Vous êtes sur la page 1sur 23

Journal of Hazardous Materials 241–242 (2012) 32–54

Contents lists available at SciVerse ScienceDirect

Journal of Hazardous Materials


journal homepage: www.elsevier.com/locate/jhazmat

Review

Flammability limits: A review with emphasis on ethanol for aeronautical


applications and description of the experimental procedure
Christian J.R. Coronado a , João A. Carvalho Jr. b,∗ , José C. Andrade c , Ely V. Cortez c ,
Felipe S. Carvalho a , José C. Santos c , Andrés Z. Mendiburu b
a
Federal University of Itajubá – UNIFEI, Mechanical Engineering Institute – IEM Av BPS 1303, Itajubá, MG CEP 37500903, Brazil
b
São Paulo State University – UNESP, Campus of Guaratinguetá – FEG Av. Ariberto P. da Cunha 333, Guaratinguetá, SP CEP 12510410, Brazil
c
National Space Research Institute – INPE, Combustion and Propulsion Laboratory – LCP Rod. Pres. Dutra, km 39, Cachoeira Paulista, SP CEP 12630-000, Brazil

h i g h l i g h t s

 Develops a comprehensive literature review on ethanol flammability limits.


 Difference in standard procedures lead to different experimental values of the flammability limits.
 Methodology for experiments to find the FL’s of ethanol for aeronautical applications.

a r t i c l e i n f o a b s t r a c t

Article history: The lower and upper flammability limits of a fuel are key tools for predicting fire, assessing the possibility
Received 22 May 2012 of explosion, and designing protection systems. Knowledge about the risks involved with the explo-
Received in revised form 23 August 2012 sion of both gaseous and vaporized liquid fuel mixtures with air is very important to guarantee safety
Accepted 16 September 2012
in industrial, domestic, and aeronautical applications. Currently, most countries use various standard
Available online 24 September 2012
experimental tests, which lead to different experimental values for these limits. A comprehensive liter-
ature review of the flammability limits of combustible mixtures is developed here in order to organize
Keywords:
the theoretical and practical knowledge of the subject. The main focus of this paper is the review of
Flammability limits
Ethanol
the flammability data of ethanol–air mixtures available in the literature. In addition, the description of
Visual criterion methodology for experiments to find the upper and lower limits of flammability of ethanol for aero-
Pressure and temperature dependence nautical applications is discussed. A heated spherical 20 L vessel was used. The mixtures were ignited
with electrode rods placed in the center of the vessel, and the spark gap was 6.4 mm. LFL and the UFL
were determined for ethanol (hydrated ethanol 96% ◦ INPM) as functions of temperature for atmospheric
pressure to compare results with data published in the scientific literature.
© 2012 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
1.1. Objectives and scope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
1.2. Flammability limits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
1.3. State of the art . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
1.3.1. Theoretical methods to determine flammability limits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
2. Standard methodology for flammability determination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
2.1. Visual criterion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
2.2. Pressure criterion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
2.3. Brief description of flammability test methods discussed in this paper. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3. Influences of temperature, pressure, turbulence and ignition energy on flammability tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.1. Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.2. Pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

∗ Corresponding author at: São Paulo State University – UNESP, Campus of Guaratinguetá – FEG, Av. Ariberto P. da Cunha 333, Guaratingueta, SP 12516-410, Brazil.
Tel.: +55 12 31232838.
E-mail addresses: joao@feg.unesp.br, joao.a.carvalho.jr@pq.cnpq.br (J.A. Carvalho Jr.).

0304-3894/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.jhazmat.2012.09.035
C.J.R. Coronado et al. / Journal of Hazardous Materials 241–242 (2012) 32–54 33

3.3. Turbulence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.4. Comparison between the limits in air and in oxygen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.5. Ignition energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.6. Flammability limits of fuel mixtures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
4. Flammability limits in the aeronautical industry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
4.1. Importance for the aeronautical industry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
4.2. Ignition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
4.3. Flammability properties of aviation fuels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
5. Tests for flammability limits of ethanol: review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
6. Experimental tests with ethanol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
6.1. Flammability apparatus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
6.2. Description of experimental procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
6.3. Mass of ethanol needed to form flammable mixtures with air . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
6.4. Description of the procedure to calculate LFL and UFL from experimental data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
6.5. Results and discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
6.5.1. Behavior of flame propagation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
6.5.2. Results for atmospheric pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
7. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

