and after (H+) the flame front, is given as flame and the normal stretched flame can exist below the
a standard flammability limit.
= e pf /2 , m= f
2a At the fuel concentration at or slightly above the standard
2 f g1 (32)
flammability limit, YF = 0.0296, Figure 4c shows that the weakly
ij 1 stretched flame island expands rapidly toward zero stretch rate.
1 yzzz 2H + 2
pf = ljjjj 2
z+ gg At this time, the weakly stretched flame can be extended to zero
j 2 f
g1 zz a f 12 stretch rate, i.e., the planar and unstretched flame. Therefore,
k {
when the Lewis number is below unity, only the extinction limit
H ijjj I1 yzz of the weakly stretched flame can be extrapolated to the standard
+ jj g3 + zzz
a j g1 z flammability limit, while this is not possible for the normally
k { (33)
stretched flame.
where gi and I1 are integral functions of flame location. In the When the fuel concentration is above the flammability limit,
limit of zero stretch rate, that is, ηf → ∞ and a → 0, eq 33 YF = 0.03, the island of the weakly stretched flame is further
F https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

extended to a higher stretch rate and merges with the island of near the FLs. The flammability exponents calculated by Wang et
the normal stretched flame [Figure 4d]. Only under this al.52 for the LFL and UFL of CH4/SG mixtures in air were
condition can the weakly stretched flame smoothly transition always different from unity. Therefore, the condition established
into the normally stretched flame by increasing the stretch rate, by Law and Egolfopoulos51 to find the FLs was not satisfactory
and conversely, a normally stretched flame can smoothly when applied to these experimental data.
transition into an unstretched flame by decreasing the stretch Two modified flammability exponents were proposed by
rate. As such, eqs 32 and 33 and Figure 4 provide a clear Liang and Law.53 In their study, Liang and Law found that, in
relationship between the normal stretched flame and the addition to competition for radicals between branching and
unstretched planar flame via a weakly stretched flame near the termination reactions, there is also competition for intermediate
flammability limit. This analytical result was validated by species involving four other competing reactions. Therefore,
numerical simulations. The relationship between the extinction they proposed two alternative modifications of the flammability
limits of weakly and normal stretched flames and the standard exponent. In the first modification of the flammability exponent,
flammability limit at different Lewis numbers is presented in shown in eq 37, two groups of competing reactions are
section 2.3. considered, while in the second modification, shown in eq 38,
2.3. Theoretical Developments with Detailed Chem- three groups of competing reactions were considered.
ical Kinetic Mechanisms. 2.3.1. Theoretical Developments ÄÅ ÉÑ
Å Ñ
lnÅÅÅÅwT2 + wB2 w +T w ÑÑÑÑ
( )
w
with Detailed Chemical Kinetic Mechanisms and without
Heat Losses. This section begins with the work of Law and ÅÇ T B ÑÖ
b( ) = ÅÄÅ ÑÉÑ
Egolfopoulos,51 in which it was concluded that for the lnÅÅÅÅwB2 w +B w ÑÑÑÑ
( )
w
propagation of a one-dimensional planar flame near the FL ÅÇ T B Ñ Ö (37)
there is a condition in which the rate of a particular chain- ÅÄÅ ÑÉ
branching reaction is equal to that of a corresponding chain-
termination reaction. Therefore, the flame can be extinguished
lnÅÅÅÅwT1 + wB1
ÅÇ ( wT2
) + w ( )(
wT2 + wB2
wB2
B1 w + w
T2 B2
wT
wT + wB )ÑÑÑÑÑÑÖ
c( )= ÄÅ ÉÑ
Å
by any perturbation of the balance of these chain-branching and
chain-termination processes.51 They determined the dominant
lnÅÅÅÅwB1
ÅÇ ( )( )ÑÑÑÑÑÑÖ
wB2
wT2 + wB2
wB
wT + wB

branching and termination reactions for the propagation of (38)


52
premixed methane−air flames using a detailed chemical kinetic In their article, Wang et al. used the refined flammability
mechanism. They then represented the net rate of the dominant exponent from eq 37 to determine the FL numerically. The
branching reaction by its maximum value, wB, and the net rate of comparison between the experimental and calculated FLs for
the corresponding termination reaction by wT. Considering the CH4/SG mixtures in air is shown in Figure 5. Although the
equivalence ratio ϕ, the following expressions were proposed.
B( ) T( )
wB and wT (34)
The relationship between these net rates is then wT ∼ wBα(ϕ),
and this allows the definition of the flammability exponent α(ϕ)
= αT(ϕ)/αB(ϕ). Therefore, the FLs are expected to be found at a
limiting ϕFL that satisfies the relationship shown in eq 35.
( FL
)= T( FL )/ B( FL ) =1 (35)
From the expressions in eqs 34 and 35, and from the
relationship assumed for α(ϕ), the flammability exponent can be
represented as shown in eq 36.
(ln wT)/ (ln ) (ln wT)
a( )= =
(ln wB)/ (ln ) (ln wB) (36)
A predictive determination of the FLs consists of solving the
adiabatic one-dimensional flame propagation model, for differ-
ent values of ϕ, using detailed chemical kinetic mechanisms.
Then, a sensitivity analysis is performed to identify the dominant
chain-branching and chain-termination reactions to determine Figure 5. Comparison of results obtained with the flammability
the exponent αa(ϕ). Finally, the condition α(ϕFL) = 1 is used to exponent and the modified flammability exponent for the FL of CH4/
find the value of the FLs by identifying ϕFL. SG mixtures in air. Reproduced with permission from ref 52. Copyright
Following this approach, Wang et al.52 investigated the lower 2021 Elsevier.
LFL and UFL of methane (CH4) in air with the addition of a
combustible gas (SG) and nitrogen (N2), where the SG was a
mixture of ethane (C2H6), ethylene (C2H4), carbon monoxide modified exponent showed better agreement than the
(CO), and hydrogen (H2) in a ratio of 5:1:1:1. Experiments unmodified one, the authors noted that large absolute errors
were performed at atmospheric pressure and temperature, and were observed between the calculated and experimental FLs.
the FLs were determined for different volumetric concentrations This error could be due to the fact that, in addition to chain-
of the SG and N2, respectively. The authors performed a branching reactions, other important chain-propagation reac-
sensitivity analysis of the mass burning rate and identified the tions must also be considered, since they also contribute
competing chain-branching and chain-termination reactions significantly to the overall reaction rate from chain branching to
G https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

Figure 6. Theories of flammability limits involving detailed chemical kinetics with radiative heat losses. (a) Inflection point for a nonadiabatic
methane−air flame at 1 atm and comparison with the adiabatic flame. Reproduced with permission from ref 54. Copyright 1992 Elsevier. (b)
Maximum flame temperature as a function of stretch rate. Reproduced with permission from ref 1. Copyright 1997 Cambridge University Press. (c)
Fundamental flammability limit found as the extension to zero stretch of weak flames. Reproduced with permission from ref 57. Copyright 1999
Cambridge University Press.

equilibrium. Therefore, even with the modified flammability point was not observed in the adiabatic case. It was concluded
exponent, it was not possible to accurately determine the FLs. that the fundamental flammability limit is explained by the
2.3.2. Theoretical Developments with Detailed Chemical effects of competition between chain-branching and chain-
Kinetic Mechanisms and with Thermal Radiation Heat termination reactions and radiative heat losses. The flame speed
Losses. In a second article, Law and Egolfopoulos54 made a and temperature at the inflection point were 2.3 cm/s and 1375
modification to the PREMIX55 routine in order to include heat
K, respectively. Note that the flame speed is of the same order of
losses due to thermal radiation, which in turn were calculated
using the expression given in eq 39. magnitude as that determined using Spalding’s theory.
Using the criterion developed in their earlier work [α(ϕFL) =
qr = 4 bKP(T 4 T 4) (39) 1], Law and Egolfopoulos found an equivalence ratio ϕLFL =
0.502, which is close to the value obtained by considering the
where inflection point (ϕLFL = 0.493). Thus, in the specific case of
KP = pCO K CO2 + pH O K H2O + pCO K CO + pCH K CH4 methane, the concepts of flammability exponent (α) and
2 2 4

(40) inflection point were shown to provide similar LFL values. The
predominant chain-branching and chain-termination reactions
The Stefan−Boltzmann constant is represented by σb, while
for the methane−air lean flames at 1 atm are shown in reactions
KP and pi are the Planck mean absorption coefficient and the
partial pressure of the radiating species i, respectively. Then they R1′ and R2′. Here H represents the hydrogen atom, O2
determined the mass flux (ṁ S) for different ϕ’s under adiabatic represents oxygen, OH represents the hydroxyl radical, HO2
and nonadiabatic conditions and found that in the latter case an represents the hydroperoxy radical, and M represents a third
inflection point for methane−air flames was reached at ϕLFL = body (any chemical species).
0.493, which is close to the experimental value of the LFL for
methane in air [Figure 6a]. On the other hand, this inflection H + O2 OH + O chain branching (R1′)

H https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

Figure 7. (a) Schematic of the flammable region for normally stretched flames and weakly stretched flames (WSF) and projected G-shaped
flammability limit diagram for Le < Lecr. (b) The G-shaped flammability limit diagram of methane/air mixture with Le = 0.97 < Lecr. (c) The K-shaped
flammability limit diagram of propane/air mixture with Le > Lecr. Reproduced with permission from ref 57. Copyright 1999 Cambridge University
Press.

H + O2 + M HO2 + M chain termination extinction limit. For a methane concentration of 4.9%, it was
(R2′) shown that at the stretch extinction limit the heat loss due to
Thus, the reactions shown in (R1′) and (R2′) dominate the radiation was less than 1% of the chemical heat release, while this
combustion process at the LFL of one-dimensional methane− loss at the radiation extinction limit was about 25% of the
air planar flames. To understand the coupling effects of flame chemical heat release.
stretch and thermal radiation with detailed chemical kinetic The maximum temperatures reached near the normally
mechanisms, Ju et al.1 numerically investigated the extinction stretched lean premixed flames at different stretch rates are
mechanism of weakly stretched and normally stretched shown in Figure 6b, where the radiation extinction limit is at
premixed methane−air flames with an optically thin radiation point d and the stretch extinction limit is at point c. The
model. The Planck constants of the radiative gases such as water existence of the radiation extinction limit is related to the
(H2O), carbon dioxide (CO2), and CO were derived from the increase in flame volume, which leads to larger radiation heat
statistical narrow-band model.56 The system of conservation losses from the flame.58
equations solved by Ju et al.1 is as follows. The stretch and radiation extinction limits were determined
by Ju et al.1,57 for various equivalence ratios, and the results are
da shown in Figure 6c. The lower burning limit of a stretched flame
=0
dt (41) is the point where the stretch and radiation extinction branches
d dU meet, which is denoted as point B in Figure 6c. The fundamental
+ = 2 G flammability limit is also shown in Figure 6c at point E. Leaner
dt dx (42)
mixtures than that at the fundamental flammability limit can be
d ij dG yz i da y burned if a suitable stretch rate is set, as shown by the branches
L(G) = jj zz G2 + jjj + a 2zzz of the curves AB and FB, which lie to the left of point E in Figure
dx k dx { k dt { (43)
6c. A flame that occurs for a mixture leaner than the fundamental
d ij dT yz
n
dT
n flammability limit is called a sublimit flame,57 and its existence
cpL(T ) = jj zz YVc
i i p,i hi iMi was confirmed by Ronney and Wachman59 in experiments
dx k dx { i=1 dx i=1 under microgravity.
+ qr As discussed later in Figure 7 and shown in Figure 6c, the
(44)
branches of the stretch extinction limit (AB) and the radiation
d
n extinction limit (FB), of normally stretched flames, extrapolated
L(Yi ) = ( YV
i ix ) + i Mi to zero stretch rate, do not lead to the fundamental flammability
dx i=1 (45) limit (E). On the other hand, the extinction limit of the weakly
where L(ϕ) = dϕ/dt + U dϕ/dx, and the boundary conditions stretched flame branch (CE) can be extrapolated to the
are given in eqs 46 and 47. fundamental flammability limit for methane. Similar conclusions
were drawn by Guo et al.60 and confirmed for hydrogen−air
x= L; T = TL ; Yi = YiL ; G=a (46) mixtures61 and methane−hydrogen−air mixtures.62
To better understand the conditions under which extrap-
x = 0; dT /dx = 0; dYi /dx = 0; dG /dx = 0; olation of the extinction limit of the normally stretched flame to
U = 0; a = a0 zero stretch can lead to the fundamental flammability limit,
(47)
computational studies were performed to observe the effects of
1,57
The numerical results obtained by Ju et al. [Figure 6b,c] the fuel Lewis number on the normally and weakly stretched
show the existence of a stretch extinction limit and a radiation premixed flames of CH4−O2−N2−He mixtures.57,63 The Lewis
I https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

Figure 8. Upper flammability limit of benzene with detailed chemical kinetic mechanism, radiative heat losses, and soot emissions. (a) Flammability
limit with sudden drop in laminar flame speed.64 (b) Flammability limit with a sudden drop in temperature.64

number of the mixture was increased by increasing the et al.65 developed a study that considered detailed chemical
concentration of helium (He) in the mixture. The results kinetic mechanisms and heat losses by thermal radiation,
showed that for a fuel Lewis number below a critical value, Lecr, without considering the emission from soot molecules at rich
as shown in Figure 7a, the existence of sublimit normally conditions. The calculated and experimental FLs of methane in
stretched flames and the near-limit weakly stretched flames air at different initial pressures were compared for validation.
(WSF) lead to a G-shaped flammability limit diagram [Figures The model proved to be very accurate for determining the LFL.
7b and 6c] for lean methane−air mixtures. However, it showed only acceptable accuracy for the
Therefore, for Le < Lecr, the stretch extinction limit of the determination of the UFL. Therefore, the comparison of the
normally stretched flames cannot be directly extrapolated to works of Bertolino et al. and Mascarenhas et al. shows that the
zero stretch rate to obtain the fundamental flammability limit, inclusion of emissions from soot is very important for the
and only the stretch extinction limit of the weakly stretched accurate prediction of the UFL.
flames can be directly extrapolated to zero stretch rate to obtain The theoretical results presented in this section could be
the fundamental flammability limit. It was also found that Lecr is extended to other fuels in air. After decades of development,
greater than unity and depends on the radiation intensity of the there are several detailed chemical kinetic mechanisms for
mixture and the product. The higher the radiation intensity, the hydrocarbons, 66,67 hydrocarbons and oxygenated com-
higher Lecr will be. pounds,68,69 and alcohols.70 The new developments in the
For the lean propane−air mixture with a Lewis number detailed chemical kinetic mechanisms for hydrogen71,72 and its
greater than Lecr, the sublimit normally stretched flames reduced or one-step mechanisms73 are of particular interest
disappear, as seen in Figure 7c, where the fundamental because it is a very important fuel that has many applications
flammability limit (E) is to the left of the ABC curve. Note
worldwide. In the cases of the United States and Brazil, ethanol
that this was not the case for the methane−air mixture in Figure
is of great importance since these countries are the first and
7b. Therefore, the result for the propane−air mixture is a K-
second ethanol producers in the world.74 There are detailed
shaped flammability limit diagram [Figure 7c] instead of the G-
chemical kinetic mechanisms for ethanol,14,75 and also a reduced
shaped diagram obtained for methane−air. Therefore, for Le >
Lecr, the stretch extinction limit of the normally stretched flames chemical kinetic mechanism was developed in 2018.76 Although
can be directly extrapolated to zero stretch rate to obtain the much progress has already been made, it is worth noting that
fundamental flammability limit, as shown by the extrapolation Brown et al.77 have shown that methods for determining
CE in Figure 7c. transport properties used in numerical simulations of flame
Another development regarding the determination of FL with propagation can still be improved.
detailed chemical kinetics and radiative heat losses was carried A possible approach to extending the results obtained with
out by Bertolino et al.64 The most notable innovation of this detailed chemical kinetic mechanisms would be to obtain one-
work was the inclusion of a soot formation model and a soot step chemical kinetic mechanisms, as has been done in the past
emissivity model to improve the predictions of the UFL. The by Westbrook and Dryer.78,79 In their work, Westbrook and
FLs were estimated by changing the equivalence ratio until Dryer fitted the parameters used to determine the reaction rates
sudden drops in laminar flame speed (SL) and flame temperature of the fuels by comparing the calculated laminar flame speeds
were achieved, as shown in Figure 8, parts a and b, respectively. with some experimental values. The laminar flame speeds were
Note that the sudden drop in flame speed occurs at about 3 cm/ plotted against the equivalence ratio, and it was expected that the
s, which again is of the same order of magnitude as the flame projection on the horizontal axis would intersect it at the
speed according to Spalding’s theory. flammability limits. The parameters obtained in this way are
The authors validated their work with experimental data for presented in some textbooks.26 However, the species in the
methanol, benzene, methane, ethane, propane, n-butane, and n- products were H2O and CO2, which is unfavorable for the
pentane. The validation included the effects of the type of analysis of rich flames, since the heat release is overestimated in
oxidant, pressure, temperature, and diluent. Also, Mascarenhas such cases.
J https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

Figure 9. Schematic representation of the experimental apparatus as presented by Coronado et al.82 Reproduced with permission from ref 82.
Copyright 2014 Elsevier.

Note that, at high CO2 and H2O concentrations, radiation


reabsorption can increase the flame temperature and extend the
flammability limit. Such radiation reabsorption effects have been
demonstrated by computational modeling of CO2-diluted
methane−air and dimethyl ether−air mixtures.56,80

3. EXPERIMENTAL DETERMINATION OF THE


FLAMMABILITY LIMITS
3.1. Description of Experimental Determination of the
FLs by the ASTM Method. In order to provide a better picture
of the experimental determination of FLs, the ASTM E68181
standard is considered. This standard was applied by Coronado
et al.82 for the determination of FLs of anhydrous ethanol and
hydrated ethanol in air at different temperatures and pressures.
A schematic representation of the experimental equipment is
shown in Figure 9. The vessel shown in Figure 9 can be placed in
a heating chamber or a cooling chamber, depending on the Figure 10. Visual criterion of flammability as given in the norm ASTM
desired initial temperature. The ignition source is located in the E681 and presented by Coronado et al.82 Reproduced with permission
center of the vessel and is supplied by two electrodes 6.44 mm from ref 82. Copyright 2014 Elsevier.
apart with an ignition voltage of 15 kV and 30 mA.81
The procedure for obtaining FLs begins by bringing the vessel internal pressure rise is about 5 or 7% of the initial pressure.
close to vacuum conditions using a vacuum pump. Then the These pressure rise values are specified by the European and
mixture is prepared to the desired pressure using the partial American standards, respectively.21
pressure method. A magnetic stirrer is activated for about 2 min Two experiments performed by Mendiburu83 in a spherical
to homogenize the mixture. Finally, the mixture is ignited, and vessel with two flammable mixtures near the FLs are shown in
the flame propagation is recorded by a video camera. The visual Figure 11. A typical flame near the LFL is shown in Figure 11a,
criterion is applied to the recorded videos to determine whether while a typical flame near the UFL is shown in Figure 11b. It is
a mixture is flammable or not. clear that both flames meet the visual criterion because the
The visual flammability criterion states that a mixture is flames form a 90° arc. However, while the flame near the LFL
flammable if the flame spreads upward and toward the wall, expands into a sphere, the flame near the UFL expands into a
forming an arc greater than 90°. This criterion is shown in Figure hemisphere. This qualitative difference is of great importance
10. The fuel concentration in the mixture is varied to determine when UFLs are theoretically determined, since the effect of
two different, but close, fuel concentrations such that one is buoyancy is obvious.
flammable and the other is not flammable under the same Standards based on the visual criterion consider a vessel that
pressure and temperature conditions. The FL is determined for vents to the atmosphere when the internal pressure increases. In
the average of the fuel content of these two mixtures. Therefore, the ASTM E681 standard, this is accomplished by using a rubber
the smaller the gap between these two compositions, the more cap to seal the vessel. This rubber is removed by the internal
reliable the FL is. Another flammability criterion is the pressure pressure of the vessel. Given this venting, which is considered in
rise criterion, which classifies a mixture as flammable when the the visual criterion, and the small pressure rise considered in the
K https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

states that a mixture is flammable if, after ignition, the flame


detaches (and spreads) a distance greater than 100 mm from the
ignition source. The second method is EN 1839(B), in which
case a 5 L spherical bomb is used. However, the flammability
criterion used is the pressure rise criterion. A mixture is classified
as flammable when the pressure increases by 5% compared to
the initial pressure.
Note that, in the ASTM E681 standard, FLs are the final
compositions that support flame propagation. This is
determined as the average fuel concentration between two
successive compositions, one showing no flame propagation and
Figure 11. Typical flames observed for flammable mixtures. (a) Slightly the other showing flame propagation. On the other hand, the EN
above the LFL of anhydrous ethanol in air at 20 kPa and 50 °C. (b) 1839 standard defines the FL as the fuel concentration for which
Slightly below the UFL of hydrated ethanol at 40 kPa and 280 °C. no flame propagation can be supported, and which follows the
Experiments performed by Mendiburu.83 last concentration that supports flame propagation. Therefore,
slightly different values for the FLs are to be expected when
pressure rise criterion (5−7%), it can be concluded that flame different standards are used. This last statement is confirmed in
propagation at constant pressure is a reasonable approximation the work of Schröder and Molnarne,85 who compared four
of the experimental conditions. Note that this is not necessarily a different methods for determining the FLs of several fuels. As an
coincidence, since the experimental methods for determining example, let us consider the FLs of CH4 in air. The LFL of CH4
FLs are intended to mimic a freely propagating flame as much as was 4.3% using EN 1839(T), 4.9% using EN 1839(B), and 3.8%
possible. using ASTM E681. On the other hand, the UFL of CH4 was
The European standard EN 183984 is also widely used for the 16.8% using EN 1839(T), 16.9% using EN 1839(B), and 16.9%
experimental determination of FLs. In this standard, there are using ASTM E681.85
two test methods for determining FLs. The first method is EN In most experimental tests to determine FLs, the composition
1839(T), which considers a vertical tube with a volume of at of the combustion gases is not measured.86 The reason for this is
least 1.52 L (D = 8 cm). The flammability criterion in this case that the main focus is to determine whether or not the mixture is

Figure 12. General characteristics of most common experimental equipment used to determine the flammability limits. (a) Experimental works with
cylindrical vessels.4,7,15,16,18,52,97−166 (b) Experimental works with spherical vessels.17,69,82,95,167−226

L https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

flammable. This is a limitation for theoretical modeling of FLs increased (slightly) and the UFL decreased with increasing
and for determining the difference between normal flame diameter.
propagation at the UFL and propagation of a cool flame (see the Similarly, Shoshin and De Goey248 observed that the LFL of
Supporting Information). CH4 and CH4−H2 mixtures decreased with increasing tube
There are several compilations of experimental data on FLs. diameter. Moreover, most CH4−H2 mixtures tested in tubes
The most complete would be that of the Design Institute for larger than 12.3 mm in diameter achieved an LFL that was
Physical Properties (DIPPR Project 801).87 However, there are independent of tube diameter. Note that the tube diameters
also some free online sources,88−90 and some handbooks used by Shoshin and De Goey were very small (6 mm to 50.2
covering various properties along with FLs are also avail- mm) compared with other studies. Therefore, it is not surprising
able.91−94 These sources may be sufficient for many applications. that the LFL was higher for the smaller diameters (6 mm to 12.3
However, to develop theoretical models, experimental con- mm). These results may be the cause of the increased flame
ditions are also of interest. surface area due to the rapid formation of the boundary layer in a
3.2. Analysis of the Available Experimental Data. As narrow tube.
mentioned earlier, there are several standard experimental Further analysis of the experimental results shows that for
methods for determining FLs. For example, Coronado et al.21 vessels of the same diameter (5 cm), the flammable range
compared the methods of the German Institute for Stand- shrinks with increasing height if the assumed flammability
ardization (DIN 51649-1) and the European standards [EN criterion requires that the flame reach all the way to the top to
1839(T), EN 1839(B)]. On the other hand, Kondo et al.95 used consider the mixture flammable. The data of Takahashi et al.230
the American Society of Heating, Refrigeration and Air for methane and propane indicate that there is a critical volume
Conditioning Engineers (ASHRAE) method and Schröder above which FLs are not significantly affected by geometry, as
and Molnarne85 compared four experimental methods for shown by the experiments they performed in a vessel of 160 L
determining FLs (as mentioned earlier). These methods are volume. On the basis of all these comparisons, Takahashi et
given by the EM 1839(T), ASTM E681-01, and DIN 5164-1 al.230 recommended that cylindrical vessels with a diameter of at
standards and gave similar values of the FL for hydrogen, least 20 cm and a height of 60 cm should be used to determine
methane, ethylene, and ammonia. FLs. The FLs determined in a 12 L spherical vessel using the
It is not the aim of this article to provide a detailed description visual criterion (90° arc) have values similar to those determined
of existing standard methods for the experimental determination in the 160 L cylindrical vessel for methane and propane.230 In
of the FLs; interested readers may wish to consult the work of addition, Kondo et al.249 compared the results for FLs of
Coronado et al.,21 Schröder and Molnarne,85 and Britton.96 The methane obtained with cylindrical vessels of 20 and 40 cm in
goal of this section is to provide a general insight into the height and 10 cm in diameter and recommended the use of the
experimental determination of the FLs. Since the experimental largest vessel to determine FLs. Thus, they confirmed to some
procedure of the ASTM E681 standard has already been extent the conclusions of Takahashi et al.230 Glukhov et al.234
presented, the evolution of the experimental determination of found that the FLs of hydrogen, methane, and propane were
FLs will now be discussed to show that researchers are currently consistent at subatmospheric pressure for vessels of 25 and 500
still using different approaches. The most commonly used L volumes. In addition, the minimum pressure at which flame
experimental configurations are cylindrical and spherical vessels. propagation was observed is lower in the work of Glukhov et al.
As can be seen in Figure 12, the use of these devices began a than in the work of Kuznetsov et al.,250 who used a spherical
century ago. The experimental works cited in Figure 12 are vessel of 8.2 L. The idea that there is a critical volume above
presented in more detail in Table S1. Some of these works have which the FLs do not show significant variations can be
determined FLs at various initial temperatures and pressures, confirmed in the work of Zlochower and Green,221 who used
taking into account diluents or fuel mixtures. For another spherical vessels of 20 and 120 L to determine the FLs of
compilation, see Table S2. In this table, the emphasis is on methane. Also, the FLs determined with the 120 L vessel for CO,
experimental work in which the FLs were determined with C2H4, and propane (C3H8) are similar to those determined by
vessels of different volumes and/or geometries. As observed in Kondo et al.184,187 in a 12 L vessel.
Figure 12, there are 73 works in which cylindrical vessels were The comparison of FLs obtained in cylindrical and spherical
used,4,7,15,16,18,52,97−166 62 works in which spherical vessels were vessels was performed by several authors. For example, Van den
used,17,69,82,95,167−226 and 10 works in which both geometries Schoor et al.231 determined the FLs of CH4−H2−air mixtures in
were used.3,227−235 Compilations of experimental data are also cylindrical (0.84 L) and spherical (4.2 L) vessels. Larger
available.8,24,236−246 flammability ranges were obtained with the cylindrical vessel.
The articles listed in Table S2 allow us to evaluate the effects Van den Schoor and Verplaetsen3 tested CH4−H2−air mixtures
of geometry on FLs. Based on the hypothesis that larger vessels and found that the UFL tended to be lower when determined in
would reduce the heat lost to the walls, it was expected that a spherical vessel (4.2 L) than when determined in a cylindrical
larger tube diameters would result in larger flammability ranges. vessel (1.50 L). Goethals et al.227 also observed a narrower
However, this hypothesis is at odds with several experimental flammability range for toluene when FLs were determined in a
results. For example, Burgoyne and Williams-Leir247 determined spherical vessel (8.0 L) compared to those determined with a
the FLs of CO, C2H4, cyclohexane (C6H12), and hexane (C6H14) cylindrical vessel (0.84 L). Markus et al.229 studied the FLs of
in tubes with diameters of 4.8 and 10.1 cm and heights of 152.4 methanol (CH3OH) mixed with CH4 and air and found higher
cm. It was observed that the UFLs of C2H4 and C6H14 decreased UFLs for a cylindrical vessel (0.8 L) than for a spherical vessel (5
with increasing diameter. On the other hand, the LFLs showed L). Note that Markus et al. used a much higher ignition energy
no significant variation with tube diameter. These observations (70 J) for the tests with spherical vessels than for the tests with
are confirmed by the results of Takahashi et al.230 for CH4 and cylindrical vessels (10 J). Therefore, the volume and geometry
C3H8 in cylindrical and spherical vessels. It was observed that for were more important than the energy of the ignition source. De
a height of 40 cm and a diameter of 5 to 150 cm, the LFL Smedt et al.232 showed that the flammable ranges for methane,
M https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