1. Introduction the following considerations: (a) visual inspection of the flame


produced by a spark (spark ignition) in an oxidant-combustible
1.1. Objectives and scope mixture inside a transparent and closed vessel (visual criteria);
(b) pressure or temperature measurement at the moment of igni-
A comprehensive review of the literature about the flammability tion inside a closed vessel (pressure criteria). In the 1950s and the
limits of combustible mixtures is developed here in order to orga- 1960s, the tests were conducted using glass tubes with diame-
nize theoretical and practical knowledge on the subject. The main ters of up to 60 mm and heights up to 300 mm, using the visual
focus of this paper is a review of flammability data of ethanol–air criteria to verify flammability of certain mixtures. Today, the visual
mixtures available in the literature. In addition, the description of criterion is still in use, as is the pressure criterion. Vessels used
an experimental method for determination of the upper and lower for the tests are generally spherical and/or cubic and of different
limits of flammability of ethanol for aeronautical applications is dis- sizes.
cussed. It also shows a diagram of the experimental prototype built There are standards for setting up experimental procedures to
to obtain the flammability limits of ethanol and the main results at determine whether a mixture is flammable or not. In all these
atmospheric pressure. standards, the variables are the same: the size of the combustion
chamber (test vessel), temperature, operation pressure, electrode
discharge time (spark ignition or fuse wire ignition), and ignition
1.2. Flammability limits
energy of the electrodes. Table 1 shows the flammability limits of
some fuels that were selected as part of the scope of this work.
Flammability limits are the main properties that represent
Among the fuels chosen are those used in the aircraft industry, such
flammability characteristics of specific fuels. These limits are the
as gasoline for commercial and military aviation and kerosene, as
borders that separate the oxidant–combustible mixture regions
well as some vehicle fuels such as gasoline, diesel, natural gas, etc.
in which flame propagation occurs and does not occur. There
It also shows the main alcohols and other alternative fuels of inter-
are two kinds of flammability limits for an oxidant–combustible
est. Most of the data shown on this table were obtained from Refs.
mixture: the minimum concentration of fuel for which flame prop-
[1–8], except those marked with letters. For readers interested in a
agation is possible (lean mixture), known as the lower flammability
particular fuel, Appendix A of the paper published by Zabetakis [2]
limit (LFL), and the maximum concentration of fuel for which
and Table 44 on page 130 of the paper published by Cowards and
flame propagation is possible (rich mixture), known as the upper
Jones [1] are recommended.
flammability limit (UFL). There is another parameter known as LOC
(limiting oxidant concentration) that is widely used along with
the limits of flammability. LOC is the concentration of oxidant in 1.3. State of the art
a fuel–oxidant–diluent mixture below which deflagration cannot
occur under specified conditions. This section will present the latest scientific studies of flamma-
Flammability limits have been thoroughly discussed in scientific bility limits. The following sections emphasize flammability limit
literature. Probably the first works on this topic were those devel- s of ethanol primarily for the aircraft industry.
oped by Coward and Jones [1] and by Zabetakis [2], both for the Naegeli and Weatherford [13] studied the flammability haz-
Mines Department of the U.S. Government. Afterwards, Kuchta [3] ard of storing pure alcohol and diesel fuel/alcohol blends in fuel
expanded upon the data obtained by Zabetakis. Nestor [4] and the tanks. The authors measured the ignition limit for the C1 through
Fuel Flammability Task Group [5] worked specifically with flamma- C4 alcohols and alkanes, ethylene, isooctane and methylal in a
bility limits for the aeronautical industry. Measurements of ignition combustion pump using an automotive-type spark plug as an igni-
energy for aeronautic fuels (Spark Ignition Energy Measurements tion source. However, the UFL of the alcohols were very different
in JET A) were reported by Shepherd et al. [6]. In addition, the from published flammability limit data, though the upper limit for
works published by Ott [7] and Kosvic et al. [8] contribute for the methanol is in the area of reported values. According these authors,
knowledge in the area. methanol’s LFL is 7.9 mol% and UFL is 26 mol%. On the other hand,
Different methods are used to evaluate flammability limits. The for ethanol LFL is 4.4 mol% and UFL is 14.3 mol%. All these data were
success of an attempt can be determined by the combination of for 0.97 atm and at a temperature of 364 K [13].
34 C.J.R. Coronado et al. / Journal of Hazardous Materials 241–242 (2012) 32–54

Table 1
Flammability limits for some fuels in air, at atmospheric pressure and 298 K, vol% [1–8].

Chemical name Chemical formula LFL UFL

1 Acetylene CH:CH 2.5 100.0


2 Ammonia NH3 15 28
3 Benzol (Benzene) C6 H6 1.3 7.9
4 Blast Furnace Gas – 35 74
5 Butane CH3 CH2 CH2 CH3 1.6 8.4
6 Butyl Alcohol CH3 (CH2 )2 CH2 OH 1.4 11.2
7 Carbon Monoxide CO 12.5 74
8 Coal Gas – 5.3 32
9 Coke-oven Gas – 4.4 34
10 Cyclepropane (CH2 )3 2.4 10.4
11 Cyclohexane C6 H12 1.3 8.0
12 Decane CH3 (CH2 )8 CH3 0.8 5.4
13 Ethane CH3 CH3 3.0 12.5
14 Ethyl Alcohol C2 H5 OH 3.3 19.0
15 Ethylene H2 C:CH2 2.7 36.0
16 Ethylene Glycol C2 H6 O2 3.2 21.6
17 Diesel (gas oil)a 0.5 5.0
18 Gasoline, Premium Automotive – 1.3–1.4 6.0–7.6
19 Gasoline, Regular Automotive – 1.3–1.4 6.0–7.6
20 Gasoline, Commercial Aviation – 1.0 6.0–7.6
21 Gasoline, Military Aviation 1.0 6.0–7.6
22 Heptane CH3 (CH2 )5 CH3 1.1 6.7
23 Hexane CH3 (CH2 )4 CH3 1.1 7.5
24 Hydrogenb H2 4.0 74.2
25 LH2 a – 4.0 75.0
26 Natural Gas – 3.8–6.5 13–17
27 Natural Gasc 5 15.6
28 LNG and CNGa – 5 15
29 Kerosene – 0.7 5
30 Methane CH4 5.0 15.0
31 Methyl Alcohol CH3 OH 6.0 36.0
32 Methyl Ether (CH3 )2 O 3.4 27.0
33 Naphtha – 0.8 5
34 Nonane C9 H20 0.8 2.9
35 Octane CH3 (CH2 )6 CH3 1.0 6.5
36 p-Dioxane OCH2 CH2 OCH2 CH2 2.0 22.0
37 Pentane CH3 (CH2 )3 CH3 1.5 7.8
38 Propanal CH3 CH2 CHO 2.6 17.0
39 Propane (LPG) CH3 CH2 CH3 2.1 9.5
40 Propylene CH2 :CHCH3 2.0 11.0
41 Propyl alcohol CH3 CH2 CH2 OH 2.2 13.7
42 Syngas (wood gas)d – 16 –
43 Toluene C6 H5 CH3 1.2 7.1
a
Ref. [9].
b
Ref. [10].
c
Ref. [11]. 96.16% CH4 by vol., 2.54% CO2 , 1.096% C2 H6 , approximately 0.189% hydrocarbons higher than C3, and the remainder including nitrogen, hydrogen sulfide, and
water is approximately 0.015%.
d
Ref. [12].

Shebeko et al. [14] proposed a new analytical method According to Brandes et al. [15], it is well known that explo-
for the calculation of flammability limits in mixtures of sive mixtures can exist at temperatures below the flash point (FP).
combustible–oxidizer–diluent. They revealed new regularities for Experiments show that the difference between flash point and LFL
LFL and the compositions of mixtures near the peak point of the may be up to 15 K and in some special cases even more. There
flammability curves. are some liquids without a flash point (pure substances as well
Van den Schoor and Verplaetsen [9] studied the UFL as mixtures) that are able to form an explosive vapor–air mixture.
of ethane–air, propane–air, n-butane–air, ethylene–air, and According this author, knowledge about LFL lets work take place at
propylene–air mixtures. The UFL were determined experimen- higher temperatures than using flash point and including a safety
tally at initial pressures up to 30 bar (this paper does not specify margin of for example 12 K or more. A good example for this case
the study at reduced pressures and negatives temperatures) and is cleaning processes using flammable liquids.
temperatures up to 250 ◦ C. The experiments were performed in Van den Schoor et al. [16] reported four different numerical
a closed spherical vessel with an internal diameter of 200 mm. methods to calculate the UFL of methane–air mixtures at initial
The mixtures were ignited by using electric current to fuse a pressures up to 10 bar and initial temperatures up to 200 ◦ C by
coiled tungsten wire that was placed at the center of the vessel. comparison with experimental data. According this author: (i)
They concluded that UFL increases linearly with initial temper- at atmospheric pressure, calculation of UFL by calculating planar
ature; however, the slope of the straight line is not a constant flames with the inclusion of radiation heat loss is satisfactory;
but depends on the initial pressure. Finally, a comparison of data (ii) at elevated pressures, calculated UFL values are significantly
for the alkanes showed that the effect of preferential diffusion high and large differences are found between the different reac-
plays an important role in near-upper flammability limit combus- tion mechanisms. Van den Schoor et al. [17] reported two different
tion. experimental methods to determine flammability limits which
C.J.R. Coronado et al. / Journal of Hazardous Materials 241–242 (2012) 32–54 35