ethane, propane, and butane determined in a spherical vessel (20 The data in Table 1 show that the FLs depend on the diameter
L) were narrower than those determined in a cylindrical vessel and height of the tube. In addition, the UFLs determined with
(0.84 L). The ignition energies were 10 and 5 J, respectively. downward flame propagation (DFP) are smaller than those
Two possible conclusions can be drawn from these results: (1) determined with upward flame propagation (UFP) due to the
vessels with larger volume yield narrower flammability ranges, buoyancy effect, while those determined with horizontal flame
regardless of the flammability criterion used; (2) experiments propagation (HFP) are intermediate between the two. This type
conducted in cylindrical vessels yield broader flammability of difference in experimental results motivated the determi-
ranges than those conducted in spherical vessels. nation of FLs of some fuels under microgravity conditions, a
Kul et al.218 considered spherical vessels with volumes ranging topic that will be discussed in section 4.
from 3.2 to 21.3 L. They found that the LFL of difluoromethane In their work, Hu et al.103 determined the LFL of H2 in pure
was higher for the smaller vessel (3.2 L) and was the same for O2 with dilution by N2 and CO2 in tube experiments and also
vessels of 5 L and above (within experimental uncertainty). considered two directions of flame propagation, namely UFP
Therefore, in this case, larger vessels did not yield narrower FLs and DFP. The results show that, for DFP, the LFL decreases
for volumes of 5 L and above. The comparison of the UFL of between 8 and 13% when O2 is diluted by CO2. On the other
propane obtained in spherical vessels of 4.2 and 8 L at different hand, for UFP, the LFL decreases between 4 and 7%, also when
initial temperatures (up to 250 °C) and at pressures between 10 O2 is diluted with CO2. A similar behavior is observed for
and 20 bar was performed by Van den Schoor220 using the dilution with N2. The influence of the flame propagation
pressure rise criterion of 1 and 5%, respectively. Since it was direction on the determination of FLs was therefore also
expected that the UFL would not depend on the volume in this demonstrated by Hu et al. In a work by Hertzberg,37 it was found
case, the author stated that the higher pressure rise criterion was that the laminar flame velocity at the limit for horizontal flame
responsible for the lower values of UFL in the 8 L vessel. These propagation is almost twice as high as the limit velocity for
results are in contradiction with conclusion 1. Therefore, upward flame propagation. Therefore, the flammable ranges
conclusion 2 should be further investigated. The larger tend to be larger for UFP than for DFP or HFP. Note that flames
flammability ranges observed in cylindrical vessels suggest that
with HFP develop inverted flame fronts in closed vessels, also
the flammability criterion and the direction of flame propagation
called tulip flames.252,253 It is not clear how this phenomenon
play important roles.
affects the experimental determination of FLs with HFP.
Kul et al.218 used the same flammability criterion and
Again considering the work of Coward and Jones,251 it was
geometry (spherical) in their work and confirmed that there is
shown that the flammability range determined with UFP is
a wall quenching effect that reduces the flammable range for
smaller vessels. Since the cylindrical vessels were always smaller larger than that determined with DFP. The experimental FL
than the spherical ones, the wall quenching effect should have values of various compounds determined with UFP and DFP are
narrowed the flammable ranges, but this was not the listed in Table 2 for comparison. All of the compounds listed in
case.227,229,231,232 This would imply that the flammability Table 2 show some dependence of UFL on the direction of
criterion and the direction of flame propagation are responsible flame propagation, and some of these compounds show strong
for increasing the flammable range in cylindrical vessels. Before
turning to the analysis of the flammability criterion and the Table 2. Flammability Limits of Some Compounds with UFP
direction of flame propagation, it is important to point out that and DFPa
Bolshova et al.213 determined the LFL of methane in air using a upward flame downward flame
spherical vessel and a countercurrent burner and found that the propagation propagation
LFLs were consistent for both methods. LFL UFL LFL UFL
In early studies where FLs were determined using flame compound formula (%) (%) (%) (%)
propagation in vertical and horizontal cylindrical tubes, the hydrogen H2 4.15 75.00 8.80 74.50
ignition source could be at the top, bottom, or center of the tube. methane CH4 5.35 14.85 5.95 13.35
Consequently, the flame would propagate downward, upward, acetylene C2H2 2.60 78.00 2.80 63.50
or in both horizontal directions, respectively. The FLs acetaldehyde C2H4O 4.21 57.00 4.36 12.80
determined with these configurations were reported by Coward propylene C3H6 2.18 9.70 2.26 7.40
and Jones251 for several compounds. Not surprisingly, the methyl ethyl ketone C4H8O 2.05 9.90 2.10 7.40
experimental values obtained under different configurations benzene C6H6 1.45 7.45 1.48 5.55
were in some cases very different. As an example, the FLs of methyl alcohol CH3OH 7.10 36.50 7.65 26.50
methane in air obtained with different experimental config- butylene C4H8 1.70 9.00 1.80 6.25
urations are shown in Table 1. toluene C7H8 1.31 6.75 1.32 4.60
carbon disulfide CS2 1.41 50.00 2.03 34.00
Table 1. Experimental FLs of Methane in Air at Atmospheric ethane C2H6 3.12 14.95 3.26 10.15
Conditions for Different Experimental Configurations251 acetone C3H6O 2.90 12.60 3.15 8.35
ethyl acetate C4H8O2 2.32 11.40 2.37 7.10
direction of flame propagation ethyl alcohol C2H5OH 4.24 18.95 4.44 11.50
pentane C5H12 1.42 8.00 1.48 4.64
tube dimensions upward downward horizontal
hydrogen sulfide H2S 4.30 45.50 5.85 21.30
diameter height LFL UFL LFL UFL LFL UFL ethylene C2H4 3.20 34.00 3.33 15.50
(cm) (cm) (%) (%) (%) (%) (%) (%)
ethyl ether C4H10O 1.93 48.00 2.15 6.15
6.0 200 5.40 14.80 6.00 13.40 5.40 14.30
acetaldehyde C2H4O 4.21 57.00 4.36 12.80
4.0 100 5.50 14.10 6.10 13.30 5.60 13.90
5.0 50 − 15.11 5.80 13.38 5.39 14.28 a
Data from Coward and Jones.251

N https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

dependence. For example, the UFLs of acetaldehyde (C2H4O), ditions produced different FLs, Jarosinski257 argued in his review
diethyl ether (C4H10O), hydrogen sulfide (H2S), and C2H4 paper that “until now it has been impossible to formulate a
determined with UFP are more than twice as high as the UFLs universal theory of extinction limits [in this review FLs] for a
determined with DFP. All compounds, except H2, showed a large range of initial and boundary conditions.”
weak dependence of the LFL on the direction of flame When the experimental environment of microgravity and
propagation. detailed kinetic models became available in the 1990s,
Sometimes a mixture could not propagate at all for DFP. For researchers used microgravity experiments and numerical
example, White254 determined the FLs of ammonia−air models to try to understand the mechanism of FLs. For
mixtures in tubes 1.5 m long and with diameters of 5.0 and example, the outward propagating spherical flame without
7.5 cm. It was observed that ammonia−air flames do not gravity was simulated by Sibulkin and Frendi258 and Lakshmisha
propagate downward at initial temperatures below 70 °C. On et al.259 with the inclusion of thermal radiation. The predicted
the other hand, Krivulin et al.255 determined the FLs of LFL for methane−air was very close to the experimental data.
ammonia−air mixtures at 25 °C with DFP in a 8 m3 cubic They also showed that there is no FL for an adiabatic flame.
combustion chamber. Thus, it was possible to measure the FLs Therefore, they concluded that heat loss by radiation is the
with DFP for ammonia−air mixtures at room temperature using mechanism that causes the flammability limit. These studies
a large vessel. motivated subsequent computational and experimental inves-
There are at least three effects that can cause this increase in tigations to understand the mechanism of FLs of stretched
UFL as the flame propagates upward: (1) natural convection flames in a counterflow system with coupling of flame stretch
due to buoyancy, (2) preheating by the burned gas, and (3) and thermal radiation.1,2,44−47,57,63,260 As shown in Figure 6, the
Lewis number effect. Natural convection promotes flame findings on the existence of sublimit stretched flames, weakly
propagation in UFP. However, in DFP, it counteracts flame stretched flames, and radiation extinction have provided a clear
propagation. Therefore, natural convection may be responsible understanding of the mechanism of FLs through the intricate
for some differences between UFLs determined with UFP and coupling between chemistry, flame stretch, and thermal
those determined with DFP. radiation.
Preheating of the unburned mixture by the burned gases in the 4.1. Flammability Limits in Microgravity. Experimental
experiments could have a greater impact on the difference determination of FLs under microgravity conditions has the
between UFLs measured in UFP and DFP experiments. In advantage of isolating the influence of buoyancy from diffusion
addition, Jarosinski et al.256 found that the differences in FLs for and convection processes in the experiments. Extensive studies
UFP and DFP also depend on the Lewis number (Le) of the of combustion in microgravity have been performed to obtain
mixture. They determined the value of Le as the ratio of the FL values that do not depend on gravity and do not depend on
thermal diffusivity (α) of the mixture to the mass diffusivity (D) the direction of flame propagation. Mixtures with Le < 1 include
of the deficient reactant. It was observed that there is a significant lean hydrogen−air, rich propane−air, and higher hydrocarbon−
difference between the UFLs for UFP and DFP when the Lewis air mixtures. In these cases, there is a significant difference
number is less than unity (Le < 1). The preferential diffusion of between FLs determined with DFP and with UFP.244 On the
such mixtures may support UFP. This observation can be related other hand, there are some mixtures with Le > 1, for example, the
to the results of the asymptotic analysis of Ju et al.,48 who rich methane−air and lean propane−air mixtures. In these cases,
showed that the lower flammability limit decreases for lower the FLs determined with DFP and UFP are almost identical.244
Lewis numbers of the fuel (deficient reactant). Jarosinski et al.261 studied the flammability limits of propane−
The results obtained by Takahashi et al.230 and Coward and air mixtures under microgravity conditions. The UFL under
Jones251 raised the question of the flammability limit criterion, microgravity conditions was 8.7% compared to the UFL
that is, the criterion applied in an experiment to determine determined under normal gravity conditions of 9.5% with
whether a mixture is flammable or nonflammable. In this regard, UFP and 6.3% with DFP. Therefore, the UFL for the propane−
Britton96 examined several flammability criteria and stated, air mixture under microgravity conditions is between the UFLs
“Although wider flammability limits might be claimed to be determined with UFP and DFP. The laminar flame speed of a
more conservative, it is not the job of a test laboratory to include propane−air mixture with a propane content of 8.4% was
safety factors in the test method. Such safety factors should be determined to be about 2.0 cm/s for microgravity conditions
included during application of data, not during measurement.” and about 2.3 cm/s for normal gravity conditions.
Lovachev et al.23 pointed out that there is confusion in Pu et al.244 found that the FLs of propane−air mixtures were
experimental determination of FLs and determination of safety almost identical under gravity and microgravity conditions. The
margins against explosions. To date, there is no unified criterion values were 2.2% for the LFL and 9.2% for the UFL. They
for the flammability limit. However, the visual criterion and the attributed the differences to insufficiently sensitive measurement
pressure rise criterion are the most commonly used. Note that techniques compared to the work of Jarosinski et al.261 However,
the temperature rise criterion is useful to determine cool flame in the work of Jarosinski et al. a nonvisible flame propagation
propagation, as applied by Cato et al.128 characterized by a pressure rise was reported for a propane−air
mixture with a propane content of 9.0%. Therefore, the
4. EFFECT OF BUOYANCY AND STUDIES IN difference in UFL values is probably related to the flammability
MICROGRAVITY criteria adopted.
As shown in section 3, the flame propagation direction affects Ronney262 studied the effects of transport properties and
the experimentally determined values of FLs, mainly due to the chemical kinetics for near-limit flames in microgravity. The
buoyancy force in normal gravity, which promotes UFP and effects of chemical kinetics were evaluated by comparing the
opposes DFP. The buoyancy effect in the experiments makes it behaviors of mixtures with similar Lewis numbers but different
difficult to understand the flame dynamics near the limiting chemical kinetic mechanisms. The mixtures studied were lean
conditions. Since different experimental methods and con- NH3−air (ammonia−air), rich C3H8−air, and lean CH4−air
O https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

Figure 13. Results of experiments under microgravity for methane−air mixtures. (a) Comparison of LFL under microgravity with those under normal
gravity for different initial pressures. Reproduced with permission from ref 59. Copyright 1985 Elsevier. (b) Stagnation velocity gradient at the
extinction limits (2U/L), for different equivalence ratios under microgravity. Reproduced with permission from ref 264. Copyright 1996 Elsevier.

with Lewis numbers slightly less than unity (Le < 1). In addition, Wachman,59 and Wang et al.,265 among others. Interestingly,
rich NH3−air and lean C3H8−air mixtures were considered the laminar flame speed determined by Wang et al.265 for a
whose Lewis numbers were greater than unity (Le > 1). At methane−air mixture at the LFL of 1.20 cm/s is consistent with
atmospheric pressure, the FLs of NH3−air mixtures in the flame speed theoretically determined by Spalding30 (using
microgravity were between the FLs with UFP and DFP in air properties).
normal gravity. To evaluate the influence of transport processes, 4.2. Flame Balls in Microgravity. One of the most
Ronney262 studied the behaviors of H2−O2−N2 and CH4−O2− important findings in research on combustion under zero gravity
N2 mixtures with different stoichiometries. In these cases, each is Ronney’s flame ball.44 To understand the competing physical
set of these mixtures has similar chemical kinetic mechanisms, processes, Ronney compared the time scales of the various key
while the effective Lewis number was changed by varying the processes in flames. The time scales for chemical reaction
stoichiometry. (tchem), for inviscid buoyant convection (tinv), for viscous
The role of transport processes was then investigated by buoyant convection (tvis), for heat loss by conduction and
comparing mixtures with similar chemical kinetic mechanisms convection to the walls (tcond), and for heat loss by radiation
but different transport properties, such as lean CH4−air (trad) are given in eqs 48−52.
compared to lean C3H8−air and lean NH3−air compared to
tchem
rich NH3−air. From the above comparisons, Ronney262 SL 2 (48)
concluded that transport properties, rather than chemical
kinetics mechanisms, play the most important role in explaining ij d yz
1/2
the behavior of near-limit flames in microgravity. Figure 13a t inv jj zz
jj g zz
shows the lower flammability limits of methane−air mixtures at k { (49)
different initial pressures under normal gravity and microgravity
1/3
conditions. Ronney263 also studied the extinction processes in jij v zyz
microgravity of lean mixtures of CH4 and NH3 and rich mixtures tvis jj 2 zz
jg z
of C3H8, all in air. It was found that for mixtures with Le ∼1 the k { (50)
FLs in microgravity are probably caused by radiative heat losses
from the product gases. d2
tcond
Experiments with premixed methane−air flames in counter- 16 (51)
flow and microgravity were performed by Maruta et al.264 It was p
observed that there are two extinction limits that appear on a trad 4
curve obtained by plotting the stagnation velocity against the 14 b (Tf T 4) (52)
extinction equivalence ratio. Moreover, the upper and lower The comparison of time scales was performed by Ronney for
branches of this curve merge at a point when the equivalence stoichiometric and near-limit flames, and the results of this
ratio is sequentially reduced. The results of the experiments are comparison are shown in Table 3. It can be seen in Table 3 that
shown in Figure 13b. The stagnation velocity gradient is defined buoyancy convection is unimportant for stoichiometric flames,
as 2U/L, where U is the flow velocity at the burner outlet and L is as can be deduced by comparing their time scales with the
the distance between the two tubes. In Figure 13b, the upper chemical reaction time scale (tinv ≫ tchem, tvis ≫ tchem). For flames
part of the curve corresponds to the stretch extinction limit, the near the limit, on the other hand, this relationship is reversed,
lower part corresponds to the radiative heat loss extinction limit, and buoyancy convection is actually very important under
and the inflection point is the flammability limit of a stretched normal gravity conditions (tchem ≫ tinv, tchem ≫ tvis). For flames
flame. near the limit, the time scales for radiation and chemical reaction
Experimental data on laminar flame speeds in zero gravity can are of the same order of magnitude. However, the time scale for
be found in the works of Ronney,262,263 Ronney and buoyant convection is much smaller than that of radiation (trad
P https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

Table 3. Comparison of Characteristic Time Scales of stable. It was Ronney who observed a stable flame ball in a
Combustion Processes for Stoichiometric and near Limit microgravity experiment.
Flamesa A schematic representation of a stationary spherical flame is
shown in Figure 14a. These flames were characterized by a
time scale
radius rf, and the fuel and oxidant were transported by diffusion
stoichiometric near limit from the surroundings into the reaction zone. At the same time,
flame flame
the combustion products and heat are being diffused outward
chemical reaction (tchem) 0.00094 s 0.250 s
from the reaction zone. In experiments conducted by Ronney267
inviscid buoyant convection (tinv) 0.071 s 0.071 s
at the NASA Lewis Research Center in microgravity, he
viscous buoyant convection (tvis) 0.012 s 0.010 s
observed the formation of flame balls for mixtures containing 7%
conduction heat loss to walls (tcond) 0.950 s 1.400 s
H2 in air with an addition of 4% bromotrifluoromethane
radiant heat loss (trad) 0.130 s 0.410 s
a (CF3Br) and characterized by a low Lewis number.
Reproduced with permission from ref 44. Copyright 1998 Elsevier. The small flame balls shown in Figure 14b had a visible radius
that ranged from 2.0 to 4.0 mm (approximately). In another
≫ tinv, trad ≫ tvis). Therefore, the effect of radiation would only study, Ronney et al.268 increased the duration of microgravity
be observable under low gravity experimental conditions. from a time of 2.2 s determined in the work of Ronney267 to a
If there is no convection and the time scales of heat/mass time 10 times longer using aircraft-based experiments. Burning
diffusion and reaction rate become comparable, a stationary velocities observed by Ronney267 in drop tower experiments
flame can exist. The possibility of the existence of stationary ranged from 0.25 to 2.5 cm/s for H2−air mixtures with H2
spherical flames was theoretically predicted by Zeldovich.266 concentrations between 5 and 7.5% (near the LFL). These
However, the theory predicted that an adiabatic flame ball is not burning velocities are of the same order of magnitude as those

Figure 14. Stationary spherical flames and flame balls observed in microgravity. (a) Schematic representation of stationary spherical flame.
Reproduced with permission from ref 44. Copyright 1998 Elsevier. (b) Flame balls observed in drop tower experiments for mixtures of 7% H2 in air
with the addition of 4% CF3Br. Reproduced with permission from ref 267. Copyright 1990 Elsevier. (c) Flame balls observed in aircraft experiments for
mixtures of 3.5% H2 in air. Reproduced with permission from ref 268. Copyright 1994 American Institute of Aeronautics and Astronautics. (d) Flame
balls observed in space shuttle experiments for mixtures of 4.9% H2, 9.8% O2, and 85.3% CO2. Reproduced with permission from ref 269. Copyright
1998 American Institute of Aeronautics and Astronautics.

Q https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

Figure 15. Experimental results for counterflow twin H2−air premixed flames, with jet spacing of 1.27 cm. (a) Stability and extinction limits. (b) Near-
planar flames for 8.26% H2 and 60 s−1 strain rate. (c) Wrinkled flames for 8.10% H2 and 110 s−1 stretch rate. (d) Moving tube flames for 6.96% H2 and
60 s−1 strain rate. (e) Stationary twin tube flames for 6.68% H2 and 60 s−1 strain rate.270 and (f) Stationary single tube flame for 6.64% H2 and 60 s−1
strain rate. (a−e) Reproduced with permission from ref 270. Copyright 2000 American Institute of Aeronautics and Astronautics.

predicted by Spalding’s theory. The flame balls were also presented in the work of Ronney et al.,269 where flame balls were
observed in aircraft experiments for various H2−air mixtures formed for mixtures of H2−O2−CO2 and H2−air. The
characterized by low Lewis numbers268 [Figure 14c]. The visible
radius of these flame balls was about 3 mm. Further confirmation maximum duration of the experiments was 500 s, and the
was provided by experiments performed in a space shuttle and flame balls for the H2−O2−CO2 mixture persisted all the time
R https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

Table 4. Comparison of Available Methods for the Determination of FLs in Air at Reference Conditions
reference limit NTot Ncorr NPred AARE (%) R2 type of compound
Methods for the Estimation of the LFL
Albahri274 LFL 472 454 18 4.10 0.9300 several
Albahri275 LFL 543 518 25 2.28 0.9998 several
Bagheri et al.276 LFL 1615 1292 323 15.29 0.9295 several
Frutiger et al.277 LFL 443 − − 11.50 0.9900 several
Gharagheizi278 LFL 1056 845 211 7.68 − several
Gharagheizi279 LFL 1057 846 211 4.62 0.9860 several
Kondo et al.280 LFL 238 − − 18.06 0.8665 several
Lazzús281 LFL 418 328 90 − 0,9865 several
Mendiburu83 LFL 324 243 81 5.32 0.9591 C−H
Mendiburu et al.33 LFL 374 273 101 5.43 0.9752 C−H−O
Pan et al.282 LFL 1038 830 208 5.60 0.9787 several
Rowley et al.283 LFL 509 309 200 10.67 − several
Seaton284 LFL 150 − − 5.13 0.9904 several
Methods for the Estimation of the UFL
Albahri274 UFL 477 464 13 11.80 0.9600 several
Frutiger et al.277 UFL 351 − − 15.90 0.9100 several
Gharagheizi285 UFL 865 693 172 9.56 0.9209 several
High and Danner286 UFL 181 − − 26.40 − several
Kondo et al.280 UFL 238 − − 19.14 0.4197 several
Lazzús281 UFL 418 328 90 7.10 0.9776 several
Mendiburu83 UFL 247 158 89 8.32 0.9202 C−H
Mendiburu et al.35 UFL 208 136 72 8.95 0.9346 C−H−O
Pan et al.287 UFL 579 465 114 19.23 0.7540 several
Seaton284 UFL 141 − − 9.68 0.8862 several

[Figure 14d]. The visible radii of the flame balls were 2.5−4.0 It is important to note that a stationary single-tube flame was
mm (approximately). observed at an H2 concentration as low as 6.11%. This result
Flame balls are more stable when the mixture is near the clearly shows that flame curvature or flame stretch has extended
flammability limits, where thermal radiation plays a critical role the flammable region. A single-tube flame with a higher stretch
in stabilizing the flame ball. Flame balls are unstable for mixtures could burn much leaner. This observation is consistent with the
with a Lewis number close to unity or greater than unity; predicted G-shaped flammability limit of a stretched flame in
therefore, these flames were not observed for lean CH4−air and Figure 6.
C3H8−air mixtures, which have Le numbers of 0.9 and 1.7,
respectively. 5. METHODS FOR ESTIMATION OF THE
These studies have shown that the Lewis number is important FLAMMABILITY LIMITS IN AIR AT REFERENCE
for the formation of the flame ball and for extending the burning CONDITIONS
limits. To better understand how flow stretch and flame Several authors have developed methods for the determination
curvature affect the burning limits of a fuel at a low Lewis of FLs in air under reference conditions. These methods are
number, Kaiser et al.270 measured the flame regimes and burning listed here, and their accuracy is compared. However, in this
limits of hydrogen flames in a counterflow slot burner under section, only the methods that have better accuracy are analyzed
normal gravity conditions. Two flame configurations for in more detail. The estimation methods are of great importance
premixed H2−O2-N2 flames were tested: (1) the single flame for industrial applications.
(premixed gas versus inert gas) and (2) the twin flame In general, methods for estimating FLs can be divided into
(premixed gas versus premixed gas). Only the results of the two groups: (1) methods that use quantitative properties of
second case are briefly discussed here. molecular structure or group contributions together with linear
The experiments were performed varying the equivalence correlation or neural networks; (2) methods that use calculated
ratio and the global strain rate σst = (Vupper + Vlower)/d, where adiabatic flame temperature (CAFT) together with linear
Vupper and Vlower are the exit velocities of the upper and lower jets, correlation and combustion theory. Note that group contribu-
tion methods are well-established for the determination of
respectively, and d is the distance between the nozzles’ exits. As
various properties of fluids, as shown in classical hand-
can be seen in Figure 15a, there are two extinction limits for a books,271,272 and several authors have applied these methods
solid fuel composition: one at high σst due to stretch extinction to develop correlations for the determination of FLs. On the
and another at low σst due to heat loss. For a fixed σst with other hand, CAFT methods are based on the theory of
decreasing H2 concentration, it was observed that the nearly equilibrium combustion. In this approach, if the adiabatic
planar twin flames [Figure 15b] gradually become wrinkled flame temperature is known and equilibrium is assumed, the
flames [Figure 15c] and multiple moving tube flames [Figure composition of the reactants can be calculated.
15d]. With even further decrease of the H2 concentration, the The available methods for determining FLs are semiempirical,
observations show stationary twin tubes [Figure 15e], stationary meaning that a total set of compounds is divided into a
single-tube flames [Figure 15f], and finally extinction. correlation set and a prediction set. The correlation set is used to
S https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

obtain a correlation or train a neural network, and the prediction where the first, second, and third order group contribution
set is used to test the results obtained in the previous step. The factors are accounted for by the terms Cj, Dk, and El, respectively.
accuracy of these methods is evaluated by the absolute value of These terms are provided by Frutiger et al.277 as Supporting
the relative error (ARE), the average value of the relative error Information in their work. Nj, Mk, and Ol are the number of
(AARE), and the squared correlation coefficient (R2), which are occurrences of a given group in the molecular structure. To
determined by eqs 53−55. know how many groups of each order are present in a
|FLcalc FLexp| compound, it is necessary to analyze the 2D structure of the
ARE = · 100% molecule.
FLexp (53) It is interesting to show how this method can be applied to a
N compound. Ethanol is considered as an example. From the
1 CAMEO Chemical Datasheet,88 the experimental LFL and UFL
AARE = AREi
N i=1 (54) of ethanol are 3.3% and 19.0%, respectively. The 2D structure of
the ethanol molecule can be found in ref 288. The groups
SSE observed in the structure are methyl radical (CH3), methylene
R2 = 1
SST (55) (CH2), and OH. Each of the groups exhibits Nj = 1. In the
The sum of squares of the errors (SSE) and the total sum of Supporting Information of Frutiger et al.,277 these are classified
squares (SST) are given in eqs 56 and 57.273 as first order groups; therefore, the terms Mk and Ol are both
N
zero. The values of Cj for the LFL are −0.24, −0.23, and 0.06 for
SSE = (FLcalc, i FLexp, i) each group; on the other hand, the values of Cj for the UFL are
i=1 (56) −1.15, −0.14, and −0.76 for each group.277 The results of
applying eqs 58 and 59 are an LFL of 3.00% and a UFL of
N 2 16.73%. As can be observed, the application of this method is
SST =
N
(FLexp, i)2
( i=1 )
FLexp, i
simple, and the accuracy is good. Frutiger et al.277 provide
i=1
N (57) another sample calculation in their paper and also develop a
method to obtain a 95% confidence interval.
The comparison of the different methods is given in Table 4. There are several group contribution methods similar to the
The total number of compounds considered is NTot, the number one developed by Frutiger et al.277 In general, these methods can
of compounds in the correlation set is Ncorr, the number of
be used without much difficulty to accurately estimate
compounds in the prediction set is Npred, the average of the
relative error of the whole data set, AARE, and the squared flammability limits under reference conditions of temperature
correlation coefficient of the whole data set, R2, are also shown in and pressure.
the Table. 5.2. Calculated Adiabatic Flame Temperature (CAFT)
Most of the methods considered rely on knowledge of the Method for FLs in Air at Reference Conditions. There is
molecular structure of the compound to directly estimate the FL another type of method that assumes that the adiabatic flame
or to find the necessary parameters for estimating the FL. In temperature can be approximated by a correlation283 or by using
some cases, it is also necessary to use the neural network an average value.289 With the use of the value of the adiabatic
developed in the specific work. Frutiger et al.277 reported that flame temperature, the FL can then be calculated by a simple
Gharagheizi,278,279,285 Pan et al.282,287 and Bagheri et al.276 used procedure assuming equilibrium in the combustion products. In
experimental and predicted values in developing their methods. this section, a method that can account for the LFL and UFL of
Therefore, these methods are not recommended. In sections 5.1
C−H and C−H−O compounds is presented. This method is a
and 5.2, methods based on group contribution properties and
another using CAFT are considered. The results presented in synthesis of previous work by Mendiburu et al.32−35 to
Table 4 show that most of the listed methods have R2 determine the FLs of C−H and C−H−O compounds. For
coefficients above 0.90 and AAREs below 20%. simplicity, a C−H compound is one that contains carbon and
5.1. Group Contribution Based Method for FLs in Air at hydrogen atoms in its molecule. A C−H−O compound is
Reference Conditions. The method developed by Frutiger et analogously defined as a compound containing carbon,
al.277 was selected from those presented in Table 4 for its hydrogen, and oxygen in its molecule.
accuracy and consistency, since the authors used only The combustion process is assumed to occur under adiabatic
experimental data to train the model. The structure of the and constant pressure conditions. The nitrogen and excess
compound must be known to determine the first, second, and oxygen are considered inert. In addition, complete combustion
third order group contributions. The FLs are calculated using is assumed in the LFL and the products of combustion are
the expressions presented in eqs 58 and 59. assumed to be in chemical equilibrium in the UFL. There are
ÄÅ É
ÅÅÅ ij yzÑÑÑÑ three possible combustion reactions: one in the LFL and two in
Å j zÑ
LFL = 4.53ÅÅÅexpjjj NC j j + Mk Dk + Ol El zzzÑÑÑ the UFL. These reactions are shown in eqs 60−62.
ÅÅ jj zzÑÑ
ÅÅÇ k j k l {ÑÑÖ (58) combustion reaction at the LFL:
ÄÅ É
ÅÅ i yzÑÑÑÑ
LFL
ÅÅ jj CxCH x HOxO + vair (O2 + 3.76N2)

UFL = 129.96ÅÅÅexpjjj NC
j j + Mk Dk + Ol El zzzÑÑÑ xH
ÅÅ jj zzÑÑ xCCO2 + LFL
H 2O + (vair s
vair LFL
)O2 + 3.76vair N2
ÅÅÇ k j k l {ÑÑÖ 2
(59) (60)

T https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

combustion reaction at the UFL, without condensate: or 62], on the compound in question, and on the equilibrium
UFL
constants, Ksc for eq 65 and Kcc for eq 66. The expressions
CxCH x HOxO + vair (O2 + 3.76N2) needed to determine the coefficients are given below.
Coefficients for the reaction without condensate are defined
nCO2CO2 + nCOCO + nH2OH 2O + nH2 H 2
in eqs 68−75.
UFL
+ 3.76vair N2 (61) av = am2 + 2am (68)
combustion reaction at the UFL, with condensate: xH
UFL
bv = 2amn + amxO + 2an (2a + 1)mxC m 2xC
CxCH x HOxO + vair (O2 + 3.76N2) 2
(69)
UFL
nCOCO + nCC + nH2OH 2O + nH2 H 2 + 3.76vair N2 xH
(62) cv = an2 (2a + 1)nxC + (an xC)xO (n xC)
2
In the case of the LFL, the energy conservation equation is + 2xC 2 (70)
used to relate the stoichiometric coefficient of air (vLFL
air ), which is
equal to the number of oxygen moles in the air, to the adiabatic hair,R 2h H 2 2 h H 2O 3.76h N2
flame temperature at the LFL (TLFL). Isolating then vLFL air , one m=
obtains the expression shown in eq 63. hCO hCO2 + h H2O hH2 (71)
s xH xN
hF,R + vair hO2 xChCO2 h h xH
LFL
vair = 2 H 2O 2 N2
n=
hF,R xChCO2 + (2xC xO)h H2O + xO ( 2 )
2xC h H2
hair hair, R (63) hCO hCO2 + h H2O h H2
where the subscript “R” refers to the reactants and all other (72)
enthalpies correspond to the products. The h̅air is defined as a=1 K sc (73)
shown in eq 64, while a similar expression determines h̅air,R. Note
that in order to evaluate the enthalpies of the products, it is UFL xH
necessary to know the value of TLFL. b = 2avair xC(2a + 1) + axO
2 (74)
hair = hO2 + 3.76h N2 (64) i x UFL y
c = xCjjj2xC + H xO 2vair zz
z
In the case of the UFL, the reaction without condensate k 2 { (75)
represents the combustion process and provides the solution
Coefficients for the reaction with condensate are defined in
when there is sufficient oxygen. On the other hand, if there is not
eqs 76−84.
enough oxygen, the reaction with condensate represents the
combustion process and provides the solution. In such cases, the av = am2 2am + 3.76(a 1)m 7.52(a 1) (76)
condensate is represented by carbon graphite (C). It is therefore
necessary to find a solution for both reactions, with and without ix y
condensate. The equilibrium condition is applied by the bv = 2amn + amjjj H xOzzz 2an + (a 1)
k 2 {
homogeneous and heterogeneous water−gas reactions. The
former is used for the reaction without condensate and the latter [3.76(n xO) x H] (77)
for the reaction with condensate. These reactions are shown in
ix y xH
eqs 65 and 66, respectively. cv = an2 + anjjj H xOzzz (a 1) xO
water−gas homogenous reaction: k 2 { 2 (78)

H 2O + CO CO2 + H 2 (65) hair,R 2h H 2 2 h H 2O 3.76h N2


m=
water−gas heterogenous reaction: hCO hCO2 + h H2O hH2 (79)
C + H 2O CO + H 2 (66) xH

The criterion for determining whether or not the solution n=


hF,R xChC xOh H2O + xO ( 2 )hH 2

should contain condensate is that the mole numbers of all hCO hH2 h H 2O hC (80)
species must be positive. In most cases, only one of the reactions
will meet this criterion. However, if both reactions meet the a = 1 + Kcc(p0 /p) (81)
criterion, then the UFL obtained in the reaction with condensate
will be greater than that in the reaction without condensate. xH UFL UFL
b= a xOa + 3.76vair (a 1) 2avair
From the point of view of safety, the first should be preferred to 2 (82)
the second. This criterion was applied by Mendiburu et al.34,35 in ÅÄÅ x
their previous work. By use of the law of conservation of energy c = (1 a)ÅÅÅÅx HvairUFL UFL
+ 3.76xOvair + H xO
and the equilibrium condition, it is possible to obtain a quadratic ÅÇ 2
equation vUFL ÑÉ
air for the two reactions under consideration. UFL 2 Ñ
+ 7.52(vair ) ÑÑÑÑ
UFL 2
av(vair UFL
) + bvvair + cv = 0 (67)
ÑÖ (83)

The coefficients of eq 67 depend on the adiabatic flame xH UFL


nT = nCO + + 3.76vair
temperature in the UFL (TUFL), on the reaction chosen [eq 61 2 (84)

U https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

Table 5. Coefficients for Use with Correlations Shown in Eqs 85 and 86


LFL UFL
C−H C−H−O C−H C−H−O
coefficient
a0 1.4774 1.5101 2.3737 2.5526
a1 1.0186 × 10−2 −6.7195 × 10−3 −2.3822 × 10−1 −1.7112 × 10−1
a2 −2.0090 × 10−3 1.3081 × 10−2 1.2076 × 10−2 −2.9203 × 10−2
a3 6.4563 × 10−2 8.4976 × 10−2 2.8293 × 10−2 −5.8587 × 10−4
a4 1.1814 × 10−2 6.4285 × 10−3 4.6652 × 10−2 −6.1402 × 10−2
a5 − − 9.8788 × 10−4 −4.1708 × 10−3
a6 − − −1.8129 × 10−4 −4.9859 × 10−5
a7 − − 5.2138 × 10−2 −4.9785 × 10−1
a8 − − −1.9116 × 10−1 1.1146 × 10−1
a9 − − 4.3725 × 10−2 −4.3060 × 10−3
a10 − − −2.0102 × 10−2 2.1873 × 10−2
a11 − − −1.3389 × 10−1 −1.6556 × 10−1
N 324 374 247 208
AARE (%) 5.53 5.57 8.46 9.24
R2 0.9608 0.9756 0.9312 0.9318

The ratio of the adiabatic flame temperature in stoichiometric implemented in Excel or Matlab to obtain a straightforward
combustion (Tstq) to TLFL is obtained from eq 85. In the same application.
manner, the ratio of Tstq to TUFL is obtained from eq 86. It is important to discuss the relationship between CAFT
Therefore, the value of Tstq must be calculated beforehand. The methods and the most fundamental methods based on chemical
expressions for xi are shown in eq 87, and the values for the kinetics and radiative heat losses discussed in section 2. The
coefficients ai are presented in Table 5. methods based only on the competition between chain-
branching and chain-termination reactions cannot be directly
Tstq
= a0 + a1x1 + a2x2 + a3x3 + a4x4 related to the CAFT methods because the latter assume
TLFL (85) chemical equilibrium and do not analyze the steps between
reactants and products.51,53,54
Tstq Nevertheless, CAFT takes into account some features of the
= a 0 + a1x1 + a 2x12 + a3x13 + a4x2 + a5x2 2 latest methods, which consider detailed chemical kinetic
TUFL
mechanisms and radiative heat losses. For example, Bertolino
+ a6x2 3 + a 7x3 + a8x32 + a 9x33 + a10x4 + a11x41/2 et al.64 introduced soot formation to determine the UFL. This
(86) feature is introduced in the CAFT method simply by assuming a
reaction containing carbon graphite in the products. In addition,
0 0 the method of Bertolino et al. identifies the FLs when there is a
hf,F hf,F
x1 = ; x2 = ; sudden drop in the temperature of the combustion products.
WF xCx H(xO + 1) The theory developed by Spalding30 predicts the FLs by
WF xC(xO + 1) assuming a one-step chemical kinetic reaction and applying
x3 = ; x4 = radiative heat losses to a propagating flame. This theory assumes
xCxH(xO + 1) xH (87)
knowledge of some properties that allow determination of
The above correlations are based on the earlier work of radiative heat losses. CAFT methods assume infinitely fast one-
Mendiburu et al.32−35 The original correlations were modified to step chemical reactions that reach equilibrium and replace
obtain two correlations: one for the LFL and one for the UFL. radiative heat losses with knowledge of an approximate adiabatic
The accuracy of the method was not significantly affected, as flame temperature. These features explain the success of the
shown by comparing the values for AARE and R2 reported in simple CAFT methods in practical applications where FLs must
Table 5 with those in Table 4. The value xO = 0 corresponds to be estimated.
C−H compounds; therefore, (xO + 1) is used in x2, x3, and x4. In Before concluding this section, a discussion of soot formation
a few cases, the method may converge to two UFL values: one is necessary because it is important to UFL. Soot can be formed
for the reaction without condensate and another for the reaction by two mechanisms: by a homogeneous reaction in the vapor
with condensate. In these cases, the two UFL values are close phase or by pyrolysis in the liquid phase. The amount of soot
and the higher value is chosen. formed from a fuel depends on several physical parameters,
A calculation example for this method is presented by including the type of flame. In premixed flames, aromatics tend
Mendiburu et al.35 in the Supporting Information of their paper. to form soot the most, followed by alkanes, alkenes, and alkynes.
Application of the method to the case of ethanol in air under It is also known that in premixed flames increasing the
reference conditions gives an LFL of 3.27% and a UFL of temperature reduces the tendency to form soot.27
18.71%. These values are close to the experimental values of The sooting limit is defined at the experimental condition that
3.3% and 19.0%, respectively.88 It is interesting to note that produces the first luminous continuum radiation on a premixed
Jarosinski et al.256 found that the adiabatic flame temperatures flame. The sooting limit of several fuels was measured by
near the FLs of propane in air are about 150 °C higher than the Takahashi and Glassman290 in a specially designed apparatus
experimental temperatures. The present method can be using a Bunsen burner. The mass flow rate of the oxidizer was
V https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

Figure 16. Effect of mixture initial temperature on FLs for different compounds in air: (a, b) hydrogen,16,291 (c) dimethyl ether,182 (d) methyl
formate,182 (e−g) ethylene;212,220,291 (h, i) methane.133,291 (a−i) Reproduced with permission from ref 292. Copyright 2017 Elsevier.

kept constant while the mass flow rate of the fuel was increased The above definition can be expressed as ψc = [(xc/2) + (xH/
until the flame emitted a luminous continuum radiation. The 4)]/nO2,exp. As the mass flow rate of the fuel is increased in the
d

critical sooting equivalence ratio (ψc) is defined as the ratio of


experiments, nO2,exp = (ṁ O2,exp/ṁ F,exp)(WF/WO2) then decreases
the stoichiometric number of moles of oxygen required to form d d d

CO and H2O in the products to the experimental number of until the sooting limit is reached. Therefore, lower values of ψc
moles of oxygen per mole of fuel at the sooting limit.290 imply a lower sooting tendency because a higher mass flow rate
W https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

of the fuel was required to find the sooting limit. Following the FLT hC,0 ij hair, T yzz
results of Takahashi and Glassman,290 it is found that at the same jj1 zz
= j
flame temperatures, ψc,ethane > ψc,ethylene > ψc,acetylene. One of the FL0 hC, T + hF, T hair, T jk FL0hC,0 z{ (90)
most important parameters for soot formation in premixed From the experimental observations, it is expected that FLT/
flames is the flame temperature, since an increase in temperature FL0 = 1 ∓ mΔT, where the minus sign (−) corresponds to the
increases the oxidation rate of the soot precursors due to the LFL and the plus sign (+) corresponds to the UFL. Therefore,
attack of OH radicals. In addition, the C/H ratio of the fuel eq 90 can be solved for the slope m. In this expression, ΔT = T −
molecule is also important, as it determines the OH T0.
concentration.
1
ji hC, T hF, T hair, T zyz jij har, T hC, T
6. TEMPERATURE AND PRESSURE EFFECTS ON FLS m = ± jjjj + zz
z
jj1
j
k hC,0 hC,0 { k FL0hC,0 hC,0
The initial temperature and pressure conditions have an
influence on the values of FLs. Therefore, it is necessary to hF, T y
hair, T zz 1
+ zz
evaluate the influence of these parameters. First, the temperature hC,0 z T
effect is considered by giving a general description in section 6.1, { (91)
a discussion of Burgess−Wheeler’s law in section 6.1.1, and a It is reasonable to assume that the term (Δh̅F,T − Δh̅air,T)/hC,0
description of available methods in section 6.1.2. Then, in is small and can be neglected compared to the other terms of eq
section 6.2, the pressure effect is considered. Finally, section 6.3 91. Moreover, the sensible enthalpy of air can be represented by
presents the available empirical equations for determining the an average heat capacity at constant pressure Δh̅air,T = cp̅ ,airΔT.
FLs of some compounds in air at various initial temperatures and Therefore, the expression for the slope can be simplified as given
pressures. in eq 92.
6.1. Temperature Effect on FLs. The effect of the initial
ij hC,0 hC, T yz cp ,air
temperature of the mixture on the FLs is determined m = ± jjjj FL0 1zzzz
experimentally by holding the pressure constant. An example j cp ,air T z FL0hC, T
of such observations is shown in Figure 16. The experimental k { (92)
data shown in Figure 16 are for hydrogen,16,291 dimethyl Substituting eq 92 into eq 90, an expression for the variation
ether,182 methyl formate,182 ethylene,212,220,291 and meth- of the FL with the initial temperature is obtained. The simple
ane.133,291 The results of the estimation method developed by expression shown in eq 93 is the modified Burgess−Wheeler law.
Mendiburu et al.292 are also shown in Figure 16. The
FLT ij hC,0 hC, T yz cp ,air
= 1 ± jjjj 1zzzz
experimental data show that the LFL decreases with the increase
FL0 T
of initial temperature, while the UFL increases with the increase FL0 j cp ,air T z FL0hC, T
of initial temperature. Moreover, the flammability limits vary k { (93)
almost linearly with the initial temperature. As Glassman et al.27 In the methods that use detailed chemical kinetic mechanisms
noted, increasing temperature accelerates chemical reactions; and include radiative heat losses, the extension of the
therefore, it is reasonable that the flammable range increases as flammability range is mainly caused by the increase in reaction
the initial temperature increases. rates of some key reactions. However, in the case of the modified
Assuming that the oxidant is air, the expression in eq 1 can be Burgess−Wheeler law, the assumption that the enthalpies of the
written as FL = 100/(1 + na/nF). From the point of view of the reactants at the FLs are constant can be related to the fact that
ratio of air to fuel (na/nF), this ratio increases in the LFL and some chain-initiation and chain-branching reactions require a
decreases in the UFL as the initial enthalpy of the mixture is certain amount of energy for a flame to propagate.
increased. This means that the reactant in excess can be 6.1.2. Methods for Determination of FLs at Different Initial
increased when the initial temperature is higher because the Temperatures. There are some methods for the determination
chemical reactions are accelerated at a higher initial temperature. of FLs at different initial temperatures and atmospheric pressure.
6.1.1. The Modified Burgess−Wheeler Law. The modified The method of Zabetakis24 follows directly from eq 93. It
Burgess−Wheeler law states the following: “For a flammable gas assumes that the heat released is constant (hC,0 = hC,T) and
the sum of the heat released at the LFL and the sensible corresponds to the heat released by complete combustion of 1
enthalpies of the reactants is constant.”24 The practical mol of fuel (HC). The values of 0.0075 kcal/(mol·K) and 10.4
implications of such a simple law are of great importance for kcal/mol can be assumed for cp̅ ,air and FL0HC, respectively. The
industrial applications. This law can be written as shown in eq expression obtained is shown in eq 94, where the minus sign (−)
88, where the index T refers to the initial temperature, hC,T is the corresponds to the LFL and the plus sign (+) to the UFL.
heat released at T, and K is a constant. FLT
=1 0.000721(T T0)
FLT[( hF, T hair, T ) + hC, T ] + hair, T = K FL0 (94)
(88)
283
The method developed by Rowley et al. uses the expression
Assuming that FL0 is a known flammability limit value at the shown in eq 95 to approximate LFLT.
reference temperature T0, Burgess−Wheeler’s law can be written
as shown in eq 89. (1 ) cp ,F air
LFLT (%) = LFL0 (T T0)
HC (95)
FL0hC,0 = K (89)
where
Dividing eq 88 by eq 89, and rearranging, the expression
cp,F air = LFL0 cp ,F, T0 + (100 LFL0) cp ,air, T0 (96)
shown in eq 90 is obtained.
X https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

Table 6. Comparison of Methods for the Determination of the FLs at Different Initial Temperatures and Constant Pressure
LFL UFL
method N AARE (%) R2 N AARE (%) R2
24
Zabetakis 180 4.82 0.9948 129 5.01 0.9748
Mendiburu et al.292 180 3.25 0.9928 129 3.60 0.9957
Britton and Frurip293 161 4.44 0.9946 − − −
Rowley et al.283 109 4.95 0.9857 − − −
Catoire and Naudet295 140 14.64 0.9003 − − −

= 0.0125xC2 0.779 (97) the fuel and the initial temperature. The adiabatic flame
293
temperatures for the stoichiometric combustion at different
Another method, developed by Britton and Frurip and initial temperatures (Tstq,T) are also needed in this method. The
Britton,294 uses the expression shown in eq 98 to approximate ratio given in eq 106 increases linearly with an increase in the
the LFLT. initial temperature of the mixture.
TLFL,0 T1 * T /TUFL,
Tstq, * T
LFLT = LFL0 UFL, T
= = 1 + kUFL(TT T0)
TLFL,0 T0 (98) * /TUFL,0
*
UFL,0 Tstq,0 (106)
The value of the adiabatic flame temperature in the LFL with
where the subscript “0” refers to the temperature T0. The slope
initial temperature T0 is denoted TLFL,0 and is determined by the
kUFL is approximated by the correlations presented in eqs 107
correlations given in eqs 99−101. Catoire and Naudet295 have
*
and 108. The value of TUFL,T is determined by eq 109.
also developed a method using the correlation shown in eq 102.
for C−H compounds and hydrogen:
hydrocarbons:294
4
ij H yz kUFL = 3.7611 × 10 3.0500 × 10 4I
TLFL,0 = 2181 7.4613jjj sC zzz
j vair z + 2.5988 × 10 6hf,F
° + 2.0798 × 10 8(hf,F
° )2
k { (99)
(107)
C−H−O−N compounds:294
for C−H−O compounds and carbon monoxide:
ij H yz
TLFL,0 = 2290 7.6944jjj sC zzz kUFL = 5.8362 × 10 3
3.8838 × 10 4I
j vair z
k { (100)
6.7817 × 10 5hf,F
° 1.4430 × 10 7(hf,F
° )2
293
C−H−O−Cl compounds: (108)
ij H yz 1
TLFL,0 = 2427 8.3846jjj sC zzz * T
Tstq, ij y
UFL, T z
j vair z * T= jj z
zz
* jj
k { (101) TUFL, z
* /TUFL,0
Tstq,0 k UFL,0 { (109)
s
LFL = 519.957(1 + 5vair ) 0.70936 xC 0.197T 0.51536 (102) Finally, the UFL is determined using the number of moles of
292
Mendiburu et al. have developed methods to estimate the oxygen obtained by eq 110, which was derived by algebraically
FLs at different initial temperatures. In the case of the LFL, a manipulating the energy conservation equation by replacing the
correlation for the slope m is obtained by using a parameter I. excess fuel with an equal mass of air. The molecular weight of air
This parameter is given in eq 103 and relates the molecular used in eq 110 is Wair = 28.96 g/mol, and the UFL is determined
weight of the fuel (WF) to the product of the heat released by UFL = 100%/(1 + 4.76vUFL air ) .
during combustion of 1 mol of fuel (HC) and the number of UFL
vair = [vair s
(WF/Wair)(hair,P hair,R )]
moles of a gaseous fuel−air mixture (nF*) occupying a volume of ÄÅ
12 L at T0. The correlation for the slope at the LFL is given in eq ÅÅÅh xH
ÅÅ F xChCO2 h H O + (WF/Wair)
104. Finally, the LFL is determined by eq 105. ÅÇ 2 2
ÉÑ
Ñ 1
3.76h N2)ÑÑÑÑ
WF s
I= (hair,P hair,R ) + vair (hair,R
*
nF HC (103) ÑÖ (110)
Note that in this method the excess fuel is treated as a heat
mLFL,calc = 8.3959 × 104 + 9.6643 × 10 5I sink. The above methods are compared in Table 6. In another
comparison of the methods for determining the LFL at different
+ 2.6402 × 10 6I 2 + 8.3413 × 10 10
HCI
initial temperatures, Mendiburu et al.292 showed that their
(104) method is more accurate for compounds with xC > 5 and for
LFLT = LFL0[1 mLFL,calc(T T0)] (105)
hydrogen. It was also found that the LFL of ammonia is not
correctly approximated by any of the three more accurate
In the case of the UFL it was assumed that the excess fuel does methods.24,292,293 In the case of the UFL at different initial
not react,292 so the mass of unreacted fuel is replaced by an equal temperatures, the method of Mendiburu et al.292 is more
mass of air in the products. This allows the determination of an accurate.
adiabatic flame temperature at the UFL that is independent of Some articles have considered the adiabatic flame temper-
the global combustion reaction (T*UFL,T) and depends only on ature in the LFL and UFL to be independent of the initial
Y https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

Figure 17. Pressure dependence of FLs of methane in air: (a) LFL and (b) UFL. Reproduced with permission from ref 299. Copyright 2016 Elsevier.
Pressure dependence of FLs of natural gas in air: (c) LFL and (d) UFL. Reproduced with permission from ref 300. Copyright 2019 Elsevier.

temperature of the mixture.296−298 This approach has given compounds show nonlinear behavior. The experimental data
good results for lower alkanes; however, the authors have found obtained by Van den Schoor220 are shown in Figure 18 and
that it leads to higher deviations for oxygenated compounds confirm the nonlinearity of the UFLs of methane and natural gas.
(C−H−O). These data show nonlinear behavior for ethane, propane, n-
In another work by Liaw and Chen,297 a method to calculate butane, and ethylene UFLs in air at 200 °C and pressures above
the FLs at different initial temperatures was developed. This atmospheric. In the case of hydrogen, Liu and Zhang109 showed
method was developed for two cases: (1) adiabatic combustion
process and (2) combustion process with radiative heat losses.
The adiabatic approach was found to be more accurate.
Moreover, in this work, the adiabatic flame temperature was
assumed to be constant and the combustion reaction in the UFL
assumes the presence of CO, CO2, H2, and H2O in the products.
However, this type of reaction cannot be satisfied for some
compounds such as ethylene.
6.2. Pressure Effect on FLs. The dependence of FLs on the
initial pressure of the mixture is discussed in this section.
Experimental data show that, for low hydrocarbons, an increase
in initial pressure increases the UFL; however, it has a weaker
effect on the LFL.156,205,212 This behavior is shown in Figure 17
for the cases of methane299 and natural gas.300 The experimental
data presented in Figure 17 confirm the stronger effect of the
initial pressure on the UFL compared to the effect on the LFL. It
can be observed that the data for methane and natural gas show a Figure 18. Flammability limits in air of ethane, propane, n-butane, and
linear behavior of the LFL for different initial pressures. ethylene at 200 °C and pressures above the atmospheric. Data taken
However, the experimental data for the UFLs of these from ref 220.