were compared, evaluated, and exemplified by determining the A general rule for estimating flammability limits is to consider
flammability limits of methane–hydrogen–air mixtures for hydro- the upper flammability limit equal to three times the stoichio-
gen fuel molar fractions of 0, 0.2, 0.4, and 0.6 at atmospheric metric value and the lower flammability limit equal to 50% of the
pressure and room temperature. stoichiometric value [26]. However, as pointed out in the same ref-
Chen at al. [18] predicted the upper/lower flammability lim- erence, there are many exceptions to the rule, as can be seen in
its of hydrocarbons diluted with inert nitrogen gas. These authors Table 1.
reported that there are linear relations between the reciprocal of Some theoretical methods in the literature are discussed in the
the upper/lower flammability limits and the reciprocal of the molar following sections. Ethanol flammability limits will be determined
fraction of hydrocarbon in the hydrocarbon/inert nitrogen mixture. for example and in order to contrast the theoretical and experi-
Also, Gharagheizi [19] studied a quantitative structure–property mental data. According to experimental data, ethanol’s LFL is 3.3%
relationship to predict the UFL of pure compounds. The obtained and UFL is 19% (vol%).
model is a five-parameter multilinear equation.
Koshiba et al. [20] characterized the explosion properties of mix- 1.3.1.1. Stoichiometric concentration. In order to determine
tures of n-pentane, diethyl ether, diethylamine, or n-butyraldehyde flammability limits using stoichiometric concentration, the
with nitrous oxide and nitrogen using three parameters: explosion following formulas are used:
limit, peak explosion pressure, and time to peak explosion pres-
LFL = 0.55 Cs , [2, 27] (1)
sure. The explosion experiments were performed in a cylindrical
vessel at atmospheric pressure and room temperature. According UFL = 3 Cs , [26] (2)
this author, measurements showed that explosion ranges of the
mixtures containing nitrous oxide were narrower than those of the UFL = 3.5 Cs , [27, 28] (3)
mixtures containing oxygen. in which Cs is stoichiometric concentration.
Shoshin and de Goey [21] reported an experimental study of the For example, ethanol’s stoichiometric reactions are: 1
LFL of methane–hydrogen–air mixtures in tubes of different diame- C2 H5 OH + 3 O2 + 11.28 N2 → 2 CO2 + 3 H2 O + 11.28 N2 . For this
ters (6.0–50.2 mm). The flames propagated upward from the open reaction, Cs = 6.54%. Using Eqs. (1) and (2), LFL = 3.59% and
bottom end of the tube to the closed upper end. According this UFL = 19.62%.
author, LFL value decreased with tube diameter for methane–air
and (90% CH4 + 10% H2 )–air mixtures for tubes. This effect has been 1.3.1.2. Combustion enthalpy. A rough estimation of LFL in air may
attributed to the stronger combined effect of preferential diffusion be obtained using the following rule of thumb from Tareq [27]. The
and flame stretch in narrower tubes for flames which resemble ris- equation to determine the lower flammability limit is:
ing bubbles. Also, Miao et al. [22] reported both LFL and UFL of
hydrogen-enriched natural gas with hydrogen concentrations of 4354
LFL = − , [27] (4)
20%, 40%, 60% and 80% as well as these of natural gas and hydro- Hcomb
gen. The experimental results showed that the flammability data of in which Hcomb is combustion enthalpy (kJ/mol) and LFL is in
methane–hydrogen mixtures are applicable to hydrogen-enriched volume percent.
natural gas mixtures. For example, ethanol has a combustion enthalpy of
Piqueras et al. [23] reported a method for the prediction of LFL at −1281 kJ/mol [2]. Applying Eq. (4), LFL is 3.39%.
high pressures for the mixtures of H2 + O2 in CO2 and N2 , between The two aforementioned methods are generally applied to
1.0 and 300 bar and 288–348 K. According to this author, the use hydrocarbons and paraffins.
of CO2 as a diluent increases the operational margin from 4.5 mol%
H2 at 1 bar up to ca. 7.0–9.0 mol% H2 at 200 bar due to the increase 1.3.1.3. Vapor pressure. The lower flammability limits can be deter-
in heat capacity. On the other hand, the use of nitrogen or air as a mined using the fuel vapor pressure and standard atmospheric
diluent only increases the margin from 5.2 mol% H2 at 1 bar up to pressure as follows:
6.0 mol% H2 at 200 bar.
pf
Rowley et al. [24] reported an estimation of the LFL of organic LFL = × 100, [23] (5)
compounds as a function of temperature. This author showed that p0
the LFL of general organic compounds may be accurately esti- in which pf is the fuel vapor pressure at flash point and p0 is standard
mated from structural contributions and the heat of formation atmospheric pressure.
of the fuel. Prediction of LFL and its dependence on temperature The flash point of ethanol is 13 ◦ C [29]. At this temperature,
closely agrees with experimental data obtained in an ASHRAE- it has a vapor pressure of 28 mmHg [10]. With p0 = 760 mmHg,
style apparatus.Kondo et al. [25] reported the flammability limits LFL = 3.68%.
of CH2 CFCF3 (HFO-1234 yf), CH2 F2 (HFC-32), and methane mea-
sured at pressures from atmospheric pressure to 2500 kPa in a 5 L 1.3.1.4. Algebraic method. The following equation allows LFL to
spherical stainless-steel vessel. In general, the changes to lower be determined from the data of the minimum number of moles
flammability limits are small compared to upper flammability lim- required for a flame propagation reaction:
its.
100
Below are some simple methods to calculate the flammability LFL = , [29] (6)
limits from stoichiometric concentration, enthalpy of combustion, 1 + na
and/or the vapor pressure of the fuel. in which na is the number of moles of air per mole of fuel in the
mixture at LFL.
For example, for ethanol, the reactants are 1 C2 H5 OH + 3
1.3.1. Theoretical methods to determine flammability limits O2 + 11.28 N2 . LFL obtained is 3.29%.
The scientific literature provides different theoretical methods The combustion reaction can be described using LFL as fol-
to determine flammability limits. In all these methods, the theoret- lows: LFL (Cn Hm Ol ) + (100 − LFL) (0.21 O2 + 0.79 N2 ) → ˛1 CO2 + ˛2
ical values for lower flammability limits fit the experimental data H2 O(v) + ˛3 O2 + ˛4 N2 , in which ˛1 = LFL × n, ˛2 = LFL × m/2,
well; on the other hand, the theoretical values for upper flamma- ˛3 = 21 − LFL (0.21 + n + m/2 − 1/2), and ˛4 = 79 − LFL × 0.79.
bility limits have some differences from experimental data. Therefore, for ethanol, the combustion reaction will be: 3.299
36 C.J.R. Coronado et al. / Journal of Hazardous Materials 241–242 (2012) 32–54

Table 2
Main characteristics of international standards for measurement of flammability limits [32].

DIN 51649-1 EN 1839 (T) EN 1839 (B) ASTM E 681-01

Ignition vessel Cylindrical glass vessel Cylindrical glass vessel Spherical or cylindrical closed Glass vessel V = 5 L
∅ = 60 mm, H = 300 mm ∅ = 80 mm, H = 300 mm vessel V > 5 L
Ignition source High tension spark, duration High tension spark; duration Fusion wire (explosion), High tension spark, duration
0.5 s; power 10 W 0.2 s; power 10 W E = 10–20 J high tension spark; 0.5 s; power 10 W
duration 0.2 s
Criteria Visual, flame detachment Visual, flame detachment and Pressure ≥ 5% of test initial Visual, flame propagation up to
propagation >100 mm, or absolute pressure. 13 mm of the wall (horizontal
without flame detachment (H) or vertical)
of H > 240 mm
Repetitions 5 4 4 1
Sample size 0.1 mol% (FL ≤ 10 mol%), 10% rel. (FL ≤ 2 mol%), 0.2 mol% 10% rel. (FL ≤ 2 mol%), 0.2 mol% Arbitrary
0.2 mol% (EL > 10 mol%) (FL >2 mol%) (FL > 2 mol%)
Flammability limit Last point without ignition Last point without ignition Last point without ignition Average values between the
first ignition point and the last
point without ignition

FL, flammability limits.

C2 H5 OH + 20.3 O2 + 76.39 N2 → 6.59 CO2 + 9.89 H2 O(v) + 10.4 2.1. Visual criterion
O2 + 76.39 N2 . This method is generally applied to organic
substances with the general formula Cn Hm Ol [30]. Test vessels must be transparent; they are incapable of resisting
a total deflagration charge. A high-speed video camera is recom-
2. Standard methodology for flammability determination mended.
The German standard utilizes vertical cylindrical tubes with
Methods of measuring flammability limits for gases have been smaller volumes than those established by the American standard.
known well for a long time, and there have been many attempts For example, DIN standard 51649 requires a cylinder diameter of
to standardize them. However, no standard method has been esti- 60 mm, with a volume of almost 0.85 L, while European standard
mated yet [31]. Modern flammability test methods to determine EN 1839 (T) requires a cylinder diameter of 80 mm, with a volume
flammability limits and oxygen concentration (LOC) are summa- of 1.5 L, and finally, American standard ASTM E681 [28] requires
rized in Fig. 1. Tables 2 and 3 show a comparison between standard spherical vessels with volumes of 5 L or 12 L, with internal diam-
test methods for flammability limits at atmospheric pressure and eters of 222 mm and 295 mm, respectively. European standard EN
298 K. 1839 offers two different visual criteria:

Table 3
Comparison of flammability data obtained using modern standards [32–35].

Substance LFL and UFL (vol%) isobar conditions LFL and UFL (vol%) isochor conditions

EN 1839 T ASTM E681 EN 1839B ASTM WK5917 and E2079

Ethylene
LFL (20 ◦ C) 2.5 ± 0.10 2.2 2.5 ± 0.10
UFL (20 ◦ C) 32.3 ± 0.40 33.3 30.8 ± 0.70
LOC (20 ◦ C) 7.7 7.5 8.1 8.5 (120 L)

Methane
LFL (20 ◦ C) 4.4 ± 0.20 3.8 4.6 ± 0.20 5.0 ± 0.1 [26] (8, 20 and 120 L test vessels)
UFL (20 ◦ C) 16.6 ± 0.20 16.9 16.8 ± 0.20 15.9 ± 0.10 (20 L) [26]15.8 ± 0.10 (120 L) [26]
LOC (20 ◦ C) 10.2 9.9 11.3 11.2 (120 L)

n-Hexane
LFL (20 ◦ C) 1.0 ± 0.10 1.0 ± 0.10
UFL (40 ◦ C) 8.3 ± 0.10 7.7 ± 0.25

1,3,5-Trimethyl-benzene
LFL (70 ◦ C) 0.8 ± 0.08 0.8 ± 0.08
UFL (130 ◦ C) 7.1 ± 0.20 6.6 ± 0.10

Hydrogen [24,27]
LFL (20 ◦ C) 3.6 3.75 4.2
UFL (20 ◦ C) 76.6 75.1 77.0
LOC (20 ◦ C) 4.4 4.1 4.7

Ammonia [24,27]
LFL (20 ◦ C) 14.3 13.3a 14.2
UFL (20 ◦ C) 32.5 32.9a 39.4

Schroder and Molnarne [32].


Ural and Zalosh [34].
Fuss et al. [35].
a
Performed in a 5 L flask instead of a 12 L flask.
C.J.R. Coronado et al. / Journal of Hazardous Materials 241–242 (2012) 32–54 37

Fig. 1. Flammability test methods [22].

• The first visual criterion is “flame detachment”, which requires the fact that results obtained in larger vessels are more reliable;
an upward movement of the detached flame from the ignition therefore, they do not allow tests in small vessels. European
spark (around 10 cm apart). standard EN 1839 does not allow tests in vessels smaller than
• The second visual criterion is that the flame reach the top of the 5 L, while the American standard permits tests in vessels down to
cylinder with a minimum distance of 24 cm. 4 L.
A mixture is considered flammable if both conditions are met. European standard EN 1839B permits two different kinds of
American standard ASTM E681 [28] gives two different visual igniters:
criteria as well:
• For the 5 L vessel, the flame front must move upward and outward • Induction spark: A series of sparks are inducted by electrodes in
from the ignition source to the vessel walls or at least 13 mm up alternate current using a high voltage transformer. The power of
the wall. the spark depends on the mixture for each test and also on the
• For the 12 L vessel, flame propagation is defined as moving details of the air gap and the circuit configuration. This standard
upward and outward from the ignition source to the vessel walls. suggests a power of 10 W. Therefore, the total energy emitted by
It forms an arc larger than 90◦ measured from the ignition point the spark is limited to 2 J per 0.2 s of average discharge time.
to the vessel walls. • Fuse wire: This kind of igniter emits more energy. The energy
fed to this wire must be between 10 and 20 J. Only a part of this
Schroder and Molnarne [32] reported comparative measure- electric energy is available for ignition in the gas phase because
ments (Table 3) using European standard EN 1839T (vol = 1.5 L) and most of the energy is used to heat and to vaporize the fuse wire.
American standard ASTM E681 (vol = 5 L). The mixture was pre-
pared outside the test vessels for both cases; the air–fuel mixture The American standard permits a continuous electric arc and the
was the same for both experiments. The ignition source was also same fuse wire as those required by the European standard. It also
the same, 10 W for 0.2 s. allows coal sparks, discrete electric sparks (almost 10 J of energy
storage in the capacitor bank), and chemical igniters.
2.2. Pressure criterion The European standard uses the criterion of 5% above the ini-
tial fixed pressure in the vessel to determine whether a fuel is
The methods that are used most often are European standard EN flammable or not. The American standard uses a 7% criterion. How-
1839B and American standard ASTM E2019. They both recognize ever, according to Brandes and Ural [33], the following remark must
38 C.J.R. Coronado et al. / Journal of Hazardous Materials 241–242 (2012) 32–54