Z https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

that increasing the initial pressure (from 0.1 to 0.4 MPa) leads to dependence of FLs on initial temperature were obtained by
a decrease in LFL and an increase in UFL, with the latter applying a linear regression to the experimental data from several
showing nonlinear behavior. articles. The results of this procedure are given in Table S3. Van
It is interesting to point out that Coward and Jones251 den Schoor211 determined the slopes for estimating the UFLs of
presented experimental FL values obtained for DFP in a 2 cm some alkanes at different temperatures and initial pressures of
diameter, 50 cm long tube and for different flame propagation the mixture. A linear relationship is also used to determine the
directions in a 5 cm diameter, 50 cm long tube. The results UFL at different initial temperatures. The coefficients for these
showed that at low pressures there is convergence of the LFL relationships are given in Table S4. Goethals et al.227 established
and UFL. However, the FLs reported by Coward and Jones251 an empirical equation to determine the change in FLs of toluene
are limits of flame propagation in a tube of a given diameter with the initial temperature of the mixture for different initial
rather than flammability limits.27 pressures. The coefficients are listed in Table S5. Vanderstraeten
In their experimental study, Coronado et al.82 determined the et al.204 experimentally studied the change in UFL of methane
FLs of anhydrous and hydrated ethanol in air at different with temperature and pressure. The coefficients for their
temperatures and reduced pressures in a 20 L spherical vessel. empirical equations are given in Table S6.
The LFL and the UFL showed no tendency to converge at low Several experimental articles presented their own empirical
pressures. There are two works in which the LFLs of some relationships to estimate the effects of initial temperature and
compounds were theoretically determined at different initial pressure on the FLs of a given compound. The information in
pressures.301,302 In these works, the adiabatic flame temper- Table S1 gives the references in which the effects of these
atures at different pressures were used to estimate the LFL. variables were determined. The dependence on pressure is
Therefore, the method used is similar to that described in generally logarithmic, as shown in eq 113.
section 5.2.
ij p yz
FLp = FL p + a0 lnjjjj zzzz
The phenomenological description of a planar flame,
presented by Law,303 is interesting to explain the behavior of 0 jp z
FLs at different initial pressures. The burning mass flux rate (ṁ S) k 0{ (113)
is related to the flame thickness (δ), the specific heat capacity To the author’s knowledge, there is no simple method
(cp), and the thermal conductivity (k) by the expression shown validated with experimental data from different data sets and
in eq 111. In addition, the burning mass flux rate is equal to the multiple fuels that allows the determination of FLs at different
product of the laminar flame speed (sL) and the unburned initial pressures. Zabetakis24 established the following empirical
mixture density (ρu). equations for the determination of FLs of natural gas (85−95%
methane and 15−5% ethane) at different initial pressures and
k /cp
mS = sL ambient temperatures in air.
u (111)
LFLp = 4.9 0.71(log p) for 1 680 atm (114)
Then, the mass flux rate can be related to the pressure by eq
112, where the dependence of the reaction rate on pressure is w0b UFLp = 14.1 + 2.04(log p) for 1 680 atm (115)
0
∼ pne−Ta/Tb , where Ta is the activation temperature, n is the
d

reaction order, and Tb is the adiabatic flame temperature. In eqs 114 and 115, p0 = 1. LFLp0 = 4.9, UFLp0 = 14.1, a0 =
d d

0.71, and a0 = 2.04, for LFLp and UFLp, respectively. The


ÄÅ ÉÑ1/2
ÅÅ i y ÑÑ experimental data available in the literature can be correlated in
ÅÅ n jj kt zz Ta / TbÑ
Ñ
mS ÅÅp jj zz e ÑÑ this way.
ÅÅ jj zz ÑÑ
ÅÅ k cp { ÑÑ
ÅÇ b ÑÖ (112)
7. DILUENT GAS EFFECTS ON FLS
The ratio k/cp is insensitive to pressure variations. Therefore, The addition of a diluent (or inert) gas to the fuel−air mixture
for reactions of order n = 2, the mass flux rate would be pressure affects the FL value. Figure 19 shows the behavior of the FLs of
independent. However, in cases where n > 0 and n ≠ 2, different methane−inert−air mixtures as a function of the
increasing the pressure would increase the mass flux rate. The concentration of the diluent gas.24 In Figure 19, the pressure is
burning mass flux rate (ṁ S) varies inversely with flame thickness atmospheric and the temperature is 25 °C. Diluent gases added
[see eq 111]. Furthermore, for a reaction order of n > 0, to fuel−air mixtures are sometimes referred to as inert gases.
increasing the initial pressure would increase the burning mass Note that, strictly speaking, CO2 and H2O are not inert gases.
flow rate (ṁ S) and decrease the flame thickness (δ). They can chemically affect the reaction process by promoting or
Since a thinner flame involves lower heat losses due to the suppressing the production of radicals. These species can be
lower volume of the flame,303 the UFL is expected to increase nitrogen, carbon dioxide, water vapor, or helium, among others.
and the LFL to decrease at higher pressures. On the other hand, The behavior shown in Figure 19 is typical of hydrocarbons (C−
increasing the pressure leads to an increase in heat loss by H) and oxygenated compounds (C−H−O) mixed with a
thermal radiation per unit volume due to an increase in the diluent gas in air. As the concentration of the diluent gas is
average absorption coefficient.54 It can be concluded that there increased, the LFL increases slightly and the UFL decreases. At a
are two competing mechanisms: the decrease in heat losses per certain point, called the fuel inertization point (FIP), both limits
unit volume associated with the decrease in flame thickness at are equal. Mixtures in the FIP are generally rich, with some
higher pressures and the increase in heat losses due to thermal exceptions (hydrogen for example).
radiation associated with the increasing mean absorption The variation in the FLs is greater for the diluent gases added,
coefficient at higher pressures. which have higher heat capacities. For example, looking at the
6.3. Empirical Formulas for FLs at Different Initial relationship for cp,i shown in eq 116 and comparing it to the
Temperatures and Pressures. The slopes for the linear variation in the FLs shown in Figure 19, we find that a greater
AA https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

LeO + (1 )LeF
Leeff,lean =
2 (117)

LeF + (1 + )LeO
Leeff,rich =
2+ (118)
In eqs 117 and 118, Φ̃ ∼ β(1 − ϕ). The Lewis number of the
fuel (LeF) and the Lewis number of the oxidizer (LeO) are
calculated by using the thermal diffusivity of the mixture and the
mass diffusivities of the fuel and oxidizer, respectively. The
equivalence ratio (ϕ) and the Zeldovich number (β) are also
used in eqs 117 and 118. Note that β = Ea(Ta − Tu)/RTb2. Fan et
al.306 calculated the effective Lewis numbers for mixtures of H2−
O2−He and H2−O2−N2 with ϕ = 0.4; for the former the value of
Leeff,lean was 0.60 and for the latter it was 0.57.
Qiao et al.29 studied the flammability limits and laminar flame
speed of methane−air mixtures in microgravity with dilution by
He, argon (Ar), N2, and CO2. They determined the effective
Lewis number as the ratio between the thermal diffusivity of the
mixture and the mass diffusivity of the methane−oxygen pair, as
shown in eq 119. They then calculated the values of the effective
Lewis numbers for mixtures of CH4−air−diluent for different
molar concentrations of the diluent. The results are summarized
in Table 7.
mix
Leeff =
DCH4 O2 (119)

Table 7. Effective Lewis Numbers for CH4−Air−Diluent


Mixturesa
diluent mole concn (%) Leeff
Figure 19. Flammability limits of methane−inert−air mixtures at 1 atm
He 0−35 0.99−3.46
and 25 °C.24
Ar 0−48 0.99−0.97
N2 0−35 0.99−0.98
CO2 0−22 0.97−0.89
variation corresponds to the added inert gases with higher values a
Reproduced with permission from ref 307. Copyright 2010 Elsevier.
of cp,i.
cp ,CCl4 > cp ,CO2 > cp ,H2O > cp,N2 > cp ,He (116) Since spherical flames have a positive flame stretch rate when
The effect of the diluent gas can be attributed to the fact that they propagate outwardly, as in the case of FL experiments, the
some of the heat released by the combustion process is used to effect of the Lewis number can be evaluated by considering the
raise the temperature of the inert species to the flame generalized stretch factor (σ), where the Karlovitz number (Ka)
temperature. This simple argument can also be used to is also needed.303 The generalized stretch factor takes into
understand the effects of the addition of He (helium) on the account the chemical and diffusive properties of the mixture
LFL. However, in the case of halogenated compounds, the through the first two factors on the right-hand side of eq 120 and
differences in heat capacities are not sufficient to explain the the system dynamics through the third factor (Ka).
behavior shown in Figure 19. Halogenated compounds inhibit ij 1 yz
= jj 1zzzKa
the chain-branching reactions of the H2−O2 system by j
j
2 k Leeff z
catalyzing the recombination of hydrogen atoms to form { (120)
molecules and radicals that are relatively less reactive.303 Therefore, flame speed and temperature decrease for mixtures
In the case of He, the LFL increases slightly with increasing with Leeff > 1 and increase for mixtures with Leeff < 1. Thus, in the
He concentration as cp,He > cp,N2. Comparing the dilution by N2
d

case of helium, flame propagation is affected by the stretch effect


with that by He, we find that the latter affects the thermal by reducing the laminar flame speed and flame temperature as
diffusivity of the mixture (α = k/ρcp) differently. For example, at the effective Lewis number of the mixtures is greater than unity.
300 K and 1 atm, the densities of N2 and He are 1.1233 and According to Glassman et al.27 the chain-termination reaction
0.1615 kg/m3, the specific heat capacities are 1.041 and 5.193 (R3) competes with the chain-branching reaction (R1) when
kJ/(kg·K), and the heat conductivities are 0.0259 and 0.152 W/ the system is rich. The third body can be an inert species such as
(m·K), respectively.31,304 Therefore He influences the value of N2, H2O, or CO2. Law and Egolfopoulos51 found that, at
the Lewis number more strongly than N2. According to atmospheric pressure for lean methane−air flames, the main
Matalon,305 the effective Lewis numbers (Leeff’s) of lean and competing reactions were the chain-termination reaction (R2)
rich mixtures can be determined by the expressions given in eqs and the chain-branching reaction (R1). Nevertheless, the
117 and 118. accuracy of these observations can be tested by performing a
AB https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

Figure 20. Flame speed sensitivity analysis for freely propagating methane−air flame near the FLs. (a) CH4 mole concentration of 4.90%; (b) CH4
mole concentration of 15.80%.

sensitivity analysis of the laminar flame speed for methane near HCO + M CO + H + M (R7)
the LFL and near the UFL. The results of this analysis are shown
in Figure 20. The simulations were performed with Cantera H 2O + O 2OH (R8)
software308 using the GRI3.0309 chemical kinetic mechanisms.
H + O2 OH + O For example, the addition of a diluent gas not only decreases the
(R1)
fuel and oxygen concentrations, but the increase in the third-
H + O2 + M HO2 + M (R2)
body coefficient also changes the reaction rates of chain-
termination and chain-propagation reactions [(R2), (R3), and
H+H+M H2 + M (R3) (R7)], so the overall reaction rate is varied, resulting in changes
in flame speeds and FLs. When H2O is added, reaction R8
The graphs in Figure 20 depict the 10 elementary reactions promotes the production of the OH radical. When CO2 is added,
that presented the highest absolute values of the sensitivity reaction R6 (reverse) suppresses the production of the H
coefficient (k/SL)(dSL/dk). The importance of the chain- radical, leading to strong chemical effects on flame speeds and
branching reaction shown in (R1) is evident. Furthermore, the FLs.
effects of N2 and H2O on the LFL are noted by promoting the The dilution effect causes the inert species to lower the flame
formation of the more stable HO2 species by consuming a more temperature. This effect is stronger in the UFL than in the LFL
reactive H radical. The recombination reactions shown in (R4) because more heat can be released in the LFL. In the case of a
and (R5) become important chain-termination reactions for the lean mixture, the termination reaction (R2) competes “more
CH4−air mixture near the UFL. efficiently” with the chain-branching reaction (R1). However,
due to the decreasing temperature, the chain-branching reaction
CH3 + H + M CH4 + M (R4)
becomes slower. For rich mixtures, the reaction rate of the chain-
2CH3 + M C2 H 6 + M termination reaction (R3) becomes greater than that of (R2)
(R5)
because it depends on the hydrogen concentration, which is
In addition, the chain-propagation reactions (R6)−(R8), expected to be higher for rich mixtures. Therefore, reaction R3
where HCO is the formyl radical, also play a crucial role in H competes “more efficiently” with the chain-branching reaction
radical production and heat generation. (R1), which at the same time becomes slower due to the lower
temperature. As a result of these interactions, the UFL is more
CO + OH CO2 + H (R6) affected by the addition of diluents than the LFL.
AC https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

Another important feature is that the rates of production of H2 radiative heat losses on the calculations. Surprisingly, they found
and HO2 by the reactions shown in (R3) and (R2), respectively, that the results obtained with the adiabatic model were more
increase with increasing concentration of the diluting species. It accurate than those obtained with the nonadiabatic model. Bade
is important to note that halogenated compounds follow a Shrestha et al.319 also applied a constant adiabatic flame
different mechanism that also inhibits the combustion process temperature approach. They considered that at the UFL the
(see refs 27 and 303). oxygen was consumed first to form CO and second to form H2O,
In addition, H2O and CO2 have a radiative effect on flames. and if any O2 was still available it would react to form CO2 from
This effect can either lower the flame temperature (optically thin the available CO. The same scheme of O2 consumption was
mixtures) or increase the flame temperature (optically thick applied in the work of Shu et al.315,316 mentioned earlier. The
mixtures). Guo et al.310 found that the addition of CO2 increases
main difficulty in applying the CAFT method for fuel−diluent−
the optical thickness of the flame because CO2 is able to emit
air mixtures is the prediction of the UFL, because in this case the
and absorb radiation. The increase in the heat capacity of the
mixture (due to the addition of CO2) causes the flame CAFT will not be constant due to the significant variations in the
temperature to drop and increases the probability that the UFL. Therefore, there is room for further simulations to
flame will extinguish due to stretch or radiative heat loss. investigate the effects of diluent addition on flammability limits.
Therefore, flammability limits are also affected by the addition of It is worth noting that Gutiérrez Velásquez et al.320 applied
CO2, as has been observed experimentally. The existence of kriging regression to predict the FLs of hydrated and anhydrous
flammability limits considering radiation reabsorption in ethanol. Therefore, dilution with H2O was somehow considered
premixed flames was demonstrated by Ju et al.311 in this work.
It is expected that the addition of a diluting species will cause 7.2. Empirical Equations to Determine FLs with Added
incomplete combustion. This is confirmed by Kumar et al.,312 Diluent Species. Wang et al.321 determined the empirical
who tested H2−H2O−air mixtures with a hydrogen concen- values of ΦFL for the determination of the UFL in air of fuels
tration of 8.0%. They found that the combustion efficiency was diluted with N2 or CO2. These empirical values are shown in
100.0% at 0.0 and 15.0% vapor concentrations; however, at Table S7. To evaluate the applicability of the empirical values of
30.0% vapor concentration, the combustion efficiency was ΦFL together with eq 122, a validation was performed by
38.0%. considering the UFL of 528 experimental data for fuel−N2−air
7.1. Methods for Determining FLs with Added Diluent mixtures and 208 experimental data for fuel−CO2−air mixtures.
Species. Some methods have been developed to determine the The results show AAREs of 4.53 and 6.07%, respectively, while
FLs of fuel−diluent−air mixtures at 1 atm and reference the R2 values were 0.9839 and 0.9748, respectively. These results
temperature. In most methods, the adiabatic flame temperature are also shown in Table S7. The experimental data considered
was assumed constant; also, a global reaction was assumed at the here are the same as specified in Tables S8 and S9.
UFL. The fuel−diluent−air mixture presents two independent
Correlations for various fuel−diluent−air mixtures can be
concentrations. The fuel molar fraction (FL) and the diluent
found in the literature. For example, Kondo et al.186,187 obtained
molar fraction (Pd) can be considered independent in order to
define a fuel-diluent mixture, characterized by the fuel molar correlations for determining the FLs of some hydrocarbons,
fraction (yF) defined in eq 121. Therefore, the diluent molar oxygenated compounds, and ammonia diluted with N2 and CO2.
fraction in the fuel−diluent mixture would be yd = 1 − yF. Also, correlations for the determination of the FLs of ammonia
diluted with HFC-134 and HFC-125 were obtained by Kondo et
FL al.95 In another work, Kondo et al.185 established correlations for
yF =
FL + Pd (121) the FLs of isobutane diluted with N2 and CO2. Li et al.136
Chen et al. 313,314
studied the FLs of hydrocarbons in air presented correlations for the determination of the FLs of
diluted with N2 and CO2. They obtained the expression shown methane−N2−air mixtures at low temperatures and 1 atm.
in eq 122, which relates the flammability limit without added In another work, Li et al.18 presented correlations to
diluent (FL0) to the flammability limit with added diluent (FLd) determine the FLs of R-290 and R-152a refrigerants diluted
by means of a factor identified as ΦFL. They theoretically with 1,1,2,2-tetrafluoroethane (C2H2F4). Zhang et al.193
determined the values of ΦFL; however, the results were not obtained correlations to calculate the FLs of dimethyl ether
entirely satisfactory. diluted with five different compounds. Molnarne et al.322
investigated the use of a method that assigns a nitrogen
ij 1 yz
1 1 jj 1zzzz
equivalence coefficient to other dilution gases. This method can
= + FL j
j
FLd/100 FL/100 jy z be used to evaluate the effectiveness of a diluent gas when it is
k F { (122)
used to reduce the flammability range of a compound. The case
313,314
In the calculations of Chen et al., the adiabatic flame of nitrogen was studied in a previous work.85
temperature was assumed to be constant and CO2, CO, H2O, A simple way to obtain empirical equations for different
and H2 were assumed to be formed at the UFL. A similar compounds is to assume that there is a simple relationship
approach was used by Shu et al.315,316 Note that compounds between the number of oxygen moles on the FL without added
with high UFLs would require a condensate (carbon) in the diluent species (vFL air,0), the number of oxygen moles on the FL
products to be considered at equilibrium. For example, in the with added diluent species (vFL air,d), and a correlation factor kFL, as
case of ethylene, the UFL in air is 31.5%186 and the reaction performed by Mendiburu et al.323 The expected relationship is
[without condensate] shown in eq 61 cannot provide a solution
shown in eq 123. Equations 124 and 125 are used to determine
for high UFL values; in these cases, the reaction [with
the values of vFL FL
air,0 and vair,d.
condensate] shown in eq 62 has been successfully applied.34,35
The assumption of a constant adiabatic flame temperature was FL FL
vair,d = yF vair,0 + yd kFL
applied by Liaw et al.,317,318 who also evaluated the effect of (123)

AD https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

1 ijj 100 yz 7.3. Methods to Determine the Fuel Inertization Point


FL
vair,0 = j 1zzz
4.76 jjk FL0 z (FIP). It is said that at the FIP both flammability limits are equal.
{ (124) However, the mixture at this point is usually rich (ϕ > 1), which
means that the FIP is actually part of the UFL curve. The
1 ijj 100 yz
FL
vair,d = j 1zzz problem of determining the FIP involves the determination of at
4.76 jjk FLd + Pd z
{ (125) least two unknowns, namely (1) the value of the equivalence
ratio at the FIP (ϕFIP) and (2) the value of the fuel molar fraction
The values of kFL were determined for several fuel−diluent− (yFIP
F ) in the fuel−diluent mixture at the FIP.
air mixtures using experimental data to solve eq 123. It was Shebeko et al.324 assumed that only CO, H2O, N2, and the
found that the dependence of the FL on the type of diluent added inert gas are present in the products. The FIP is obtained
added can be approximated by the average values of kFL as shown assuming that the number of moles of oxygen is exactly that
in eqs 126 and 127. The FIP is part of the UFL curve since it needed to obtain the aforementioned products. However, the
corresponds to a rich mixture (except hydrogen). determination of the fuel molar fraction or of the adiabatic flame
N 1 temperature is not specified. Bade Shrestha et al.319 developed
av 1 an approximation for the fuel molar fraction at the FIP. This
kLFL = kLFL
N i=1 (126) approximation uses empirically determined constants that
measure the influence of the added inert at the LFL (aLFL)
N
av 1 and UFL (aUFL). The equation to be solved is shown in eq 129.
kUFL = kUFL The authors have not provided a method for determining aFL
N i=1 (127) when experimental data are not available for a particular fuel.
323
The values of kav
FL were obtained by Mendiburu et al. for 1
1 i 100 100 yz
different fuel−N2−air and fuel−CO2−air mixtures at 1 atm and = 1 + (aLFL aUFL)jjj + zz
25 °C. The results of this procedure are given in Tables S8 and 1 yFFIP k UFL LFL { (129)
S9. Once the value of vFL air,d is obtained, the FL is determined by
applying eq 128. The accuracy of the empirical approximations Razus et al.325,326 have developed a method that assumes an
is shown in Table 8, where it can be seen that the AARE is less average value of the equivalence ratio on the FIP (ϕav FIP) for the
than 5.0% and the R2 is greater than 0.989. cases of addition of N2, CO2, or H2O. Then, the value of the
yF adiabatic flame temperature at the FIP (TFIP) is approximated by
FL = FL
·100% a linear correlation with the adiabatic flame temperature at the
1 + 4.76vair,d (128) LFL without the added inert (TLFL). Hansen327 and Hansen and
Crowl328 assumed a constant adiabatic flame temperature and
Table 8. Accuracy of Empirical Determination of the FL of assumed that excess oxygen in the products was not converted at
Fuel−Inert−Air Mixtures with Values of kav
FL Obtained from the UFL. In this way, CO2 and H2O could be assumed to be part
Experimental Dataa of the products. An empirical model was fitted with data for 38
compounds diluted with nitrogen. This model showed a low
ARE
value of squared correlation coefficient (R2 = 0.672).
added inert N AARE (%) R 2
≤10% 10−20% >20% Van den Schoor et al.203 experimentally determined the FIP
Lower Flammability Limits points of C2H4−N2−air, H2−N2−air, and C2H4−H2−N2−air.
N2 461 1.45 0.9973 457 6 0 The FIP was approximated by the following considerations: (a)
CO2 170 3.08 0.9895 159 8 3 the LFL does not change with the addition of N2; (b) at the FIP
Upper Flammability Limits the equivalence ratio is equal to unity. However, as shown in
N2 521 3.29 0.9973 487 23 11 Table S8, the equivalence ratios at the FIP points of fuel−N2−air
CO2 201 4.21 0.9959 178 15 8 mixtures are not equal to unity. In a later work, it was found that
a
Adapted from Mendiburu et al.323 this approximation is inaccurate for mixtures containing H2−
CO−N2−air.137
Mendiburu et al.323 proposed a method for determining the
molar fraction of diluent in a fuel−diluent mixture, assuming
As an example, the UFL of an ethanol−CO2−air mixture can that the adiabatic flame temperature at the FIP (TFIP) can be
be determined. The experimental UFL of an ethanol−CO2 determined and used to determine yFIP by assuming a global
mixture in air is 9.5% when the fuel−diluent mixture contains reaction and performing equilibrium calculations as explained
32.2% ethanol and 67.8% CO2. The values kav UFL = 0.3410 and for the CAFT method in section 5.2. The adiabatic flame
vUFL
air,0 = 0.913 are taken from Table S9. Applying eq 123, the value temperatures at the FIP and at the LFL (without diluent) were
vUFL
air,d = 0.256 is determined. Finally, from eq 128, the calculated related as shown in eq 130.
UFL is 9.2%. In this case, the ARE is 3.16%.
It is important to note that, to determine the FLs of a fuel− TFIP
d =
diluent−air mixture, the concentration of the diluent added TLFL (130)
must be kept constant. Therefore, it is easier to use the
concentrations in the fuel−diluent mixture that are derived from Mendiburu et al.323 found that, for the fuel−diluent−air
the actual concentrations in the fuel−diluent−air mixture. mixtures listed in Tables S8 and S9, the θd value ranged from
Although there is a reasonable amount of experimental data on 0.78 to 1.17 and the ϕFIP value ranged from 0.841 to 1.847. The
FLs for fuels diluted with N2 and CO2, data for fuels diluted with exception was H2, for which the ϕFIP values, when the mixtures
H2O are still sparse. Moreover, other diluents have rarely been were diluted with N2 and CO2, were 0.463 and 0.468,
considered in experiments. respectively. The equivalence ratio at the FIP (ϕFIP) is related
AE https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

to the mole fraction of fuel at the FIP (yFIP


F ) and to the moles of hd, R hd,P
oxygen at the FIP (vFIP n=
air ) by eq 131. hCO hCO2 + h H2O hH2 (141)
yFFIP i x xO yz
FIP
vair = j
jjxC + H zz Note that the eq 134 can provide two values for yFIP
F , but only
FIP k 4 2 { (131) one of them satisfies the mass balance of the global reaction [eq
In the same article, a correlation was proposed to obtain θd, 133]. The method developed by Mendiburu et al.323 showed
ϕFIP, kav av
LFL, and kUFL, as shown in eq 132.
AAREs of 0.82 and 1.95% for the FIPs of fuel−N2−air and fuel−
CO2−air mixtures, respectively. The Supporting Information of
L = a 0 + a1z1 + a 2z 2 + a3(z1z 2) + a4z12 + a5z 2 2 the original article contains sample calculations for the
application of this method. For a discussion of flammability
+ a6(z1z 2)2 (132) limits in oxygen see the Supporting Information.
where L might be θd, ϕFIP, kav
LFL, or kav
UFL. The zi are given in eq
133. 8. THE FLAMMABILITY LIMITS OF FUEL MIXTURES
LHVi LHVav FLA, i FLA,av Le Chatelier’s law can be used to determine the FLs in air of a
z1 = ; z2 = mixture containing more than one fuel based on the FL data for
SLHV SFL (133)
each fuel. The mathematical expression for this law is very
The LHVav and FLA,av are average values of the lower heating simple and is shown in eq 142.
value and of the flammability limit of the compounds used for
the correlation set. The SLHV and SFL are the standard deviations 1
FLmix = yi
of these variables for the correlation set. Table S10 presents the
coefficients and the values of the averages and standard FLi (142)
deviations necessary to apply the correlation in eq 132. In the
cases of θd and ϕFIP the determination of z2 involves the LFLA,i, There are a total of four different ways to derive Le Chatelier’s
LFLA,av and SLFL for determining z2. Assuming the global law; these are presented by Mashuga and Crowl,329 Carvalho et
reaction shown in eq 134, the necessary equations to apply the al.,330 Chen et al.,331 and more recently by Mendiburu et al.5 The
method are obtained. following can be stated as Le Chatelier’s law: A mixture of N
combustible gases at the FL in air requires a number of moles of
yFFIP CxCH x HOxO + ydFIP Md + vair
FIP
(O2 + 3.76N2) oxygen equal to the weighted sum of the number of moles of
oxygen required by each gas at its respective FL in air, where the
nCO2CO2 + nCOCO + n H2OH 2O + n H2 H 2 mole fractions are the weighting factors. This section briefly
FIP
+ 3.76vair N2 + ydFIP Md presents the derivation of Le Chatelier’s law as given by
(134) Mendiburu et al.5
The value of yFIP
Fis obtained from eq 135, and the required Consider the combustible gases A and B and their mixture,
coefficients are obtained from eqs 136−141. The Keq is from the denoted C. Given the composition of the FLs of A, there are vFL air,A
homogeneous water−gas reaction [eq 65]. moles of oxygen per mole of species A. Assuming an adiabatic
combustion process at constant pressure, that the nitrogen and
av(yFFIP )2 + bvyFFIP + cv = 0 (135) excess oxygen are inert, and applying conservation of energy to
the global combustion reaction for species A, eq 143 is obtained.
a=1 Keq (136) * = vFL
In this equation, the excess oxygen is represented by vair,A air,A
s
s − vair,A. Moreover, in the LFL, vair,A
* > 0 and α = 1. In the UFL,
var,1 * < 0 and α = 0.
however, vair,A
av = am2 + 2 (am xC) + xC[2xC (2a + 1)m]
FIP
xH hA njhj * hO ,A
vair,A
FL A 2
+ xO(am xC) + (xC m) vair,A =
2 (137) 3.76h N2,A (hO2,R + 3.76h N2,R ) (143)
s
var,1 xH For 1 mol of A, FLA can be determined as follows FLA =
bv = 2amn + 2an (2a + 1)nxC + anxO n
FIP 2 100%/(1 + 4.76vFL air,A). This gives the expression shown in eq
(138) 144. In this expression, the heat released in the combustion
process of A is hC,A = h̅0f,A − ∑A njh̅0f,j.
cv = an2 (139)
ÄÅ É
ÄÅ yA jij hC,A + hA ÅÅ
ÅÅÇ n * hO ,A ÑÑÑÑ zyz
hj + vair,A 2 Ñ
ÅÅ = yA + 4.76yA jjjj Ö zzz
s A j
vair,1
m = ÅÅÅÅhF + jj zz
(har,R 3.76h N2 + 2h H2 2h H2O) FLA 3.76h N2,A (hO2,R + 3.76h N2,R ) z
ÅÅ k {
ÅÇ FIP
(144)
xH
h H + xC(2h H2O 2h H 2 hCO2)
2 2 Note that it is always possible to choose the reference
ÉÑ
ÑÑ temperature equal to the initial temperature of the reactants.
+ xO(h H2 h H2O) + (hd,P hd,R )ÑÑÑÑ Thus, writing eq 144 for the combustible gas B, considering the
ÑÑ
ÑÖ initial temperature to be equal to the reference temperature for
1 gases A and B at the FL (∑njΔh̅j = 0), and adding these
(hCO hCO2 + h H2O h H 2) (140) expressions, the relationship shown in eq 145 is obtained.
AF https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

Figure 21. Results obtained by applying the law of Le Chatelier to the determination of (a) LFLs of mixtures and (b) UFLs of mixtures. Reproduced
with permission from ref 5. Copyright 2020 Elsevier.

yA yB N
+ = yA FL
var,mix = yv FL
FLA FL B i ar, i
ÄÅ É i=1 (149)
ij h
j C,A
ÅÅ
ÅÅÇ * hO ,A ÑÑÑÑ yzz
n hj + vair,A 2 Ñ
+ 4.76yA jjjj Ö zz + y
A j
zz The above expressions are consistent with Le Chatelier’s law,
jj 3.76h N2,A zz B
although they are formulated in terms of the number of moles of
k {
oxygen at the FLs. In their work, Mendiburu et al.5 compared the
ij h ÅÄÅ É
* hO ,BÑÑÑÑ yzz
j C,B ÅÅ n hj + vair,B FLs calculated according to Le Chatelier’s law with experimental
+ 4.76yB jjjj ÅÇ B j 2 Ñ Ö zz
zz data from Coward and Jones,251 Hustad and Sonju,147 Karim et
jj 3.76h N2,B zz
k { (145) al.,130 Kondo et al.,95,184,185 Shoshin and de Goey,248 White,123
Zhao et al.,99 Wierzba and Wang,133 and Wierzba et al.135 It is
A similar procedure can be applied to C (mixture of A and B) important to note that most experimental data correspond to
to obtain the expression in eq 146. binary mixtures of fuels in air. For ternary mixtures, there were
23 experimental data for LFL and 14 for UFL. No data were
ÄÅ É
ij h
j C,C
ÅÅ
ÅÅÇ n * hO ,CÑÑÑÑ yzz
hj + vair,C found for mixtures of more than three flammable compounds.
2 Ñ
= 1 + 4.76jjjj Ö zz
1 C j The validation results presented in this section are sufficient to
zz
FLC jj 3.76h N2,C zz understand the accuracy of this law. Nevertheless, the lack of
k { data for mixtures (or blends) of three or more fuels is a gap that
(146) should be addressed in future studies of conventional and
Since the gas C is a mixture of the gases A and B (yA + yB = 1), renewable fuels.
the law of Le Chatelier requires that the equality shown in eq 147 A total of 272 experimental data for the LFLs and 217 for the
UFLs were used for validation. The AAREs were 2.76% for the
is fulfilled.
LFLs and 8.28% for the UFLs. The values of R2 were 0.9870 for
ÄÅ É
* hO ,CÑÑÑÑ
the LFLs and 0.8649 for the UFLs. Looking at the results
hC,C ÅÅÅÅ C nj hj + vair,C 2 Ñ
Ç Ö presented in Figure 21, it is clear that Le Chatelier’s law predicts
3.76h N2,C LFLs with great accuracy, while it is only acceptably accurate in
ÄÅ É predicting UFLs. The experimental data used for this validation
ij h Å
jj C,A ÅÅÅÇ A nj hj + vair,A * hO ,A ÑÑÑÑ yzz are listed in Tables S11 and S12.
2 Ñ Ö zz
= yA jjj zzz After an analysis of the above derivation, it can be
jj 3.76h N2,A z
k { hypothesized that Le Chatelier’s law applies to fuels that react
ÄÅ ÉÑ according to their own chemical kinetic mechanisms, as if the
ij h ÅÅ Ñy
jj C,B ÅÅÇ B nj hj + vair,BhO2,BÑÑÑÖ zzz
* other fuels were not present in the mixture. This hypothesis will
+ yB jjj zzz be partially tested in the remainder of this section.
jj 3.76h N2,B zz
k { (147) In their simulations, Bertolino et al.64 used laminar flame
speed as an indicator to identify FLs. Moreover, Liang and Law53
The equality shown in eq 147 is possible only if the reaction assumed that three pairs of competing reactions are relevant in
kinetic paths of A and B are virtually unchanged, even if these FLs. Therefore, the hypothesis can be tested by performing a
compounds react in the same combustible mixture. In such a sensitivity analysis of laminar flame speed and identifying up to
case, the global effect of the reaction kinetic mechanisms is three pairs of competing reactions. It is expected that the three
accounted for by eq 146. Moreover, the global effect can be pairs of major competing reactions identified for each fuel at its
represented in a simpler expression by substituting eq 143 into FLs must have a similar degree of importance (as measured by
eq 147. Thus, eq 148 is obtained. The same procedure can be their sensitivity coefficients) when the fuel mixture at the FLs is
considered.
applied to any number of combustible gases forming a mixture.
Consider the mixtures containing methane, propane, and
In this case, the expression shown in eq 149 is obtained.
dimethyl ether (C2H6O), namely CH4−C3H8−air and CH4−
FL FL FL C2H6O−air. In both cases, Le Chatelier’s law accurately predicts
vair,C = yA vair,A + yB vair,B (148) the LFLs. However, the UFL of CH4−C2H6O−air cannot be
AG https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

Figure 22. Flame speed sensitivity analysis for mixtures 50% CH4−50% C3H8−air (mix 1) at the FLs.