be considered: the criterion of 7% may not be appropriate for certain


mixtures. This is the case when vessel volume is small or ignition
energy is substantially larger than 10 J. It is a good idea to perform
an exploratory test in the vicinity of limit mixtures to evaluate the
validity of the selected pressure rise criterion. In fact, users are free
to choose any pressure rise criterion as long as it is backed up by
appropriate baseline tests [33].
The authors state that when collecting data from flammabil-
ity tests, researchers must keep in mind that tests carried out in
small volumes or with high ignition energies tend to result in
wider flammability intervals. Considering the fact that tests car-
ried out with small ignition energies or small vessels may not have
noticeable flammability with certain mixtures, it is important to use
sufficient ignition energy and vessel volumes that are large enough
for ignition.
Table 4 shows a comparison of flammability limits obtained
experimentally using different standard methods (criteria) for the
most important combustible gases.
To sum up, tests carried out in small volumes with high ignition Fig. 2. Temperature effect on flammability limits [1].
energies result in wider intervals of flammability and thus more
conservative results at the expense of higher safety costs. Consid-
is used to establish flammability. Draft WK5917 employs the same
ering the fact that tests carried out with small ignition energies
concept as E2079 and is available for the LFL and UFL determina-
or small vessel volumes might not have a noticeable flammability
tions [33,38]. EN1839B. The ignition procedure is identical to the
for certain combustibles, it is important to use sufficient ignition
previous method. It uses a tube apparatus with an inner closed
energy and large enough volumes. A common practice is to use
spherical vessel with a volume of 14 dm3 . The mixture is prepared
high-speed video cameras in order to analyze the film from differ-
with partial pressure and the ignition source is a fusing (explod-
ent points of view in the same laboratory. According to Takahashi
ing) wire. The criterion for a successful ignition is a pressure rise
et al. [36], if the vessel is large enough, the experimental values
of at least 5% of the starting absolute pressure. The criterion for
of flammability limits may approach the values that would be
flammability is a 5% minimum pressure rise after ignition [31].
obtained in open space.

2.3. Brief description of flammability test methods discussed in 3. Influences of temperature, pressure, turbulence and
this paper ignition energy on flammability tests

US bureau of mines flammability tubes. This method used a visual 3.1. Temperature
criterion. The test vessel is made of glass (50 mm diameter and
1000 mm height, minimum) with a pressure release at the bot- Heat released by a reaction mixture and heat transfer to the
tom. The criterion for successful ignition (go) is when the flame surroundings are both important in determining the course of the
travels up from the tips of the electrodes to at lest half way combustion process. According to Jones [39], it is not possible to
through the tube [1,2,33]. EN1839T. This method used a visual cri- assign an ignition temperature for a particular mixture, since igni-
terion. It uses an 8 cm wide, 30 cm tall, open top glass cylinder tion depends not only on the reagents but also on heat exchanged
with a spark igniter at the bottom (0.2 s and 10 W). The crite- with the surroundings. However, this parameter is found in the
rion for flammability is propagation of the flame 10 cm vertically combustion literature [40].
above the igniter or 12 cm horizontally at any point on the flame’s A low ignition temperature means that the mixture is poten-
path. [31]. ASTM E681. This test method covers the determina- tially hazardous. The higher the activation energy of the reaction,
tion of the lower and upper concentration flammability limits for the higher the ignition temperature. For mixtures with a high
chemicals with sufficient vapor pressure to form flammable mix- activation energy, a spark is needed to initiate combustion. After
tures in air at atmospheric pressure at the test temperature. This ignition by spark, the heat generated is usually sufficient for the
test method is based on electrical ignition and visual observations reaction to be self sustainable [40].
of flame propagation [28]. ASTM E918. This method is designed If the temperature increases, the lower flammability lim-
to determine flammability limits at elevated temperatures (up its should decrease, while the upper flammability limits should
to 200 ◦ C) and initial pressure of up to 200 psia (1.38 MPa). This increase. In other words, flammability interval widths increase
practice is limited to mixtures that have explosion pressures less when initial mixture temperature increases. This can be seen in
than 2000 psia (13.79 MPa). The pressure resistant metal test ves- Fig. 2.
sel is a vertical cylinder, only 1 L in size, which is much smaller As has been discussed, LFL values decrease when temperature
than that used in E681. Flame propagation is defined as a com- increases; consequently, less combustion energy will be necessary
bustion reaction that produces at least a 7% rise of the initial to ignite the flame. Drysdale [41] stated that LFLT and UFLT for any
absolute pressure [33,37]. ASTM E2079 and draft standard WK5917. temperature T at 1 atm pressure can be calculated based on the
These test methods cover the determination of the limiting oxy- limits determined for 298 K using the following equation:
gen (oxidant) concentration of mixtures of oxygen (oxidant) and
LFLT = LFL298 [1 − 0.00078(T − 298)], [2] (7)
inert gases with flammable gases and vapors at a specified initial
pressure and initial temperature. They can also be used to deter- UFLT = UFL298 [1 + 0.000721(T − 298)]. [41] (8)
mine the limiting concentration of oxidizers other than oxygen.
This method has been designed to address problems associated The flammability limits of ethanol at 298 K and 1 atm are 3.3% for
with the E681 and E918 methods. The test vessel required is LFL and 19% for UFL [1,2], both in volume concentration, so Fig. 3
larger than 4 L and nearly spherical. A 7% pressure rise criterion can be made from Eqs. (7) and (8) for different temperatures. Later
C.J.R. Coronado et al. / Journal of Hazardous Materials 241–242 (2012) 32–54 39

Table 4
Flammability limits of hydrogen, ethylene, methane and ammonia in air, T = 20 ◦ C [32].