Figure 23. Flame speed sensitivity analysis for mixtures 50% CH4−50% C2H6O−air (mix 2) at the FLs.

predicted accurately. The flame speed sensitivity analysis was CH4 + H CH3 + H 2 (R14)
performed using Cantera.308 The relevant elementary reactions
include all the reactions presented so far in this article as well as C3H6 + H C3H5 + H 2 (R15)
the reactions listed below.
8.1. The CH4−C3H8−Air Mixture. In this case, Le C3H8 + H H 2 + i C3H 7 (R16)
Chatelier’s law accurately predicts the FLs of the mixture 50% The results of the flame speed sensitivity analysis are shown in
CH4 + 50% C3H8 (mix 1). The San Diego mechanism67 was Figure 22. It is noticeable that the relative importance of the
used together with Cantera.308 At the LFL, the three main chain- competing elementary reactions is consistent with the
branching (or initiation) reactions are reactions R1, R7, and R6 hypothesis resulting from the derivation of Le Chatelier’s law.
for CH4, C3H8, and mix 1. The main chain-termination (or That is, Le Chatelier’s law is true if the chemical kinetic
recombination) reactions are reactions R2, R9, and R10 for mechanisms of the individual fuels are not affected by the
CH4; reactions R2, R9, and R11 for C3H8; and reactions R2, R9, presence of the other fuels in the mixture. This is confirmed by
R10, and R11 for mix 1. the fact that the main competing reactions of the individual fuels
HCO + O2 CO + HO2 (R9) at the FLs are also very important for the fuel mixture at the FLs.
8.2. The CH4−C2H6O−Air Mixture. In this case, Le
CH3OH + M CH3 + OH + M (R10)
Chatelier’s law accurately predicts the LFL of the mixture 50%
CH4 + 50% C2H6O (mix 2). However, it is inaccurate in
C2H4 + O CH3 + HCO (R11)
predicting the UFL of mix 2. The mechanism developed by Liu
et al.332 was used in conjunction with Cantera.308 At the LFL,
At the UFL, the major chain-branching (or initiation) the major chain-branching (or initiation) reactions are reactions
reactions are reactions R1, R12, and R13 for CH4; reactions R1, R6, and R7 for CH4, C2H6O, and mix 2. The major chain-
R1, R13, and R14 for C3H8; and reactions R1, R13, R14, and termination (or recombination) reactions are reactions R2, R9,
R12 for mix 1. The major chain-termination (or recombination) and R17 for CH4, C2H6O, and mix 2.
reactions are reactions R4, R5, and R14 for CH4; reactions R5, In the UFL, the main chain-branching (or initiation) reactions
R15, and R16 for C3H8; and reactions R5, R15, R4, R12, and are reactions R1, R6, and R18 for CH4 and reactions R20, R21,
R16 for mix 1. and R22 for C2H6O. However, for mix 2, reactions R6, R18, and
R21 are not of the same relative importance as for the individual
CH3 + O2 CH 2O + OH (R12) fuels. The most important chain-termination (or recombina-
tion) reactions are reactions R4, R5, and R14 for CH4 and
C2H3 + O2 CH 2CHO + O (R13) reactions R7, R22, and R23 for C2H6O. However, for mix 2,
AH https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

reactions R4, R5, R7, and R14 are not of the same relative flames at the UFL. Note that reactions R19−R22 are reactions
importance as for the individual fuels. that characterize combustion at low temperatures.333 For further
HO2 + OH H 2O + O2 discussion on the chemistry of cool flames, see the Supporting
(R17)
Information.
2CH3 C2H5 + H (R18) Systematic testing of the hypothesis of the applicability of Le
Chatelier’s law requires a sensitivity analysis of the chemical
CH 2OCH 2O2 H + O2 O2 CH 2OCH 2O2 H (R19) kinetics of fuel mixtures for which experimental data are
available (see Tables S11 and S12). However, such a study is
CH3OCH3 + OH CH3OCH 2 + H 2O (R20) beyond the scope of this section and is proposed as a future
research prospect.
CH3OCH 2O2 CH 2OCH 2O2 H (R21)
As Law303 states, Le Chatelier’s law represents the possibility
CH 2OCH 2O2 H 2CH 2O + OH (R22) of a unifying concept for the phenomena of flammability limits.
To overcome the experimental difficulties in the UFL, it would
CH 2O + OH H 2O + HCO (R23) be necessary to perform experiments with UFP and DFP in
microgravity and compare the results. However, such experi-
The results of the flame speed sensitivity analysis are shown in ments would be very complex.
Figure 23. Note that, in this case, some of the competing
elementary reactions that are very important in the UFLs of the 9. FLAMMABILITY LIMITS OF RENEWABLE FUELS
individual fuels have a lower relative importance in the UFL of AND FUTURE PERSPECTIVE
mix 2. This observation is consistent with the fact that Le
Chatelier’s law is not accurate for determining the UFL of mix 2. This section focuses on the FLs of renewable and alternative
Analyzing the available experimental data, the authors have fuels. Knowledge of FLs is important for the safe use of these
shown in a previous work5 that Le Chatelier’s law is accurate for fuels. Up to this point, the theory and estimation methods for
estimating the experimental LFLs and UFLs determined with FLs have been discussed. It is also interesting to discuss
downward flame propagation (DFP). Moreover, higher AREs strategies to establish reliable databases for new fuels by
are observed when Le Chatelier’s law is used to estimate the minimizing the need to conduct new experiments. Therefore,
experimental UFLs determined with spherical flame propaga- estimation methods should be used when possible.
tion (SFP) and with upward flame propagation (UFP). 9.1. The FLs of Emerging Renewable Fuels. The use of
Therefore, the effect of buoyancy is also important for the hydrogen as an energy carrier has become very important for
accuracy of Le Chatelier’s law. Note that, for UFP, buoyancy society. However, production methods334 and storage strat-
promotes flame propagation. Also, for SFP, the formation of the egies335 are still under investigation. Hydrogen can be produced
top half of a sphere is sufficient to classify the mixture as from natural gas336 and biomass,337 among others. Other
flammable (see Figure 11), i.e., buoyancy promotes flame important renewable and alternative fuels include ammonia,
propagation at the UFL for SFP. ethanol, methanol, syngas, and biogas. For aerospace
Looking at experiments with SFP, the fuel mixtures for which applications, there is interest in Fischer−Tropsch isoparaffinic
Le Chatelier’s law shows low accuracy in UFLs are the following: kerosene (FT-IPK), hydroprocessed esters and fatty acids
(HEFA), and synthesized isoparaffins from hydroprocessed
• C2H6O−C2H4, C2H6O−NH3, C2H6O−CO, C2H6O− fermented sugars (SIP). The list of experimental works in which
C3H8−C2H4, C2H6O−C3H8−CO, C2H6O−C2H4−CO, FLs were determined can be found in Tables S1 and S2.
C2H4−NH3, C2H4−C3H6, C2H4−CO, and C2H4− However, Table 10 provides a compilation of the pressure and
C3H8−CO temperature ranges for which experimental FLs are available for
On the other hand, Le Chatelier’s law shows low accuracy for these renewable and alternative fuels.
UFLs in the experiments with UFP for the following: As can be seen in Table 10, the experimental FLs of several
• H2−H2S, H2−C2H4, H2−C3H8, H2−NH3, CH4−C2H2, alternative fuels are available for a wide range of temperatures
and CH4−H2S and pressures. For example, the FLs of hydrogen have been
determined for temperatures between 143 and 673 K and for
As shown in Table 9, some oxygenated fuels such as dimethyl
pressures between 10.1 and 450.3 kPa. On the other hand, there
ether have manifested cool flames near the UFL. Since cool
are fuels for which there is not as much data, such as alternative
flames are characterized by low temperature reactions, Le
aviation fuels and fuel blends. The experimental determination
Chatelier’s law is not expected to give accurate results for
of FLs of liquid fuels is associated with the difficulty of
mixtures containing a fuel with a tendency to develop cool
vaporizing them before testing. Therefore, most of the available
experimental data on alternative aviation fuels are related to
Table 9. Compounds for Which Cool Flames Have Been temperatures of 373 K or higher. Moreover, no experiments are
Reported near or at the UFL in Air available at pressures above 101 kPa.
As discussed in sections 2 and 8, chemical kinetic mechanisms
compound formula UFL (%)
190
are important to perform theoretical studies on FLs. There are
dimethyl ether C2H6O 26.2 comprehensive chemical kinetic mechanisms for alternative
ethyl formate190 C3H6O2 15.7 fuels such as hydrogen, carbon monoxide, ammonia, methanol,
methyl formate190 C2H4O2 22.7 and ethanol. However, to the authors’ knowledge, the chemical
methyl acetate190 C3H6O2 14.0 kinetic mechanisms for hydrogen sulfide, HEFA, FT-IPK, and
acetaldehyde162,190 C2H4O 57.0, 57.0 SIP need further development. In the case of alternative aviation
methylal190 C3H8O2 18.5 fuels, however, these are modeled by fuel surrogates.338 In
ethyl ether162,190 C4H10O 46.0, 48.0 addition to FLs, detonation limits339 and fast flame limits340 are
ethyl nitrite162 C2H5NO2 >50.0 also important for the application of alternative fuels.
AI https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

9.2. Estimation of the FLs of Renewable and

4, 7, 16, 105, 111, 115, 123, 130−132, 134, 135, 147, 165, 166, 215, 217, 221, 291
Alternative Fuels. If the FLs are not available for a renewable
fuel at 298 K and atmospheric pressure, it is possible to apply the
methods described in section 5. In the case of the CAFT
method, the data required for its application are the molecular
weight, the enthalpy of formation, and the number of moles of
carbon, hydrogen, and oxygen in the molecule of the fuel. The
implementation of such a method can be done in any
programming language or even in spreadsheet software.
The effect of temperature on FLs can be estimated using the
methods in section 6.1. The simplest method for estimating FLs
15, 105, 122, 130, 133, 137, 141, 159, 202, 221

at different initial temperatures is to use eq 94 derived from the


ref

16, 95, 123, 126, 181, 182, 184, 217, 254

modified Burgess−Wheeler’s law. As observed experimentally,


105, 115, 131, 147, 182, 184, 221, 291

the FLs vary linearly with temperature. Therefore, an improve-


ment to eq 94 is to allow for a variable slope. In the case of the
LFL, this is achieved by using eq 104. However, for the UFL, the

Tables S1 and S2 cite the experimental works used to construct this table. bSyngas presents different compositions involving at least H2 and CO.
improvement is introduced by using a CAFT approach, as
shown in eq 109. To implement these enhancements, the
167, 174, 175, 226

molecular weight of the fuel, the enthalpy of formation, the heat


82, 176, 178, 216
Table 10. Pressure and Temperature Ranges of the Available Experimental Data for the FLs of Several Alternative Fuels

released during combustion per mole of fuel, and the number of


16, 108, 217

moles of fuel and oxidant (for a volume of 12 L at T0) must be


168, 169

123, 140

167, 226
167, 226

174, 175

known.
The effect of added diluents on FLs can be estimated using the
176

methods in section 7. The simplest approach in this case is to use


eq 123 together with the average correlation factor (kav FL) given
pmax (kPa)

by the correlation in eq 132. The fuel inertization point can also


405.3
101.3
101.3
101.3
101.3
101.3
689.4
101.3

101.3
101.3
101.3
101.3

be calculated using eq 131 and the correlation in eq 132, but with


different coefficients. In this case, the data required are the lower
heating value and the flammability limit of the fuel without
upper flammability limit
pmin (kPa)

diluent. Note that this method was developed for N2 and CO2
101.3
101.3
101.3
101.3
101.3

101.3
10.1

20.0

50.7
50.7
20.0
20.0

only. However, it can be extended to other diluents.


The LFLs of fuel mixtures can be determined by applying Le


Chatelier’s law. As stated in section 8, Le Chatelier’s law is
Tmax (K)

accurate for determining LFLs, but has inaccuracies when used


673.0
873.0
295.0
673.0
573.0
473.0
523.0
292.0

415.0
385.0
465.0
465.0

to determine UFLs. To the authors’ knowledge, there is no other


method for determining the FLs of fuel mixtures.


Furthermore, to the authors’ knowledge, there is no method
Tmin (K)
173.0
278.0
295.0
278.0
291.0
373.0
298.0
292.0

450.0
415.0
425.0
425.0

for estimating FLs as a function of initial pressure. Therefore, the


FLs of renewable and alternative fuels must be determined


experimentally at different initial pressures.
9.3. Future Perspective for Research on FLs. In
pmax (kPa)
405.3
250.0
250.0
101.3
911.9
101.3
689.4
101.3

101.3
101.3
101.3
101.3
101.3

anticipation of future applications of renewable and alternative


fuels such as hydrogen and ammonia and given the need to
explore the flammability of fuel mixtures, more experimental
lower flammability limit

data are needed on the FLs of mixtures with three or more fuels.
pmin (kPa)
10.1
50.0
50.0
101.3
101.3
101.3
20.0
101.3

40.0
50.7
50.7
20.0
20.0

Le Chatelier’s law has been shown to be very accurate for


determining LFLs. However, it shows inaccuracies in determin-
ing the UFLs of some fuel mixtures.5 In this case, the influence of
the direction of flame propagation needs to be further
Tmax (K)
723.0
873.0
573.0
673.0
573.0
473.0
523.0
373.0

423.0
385.0
385.0
460.0
470.0

investigated, and new experiments on UFLs should include


sampling of the combustion products to determine their
composition.
Tmin (K)

The one-dimensional simulation approach could be improved


143.0
278.0
293.0
143.0
291.0
373.0
273.0
292.0

373.0
415.0
415.0
415.0
415.0

with two different objectives. One objective would be to make it


more realistic by including more detailed schemes for soot
formation and thermal radiation losses to deepen the under-
hydrogen + ammonia

jet fuel + ethanol

standing of the phenomenon. The other objective would be to


a
alternative fuel

Jet A + HEFA
carbon monoxide

Jet fuel + SIP


hydrogen sulfide

make the one-dimensional simulations more practical for the


aviation fuels

prediction of FLs. In this case, the simulation of fuels


FT-IPK
hydrogen

methanol
ammonia

characterized by more complex molecules could be performed


syngasb

ethanol

SIP

by adopting reduced chemical kinetic mechanisms76 or the


approach known as flamelet generated manifolds.341
a

AJ https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

Methods for estimating FLs of pure fuels have shown great measured at various oxygen dilutions and concentrations
accuracy in determining LFLs and only good accuracy in because they are important for the safe operation of industrial
determining UFLs. In the development of new estimation applications.
methods for the determination of UFLs, the critical sooting The mechanism of flammability has been studied theoretically
equivalence ratio (ψc) could be included as one of the relevant and computationally by asymptotic and numerical methods,
parameters to obtain new correlations for CAFT or directly for respectively, with detailed and one-step chemical kinetics and
UFL. optically thin and thick radiation models in various flame
There is evidence that some fuels, such as dimethyl ether, geometries. These studies have shown that the fundamental
develop cool flames at the UFL.190 In addition, for some fuels, flammability limits are determined by the heat loss of thermal
UFLs determined with UFP are higher than those determined radiation in the suppression of chain-branching reactions. It has
with DFP.251 Therefore, experimental studies are needed to also been shown that, for a stretched flame, the flammable region
clarify the influence of the direction of flame propagation. These is bounded by a stretch extinction limit and a radiation
experiments should include sampling the combustion products extinction limit and may lie outside the region bounded by the
at the UFL to determine whether they are typical products of fundamental flammability limits.
normal or cool flames. Note that the results of such experimental There is a critical Lewis number below and above which the
studies could affect the methods used to determine the burning limits of a stretched flame and the fundamental
fundamental flammability limits, which should depend only on flammability limit can generally be described by a G-curve and
the fuel concentration and not on the experimental config- a K-curve, respectively. These curves determine whether the
uration or flammability criterion. extinction limits of a stretched flame can be directly extrapolated
As far as the authors are aware, there are no two- or three- to the fundamental flammability limit at zero stretch. At a very
dimensional numerical studies of flame propagation near the low Lewis number, a flame ball consisting only of diffusion and
flammability limits. That is, the field of flammability limits has thermal radiation can exist well below the fundamental
not benefited from recent advances in computational fluid flammability limit.
dynamics and computational capabilities. Such studies could be Experimental determination of flammability limits is detailed
useful to understand the influence of wall quenching on FLs, by several standards, namely DIN 51649-1, EN 1839(T), EN
which in turn could be useful for developing experimental 1839(B), ASHRAE, ASTM E681-01, and others. The most
devices to determine the FLs. In addition, the two- and three- commonly used test configurations are the vertical tube and
dimensional simulations could help clarify the role of buoyancy spherical vessel, which have been used for more than a century.
in experiments. Reduced chemical kinetic mechanisms or It should be noted that flammability limits depend on the
flamelet generated manifolds could also be used for this type volume of the vessel up to a critical volume value. For spherical
of study. vessels, for example, there is a slight variation in experimental
From this review article, it is also clear that the effect of initial flammability limits for volumes of 12 L or more.
pressure on FLs needs to be further investigated, not only to Other factors such as the direction of flame propagation also
improve the understanding of the phenomenon, but also affect the experimental values of flammability limits. The effect
because there is no reliable estimation method that takes into of buoyancy is evident when upper flammability limits (UFLs)
account the effect of pressure on FLs. In addition, the estimation are determined, especially for upward flame propagation (UFP)
methods for FLs of fuels with diluents also need to be improved. inside tubes. For lower flammability limits (LFLs), there is not
Finally, the theoretical approach developed by Spalding30 much difference between experimental values determined with
could be improved by removing some of the considerations he UFP and downward flame propagation (DFP). Experiments
made in the 1950s. For example, the dimensionless reaction rate under microgravity conditions have shown that buoyancy can
assumed by Spalding was φ = αθn. Therefore, the adoption of an significantly affect the limits. The UFLs determined under
Arrhenius form for this reaction rate would improve the current microgravity conditions are usually somewhere between the
results. In addition, the heat loss term was not explicitly given by values determined with DFP and UFP.
Spalding. It could be written using Planck’s mean absorption The methods developed to determine the flammability limits
coefficients for optically thin conditions. Then the relationship under reference conditions are generally classified as group
between the thermal radiation and the critical value of the heat contribution methods and calculated adiabatic flame temper-
loss term could be determined. It is clear that this approach ature (CAFT) methods. The latter are most commonly used
complicates the search for an analytical solution to the problem. because they relate the fundamental theory of equilibrium
In case this problem cannot be solved by purely analytical combustion to the flammability limits. These methods show
methods, a mixed approach of analytical and numerical methods high accuracy in determining the LFL and good accuracy in
could be an alternative. determining the UFL. Temperature dependence can also be
estimated using the available methods. However, there is no
10. CONCLUSIONS available method that accounts for the pressure dependence of
The fundamental flammability limits of a planar and unstretched the flammability limits. The effect of diluents such as N2 and
flame have been distinguished from the extinction limits of a CO2, including the fuel inertization point, can be estimated
stretched flame and from measured empirical flammability using published methods. Nevertheless, the impact of other
limits. Although theoretical studies have shown that the diluents such as water vapor or helium need further study.
fundamental flammability limits are determined solely by fuel Experimental flammability limits of several renewable and
chemistry, temperature, pressure, and thermal radiation, the alternative fuels are available at various initial temperatures and
extinction limits of a flame can be affected by flame geometry, pressures. The flammability limits of hydrogen have been
flame length, and the Lewis number of the mixture. Despite determined experimentally over a wide range of initial
these differences, the empirically determined flammability limits temperatures and pressures. However, the available database
of single-component fuels and fuel mixtures are still being for other renewable fuels is still under development. Alternative
AK https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

aviation fuels are an example of a database under development. Notes


In addition, flammability limits of fuel mixtures may not be The authors declare no competing financial interest.
available. In such cases, Le Chatelier’s law may be used. Biographies
Although this law is very accurate in determining the LFLs, it is
Andrés Z. Mendiburu is a professor at the Federal University of Rio
less accurate in determining the UFLs of fuel mixtures. Available
Grande do Sul (UFRGS), Brazil. He earned his bachelor’s degree in
methods based on the CAFT can also be used to estimate the
energy engineering from Santa University (Peru) in 2008 and his Ph.D.
flammability limits of renewable and alternative fuels.
in mechanical engineering from São Paulo State University (Brazil) in
In light of the issues discussed in this review, the authors argue
2016. He is interested in deflagrations, detonations, DDT, and
that future studies of flammability limits could incorporate more
alternative fuels. Andrés coordinates the International Research
realistic Arrhenius-type reactions and radiative heat losses to
Group for Energy Sustainability (IRGES). He is a coauthor of the
promote theoretical developments. To improve the one-
book Applied Combustion (in Portuguese).
dimensional simulations, it is also suggested that more detailed
mechanisms for soot formation be included. Experimental João A. de Carvalho, Jr., is a professor at the Department of Chemistry
determination of flammability limits should include the and Energy of São Paulo State University (UNESP). He coordinated
determination of product composition and verification of the (supervised) several research projects, master’s dissertations, and
UFLs of some fuel mixtures with respect to the inaccuracies of doctoral theses. He received his Ph.D. in aerospace engineering from
Le Chatelier’s law in the UFLs. In addition, two- and three- Georgia Institute of Technology in 1983. João was head of the
dimensional simulations could be helpful in understanding the Combustion and Propulsion Associate Laboratory of the Brazilian
effects of wall quenching and buoyancy. Space Research Institute (INPE) for 8 years. He is a coauthor of the
Regarding estimation methods, one-dimensional simulations book Applied Combustion (in Portuguese).
need to be made more practical to determine the flammability Yiguang Ju is the Robert Porter Patterson Professor at Princeton
limits of more complex fuels. This can be accomplished by using University. He received his bachelor’s degree in engineering
techniques to simplify chemical kinetic mechanisms. Note that thermophysics from Tsinghua University in 1986 and his Ph.D. degree
improvement in methods for determining UFLs could be in mechanical and aerospace engineering from Tohoku University
achieved by including the critical sooting equivalence ratio in the (Japan) in 1994. He is an associate editor for AIAA Journal, Proceedings
correlations. In addition, methods for determining flammability of the Combustion Institute, and Combustion Science and Technology. He
limits at different initial pressures need to be developed and received the Bessel Research Award from the von Humboldt
methods for estimating flammability limits of dilute fuels need to Foundation and the 2021 AIAA Propellants and Combustion Award.
be improved.

■ ASSOCIATED CONTENT ■ ACKNOWLEDGMENTS


The authors are grateful to CNPq (Conselho Nacional de
*
sı Supporting Information
Desenvolvimento Científico e Tecnológico) for supporting this
The Supporting Information is available free of charge at work through Projects 423369/2018-0 and 308915/2022-4, and
https://pubs.acs.org/doi/10.1021/acs.energyfuels.2c03598. to FAPERGS (Fundação de Amparo à Pesquisa do Estado de
List of experimental works for different geometries; Rio Grande do Sul) for supporting this work through Project
coefficients and information on the use of the main 21/2551-0000677-3. The authors would also like to thank the
correlations listed in the review article; compilation of grant support from DOE BES award DE-SC0021135 and the
experimental results for fuel mixtures; flammability limits ExxonMobil fuel research grant.
with oxygen and cool flame chemistry (PDF)
■ NOMENCLATURE

■ AUTHOR INFORMATION
Corresponding Author
a = stretch rate, s−1
e = Euler’s number
Ea = activation energy, kJ/kmol
Andrés Z. Mendiburu − Department of Mechanical cp = heat capacity at constant pressure, kJ/kg·K
Engineering, Federal University of Rio Grande do Sul T = temperature, K
(UFRGS), Porto Alegre, Rio Grande do Sul 90050-170, d = characteristic length, m
Brazil; International Research Group for Energy Sustainability D = ordinary diffusion coefficient, m2/s
(IRGES), Porto Alegre, Rio Grande do Sul 90050-170, G = combined function of stretch rate and the stream function
Brazil; orcid.org/0000-0003-4733-625X; h0f,i = enthalpy of formation of the species i, kJ/kg
Email: andresmendiburu@ufrgs.br h̅0f,i = enthalpy of formation of the species i, kJ/mol
h̅i = absolute enthalpy of species i, kJ/mol
Authors
Δh̅i = sensible enthalpy of species i, kJ/mol
João A. Carvalho Jr. − Department of Chemistry and Energy, hC,i = heat released at the combustion process of i, kJ/mol
São Paulo State University (UNESP), Guaratinguetá, São HC = heat released at the complete combustion of 1 mol of
Paulo 12510-410, Brazil; International Research Group for fuel, kJ
Energy Sustainability (IRGES), Porto Alegre, Rio Grande do k = thermal conductivity coefficient, W/m·K
Sul 90050-170, Brazil KP = Planck’s mean absorption coefficient for species i
Yiguang Ju − Department of Mechanical and Aerospace l = normalized Lewis number deviation from unity
Engineering, Princeton University, Princeton, New Jersey L = heat losses, kJ/m3
08544, United States L̅ = mean value of the heat loss, kJ/m3
Complete contact information is available at: ṁ S = mass flux rate, kg/m2·s
https://pubs.acs.org/10.1021/acs.energyfuels.2c03598 nF = number of moles of fuel, mol
AL https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

n*F = number of moles of gaseous fuel which occupies 12 L at HEFA50 = hydroprocessed esters and fatty acids
T0, mole Jet A = Jet A fuel
nox = number of moles of oxidant, mol LFL = lower flammability limit
p = pressure, Pa UFL = upper flammability limit
pi = partial pressure of ith species in a gas mixture, Pa UFP = upward flame propagation
ΔhC = heat of combustion, kJ/kg fuel HFP = horizontal flame propagation
R2 = squared correlation coefficient SIP = synthesized isoparaffins from hydroprocessed fer-
R = radial coordinate, m mented sugars
SL = laminar flame speed, m/s
T∞ = ambient temperature, K
Tb = adiabatic flame temperature, K
■ REFERENCES
(1) Ju, Y.; Guo, H.; Maruta, K.; Liu, F. On the Extinction Limit and
Tf = flame temperature, K Flammability Limit of Non-Adiabatic Stretched Methane−Air
tchem = time scale for chemical reaction, s Premixed Flames. J. Fluid Mech. 1997, 342, 315−334.
tinv = time scale for inviscid buoyant convection, s (2) Ju, Y.; Maruta, K.; Niioka, T. Combustion Limits. Appl. Mech. Rev.
tvis = time scale for viscous buoyant convection, s 2001, 54 (3), 257−277.
tcond = time scale for conduction heat loss to the walls, s (3) Van den Schoor, F.; Verplaetsen, F. The Upper Flammability
trad = time scale for radiation heat loss, s Limit of Methane/Hydrogen/Air Mixtures at Elevated Pressures and
u = axial component of velocity, m/s Temperatures. Int. J. Hydrogen Energy 2007, 32 (13), 2548−2552.
U = axial mass flow rate, kg/s (4) Miao, H.; Lu, L.; Huang, Z. Flammability Limits of Hydrogen-
vair = number of moles of oxygen in the oxidant, mol Enriched Natural Gas. Int. J. Hydrogen Energy 2011, 36 (11), 6937−
6947.
vsair = number of moles of oxygen in oxidant for stoichiometric (5) Mendiburu, A. Z.; Coronado, C. R.; de Carvalho, J. A. Difficulties
reaction, mol on the Determination of the Flammability Limits of Fuel Mixtures by
* = number of moles of oxygen in excess, mol
vair the Law of Le Chatelier. Process Saf. Environ. Prot. 2020, 142, 45−55.
v = radial component of velocity, m/s (6) Ju, Y.; Reuter, C. B.; Yehia, O. R.; Farouk, T. I.; Won, S. H.
Vix = diffusion velocity of species i in the x-direction Dynamics of Cool Flames. Prog. Energy Combust. Sci. 2019, 75, 100787.
Wi = molecular weight of species i, kg/kmol (7) Wierzba, I.; Ale, B. B. Rich Flammability Limits of Fuel Mixtures
w = net rate of reaction, kmol/m3·s Involving Hydrogen at Elevated Temperatures. Int. J. Hydrogen Energy
x = axial coordinate 2000, 25 (1), 75−80.
xC = number of moles of carbon (C) in fuel, mol (8) Pfahl, U. J.; Ross, M. C.; Shepherd, J. E.; Pasamehmetoglu, K. O.;
xH = number of moles of hydrogen (H) in fuel, mol Unal, C. Flammability Limits, Ignition Energy, and Flame Speeds in
H2-CH4-NH3- N2O-O2-N2Mixtures. Combust. Flame 2000, 123 (1−
xO = number of moles of oxygen (O) in fuel, mol 2), 140−158.
yi = mole fraction of species i (9) Rehbein, M. C.; Meier, C.; Eilts, P.; Scholl, S. Mixtures of
Yi = mass fraction of species i Ammonia and Organic Solvents as Alternative Fuel for Internal
Greek Symbols Combustion Engines. Energy Fuels 2019, 33 (10), 10331−10342.
γ = specific heat ratio (10) Siddiqui, O.; Dincer, I. Design and Optimization of a Dual
σb = Stefan−Boltzmann constant (5.670367 × 10−8 W·m−2· Renewable Energy-Based Plant Utilizing Integrated Hydrogen to
Ammonia and Fuel Cell Systems. Energy Fuels 2021, 35 (1), 670−689.
K−4) (11) Valera-Medina, A.; Amer-Hatem, F.; Azad, A. K.; Dedoussi, I. C.;
α(ϕ) = flammability limit exponent De Joannon, M.; Fernandes, R. X.; Glarborg, P.; Hashemi, H.; He, X.;
αt = thermal diffusivity, m2 Mashruk, S.; et al. Review on Ammonia as a Potential Fuel: From
ϕ = equivalence ratio Synthesis to Economics. Energy Fuels 2021, 35 (9), 6964−7029.
φ = dimensionless reaction rate (12) Liu, Z.; Lian, T.; Li, W.; Cao, C.; Xiong, H.; Li, Y. Mini Review of
ρ = density, kg/m3 Current Combustion Research Progress of Biodiesel and Model
δ = flame thickness, m Compounds for Gas Turbine Application. Energy Fuels 2021, 35 (17),
ω̇ i = rate of production of species i, kmol/m3·s 13569−13584.
3 (13) Al Qubeissi, M.; Sazhin, S. S.; Al-Esawi, N.; Kolodnytska, R.;
f = mean value of fuel’s rate of destruction, kmol/m ·s
Khanal, B.; Ghaleeh, M.; Elwardany, A. Heating and Evaporation of
ψc = critical sooting equivalence ratio Droplets of Multicomponent and Blended Fuels: A Review of Recent
χ = dimensionless heat losses Modeling Approaches. Energy Fuels 2021, 35 (22), 18220−18256.
Subscripts (14) Mendiburu, A. Z.; Lauermann, C. H.; Hayashi, T. C.; Mariños, D.
ad = adiabatic J.; Rodrigues da Costa, R. B.; Coronado, C. J. R.; Roberts, J. J.; de
B = chain-branching reaction Carvalho, J. A. Ethanol as a Renewable Biofuel: Combustion
Characteristics and Application in Engines. Energy 2022, 257,
b = burned
No. 124688.
f = fuel (15) Wierzba, I.; Kilchyk, V. Flammability Limits of Hydrogen-
T = chain-termination reaction Carbon Monoxide Mixtures at Moderately Elevated Temperatures. Int.
u = unburned J. Hydrogen Energy 2001, 26 (6), 639−643.
eff = effective (16) Ciccarelli, G.; Jackson, D.; Verreault, J. Flammability Limits of
Abbreviations NH3-H2-N2-Air Mixtures at Elevated Initial Temperatures. Combust.
Flame 2006, 144 (1−2), 53−63.
ARE = absolute value of relative error, % (17) Kondo, S.; Takizawa, K.; Takahashi, A.; Tokuhashi, K.;
AARE = average of relative error, % Mizukado, J.; Sekiya, A. Flammability Limits of Olefinic and Saturated
CAFT = calculated adiabatic flame temperature Fluoro-Compounds. J. Hazard. Mater. 2009, 171 (1−3), 613−618.
DFP = downward flame propagation (18) Li, Z.; Gong, M.; Wu, J.; Zhou, Y. Flammability Limits of
FL = flammability limits Refrigerant Mixtures with 1,1,2,2-Tetrafluoroethane. Exp. Therm. Fluid
FT-IPK = Fischer−Tropsch isoparaffinic kerosene Sci. 2011, 35 (6), 1209−1213.

AM https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

(19) National Transportation Safety Board. In-Flight Breakup Over the (40) Hertzberg, M. The Theory of Flammability Limits: Flow Gradient
Atlantic Ocean Trans World Airlines Flight 800 Boeing 747-131, N93119 Effects and Flame Stretch; Bureau of Mines, U.S. Department of the
Near East Moriches, New York July 17, 1996. Defense Technical Interior: Washington, DC, 1984; 41 pp.
Information Center, August 23, 2000. https://apps.dtic.mil/sti/ (41) Joulin, G.; Clavin, P. Analyse Asymptotique Des Conditions
citations/ADA388166. d’extinction Des Flammes Laminaires. Acta Astronaut. 1976, 3 (3−4),
(20) Westbrook, C.; Heufer, K. A.; Wildenberg, A. Key Chemical 223−240.
Kinetic Steps in Reaction Mechanisms for Fuels from Biomass: A (42) Buckmaster, J. The Quenching of Deflagration Waves. Combust.
Perspective. Energy Fuels 2021, 35 (19), 15339−15359. Flame 1976, 26 (C), 151−162.
(21) Coronado, C.; Carvalho, J. A.; Andrade, J. C.; Cortez, E. V.; (43) Mitani, T. Asymptotic Theory for Extinction of Curved Flames
Carvalho, F. S.; Santos, J. C.; Mendiburu, A. Z. Flammability Limits: A with Heat Loss. Combust. Sci. Technol. 1980, 23 (3−4), 93−101.
Review with Emphasis on Ethanol for Aeronautical Applications and (44) Ronney, P. D. Understanding Combustion Processes through
Description of the Experimental Procedure. J. Hazard. Mater. 2012, Microgravity Research. Symp. Combust. 1998, 27 (2), 2485−2506.
241−242, 32−54. (45) T’ien, J. S. Diffusion Flame Extinction at Small Stretch Rates:
(22) Qi, C.; Yan, X.; Wang, Y.; Ning, Y.; Yu, X.; Hou, Y.; Lv, X.; Ding, The Mechanism of Radiative Loss. Combust. Flame 1986, 65 (1), 31−
J.; Shi, E.; Yu, J. Flammability Limits of Combustible Gases at Elevated 34.
Temperatures and Pressures: Recent Advances and Future Perspec- (46) Chao, B. H.; Law, C. K.; T’ien, J. S. Structure and Extinction of
tives. Energy Fuels 2022, 36 (21), 12896−12916. Diffusion Flames with Flame Radiation. Symp. Combust. 1991, 23 (1),
(23) Lovachev, L. A.; Babkin, V. S.; Bunev, V. A.; V’Yun, A. V.; 523−531.
Krivulin, V. N.; Baratov, A. N. Flammability Limits: An Invited Review. (47) Maruta, K.; Yoshida, M.; Guo, H.; Ju, Y.; Niioka, T. Extinction of
Combust. Flame 1973, 20 (2), 259−289. Low-Stretched Diffusion Flame in Microgravity. Combust. Flame 1998,
(24) Zabetakis, M. G. Flammability Characteristics of Combustible 112 (1−2), 181−187.
Gases and Vapors; BM-BULL-627; Bureau of Mines, U.S. Department (48) Ju, Y.; Masuya, G.; Liu, F.; Hattori, Y.; Riechelmann, D.
of the Interior: 1965; 121 pp. DOI: 10.2172/7328370. Asymptotic Analysis of Radiation Extinction of Stretched Premixed
(25) Kuo, K. K. Principles of Combustion, 2nd ed.; John Wiley & Sons, Flames. Int. J. Heat Mass Transfer 2000, 43 (2), 231−239.
Inc.: 2005; 732 pp. (49) Buckmaster, J. The Effects of Radiation on Stretched Flames.
(26) Turns, S. R. An Introduction To Combustion: Concepts and Combust. Theory Model. 1997, 1, 1−11.
Applications, 3rd ed.; McGraw Hill: New York, 2011; 752 pp. (50) Ju, Y.; Minaev, S. Dynamics and Flammability Limit of Stretched
(27) Glassman, I.; Yetter, R. A.; Glumac, N. G. Combustion, 5th ed.; Premixed Flames Stabilized by a Hot Wall. Proc. Combust. Inst. 2002, 29
Elsevier: New York, 2015; 757 pp. (1), 949−956.
(28) Date, A. W. Analytic Combustion, 2nd ed.; Springer Nature (51) Law, C. K.; Egolfopoulos, F. N. A Kinetic Criterion of
Singapore: Singapore, 2020; 405 pp. DOI: 10.1007/978-981-15-1853- Flammability Limits: The C-H-O-Inert System. Symp. Combust.
9. 1991, 23 (1), 413−421.
(29) Qiao, L.; Gan, Y.; Nishiie, T.; Dahm, W. J. A.; Oran, E. S. (52) Wang, T.; Liang, H.; Luo, Z.; Su, B.; Liu, L.; Su, Y.; Wang, X.;
Extinction of Premixed Methane/Air Flames in Microgravity by Cheng, F.; Deng, J. Near Flammability Limits Behavior of Methane-Air
Diluents: Effects of Radiation and Lewis Number. Combust. Flame Mixtures with Influence of Flammable Gases and Nitrogen: An
2010, 157 (8), 1446−1455. Experimental and Numerical Research. Fuel 2021, 294, 120550.
(30) Spalding, D. B. A Theory of Inflammability Limits and Flame- (53) Liang, W.; Law, C. K. Extended Flammability Limits of N-
Quenching. Proc. R. Soc. London. Ser. A. Math. Phys. Sci. 1957, 240, 83− Heptane/Air Mixtures with Cool Flames. Combust. Flame 2017, 185,
100. 75−81.
(31) Bergman, T. L.; Lavine, A. S. Fundamentals of Heat and Mass (54) Law, C. K.; Egolfopoulos, F. N. A Unified Chain-Thermal
Transfer, 8th ed.; Wiley: 2017; 1046 pp. Theory of Fundamental Flammability Limits. Symp. Combust. 1992, 24
(32) Mendiburu, A. Z.; de Carvalho, J. A.; Coronado, C. R. Estimation (1), 137−144.
of Lower Flammability Limits of C-H Compounds in Air at (55) Kee, R. J.; Grcar, J. F.; Smooke, M. D.; Miller, J. A.; Meeks, E.
Atmospheric Pressure, Evaluation of Temperature Dependence and PREMIX: A FORTRAN Program for Modeling Steady, Laminar, One-
Diluent Effect. J. Hazard. Mater. 2015, 285, 409−418. Dimensional Premixed Flames; Reaction Design: San Diego, CA, 1998.
(33) Mendiburu, A. Z.; de Carvalho, J. A.; Coronado, C. R.; (56) Ju, Y.; Masuya, G.; Ronney, P. D. Effects of Radiative Emission
Chumpitaz, G. A. Determination of Lower Flammability Limits of C− and Absorption on the Propagation and Extinction of Premixed Gas
H−O Compounds in Air and Study of Initial Temperature Depend- Flames. Symp. (Int.) Combust. 1998, 27 (2), 2619−2626.
ence. Chem. Eng. Sci. 2016, 144, 188−200. (57) Ju, Y.; Guo, H.; Liu, F.; Maruta, K. Effects of the Lewis Number
(34) Mendiburu, A. Z.; de Carvalho, J. A.; Coronado, C. R. Estimation and Radiative Heat Loss on the Bifurcation and Extinction of CH4/O2-
of Upper Flammability Limits of C-H Compounds in Air at Standard N2-He Flames. J. Fluid Mech. 1999, 379, 165−190.
Atmospheric Pressure and Evaluation of Temperature Dependence. J. (58) Ju, Y.; Guo, H.; Maruta, K.; Niioka, T. Determination of Burning
Hazard. Mater. 2016, 304, 512−521. Velocity and Flammability Limit of Methane/Air Mixture Counterflow
(35) Mendiburu, A. Z.; de Carvalho, J. A.; Coronado, C. R. Flames Using Determination of Burning Velocity and Flammability
Determination of Upper Flammability Limits of C-H-O Compounds Limit of Methane/Air Mixture Using Counterflow Flames. Jpn. J. Appl.
in Air at Reference Temperature and Atmospheric Pressure. Fuel 2017, Phys. 1999, 38 (2), 961−967.
188, 212−222. (59) Ronney, P. D.; Wachman, H. Y. Effect of Gravity on Laminar
(36) Williams, F. Combustion Theory: The Fundamental Theory of Premixed Gas Combustion I: Flammability Limits and Burning
Chemically Reacting Flow Systems, 2nd ed.; The Benjamin/Cummings Velocities. Combust. Flame 1985, 62 (2), 107−119.
Publishing Co., Inc.: 1985; 665 pp. (60) Guo, H.; Ju, Y.; Maruta, K.; Niioka, T.; Liu, F. Radiation
(37) Hertzberg, M. The Theory of Flammability Limits: Natural Extinction Limit of Counterflow Premixed Lean Methane-Air Flames.
Convection; Bureau of Mines, U.S. Department of the Interior: Combust. Flame 1997, 109 (4), 639−646.
Washington, DC, 1976; 19 pp. (61) Guo, H.; Ju, Y.; Niioka, T. Effects of Radiative Heat Loss on the
(38) Hertzberg, M. The Theory of Flammability Limits: Conductive- Extinction of Counterflow Premixed H2-Air Flames. Combust. Theory
Convective Wall Losses and Thermal Quenching; Bureau of Mines, U.S. Model. 2000, 4 (4), 459−475.
Department of the Interior: Washington, DC, 1980; 31 pp. (62) Guo, H.; Smallwood, G. J.; Liu, F.; Ju, Y.; Gülder, Ö . L. The Effect
(39) Hertzberg, M. The Theory of Flammability Limits: Radiative Losses of Hydrogen Addition on Flammability Limit and NOx Emission in
and Selective Diffusional Demixing; Bureau of Mines, U.S. Department of Ultra-Lean Counterflow CH4/Air Premixed Flames. Proc. Combust.
the Interior: Washington, DC, 1982; 42 pp. Inst. 2005, 30 (1), 303−311.

AN https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

(63) Maruta, K.; Ju, Y.; Honda, A.; Niioka, T. Lewis Number Effect on Ph.D. Thesis, Sao Paulo State University, 2016; 272 pp. https://
Extinction Characteristics of Radiative Counterflow CH4−O2−N2− repositorio.unesp.br/handle/11449/141531.
He Flames. Twenty-Seventh Symp. Combust. 1998, 27 (2), 2611−2617. (84) EN 1839:2017 Determination of the Explosion Limits and the
(64) Bertolino, A.; Stagni, A.; Cuoci, A.; Faravelli, T.; Parente, A.; Limiting Oxygen Concentration (LOC) for Flammable Gases and
Frassoldati, A. Prediction of Flammable Range for Pure Fuels and Vapours; European Union: 2017.
Mixtures Using Detailed Kinetics. Combust. Flame 2019, 207, 120−133. (85) Schröder, V.; Molnarne, M. Flammability of Gas Mixtures: Part
(65) Mascarenhas, V. J.; Weber, C. N.; Westmoreland, P. R. 1: Fire Potential. J. Hazard. Mater. 2005, 121 (1−3), 37−44.
Estimating Flammability Limits through Predicting Non-Adiabatic (86) Macek, A. Flammability Limits: A Re-Examination. Combust. Sci.
Laminar Flame Properties. Proc. Combust. Inst. 2021, 38 (3), 4673− Technol. 1979, 21 (1−2), 43−52.
4681. (87) Design Institute for Physical Properties. DIPPR Project 801
(66) Simmie, J. M. Detailed Chemical Kinetic Models for the Database. https://www.aiche.org/dippr (accessed 2023-01-26).
Combustion of Hydrocarbon Fuels. Prog. Energy Combust. Sci. 2003, 29 (88) NOAA. Cameo Chemicals. https://cameochemicals.noaa.gov/
(6), 599−634. search/simple (accessed 2023-01-26).
(67) The San Diego Mechanism. University of California at San Diego. (89) NCBI. PubChem Compound Database. https://www.ncbi.nlm.
https://web.eng.ucsd.edu/mae/groups/combustion/mechanism.html nih.gov/pccompound (accessed 2023-01-26).
(accessed 2022-12-16). (90) ChemicalBook. Chemical Book. http://www.chemicalbook.com
(68) Metcalfe, W. K.; Burke, S. M.; Ahmed, S. S.; Curran, H. J. A (accessed 2023-01-26).
Hierarchical and Comparative Kinetic Modeling Study of C1 - C2 (91) CRC Handbook of Chemistry and Physics, 97th ed.; Haynes, W.
Hydrocarbon and Oxygenated Fuels. Int. J. Chem. Kinet. 2013, 45 (10), M., Lide, D. R., Bruno, T. J., Eds.; CRC Press: Boca Raton, FL, 2016;
638−675. 2641 pp. DOI: 10.1201/9781315380476.
(69) Rarata, G.; Szymczyk, J.; Wolanski, P. Experiments on the Upper (92) Hazardous Chemicals Handbook, 2nd ed.; Carson, P., Mumford,
Explosion Limits of Gaseous Alkanes-Oxygen Mixtures at Elevated C., Eds.; Butterworth Heinemann: London, 2002; 608 pp.
Conditions of T and P in a Spherical Vessel. Trans. Inst. Aviat. 2009, (93) The John Zink Combustion Hand Book; Baukal, C. E. J., Ed.; CRC
No. 200, 160−167. Press: 2001; 730 pp.
(70) Sarathy, S. M.; Oßwald, P.; Hansen, N.; Kohse-Höinghaus, K. (94) Bretherick’s Handbook of Reactive Chemical Hazards, 7th ed.;
Alcohol Combustion Chemistry. Prog. Energy Combust. Sci. 2014, 44, Urben, P. G., Ed.; Elsevier: 2007; 2682 pp.
40−102. (95) Kondo, S.; Takizawa, K.; Tokuhashi, K. Flammability Limits of
(71) Sánchez, A. L.; Williams, F. A. Recent Advances in Under- Binary Mixtures of Ammonia with HFO-1234yf, HFO-1234ze, HFC-
standing of Flammability Characteristics of Hydrogen. Prog. Energy 134a, and HFC-125. J. Fluor. Chem. 2013, 149, 18−23.
Combust. Sci. 2014, 41, 1−55. (96) Britton, L. G. Two Hundred Years of Flammable Limits. Process
(72) Konnov, A. A. Yet Another Kinetic Mechanism for Hydrogen Saf. Prog. 2002, 21 (1), 1−11.
Combustion. Combust. Flame 2019, 203, 14−22. (97) Zhao, F. Inert Gas Dilution Effect on Flammability Limits of
(73) Fernández-Galisteo, D.; Sánchez, A. L.; Liñán, A.; Williams, F. A. Hydrocarbon Mixtures. Ph.D. Dissertation, Texas A&M University,
One-Step Reduced Kinetics for Lean Hydrogen-Air Deflagration. 2011.
Combust. Flame 2009, 156 (5), 985−996. (98) Zhao, F.; Rogers, W. J.; Mannan, M. S. Calculated Flame
(74) Annual World Fuel Ethanol Production. Renewable Fuels Temperature (CFT) Modeling of Fuel Mixture Lower Flammability
Association. https://ethanolrfa.org/markets-and-statistics/annual- Limits. J. Hazard. Mater. 2010, 174 (1−3), 416−423.
ethanol-production (accessed 2023-01-26). (99) Zhao, F.; Rogers, W. J.; Sam Mannan, M. Experimental
(75) Marinov, N. A Detailed Chemical Kinetic Model for High Measurement and Numerical Analysis of Binary Hydrocarbon Mixture
Temperature Ethanol Oxidation. Int. J. Chem. Kinet. 1999, 31 (3), 183− Flammability Limits. Process Saf. Environ. Prot. 2009, 87 (2), 94−104.
220. (100) Zhao, F. Experimental Measurements and Modeling Prediction of
(76) Millán-Merino, A.; Fernández-Tarrazo, E.; Sánchez-Sanz, M.; Flammability Limits of Binary Hydrocarbon Mixtures. M.S. Thesis, Texas
Williams, F. A. A Multipurpose Reduced Mechanism for Ethanol A&M University, 2008; 108 pp.
Combustion. Combust. Flame 2018, 193, 112−122. (101) Jones, G. W.; Beattie, B. B. Explosive Properties of Divinyl
(77) Brown, N. J.; Bastien, L. A. J.; Price, P. N. Transport Properties Ether. Ind. Eng. Chem. 1934, 26 (5), 557−560.
for Combustion Modeling. Prog. Energy Combust. Sci. 2011, 37 (5), (102) Jones, G. W.; Miller, W. E.; Seaman, H. Explosive Properties of
565−582. Methyl Formate-Air Mixtures. Ind. Eng. Chem. 1933, 25 (6), 694−696.
(78) Westbrook, C. K.; Dryer, F. L. Simplified Reaction Mechanisms (103) Hu, X.; Yu, Q.; Sun, N.; Qin, Q. Experimental Study of
for the Oxidation of Hydrocarbon Fuels in Flames. Combust. Sci. Flammability Limits of Oxy-Methane Mixture and Calculation Based
Technol. 1981, 27 (1−2), 31−43. on Thermal Theory. Int. J. Hydrogen Energy 2014, 39 (17), 9527−9533.
(79) Westbrook, C. K.; Dryer, F. L. Chemical Kinetic Modeling of (104) Hu, X.; Yu, Q.; Sun, N.; Liu, J. Effects of High Concentrations of
Hydrocarbon Combustion. Prog. Energy Combust. Sci. 1984, 10 (1), 1− CO2 on the Lower Flammability Limits of Oxy-Methane Mixtures.
57. Energy Fuels 2016, 30 (5), 4346−4352.
(80) Chen, Z.; Qin, X.; Xu, B.; Ju, Y.; Liu, F. Studies of Radiation (105) Shang, R.; Li, G.; Wang, Z.; Wang, Z. Lower Flammability Limit
Absorption on Flame Speed and Flammability Limit of CO2 Diluted of H2/CO/Air Mixtures with N2 and CO2 Dilution at Elevated
Methane Flames at Elevated Pressures. Proc. Combust. Inst. 2007, 31 Temperatures. Int. J. Hydrogen Energy 2020, 45 (16), 10164−10175.
(2), 2693−2700. (106) Huang, L.; Wang, Y.; Pei, S.; Cui, G.; Zhang, L.; Ren, S.; Zhang,
(81) American Society for Testing and Materials. ASTM E681-09. Z.; Wang, N. Effect of Elevated Pressure on the Explosion and
Standard Test Method for Concentration Limits of Flammability of Flammability Limits of Methane-Air Mixtures. Energy 2019, 186,
Chemicals (Vapors and Gases); ASTM: 2015. DOI: 10.1520/E0681- 115840.
09R15. (107) Cui, G.; Yang, C.; Li, Z.; Zhou, Z.; Li, J. Experimental Study and
(82) Coronado, C. J. R.; Carvalho, J. A.; Andrade, J. C.; Mendiburu, A. Theoretical Calculation of Flammability Limits of Methane/Air
Z.; Cortez, E. V.; Carvalho, F. S.; Gonçalves, B.; Quintero, J. C.; Mixture at Elevated Temperatures and Pressures. J. Loss Prev. Process
Velásquez, E. I. G.; Silva, M. H.; et al. Flammability Limits of Hydrated Ind. 2016, 41, 252−258.
and Anhydrous Ethanol at Reduced Pressures in Aeronautical (108) Lesmana, H.; Zhu, M.; Zhang, Z.; Gao, J.; Wu, J.; Zhang, D.
Applications. J. Hazard. Mater. 2014, 280, 174−184. Experimental and Kinetic Modelling Studies of Flammability Limits of
(83) Mendiburu, A. Study of the Flammability Limits of Hydrocarbons Partially Dissociated NH3 and Air Mixtures. Proc. Combust. Inst. 2021,
and Alcohols in Fuel-Air and Fuel-Diluent-Air Mixtures (in Portuguese). 38 (2), 2023−2030.