DIN 51649-1 (mol%) EN 1839 (T) (mol%) EN 1839 (B) (mol%) ASTM E681-01 (mol%)

LFL (H2 –air) 3.8 3.6 4.2 3.75


LFL (H2 –air) 75.8 76.6 77.0 75.1
LFL (H2 –40% N2 –air) 3.6 3.6 4.4 3.65
UFL (H2 –40% N2 –air) 38.2 38.4 38.2 37.3

LFL (C2 H4 –air) 2.3 2.4 2.6 2.15


UFL (C2 H4 –air) 33.0 32.6 27.4 33.3
LFL (C2 H4 –40% N2 –air) 2.4 2.4 2.6 2.85
UFL (C2 H4 –40% N2 –air) 8.4 8.2 6.9 8.05

LFL (CH4 –air) 4.2 4.3 4.9 3.8


UFL (CH4 –air) 16.6 16.8 16.9 16.9
LFL (CH4 –40% N2 –air) 4.3 4.5 5.1 4.15
UFL (CH4 –40% N2 –air) 6.5 6.4 5.7 6.35

LFL (NH3 –air) 14.3 14.3 14.2 13.3


UFL (NH3 –air) 31.7 32.5 39.4 32.9
LFL (NH3 –20% N2 –air) 15.4 15.2 16.2 14.1
UFL (NH3 –20% N2 –air) 19.8 20.4 21.3 20.9

this same figure will be compared with experimental data of this In airplanes, the temperature of the fuel and vapor depends on
work for ethanol. the rate of heat transfer to the fuel from the surroundings. Fuel
Kondo et al. [42] reported that flammability limits of gases are temperature is not normally measured in flight but can be esti-
dependent on temperature at temperatures from 5 to 100 ◦ C in a mated from a heat transfer model or measured in a special flight
12 L spherical flask basically following ASHRAE method. The mea- test. A heat transfer model must include heat transfer from the
surements were done for methane, propane, isobutane, ethylene, heated cabin (nominally at 21 ◦ C), heat transfer from the air packs,
propylene, dimethyl ether, methyl formate, 1,1-difluoroethane, heat losses through the rear spar to the ambient atmosphere in the
ammonia, and carbon monoxide. As a result, the linear tempera- wheel well, and heat transfer through the side-of-body ribs from
ture dependence of LFL was found to be adequately predicted using the adjacent fuel in the wing. The amount of fuel in the tank, the rate
a limiting flame temperature concept that is constant independent at which fuel is removed, and the rate of climb or descent are addi-
of the experimental temperature. In addition, for compounds like tional factors that must be considered in addition to the properties
methane, propane, isobutane, propylene, methyl formate, and 1,1- of the particular batch of fuel [43].
difluoroethane, a limiting flame temperature concept along with
an assumption that geometric mean “G” [37] is independent of 3.2. Pressure
temperature enables one to predict the temperature dependence
of UFL reasonably well, at least in the temperature range exam- When compared to the effect of fuel concentration, atmospheric
ined. However, this author mentions that the UFL of ethylene, pressure variations do not have a considerable effect on flamma-
dimethyl ether, and carbon monoxide deviate from this predicted bility limits. The effect of pressure variation is neither simple nor
temperature dependence. Also, ethylene and dimethyl ether have uniform, but it is specific for each mixture. The variation is generally
special burning characteristics at high concentrations. In addition, unnoticeable for the first few hundred mmHg, which decreases due
the flammable range of CO may be too wide to apply the present to atmospheric pressure; the effect increases until the flame cannot
model to UFL [42]. propagate at low pressure [1].
As the temperature rises, lower flammability limits are gradu- Increases above atmospheric pressure do not always increase
ally shifted down and upper limits are shifted up. the interval width of flammability limits. For some mixtures, the

Fig. 3. Temperature effect on flammability limits of ethanol at 1 atm [2,36].


40 C.J.R. Coronado et al. / Journal of Hazardous Materials 241–242 (2012) 32–54

Fig. 4. Effect of pressure and concentration on flammability limits at 298 K for avi- Fig. 6. Effect of pressure and temperature on flammability limits for saturated avi-
ation gasoline [1]. ation gasoline vapor–air mixtures [1].

interval decreases as pressure increases. So a mixture that has flame about 0.25 psi) when cruising since the vents are located in a low
propagation at atmospheric pressure may not have it at higher pressure region on the wing tips [43].
pressures. From tests conducted at the FAA William J. Hughes Technical
A moderate pressure increase above atmospheric pressure Center, Atlantic City International Airport, Atlantic City, NJ, it was
decreases the interval width of flammability intervals for determined that at sea level through 10 kft, LOC is approximately
hydrogen–air or carbon monoxide–air mixtures. In the same con- 12% O2 , with a linear increase from 12% at 10 kft to approximately
ditions, the flammability interval width for paraffin gases mixed 14.5% at 40 kft. There was little variation in tests with various sparks
with air decreases on the LFL side and increases on the UFL side [1]. and arcs as ignition sources at sea level, with LOC ranging from
Fig. 4 shows the effect of pressure and concentration on flamma- 12.0% to 12.8%. Also, a heated surface capable of igniting a fuel air
bility limits at 298 K for aviation gasoline. Fig. 5 presents the effect mixture proved insufficient for ignition in a tank inerted to just
of pressure and temperature on flammability limits for saturated 14%. Peak pressures resulting from ignition at oxygen concentra-
aviation gasoline vapor–air mixtures [1]. tions 1–1.5% above LOC values decreased as the altitude increased
LFL has a small decrease as pressure rises, but UFL has a sig- to 30 kft, while time to reach peak pressure increased. According
nificant increase [2]. Below atmospheric pressure, as pressure to Summer [46], further experiments to examine the trend of peak
decreases, the values of the two limits approach each other and the pressure rise as a function of both altitude and oxygen concentra-
gap between them is narrowed until a certain pressure is reached tion are needed.
(for methane, 125 mmHg at 293 K) [1,44]; at low pressures, the According to Nestor [4], for the LFL region, the color of the flame
flame cannot propagate through the mixture; according to Stull is blue at an altitude between 0 and 20 kft, the flame is blue-green
[45], this is due to the fact that the gas (fuel) concentration is too at an altitude between 20 and 40 kft, and the flame is green at an
low to sustain combustion. altitude between 40 and 60 kft. For the UFL region, the flame is yel-
Fuel in an airplane is exposed to a range of conditions during low at an altitude between 0 and 20 kft, yellow-green at an altitude
flight. The pressure and temperature in the atmosphere are a func- between 20 and 40 kft, and green at an altitude between 40 and
tion of altitude (see Fig. 6). The pressure within the fuel tank is 60 kft.
slightly lower than ambient pressure (the difference in pressure is
3.3. Turbulence