AO https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

(109) Liu, X.; Zhang, Q. Influence of Initial Pressure and Temperature (130) Karim, G.; Wierzba, I.; Boon, S. Some Considerations of the
on Flammability Limits of Hydrogen-Air. Int. J. Hydrogen Energy 2014, Lean Flammability Limits of Mixtures Involving Hydrogen. Int. J.
39 (12), 6774−6782. Hydrogen Energy 1985, 10 (2), 117−123.
(110) Hu, X.; Xie, Q.; Yu, Q.; Liu, H.; Yan, F. Effect of Carbon Dioxide (131) Karim, G. a.; Wierzba, I.; Boon, S. The Lean Flammability
on the Lower Flammability Limit of Propane in O2/CO2 atm. Energy Limits in Air of Methane, Hydrogen and Carbon Monoxide at Low
Fuels 2020, 34 (4), 4993−4998. Temperatures. Cryogenics 1984, 24 (6), 305−308.
(111) Hu, X.; Xie, Q.; Zhang, J.; Yu, Q.; Liu, H.; Sun, Y. Experimental (132) Terpstra, M. A. Flammability Limits of Hydrogen-Diluent
Study of the Lower Flammability Limits of H2/O2/CO2Mixture. Int. J. Mixtures in Air. M.S. Thesis, The University of Calgary, 2012; 72 pp.
Hydrogen Energy 2020, 45 (51), 27837−27845. DOI: 10.11575/PRISM/26185.
(112) Wu, X.; Yang, Z.; Wang, X.; Lin, Y. Experimental and (133) Wierzba, I.; Wang, Q. The Flammability Limits of H2-CO-
Theoretical Study on the Influence of Temperature and Humidity on CH4Mixtures in Air at Elevated Temperatures. Int. J. Hydrogen Energy
the Flammability Limits of Ethylene (R1150). Energy 2013, 52, 185− 2006, 31 (4), 485−489.
191. (134) Wierzba, I.; Ale, B. B. Effects of Temperature and Time of
(113) Luo, Z.; Su, B.; Cheng, F.; Wang, T.; Shu, C.; Li, Y. Influences of Exposure on the Flammability Limits of Hydrogen-Air Mixtures. Int. J.
Ethane on the Flammable Limits and Explosive Oxygen Concentration Hydrog. EnergyInternational J. Hydrog. Energy 1998, 23 (12), 1197−
of Methane with Nitrogen Dilution. J. Loss Prev. Process Ind. 2018, 56,
1202.
478−485.
(135) Wierzba, I.; Harris, K.; Karim, G. Effect of Low Temperature on
(114) Luo, Z.; Liang, H.; Wang, T.; Cheng, F.; Su, B.; Liu, L.; Liu, B.
the Rich Flammability Limits in Air of Hydrogen and Some Fuel
Evaluating the Effect of Multiple Flammable Gases on the Flammability
Mixtures Containing Hydrogen. Int. J. Hydrogen Energy 1992, 17 (2),
Limit of CH4: Experimental Study and Theoretical Calculation. Process
Saf. Environ. Prot. 2021, 146, 369−376. 149−152.
(115) Wang, T.; Zhou, Y.; Luo, Z.; Wen, H.; Zhao, J.; Su, B.; et al. (136) Li, Z.; Gong, M.; Sun, E.; Wu, J.; Zhou, Y. Effect of Low
Flammability Limit Behavior of Methane with the Addition of Gaseous Temperature on the Flammability Limits of Methane/Nitrogen
Fuel at Various Relative Humidities. Process Saf. Environ. Prot. 2020, Mixtures. Energy 2011, 36 (9), 5521−5524.
140, 178−189. (137) Van den Schoor, F.; Norman, F.; Vandermeiren, K.;
(116) Hao, Q.; Luo, Z.; Wang, T.; Xie, C.; Zhang, S.; Bi, M.; Deng, J. Verplaetsen, F.; Berghmans, J.; Van den Bulck, E. Flammability Limits,
The Flammability Limits and Explosion Behaviours of Hydrogen- Limiting Oxygen Concentration and Minimum Inert Gas/Combustible
Enriched Methane-Air Mixtures. Exp. Therm. Fluid Sci. 2021, 126, Ratio of H2/CO/N2/Air Mixtures. Int. J. Hydrogen Energy 2009, 34
110395. (4), 2069−2075.
(117) Lv, Z.; Yang, Z.; Chen, Y.; Zhang, Y. Experimental Studies on (138) Yu, X.; Yu, J.; Ji, W.; Lv, X.; Yan, X. A Research on Flammability
the Flammability Limit and Burning Velocity of the Mixtures of 2,3,3,3- Limits of the Refrigerant HCFC-22/Air Mixtures at Elevated Pressures.
Tetrafluoroprop-1-Ene. Fuel 2021, 298, No. 120698. J. Loss Prev. Process Ind. 2019, 61, 89−93.
(118) Jones, G. W.; Kennedy, R. E. Extinction of Propylene Flames by (139) Li, P.; Liu, Z.; Li, M.; Huang, P.; Zhao, Y.; Li, X.; Jiang, S.
Diluting with Nitrogen and Carbon Dioxide and Some Observations on the Experimental Study on the Flammability Limits of Natural Gas/Air
Explosive Properties of Propylene; RI 3395; Bureau of Mines, U.S. Mixtures at Elevated Pressures and Temperatures. Fuel 2019, 256,
Department of the Interior: Washington, DC, 1938. No. 115950.
(119) Jones, G. W.; Kennedy, R. E. Extinction of Ethylene Flames by (140) Li, P.; Li, M.; Liu, Z.; Zhao, Y.; Qian, X.; Huang, P. Effect of
Carbon Dioxid and Nitrogen. Anesth. Analg. 1930, 9 (1), 6−11. High Temperature and Sulfur Vapor on the Flammability Limit of
(120) Muth, W. A. Effect of Pressure on the Flammable Limits of Some Hydrogen Sulfide. J. Clean. Prod. 2022, 337, 130579.
Hydrocarbon-Air Mixtures. Ph.D. Dissertation, Iowa State University, (141) Li, R.; Liu, Z.; Li, P.; Li, M.; Zhao, Y. Investigation on the
1963; 248 pp. DOI: 10.31274/rtd-180813-1904. Flammability Limit and Limiting Oxygen Concentration of N2-Diluted
(121) Zabetakis, M. G.; Scott, G. S.; Jones, G. W. Limits of H2/CO/Air Mixtures at High Temperature and Pressure. Fuel 2022,
Flammability of Paraffin Hydrocarbons in Air. Ind. Eng. Chem. 1951, 43 308, No. 121955.
(9), 2120−2124. (142) Koshiba, Y.; Hasegawa, T.; Kim, B.; Ohtani, H. Flammability
(122) Heffington, W. M.; Gaines, W. R.; Renfroe, D. A. Flammability Limits, Explosion Pressures, and Applicability of Le Chatelier’s Rule to
Limits of Coal-Derived Low-BTU Gas Mixtures Containing Large Binary Alkane - Nitrous Oxide Mixtures. J. Loss Prev. Process Ind. 2017,
Amounts of Inert Gases. Combust. Sci. Technol. 1984, 36 (3−4), 191− 45, 1−8.
197. (143) Koshiba, Y.; Takigawa, T.; Matsuoka, Y.; Ohtani, H. Explosion
(123) White, A. G. VIII. Limits for the Propagation of Flame in
Characteristics of Flammable Organic Vapors in Nitrous Oxide
Inflammable Gas-Air Mixtures. Part II. Mixtures of More than One Gas
Atmosphere. J. Hazard. Mater. 2010, 183 (1−3), 746−753.
and Air. J. Chem. Soc. Trans. 1925, 127, 48−61.
(144) Koshiba, Y.; Nishida, T.; Morita, N.; Ohtani, H. Explosion
(124) Zhao, Y.; Ting, W.; Xihong, L. Experimental Studies and
Behavior of N-Alkane/Nitrous Oxide Mixtures. Process Saf. Environ.
Estimates of the Explosion Limit of Some Environmentally Friendly
Refrigerants. Combust. Sci. Technol. 2005, 177 (3), 613−626. Prot. 2015, 98 (3), 11−15.
(125) Jones, G. W.; Kennedy, R. E. Prevention of Butadiene-Air (145) Koshiba, Y.; Hasegawa, T.; Ohtani, H. Numerical and
Explosions by Addition of Nitrogen and Carbon Dioxide; RI 3691; Bureau Experimental Study of the Explosion Pressures and Flammability
of Mines, U.S. Department of the Interior: 1943. Limits of Lower Alkenes in Nitrous Oxide Atmosphere. Process Saf.
(126) Rolingson, W. R.; MacPherson, J.; Montgomery, P. D.; Environ. Prot. 2018, 118, 59−67.
Williams, B. L. Effect of Temperature on the Upper Flammable Limit (146) Mohler dos Santos, C. C.; Zanoelo, E. F. Flammability Limits of
of Methane, Ammonia, and Air Mixtures. J. Chem. Eng. Data 1960, 5 Iso-Butanol/Iso-Octane/n-Heptane Blends. Fire Saf. J. 2017, 88, 40−
(3), 349−351. 44.
(127) Mishra, D. P.; Rahman, A. An Experimental Study of (147) Hustad, J. E.; Sonju, O. K. Experimental Studies of Lower
Flammability Limits of LPG/Air Mixtures. Fuel 2003, 82, 863−866. Flammability Limits of Gases and Mixtures of Gases at Elevated
(128) Cato, R. J.; Gilbert, W. H.; Kuchta, J. M. Effect of Temperature Temperatures. Combust. Flame 1988, 71 (3), 283−294.
on Upper Flammability Limits of Hydrocarbon Fuel Vapors in Air. Fire (148) Ohtani, H. Flammability Limits of Silane/Perfluorocarbon/
Technol. 1967, 3 (1), 14−19. Nitrogen Mixtures. J. Loss Prev. Process Ind. 2004, 17 (5), 381−383.
(129) Cui, G.; Li, Z.; Yang, C. Experimental Study of Flammability (149) Ohtani, H. Experimental Study on Flammability Limits of
Limits of Methane/Air Mixtures at Low Temperatures and Elevated Perfluorocarbons in a Fluorine Atmosphere. J. Loss Prev. Process Ind.
Pressures. Fuel 2016, 181, 1074−1080. 2004, 17 (5), 377−379.

AP https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

(150) Ohtani, H.; Horiguchi, S.; Urano, Y.; Iwasaka, M.; Tokuhashi, (171) Flasinska, P.; Fraczak, M.; Piotrowski, T. Explosion Hazard
K.; Kondo, S. Flammability Limits of Arsine and Phosphine. Combust. Evaluation and Determination of the Explosion Parameters for Selected
Flame 1989, 76 (3−4), 307−310. Hydrocarbons C6 - C8. Cent. Eur. J. Energ. Mater. 2012, 9 (4), 399−
(151) Scott, B. G. S.; Kennedy, R. E.; Spolan, I.; Zabetakis, M. G.; 409.
Stanley, G. Flammability Characteristics of Ethylene; Bureau of Mines, (172) Wen, P. J.; Shu, C. M.; Chang, Z. C.; Chen, S. C.; Shyu, M. L.
U.S. Department of the Interior: Washington, DC, 1965. Effects of Initial Pressure on the Flammability Limit of OX-Air Mixture
(152) Jones, G. W.; Spolan, I.; Kennedy, R. E. Effect of High Pressures with 20-L-Apparatus. In The 17th Annual Conference of Asia Pacific
on the Flammability of Natural Gas-Air-Nitrogen Mixtures; Bureau of Occupational Safety & Health Organization; APOSHO: 2001; pp 443−
Mines, U.S. Department of the Interior: Washington, DC, 1949. 465.
(153) Huang, L.; Wang, Y.; Zhang, L.; Su, Y.; Zhang, Z.; Ren, S. (173) Bai, C.; Liu, N.; Zhang, B. Experimental Investigation on the
Journal of Loss Prevention in the Process Industries Influence of Lower Flammability Limits of Diethyl Ether/n-Pentane/Epoxypro-
Pressure on the Flammability Limits and Explosion Pressure of Ethane/ pane-Air Mixtures. J. Loss Prev. Process Ind. 2019, 57, 273−279.
Propane-Air Mixtures in a Cylinder Vessel. J. Loss Prev. Process Ind. (174) Barbosa, J. A.; Coronado, C. J. R.; Tuna, C. E.; Silva, M. H.;
2022, 74, No. 104638. Mendiburu, A. Z.; Carvalho, J. A.; de Andrade, J. C. Experimental
(154) Naegeli, D. W.; Weatherford, W. D. Practical Ignition Limits for Determination of Lower Flammability Limits of Synthesized Iso-
Low Molecular Weight Alcohols. Fuel 1989, 68 (1), 45−48. Paraffins (SIP), Jet Fuel and Mixtures at Atmospheric and Reduced
(155) Crescitelli, S.; Russo, G.; Tufano, V. The Influence of Different Pressures with Air. Fire Saf. J. 2021, 121, 103276.
Diluents on the Flammability Limits of Ethylene at High Temperatures (175) Barbosa, J. A.; Coronado, C. J. R.; de Andrade, J. C.; Tuna, C. E.;
and Pressures. J. Hazard. Mater. 1979, 3 (2), 167−175. Silva, M. H.; Carvalho Junior, J. A.; Mendiburu, A. Z. Experimental
(156) Payman, W.; Wheeler, R. V. The Effect of Pressure on the Limits Determination of Upper Flammability Limits of Synthesized Iso-
of Inflammability of Mixtures of Paraffin Hydrocarbons with Air. Chem. Paraffins (SIP), Jet Fuel and Their Mixtures with Air at Atmospheric
Soc. J. 1923, 123, 426−434. and Sub-Atmospheric Pressures. Process Saf. Environ. Prot. 2022, 160,
(157) Marmentini Vivas, B. M.; Zanoelo, É . F. An Experimental 102−115.
Investigation of Flammability Limits and Autoignition Temperatures of (176) Rios Escalante, E. S.; Rodriguez Coronado, C. J.; de Carvalho, J.
Petrofuels and Biofuels in a Tubular Burner. Combust. Sci. Technol. A., Jr. A Detailed Experimental and Numerical Assessment of the
2011, 183 (12), 1433−1444. QAV−1/Anhydrous Ethanol Blends in Their Lower Flammability
(158) Hu, X.; Yu, Q.; Sun, Y. Effects of Carbon Dioxide on the Upper Limits. Fuel 2022, 311, No. 122531.
Flammability Limits of Methane in O2/CO2 atm. Energy 2020, 208, (177) Qin, S. H.; Sun, X. X.; Lin, W. C.; Shu, C. M.; You, F.; Ho, S. C.
No. 118417. Experimental and Computational Approaches for CH4 and C2H4
(159) Jaimes, D.; McDonell, V. G.; Samuelsen, G. S. Lean Flammability Zones. Energy Fuels 2017, 31 (9), 9950−9956.
Flammability Limits of Syngas/Air Mixtures at Elevated Temperatures (178) Gibbon, H. J.; Wainwright, J.; Rogers, R. L. Experimental
and Pressures. Energy Fuels 2018, 32 (10), 10964−10973. Determination of Flammability Limits of Solvents At Elevated
(160) Bee, A.; Börner, M. Laminar Burning Speeds and Flammability Temperatures and Pressures. IChemE Symp. Ser. 1994, 134, 1−12.
Limits of CH4/O2Mixtures With Varying N2 Dilution at Sub- (179) Fu, W.; Zhang, K.; Wu, J. Flammability Limits of Benzene,
Atmospheric Conditions. Combust. Sci. Technol. 2021, 1−20. Toluene, Xylenes from 373 to 473 K and Flame-Retardant Effect of
(161) Furno, A. L.; Martindill, G. H.; Zabetakis, M. G. Limits of Steam on Benzene Series. Process Saf. Environ. Prot. 2020, 137, 328−
Flammability of Hydrazine-Hydrocarbon Vapor Mixtures. J. Chem. Eng. 339.
Data 1962, 7 (3), 375−376. (180) Chen, Q.; Yan, J.; Chen, G.; Zhao, Y.; Shi, Y.; Zeng, Z.; Pan, Q.
(162) White, A. G. CXLVIII.�Limits for the Propagation of Flame in Experimental Studies on the Flammability of Mixtures of Dimethyl
Vapour−Air Mixtures. Part I. Mixtures of Air and One Vapour at the Ether. J. Fluor. Chem. 2015, 176 (1), 40−43.
Ordinary Temperature and Pressure. J. Chem. Soc., Trans. 1922, 121, (181) Kondo, S.; Takizawa, K.; Tokuhashi, K. Effects of Temperature
1244−1270. and Humidity on the Flammability Limits of Several 2L Refrigerants. J.
(163) White, A. G.; Price, T. W. CXXXVIII. The Ignition of Ether- Fluor. Chem. 2012, 144, 130−136.
Alcohol-Air and Acetone-Air Mixtures in Contact with Heated (182) Kondo, S.; Takizawa, K.; Takahashi, A.; Tokuhashi, K. On the
Surfaces. J. Chem. Soc. Trans. 1919, 115, 1462−1505. Temperature Dependence of Flammability Limits of Gases. J. Hazard.
(164) Holmstedt, G. S. The Upper Limit of Flammability of Hydrogen Mater. 2011, 187 (1−3), 585−590.
in Air, Oxygen, and Oxygen-Inert Mixtures at Elevated Pressures. (183) Kondo, S.; Takizawa, K.; Takahashi, A.; Tokuhashi, K.; Sekiya,
Combust. Flame 1971, 17 (3), 295−301. A. Flammability Limits of Five Selected Compounds Each Mixed with
(165) Le, H.; Liu, Y.; Mannan, M. S. Lower Flammability Limits of HFC-125. Fire Saf. J. 2009, 44 (2), 192−197.
Hydrogen and Light Hydrocarbons at Subatmospheric Pressures. Ind. (184) Kondo, S.; Takizawa, K.; Takahashi, A.; Tokuhashi, K.; Sekiya,
Eng. Chem. Res. 2013, 52 (3), 1372−1378. A. A Study on Flammability Limits of Fuel Mixtures. J. Hazard. Mater.
(166) Le, H.; Nayak, S.; Mannan, M. S. Upper Flammability Limits of 2008, 155 (3), 440−448.
Hydrogen and Light Hydrocarbons in Air at Subatmospheric Pressures. (185) Kondo, S.; Takizawa, K.; Takahashi, A.; Tokuhashi, K.; Sekiya,
Ind. Eng. Chem. Res. 2012, 51 (27), 9396−9402. A. Flammability Limits of Isobutane and Its Mixtures with Various
(167) Li, A. Flammability Limits of Alternative Aviation Fuels. M.S. Gases. J. Hazard. Mater. 2007, 148 (3), 640−647.
Thesis, Purdue University, 2016; 86 pp. (186) Kondo, S.; Takizawa, K.; Takahashi, A.; Tokuhashi, K.
(168) Chang, Y. M.; Lee, J. C.; Chen, J. R.; Liaw, H. J.; Shu, C. M. Extended Le Chatelier’s Formula and Nitrogen Dilution Effect on the
Flammability Characteristics Studies on Toluene and Methanol Flammability Limits. Fire Saf. J. 2006, 41 (5), 406−417.
Mixtures with Different Vapor Mixing Ratios at 1 Atm and 150°C. J. (187) Kondo, S.; Takizawa, K.; Takahashi, A.; Tokuhashi, K.
Therm. Anal. Calorim. 2008, 93 (1), 183−188. Extended Le Chatelier’s Formula for Carbon Dioxide Dilution Effect
(169) Chang, Y. M.; Hu, K. H.; Chen, J. K.; Shu, C. M. Flammability on Flammability Limits. J. Hazard. Mater. 2006, 138 (1), 1−8.
Studies of Benzene and Methanol with Different Vapor Mixing Ratios (188) Kondo, S.; Urano, Y.; Takizawa, K.; Takahashi, A.; Tokuhashi,
under Various Initial Conditions. J. Therm. Anal. Calorim. 2006, 83 (1), K.; Sekiya, A. Flammability Limits of Multi-Fluorinated Compounds.
107−112. Fire Saf. J. 2006, 41 (1), 46−56.
(170) Chang, Y.; Lee, J.; Wu, S.; Chen, C.; Shu, C. Elevated Pressure (189) Kondo, S.; Takizawa, K.; Takahashi, A.; Tokuhashi, K.
and Temperature Effects on Flammability Hazard Assessment for Measurement and Numerical Analysis of Flammability Limits of
Acetone and Water Solutions. J. Therm. Anal. Calorim. 2009, 95 (2), Halogenated Hydrocarbons. J. Hazard. Mater. 2004, 109 (1−3), 13−
525−534. 21.

AQ https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

(190) Kondo, S.; Takahashi, A.; Tokuhashi, K. Experimental (210) Van den Schoor, F.; Verplaetsen, F.; Berghmans, J.
Exploration of Discrepancies in F-Number Correlation of Flammability Experimental and Numerical Study of the Influence of Pressure and
Limits. J. Hazard. Mater. 2003, 100 (1−3), 27−36. Temperature on the Flammability Limits of Combustible Gases in Air.
(191) Rowley, J. R.; Rowley, R. L.; Wilding, W. V. Experimental In European Combustion Meeting 2007; European Union: 2007; pp 1−7.
Determination and Re-Examination of the Effect of Initial Temperature (211) Van Den Schoor, F.; Verplaetsen, F. The Upper Explosion Limit
on the Lower Flammability Limit of Pure Liquids. J. Chem. Eng. Data of Lower Alkanes and Alkenes in Air at Elevated Pressures and
2010, 55 (9), 3063−3067. Temperatures. J. Hazard. Mater. 2006, 128 (1), 1−9.
(192) Yang, Z.; Wu, X.; Tian, T. Flammability of Trans-1, 3, 3, 3- (212) Craven, A. D.; Foster, M. G. The Limits of Flammability of
Tetrafluoroprop-1-Ene and Its Binary Blends. Energy 2015, 91, 386− Ethylene in Oxygen, Air and Air-Nitrogen Mixtures at Elevated
392. Temperatures and Pressures. Combust. Flame 1966, 10 (2), 95−100.
(193) Zhang, K.; Meng, X.; Wu, J. Flammability Limits of Binary (213) Bolshova, T. A.; Bunev, V. A.; Knyazkov, D. A.; Korobeinichev,
Mixtures of Dimethyl Ether with Five Diluent Gases. J. Loss Prev. Process O. P.; Chernov, A. A.; Shmakov, A. G.; Yakimov, S. A. Dependence of
Ind. 2014, 29 (1), 138−143. the Lower Flammability Limit on the Initial Temperature. Combust.
(194) Tian, H.; Wu, M.; Shu, G.; Liu, Y.; Wang, X. Experimental and Explos. Shock Waves 2012, 48 (2), 125−129.
Theoretical Study of Flammability Limits of Hydrocarbon−CO2M- (214) Burgess, M. J.; Wheeler, R. V. CCXXVIII. The Lower Limit of
ixture. Int. J. Hydrogen Energy 2017, 42 (49), 29597−29605.
Inflammation of Mixtures of the Paraffin Hydrocarbons with Air. J.
(195) Wu, M.; Shu, G.; Chen, R.; Tian, H.; Wang, X.; Wang, Y. A New
Chem. Soc. Trans. 1911, 99, 2013−2030.
Model Based on Adiabatic Flame Temperature for Evaluation of the
(215) Zhang, K.; Luo, T.; Li, Y.; Zhang, T.; Li, X.; Zhang, Z.; Shang, S.;
Upper Flammable Limit of Alkane-Air-CO2Mixtures. J. Hazard. Mater.
Zhou, Y.; Zhang, C.; Chen, X.; et al. Effect of Ignition, Initial Pressure
2018, 344, 450−457.
(196) Zhai, R.; Yang, Z.; Feng, B.; Sun, S. Effect of Environmental and Temperature on the Lower Flammability Limit of Hydrogen/Air
Condition on the Flammability Limits of Two Isomers of Mixture. Int. J. Hydrogen Energy 2022, 47 (33), 15107−15119.
Tetrafluoropropene. Combust. Flame 2019, 207, 295−301. (216) Zhang, K.; Gao, W.; Li, Y.; Zhang, Z.; Shang, S.; Zhang, C.;
(197) Azatyan, V.; Shebeko, Y.; Bolod’yan, I.; Shebeko, A.; Navtsenya, Chen, X.; Sun, K. Journal of Loss Prevention in the Process Industries
V.; Tomilin, A. Concentration Limits of Flammability of H2-N2-O2- Lower Flammability Limits of Ethanol, Acetone and Ethyl Acetate
Inhibitor Mixtures. Russ. J. Phys. Chem. B 2008, 2 (4), 568−574. Vapor Mixtures in Air. J. Loss Prev. Process Ind. 2022, 74, No. 104676.
(198) Zhai, R.; Yang, Z.; Zhang, Y.; Lv, Z.; Feng, B. Effect of (217) Zhang, K.; Shang, S.; Li, X.; Gao, W. Lower Flammability Limits
Temperature and Humidity on the Flammability Limits of Hydro- of NH3/H2Mixtures under Different Initial Temperatures and Initial
carbons. Fuel 2020, 270, No. 117442. Pressures. Fuel 2023, 331 (P2), 125982.
(199) Zhong, Q.; Huang, Y.; Zhao, H.; Wang, X.; Zhang, Y.; Shen, J. (218) Kul, I.; Gnann, D. L.; Beyerlein, A. L.; DesMarteau, D. D. Lower
Experimental Study on the Influence of Trifluoroiodomethane on the Flammability Limit of Difluoromethane and Percolation Theory. Int. J.
Flammability of Difluoromethane and Propane. Int. J. Refrig. 2022, 135, Thermophys. 2004, 25 (4), 1085−1095.
14−19. (219) Bui-pham, M. N.; Lutz, A. E.; Miller, J. A.; Desjardin, M.;
(200) Wu, X.; Zhang, X.; Tang, F.; Xu, S.; Dang, C.; Sun, D. O’Shaughnessey, D. M.; Zondlak, R. J. Rich Flammability Limits in
Flammability Inhibition Effects of R227ea, R125, R134a, and R744 On. CH3OH/CO/Diluent Mixtures. Combust. Sci. Technol. 1995, 109 (1−
Int. J. Refrig. 2022, 134 (2), 55−63. 6), 71−91.
(201) Askari, M. H.; Ashjaee, M. Experimental Measurement of (220) Van den Schoor, F. Influence of Pressure and Temperature on
Laminar Burning Velocity and Flammability Limits of Landfill Gas at Flammability Limits of Combustible Gases in Air. D.Ing. Thesis,
Atmospheric and Elevated Pressures. Energy Fuels 2017, 31 (3), 3196− Katholieke Universiteit Leuven, 2007; 209 pp.
3205. (221) Zlochower, I. A.; Green, G. M. The Limiting Oxygen
(202) Grune, J.; Breitung, W.; Kuznetsov, M.; Yanez, J.; Jang, W.; Concentration and Flammability Limits of Gases and Gas Mixtures. J.
Shim, W. Flammability Limits and Burning Characteristics of CO− Loss Prev. Process Ind. 2009, 22 (4), 499−505.
H2−H2O−CO2−N2Mixtures at Elevated Temperatures. Int. J. (222) Wang, Y.; Qi, C.; Ning, Y.; Lv, X.; Yu, X.; Yan, X.; Yu, J.
Hydrogen Energy 2015, 40 (31), 9838−9846. Experimental Determination of the Lower Flammability Limit and
(203) Van den Schoor, F.; Norman, F.; Tangen, L.; Sæter, O.; Limiting Oxygen Concentration of Propanal/Air Mixtures under
Verplaetsen, F. Explosion Limits of Mixtures Relevant to the Elevated Temperatures and Pressures. Fuel 2022, 326, 124882.
Production of 1,2-Dichloroethane (Ethylene Dichloride). J. Loss Prev. (223) Qi, C.; He, M.; Ning, Y.; Chen, S.; Yan, X.; Wang, Y.; Yu, X.; Yu,
Process Ind. 2007, 20 (3), 281−285. J. Experimental Measurement and Theoretical Prediction for Lower
(204) Vanderstraeten, B.; Tuerlinckx, D.; Berghmans, J.; Vliegen, S.; Flammability Limits of Ternary Hydrocarbon Mixtures. Process Saf.
Van't Oost, E.; Smit, B. Experimental Study of the Pressure and
Prog. 2022, 41 (3), 581−590.
Temperature Dependence on the Upper Flammability Limit of
(224) Akram, M. Z. Study of Hydrogen Impact on Lean Flammability
Methane/Air Mixtures. J. Hazard. Mater. 1997, 56 (3), 237−246.
Limit and Burning Characteristics of a Kerosene Surrogate. Energy
(205) Kondo, S.; Takahashi, A.; Takizawa, K.; Tokuhashi, K. On the
2021, 231, 120925.
Pressure Dependence of Flammability Limits of CH2 = CFCF3,
(225) Gomez Casanova, C. A.; Othen, E.; Sorensen, J. L.; Levin, D. B.;
CH2F2 and Methane. Fire Saf. J. 2011, 46 (5), 289−293.
(206) Jia, Z.; Pan, Y.; Yang, L.; Liu, Y.; Jiang, J. Journal of Loss Birouk, M. Measurement of Laminar Flame Speed and Flammability
Prevention in the Process Industries Lower Flammability Limits of Limits of a Biodiesel Surrogate. Energy Fuels 2016, 30 (10), 8737−
Flammable Ternary Organic Mixtures: Synergistic Behavior. J. Loss 8745.
Prev. Process Ind. 2020, 66, No. 104173. (226) Li, A.; Vozka, P.; Mastrean, A.; Kilaz, G.; Qiao, L. Lean
(207) Liaw, H. J.; Li, Z. H. Mathematical Model for Describing the Flammability Limits of Alternative Aviation Fuels. Fire Saf. J. 2019, 108,
Influence of Initial Pressure on the Flammability Limits of Light 102851.
Hydrocarbons at Subatmospheric Pressures. J. Loss Prev. Process Ind. (227) Goethals, M.; Vanderstraeten, B.; Berghmans, J.; De Smedt, G.;
2022, 77 (1), No. 104776. Vliegen, S.; Van’t Oost, E. Experimental Study of the Flammability
(208) Azatyan, V.; Shebeko, Y.; Bolod'yan, I.; Navtsenya, V. Effect of Limits of Toluene−Air Mixtures at Elevated Pressure and Temperature.
Diluents of Various Chemical Nature On Flammability Limits of Gas J. Hazard. Mater. 1999, 70 (3), 93−104.
Mixtures. Combust. Explos. Shock Waves 2006, 42 (6), 708−714. (228) Shebeko, Y.; Tsarichenko, S.; Korolchenko, A.; Trunev, A.;
(209) Shebeko, A.; Shebeko, Y.; Zuban’, A.; Navtsenya, V.; Azatyan, V. Navzenya, V.; Papkov, S.; Zaitzev, A. Burning Velocities and
Influence on Fluorocarbons Flammability Limits in the Mixtures of H2- Flammability Limits of Gaseous Mixtures at Elevated Temperatures
N2O and CH4-N2O. Russ. J. Phys. Chem. B 2014, 8 (1), 65−70. and Pressures. Combust. Flame 1995, 102 (4), 427−437.

AR https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

(229) Markus, D.; Schildberg, H.-P.; Wildner, W.; Krdzalic, G.; Maas, (248) Shoshin, Y. L.; de Goey, L. P. H. Experimental Study of Lean
U. Flammability Limits of Premixed Methane/Methanol/Air Flames. Flammability Limits of Methane/Hydrogen/Air Mixtures in Tubes of
Combust. Sci. Technol. 2003, 175 (11), 2095−2112. Different Diameters. Exp. Therm. Fluid Sci. 2010, 34 (3), 373−380.
(230) Takahashi, A.; Urano, Y.; Tokuhashi, K.; Kondo, S. Effect of (249) Kondo, S.; Urano, Y.; Takahashi, A.; Tokuhashi, K.
Vessel Size and Shape on Experimental Flammability Limits of Gases. J. Reinvestigation of Flammability Limits Measurement of Methane by
Hazard. Mater. 2003, 105 (1−3), 27−37. the Conventional Vessel Method with AC Discharge Ignition. Combust.
(231) Van den Schoor, F.; Hermanns, R. T. E.; van Oijen, J. A.; Sci. Technol. 1999, 145 (1−6), 1−15.
Verplaetsen, F.; de Goey, L. P. H. Comparison and Evaluation of (250) Kuznetsov, M.; Kobelt, S.; Grune, J.; Jordan, T. Flammability
Methods for the Determination of Flammability Limits, Applied to Limits and Laminar Flame Speed of Hydrogen−Air Mixtures at Sub-
Methane/Hydrogen/Air Mixtures. J. Hazard. Mater. 2008, 150 (3), Atmospheric Pressures. Int. J. Hydrogen Energy 2012, 37 (22), 17580−
573−581. 17588.
(232) De Smedt, G.; De Corte, F.; Notelé, R.; Berghmans, J. (251) Coward, H. F.; Jones, G. W. Limits of Flammability of Gases and
Comparison of Two Standard Test Methods for Determining Vapors; Bulletin 503; Bureau of Mines, U.S. Department of the Interior:
Explosion Limits of Gases at Atmospheric Conditions. J. Hazard. 1952; 155 pp.
Mater. 1999, 70 (3), 105−113. (252) Mendiburu, A. Z.; Serra, A. M.; Andrade, J. C.; Silva, L. M.;
(233) Heinonen, E. W.; Tapscott, R. E.; Crawford, F. R. Methods Santos, J. C.; de Carvalho, J. A. Characterization of the Flame Front
Development for Measuring and Classifying Flammability/Combustibility Inversion of Ethanol−Air Deflagrations inside A Closed Tube. Energy
of Refrigerants. Final Report; U.S. Department of Energy: 1994; 93 pp. 2019, 187, 115932.
DOI: 10.2172/114574. (253) Serra, A. M.; Andrade, J. C.; Silva, L. M.; Santos, J. C.; Silveira, J.
(234) Glukhov, I. S.; Shebeko, Y. N.; Shebeko, A. Y.; Zuban, A. V. C.; de Carvalho, J. A.; Mendiburu, A. Z. Experimental Observation of
Journal of Loss Prevention in the Process Industries An Investigation of Ethanol−Air Premixed Flames Propagating inside a Closed Tube with
Flammability Limits of Gaseous Mixtures at Reduced Pressures. J. Loss High Aspect Ratio. J. Brazilian Soc. Mech. Sci. Eng. 2023, 45 (2), 80.
Prev. Process Ind. 2020, 66, 104195. (254) White, A. G. Limits for the Propagation of Flame at Various
(235) Yu, X.; Yan, X.; Ji, W.; Luo, C.; Yao, F.; Yu, J. Effect of Super- Temperatures in Mixtures of Ammonia with Air and Oxygen. J. Chem.
Ambient Conditions on the Upper Explosion Limit of Ethane/Oxygen Soc. Trans. 1922, 121, 1688−1695.
and Ethylene/Oxygen Mixtures. J. Loss Prev. Process Ind. 2019, 59, (255) Krivulin, V. N.; Lovachev, L. A.; Kudryavtsev, E. A.; Baratov, A.
100−105. N. Flammability Limits Ammonia Air Mixtures (Translated). Fiz.
(236) Knyazkov, D. A.; Bolshova, T. A.; Shvartsberg, V. M.; Goran. i Vzryva 1975, 11 (6), 759−764.
Gerasimov, I. E.; Shmakov, A. G.; Korobeinichev, O. P. Effect of (256) Jarosinski, J.; Podfilipski, J.; Fodemski, T. Properties of Flames
Inhibitors on Flammability Limits of Dimethyl Ether/Air Mixtures. Propagating in Propane-Air Mixtures near Flammability and Quench-
Proc. Combust. Inst. 2019, 37 (3), 4267−4275. ing Limits. Combust. Sci. Technol. 2002, 174 (1), 167−187.
(237) Li, S.; Zhang, Y.; Qiu, X.; Li, B.; Zhang, H. Effects of Inert (257) Jarosinski, J. A Survey of Recent Studies on Flame Extinction.
Dilution and Preheating Temperature on Lean Flammability Limit of Prog. Energy Combust. Sci. 1986, 12, 81−116.
Syngas. Energy Fuels 2014, 28 (5), 3442−3452. (258) Sibulkin, M.; Frendi, A. Prediction of Flammability Limit of an
(238) Ale, B. J. M.; Bruning, F.; Koenders, H. A. A. The Limits of Unconfined Premixed Gas in the Absence of Gravity. Combust. Flame
Flammability of Mixtures of Ammonia, Hydrogen and Methane in 1990, 82 (3−4), 334−345.
Mixtures of Nitrogen and Oxygen at Elevated Temperatures and (259) Lakshmisha, K. N.; Paul, P. J.; Mukunda, H. S. On the
Pressures. J. Hazard. Mater. 1981, 4 (3), 283−289. Flammability Limit and Heat Loss in Flames with Detailed Chemistry.
(239) Crowl, D. A.; Louvar, J. F. Chemical Process Safety, 2nd ed.; Symp. Combust. 1991, 23, 433−440.
Prentice Hall PTR: 2002; 625 pp. (260) Ronney, P. D. Effect of Chemistry and Transport Properties on
(240) Kuchta, J. M. Investigation of Fire and Explosion Accidents in the Near-Limit Flames at Microgravity. Combust. Sci. Technol. 1988, 59 (1−
Chemical, Mining, and Fuel-Related Industries: A Manual; Bulletin 680; 3), 123−141.
Bureau of Mines, U.S. Department of the Interior: 1985; 84 pp. (261) Jarosinski, J.; Podfilipski, J.; Gorczakowski, A.; Veyssiere, B.
(241) Liao, S. Y.; Jiang, D. M.; Huang, Z. H.; Cheng, Q.; Gao, J.; Hu, Experimental Study of Flame Propagation in Propane-Air Mixture near
Y. Approximation of Flammability Region for Natural Gas-Air-Diluent Rich Flammability Limits in Microgravity. Combust. Sci. Technol. 2002,
Mixture. J. Hazard. Mater. 2005, 125 (1−3), 23−28. 174 (9), 21−48.
(242) Liao, S. Y.; Cheng, Q.; Jiang, D. M.; Gao, J. Experimental Study (262) Ronney, P. D. Effect of Chemistry and Transport Properties on
of Flammability Limits of Natural Gas-Air Mixture. J. Hazard. Mater. Near-Limit Flames at Microgravity. Combust. Sci. Technol. 1988, 59 (1−
2005, 119 (1−3), 81−84. 3), 123−141.
(243) Liao, S. Y.; Jiang, D. M.; Cheng, Q.; Gao, J.; Hu, Y. (263) Ronney, P. D. On the Mechanisms of Flame Propagation Limits
Approximations of Flammability Characteristics of Liquefied Petro- and Extinguishment-Processes at Microgravity. Symp. Combust. 1989,
leum Gas−Air Mixture with Exhaust Gas Recirculation (EGR). Energy 22 (1), 1615−1623.
Fuels 2005, 19 (1), 324−325. (264) Maruta, K.; Yoshida, M.; Ju, Y.; Niioka, T. Experimental Study
(244) Pu, Y.; Hu, J.; Jarosinski, J. Experimentally Determined Flame on Methane-Air Premixed Flame Extinction at Small Stretch Rates in
Properties near Flammability Limits Under Gravity and Microgravity Microgravity. Symp. Combust. 1996, 26 (1), 1283−1289.
Conditions. Combust. Sci. Technol. 2009, 181 (12), 1431−1442. (265) Wang, S. F.; Zhang, H.; Jarosinski, J.; Gorczakowski, A.;
(245) Liekhus, K. J.; Zlochower, I. A.; Cashdollar, K. L.; Djordjevic, S. Podfilipski, J. Laminar Burning Velocities and Markstein Lengths of
M.; Loehr, C. A. Flammability of Gas Mixtures Containing Volatile Premixed Methane/Air Flames near the Lean Flammability Limit in
Organic Compounds and Hydrogen. J. Loss Prev. Process Ind. 2000, 13 Microgravity. Combust. Flame 2010, 157 (4), 667−675.
(3−5), 377−384. (266) Zeldovich, Y. B. Selected Works of Yakov Borisovich Zeldovich;
(246) Knyazkov, D. A.; Yakimov, S. A.; Korobeinichev, O. P.; Ostriker, J. P., Barenblatt, G. I., Sunyaev, R. A., Eds.; Princeton
Shmakov, A. G. Effect of Trimethylphosphate Additives on the University Press: 1992; 479 pp.
Flammability Concentration Limits of Premixed Methane-Air (267) Ronney, P. D. Near-Limit Flame Structures at Low Lewis
Mixtures. Combust. Explos. Shock Waves 2008, 44 (1), 9−17. Number. Combust. Flame 1990, 82 (1), 1−14.
(247) Burgoyne, J. H.; Williams-Leir, G. The Influence of (268) Ronney, P. D.; Whaling, K. N.; Abbud-Madrid, A.; Gatto, J. L.;
Incombustible Vapours on the Limits of Inflammability of Gases and Pisowicz, V. L. Stationary Premixed Flames in Spherical and Cylindrical
Vapours in Air. Proc. R. Soc. A Math. Phys. Eng. Sci. 1948, 193, 525−539. Geometries. AIAA J. 1994, 32 (3), 569−577.

AS https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

(269) Ronney, P. D.; Wu, M.-S.; Pearlman, H. G.; Weiland, K. J. Compounds in Air at Atmospheric Pressure. Energy 2017, 118, 414−
Experimental Study of Flame Balls in Space: Preliminary Results from 424.
STS-83. AIAA J. 1998, 36 (8), 1361−1368. (293) Britton, L. G.; Frurip, D. J. Further Uses of the Heat of
(270) Kaiser, C.; Liu, J.; Ronney, P. D. Diffusive-Thermal Instability of Oxidation in Chemical Hazard Assessment. Process Saf. Prog. 2003, 22
Counterflow Flames at Low Lewis Number. In 38th Aerospace Sciences (1), 1−19.
Meeting and Exhibit; AIAA: 2000; pp 1−14. DOI: 10.2514/6.2000-576. (294) Britton, L. G. Using Heats of Oxidation to Evaluate
(271) Perry’s Chemical Engineers’ Handbook, 9th ed.; Green, D. W., Flammability Hazards. Process Saf. Prog. 2002, 21 (1), 31−54.
Southard, M. Z., Eds.; McGraw Hill: New York, 2019; 2274 pp. (295) Catoire, L.; Naudet, V. Estimation of Temperature-Dependent
(272) Poling, B. E.; Prausnitz, J. M.; O’Connell, J. P. The Properties of Lower Flammability Limit of Pure Organic Compounds in Air at
Gases and Liquids, 5th ed.; McGraw-Hill: New York, 2001; 803 pp. Atmospheric Pressure. Process Saf. Prog. 2005, 24 (2), 130−137.
DOI: 10.1036/0070116822. (296) Zlochower, I. A. Experimental Flammability Limits and
(273) Montgomery, D. C. Design and Analysis of Experiments, 5th ed.; Associated Theoretical Flame Temperatures as a Tool for Predicting
John Wiley & Sons, Inc.: 2001; 684 pp. the Temperature Dependence of These Limits. J. Loss Prev. Process Ind.
(274) Albahri, T. a. Flammability Characteristics of Pure Hydro- 2012, 25 (3), 555−560.
carbons. Chem. Eng. Sci. 2003, 58 (16), 3629−3641. (297) Liaw, H. J.; Chen, K. Y. A Model for Predicting Temperature
(275) Albahri, T. A. Prediction of the Lower Flammability Limit Effect on Flammability Limits. Fuel 2016, 178, 179−187.
Percent in Air of Pure Compounds from Their Molecular Structures. (298) Wan, X.; Zhang, Q.; Lian, Z. Estimation of the Upper
Fire Saf. J. 2013, 59, 188−201. Flammability Limits of Hydrocarbons in Air at Elevated Temperatures
(276) Bagheri, M.; Rajabi, M.; Mirbagheri, M.; Amin, M. BPSO-MLR and Atmospheric Pressure. Ind. Eng. Chem. Res. 2016, 55 (30), 8472−
and ANFIS Based Modeling of Lower Flammability Limit. J. Loss Prev. 8479.
Process Ind. 2012, 25 (2), 373−382. (299) Cui, G.; Yang, C.; Li, Z.; Zhou, Z.; Li, J. Experimental Study and
(277) Frutiger, J.; Marcarie, C.; Abildskov, J.; Sin, G. Group- Theoretical Calculation of Flammability Limits of Methane/Air
Contribution Based Property Estimation and Uncertainty Analysis for Mixture at Elevated Temperatures and Pressures. J. Loss Prev. Process
Flammability-Related Properties. J. Hazard. Mater. 2016, 318, 783− Ind. 2016, 41, 252−258.
793. (300) Li, P.; Liu, Z.; Li, M.; Huang, P.; Zhao, Y.; Li, X.; Jiang, S.
(278) Gharagheizi, F. Quantitative Structure-Property Relationship Experimental Study on the Flammability Limits of Natural Gas/Air
for Prediction of the Lower Flammability Limit of Pure Compounds. Mixtures at Elevated Pressures and Temperatures. Fuel 2019, 256,
Energy Fuels 2008, 22 (5), 3037−3039. 115950.
(279) Gharagheizi, F. A New Group Contribution-Based Model for (301) Wan, X.; Zhang, Q. Numerical Study of Influence of Initial
Estimation of Lower Flammability Limit of Pure Compounds. J. Pressures and Temperatures on the Lower Flammability Limits of
Oxygenated Fuels in Air. J. Loss Prev. Process Ind. 2016, 41, 40−47.
Hazard. Mater. 2009, 170 (2−3), 595−604.
(302) Wan, X.; Zhang, Q.; Shen, S. L. Theoretical Estimation of the
(280) Kondo, S.; Urano, Y.; Tokuhashi, K.; Takahashi, A.; Tanaka, K.
Lower Flammability Limit of Fuel-Air Mixtures at Elevated Temper-
Prediction of Flammability of Gases by Using F-Number Analysis. J.
atures and Pressures. J. Loss Prev. Process Ind. 2015, 36, 13−19.
Hazard. Mater. 2001, 82 (2), 113−128.
(303) Law, C. K. Combustion Physics; Cambridge University Press:
(281) Lazzús, J. a. Neural Network/Particle Swarm Method to Predict
New York, 2006; 722 pp.
Flammability Limits in Air of Organic Compounds. Thermochim. Acta
(304) Borgnakke, C.; Sonntag, R. E. Fundamentals of Thermodynamics,
2011, 512 (1−2), 150−156.
8th ed.; Wiley: 2013; 894 pp.
(282) Pan, Y.; Jiang, J.; Wang, R.; Cao, H.; Cui, Y. A Novel QSPR
(305) Matalon, M. Flame Dynamics. Proc. Combust. Inst. 2009, 32 (1),
Model for Prediction of Lower Flammability Limits of Organic 57−82.
Compounds Based on Support Vector Machine. J. Hazard. Mater. (306) Fan, A.; Xiang, Y.; Yang, W.; Li, L. Enhancement of Hydrogen
2009, 168 (2−3), 962−969. Combustion Efficiency by Helium Dilution in a Micro-Combustor with
(283) Rowley, J. R.; Rowley, R. L.; Wilding, W. V. Estimation of the Wall Cavities. Chem. Eng. Process. - Process Intensif. 2018, 130, 201−207.
Lower Flammability Limit of Organic Compounds as a Function of (307) Qiao, L.; Gan, Y.; Nishiie, T.; Dahm, W. J. A.; Oran, E. S.
Temperature. J. Hazard. Mater. 2011, 186 (1), 551−557. Extinction of Premixed Methane/Air Flames in Microgravity by
(284) Seaton, W. H. Group Contribution Method for Predicting the Diluents: Effects of Radiation and Lewis Number. Combust. Flame
Lower and the Upper Flammable Limits of Vapors in Air. J. Hazard. 2010, 157 (8), 1446−1455.
Mater. 1991, 27 (2), 169−185. (308) Goodwin, D. G.; Moffat, H. K.; Schoegl, I.; Speth, R. L.; Weber,
(285) Gharagheizi, F. Prediction of Upper Flammability Limit Percent B. W. Cantera: An object-oriented sof tware toolkit for chemical kinetics,
of Pure Compounds from Their Molecular Structures. J. Hazard. Mater. thermodynamics, and transport processes. Zenobo, February 12, 2021.
2009, 167 (1−3), 507−510. DOI: 10.5281/zenodo.4527812.
(286) High, M. S.; Danner, R. P. Prediction of Upper Flammability (309) GRI-Mech Home Page. GRIMECH3.0. http://combustion.
Limit by a Group Contribution Method. Ind. Eng. Chem. Res. 1987, 26 berkeley.edu/gri-mech/index.html (accessed 2021-08-30).
(7), 1395−1399. (310) Guo, H.; Ju, Y.; Maruta, K.; Niioka, T.; Liu, F. Numerical
(287) Pan, Y.; Jiang, J.; Wang, R.; Cao, H.; Cui, Y. Prediction of the Investigation of CH4/CO2/Air and CH4/CO2/O2 Counterflow
Upper Flammability Limits of Organic Compounds from Molecular Premixed Flames with Radiation Reabsorption. Combust. Sci. Technol.
Structures. Ind. Eng. Chem. Res. 2009, 48 (10), 5064−5069. 1998, 135 (1−6), 49−64.
(288) ChemSpider. http://www.chemspider.com/ (accessed 2023-01- (311) Ju, Y.; Masuya, G.; Ronney, P. D. Effects of Radiative Emission
26). and Absorption on the Propagation and Extinction of Premixed Gas
(289) Vidal, M.; Rogers, W. J.; Holste, J. C.; Mannan, M. S. A Review Flames. Symp. Combust. 1998, 27 (2), 2619−2626.
of Estimation Methods for Flash Points and Flammability Limits. (312) Kumar, R. K.; Tamm, H.; Harrison, W. C. Combustion of
Process Saf. Prog. 2004, 23 (1), 47−55. Hydrogen-Steam-Air Mixtures near Lower Flammability Limits.
(290) Takahashi, F.; Glassman, I. Sooting Correlations for Premixed Combust. Sci. Technol. 1983, 33 (1−4), 167−178.
Flames. Combust. Sci. Technol. 1984, 37 (1−2), 1−19. (313) Chen, C. C.; Wang, T. C.; Liaw, H. J.; Chen, H. C. Nitrogen
(291) White, A. G. XCVI.�Limits for the Propagation of Flame in Dilution Effect on the Flammability Limits for Hydrocarbons. J.
Inflammable Gas−Air Mixtures. Part III. The Effects of Temperature Hazard. Mater. 2009, 166 (2−3), 880−890.
on the Limits. J. Chem. Soc., Trans. 1925, 127, 672−684. (314) Chen, C. C.; Liaw, H. J.; Wang, T. C.; Lin, C. Y. Carbon Dioxide
(292) Mendiburu, A. Z.; de Carvalho, J. A.; Coronado, C. R.; Roberts, Dilution Effect on Flammability Limits for Hydrocarbons. J. Hazard.
J. J. Flammability Limits Temperature Dependence of Pure Mater. 2009, 163 (2−3), 795−803.

AT https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels pubs.acs.org/EF Review

(315) Shu, G.; Long, B.; Tian, H.; Wei, H.; Liang, X. Flame (334) Ishaq, H.; Dincer, I.; Crawford, C. A Review on Hydrogen
Temperature Theory-Based Model for Evaluation of the Flammable Production and Utilization: Challenges and Opportunities. Int. J.
Zones of Hydrocarbon-Air-CO2Mixtures. J. Hazard. Mater. 2015, 294, Hydrogen Energy 2022, 47 (62), 26238−26264.
137−144. (335) Modisha, P. M.; Ouma, C. N. M.; Garidzirai, R.; Wasserscheid,
(316) Shu, G.; Long, B.; Tian, H.; Wei, H.; Liang, X. Evaluating Upper P.; Bessarabov, D. The Prospect of Hydrogen Storage Using Liquid
Flammability Limit of Low Hydrocarbon Diluted with an Inert Gas Organic Hydrogen Carriers. Energy Fuels 2019, 33 (4), 2778−2796.
Using Threshold Temperature. Chem. Eng. Sci. 2015, 138, 810−813. (336) Ehlig-Economides, C. A.; Hatzignatiou, D. G. Energy Analysis
(317) Liaw, H. J.; Chen, C. C.; Chang, C. H.; Lin, N. K.; Shu, C. M. of Low Carbon Hydrogen from Methane and End Use Implications.
Model to Estimate the Flammability Limits of Fuel-Air-Diluent Energy Fuels 2022, 36 (16), 8886−8899.
Mixtures Tested in a Constant Pressure Vessel. Ind. Eng. Chem. Res. (337) Zhou, D.; Li, D.; Yuan, S.; Chen, Z. Recent Advances in
2012, 51 (6), 2747−2761. Biomass-Based Photocatalytic H2 Production and Efficient Photo-
(318) Liaw, H. J.; Chen, C. C.; Lin, N. K.; Shu, C. M.; Shen, S. Y. catalysts: A Review. Energy Fuels 2022, 36 (18), 10721−10731.
(338) Kathrotia, T.; Oßwald, P.; Zinsmeister, J.; Methling, T.; Köhler,
Flammability Limits Estimation for Fuel-Air-Diluent Mixtures Tested
M. Combustion Kinetics of Alternative Jet Fuels, Part-III: Fuel
in a Constant Volume Vessel. Process Saf. Environ. Prot. 2016, 100,
Modeling and Surrogate Strategy. Fuel 2021, 302, 120737.
150−162. (339) Mendiburu, A. A. Z.; Ciccarelli, G.; Carvalho, J. A. DDT Limits
(319) Bade Shrestha, S. O.; Wierzba, I.; Karim, G. A. Prediction of the of Ethanol−Air in an Obstacles-Filled Tube. Combust. Sci. Technol.
Extent of Diluents Concentrations in Flammability Limited Gaseous 2018, 1−16.
Fuel-Diluent Mixtures in Air. Appl. Therm. Eng. 2009, 29 (11−12), (340) Ciccarelli, G.; Chaumeix, N.; Mendiburu, A. Z.; N'Guessan, K.;
2574−2578. Comandini, A. Fast- Flame Limit for Hydrogen/Methane-Air Mixtures.
(320) Gutiérrez Velásquez, E. I.; Coronado, C. J. R.; Quintero Proc. Combust. Inst. 2019, 37 (3), 3661−3668.
Cartagena, J. C.; Carvalho, J. A.; Mendiburu, A. Z.; Andrade, J. C.; (341) van Oijen, J. A.; Donini, A.; Bastiaans, R. J. M.; ten Thije
Cortez, E. V.; Santos, J. C. Prediction of Flammability Limits for Boonkkamp, J. H. M.; de Goey, L. P. H. State-of-the-Art in Premixed
Ethanol-Air Blends by the Kriging Regression Model and Response Combustion Modeling Using Flamelet Generated Manifolds. Prog.
Surfaces. Fuel 2017, 210, 410−424. Energy Combust. Sci. 2016, 57, 30−74.
(321) Wang, T.; Chen, C.; Chen, H. Nitrogen and Carbon Dioxide
Dilution Effect on Upper Flammability Limits for Organic Compound
Containing Carbon, Hydrogen and Oxygen Atoms. J. Taiwan Inst.
Chem. Eng. 2010, 41 (4), 453−464.
(322) Molnarne, M.; Mizsey, P.; Schröder, V. Flammability of Gas
Mixtures. Part 2: Influence of Inert Gases. J. Hazard. Mater. 2005, 121
(A1), 45−49.
(323) Mendiburu, A. Z.; de Carvalho, J. A.; Coronado, C. R. Method
for Determination of Flammability Limits of Gaseous Compounds
Diluted with N2 and CO2 in Air. Fuel 2018, 226, 65−80.
(324) Shebeko, Y.; Fan, W.; Bolodian, I.; Navzenya, V. An Analytical
Evaluation of Flammability Limits of Gaseous Mixtures of Combus-
tible−Oxidizer−Diluent. Fire Saf. J. 2002, 37 (6), 549−568.
(325) Razus, D.; Molnarne, M.; Fuß, O. Limiting Oxygen
Concentration Evaluation in Flammable Gaseous Mixtures by Means
of Calculated Adiabatic Flame Temperatures. Chem. Eng. Process.
Process Intensif. 2004, 43 (6), 775−784.
(326) Razus, D.; Molnarne, M.; Movileanu, C.; Irimia, A. Estimation
of LOC (Limiting Oxygen Concentration) of Fuel-Air-Inert Mixtures
at Elevated Temperatures by Means of Adiabatic Flame Temperatures.
Chem. Eng. Process. Process Intensif. 2006, 45 (3), 193−197.
(327) Hansen, T. J. Estimation of the Flammability Zone Boundaries
with Thermodynamic and Empirical Equations. M.S. Thesis, Michigan
Technological University, 2009; 139 pp.
(328) Hansen, T. J.; Crowl, D. A. Estimation of the Flammability Zone
Boundaries for Flammable Gases. Process Saf. Prog. 2010, 29 (3), 209−
215.
(329) Mashuga, C. V.; Crowl, D. A. Derivation of Le Chatelier’s
Mixing Rule for Flammable Limits. Process Saf. Prog. 2000, 19 (2),
112−117.
(330) Carvalho, J. A.; Mendiburu, A. Z.; Coronado, C. R.; McQuay,
M. Q. Combustão Aplicada; Editora da UFSC: Florianópolis, 2018; 372
pp.
(331) Chen, C. C.; Liu, S. H.; Kang, X. Evaluating Lower Flammability
Limit of Flammable Mixtures Using Threshold Temperature Approach.
Chem. Eng. Sci. 2018, 185, 84−91.
(332) Liu, D.; Santner, J.; Togbé, C.; Felsmann, D.; Koppmann, J.;
Lackner, A.; Yang, X.; Shen, X.; Ju, Y.; Kohse-höinghaus, K. Flame
Structure and Kinetic Studies of Carbon Dioxide-Diluted Dimethyl
Ether Flames at Reduced and Elevated Pressures. Combust. Flame 2013,
160 (12), 2654−2668.
(333) Compton, R. G.; Hancock, G. Low-Temperature Combustion and
Autoignition; Elsevier: New York, NY, 1997; 794 pp.

AU https://doi.org/10.1021/acs.energyfuels.2c03598
Energy Fuels XXXX, XXX, XXX−XXX

You might also like