Few observations have been made in relation to the effect of tur-


bulence on flammability limits. However, it has been shown that
the LFL for methane and ethane in air are slightly reduced due to
turbulence generated by a fan or a mixer [1]. Additionally, flamma-
bility limit interval widths are slightly increased due to an increase
in mixture flow rate.
Turbulence has an influence on mixtures of methane and air.
When a small fan was set at the right speed (fast enough but not
too fast) in a methane–air mixture contained in a 4 L globe, the
LFL of methane was 5.0% compared with 5.6% observed for a still
mixture in the same vessel. A streaming movement of the gas
mixture produces similar effects on LFL. At a speed of 35–65 cm/s,
flame was propagated in a 5.02% methane–air mixture but not at
Fig. 5. Temperature and pressure as functions of altitude, standard atmosphere [43]. any speed in a 5.00% mixture. The same figure was obtained when
C.J.R. Coronado et al. / Journal of Hazardous Materials 241–242 (2012) 32–54 41

Table 5
Flammability limits in air and in oxygen [1,32].

Fuel LFL (%) UFL (%)

Air O2 Air O2

Hydrogen 4 4 74 94
Carbon monoxide 12 16 74 94
Ammonia 15 15 28 79
Methane 5 5 15 61
Propane 2 2 10 55
Ethane 3 3 12.5 66
Isobutene 1.8 1.8 8.4 48
Ethylene 2.7 2.9 36 80
Propylene 2 2 11 53
Coal gas 5.3 5.3 32 70

Fig. 7. Minimum ignition energy [41].

movement of the mixture was produced by expansion caused by


its own combustion in experiments on the propagation of flame ASTM E2079-07 standard test methods cover the determina-
from the closed end to the open end of a large vessel [1]. tion of the limiting oxygen (oxidant) concentration of mixtures of
For a mixture of ethane in air, in a 4 L globe, the LF for a still oxygen (oxidant) and inert gases with flammable gases and vapor
mixture was 3.10% ethane. With a fan running at high speed inside at a specified initial pressure and initial temperature. These test
the vessel, a 3.2% mixture did not ignite, but when the fan was run methods may also be used to determine the limiting concentra-
at a moderate speed, a 3.0% mixture exploded, producing a pressure tion of oxidizers other than oxygen. These test method are not
of 4.3 atm [1]. applicable to mixtures that undergo spontaneous reaction before
Coward and Jones [1] reported that the LFL of methane, ethane ignition is attempted, nor to mixtures that have maximum defla-
and natural gas air mixtures can be reduced by about 12% by a gration pressures below the maximum operating pressure of the
suitable amount of turbulence produced either by a fan or by the test apparatus [51].
movement of the mixture itself. The effect disappears at higher Zlochower and Green [52] reported data on limiting (minimum)
velocities, at which no flame is observed [47]. Sokolik et al. [48] oxygen concentration (LOC) with N2 added for methane, propane,
showed that for each mixture there is a certain maximum turbu- ethylene, carbon monoxide, and hydrogen and some of their binary
lence level for turbulent flame front propagation velocity. A further mixtures. They also address the issue of the flammability limits
increase in turbulence causes a drop in flame velocity and finally of these pure gases in air. Their results had excellent agreement
the flame goes out. Karpov and Severin [49] observed the same between the 120 L and 12 L results, good agreement with the 20 L
effect for mass burning rate. Lockwood and Megahed [50], in a results, and reasonable agreement with the earlier flammability
study of flame lift above small burners, concluded that flames tube values for LFL. However, they differ from the more conserva-
can be quenched by turbulence when the molecular diffusion rate tive European values. Several studies give examples of LOC data
between the microscale turbulent vortices exceeds the chemical applications [53–56]. Much of the previous literature LOC data
kinetic rate. However, Chomiak and Jarosinski [47] raised some [1–3] were measured in a flammability tube.
objections to this.
3.5. Ignition energy
3.4. Comparison between the limits in air and in oxygen
Explosion hazards are quantifiable in terms of the product of
Knowledge about LOC is needed to make some chemical and the probability of existence of a flammable mixture and the sepa-
industrial processes safe. This information may be needed in order rate probability of the presence of an adequate ignition source [57].
to start up or operate a reactor while avoiding the creation of LFL and UFL are thus boundaries defining the extent of a compo-
flammable gas compositions therein or to store or ship materials sition’s domain of explosibility at room temperature and ambient
safely. The NFPA (National Fire Protection Association) has guide- pressure within which explosion is possible in the presence of an
lines for the practical use of LOC data, including the appropriate adequate ignition source, but beyond which explosion is impossible
safety margin to use [51]. regardless of the strength of the ignition source [57].
In general, the lower limit is almost the same in oxygen and in air A potential ignition source in an accidental fire and explosions
for a gas. Until 1960, very few comparative results were available is an electric shock or spark. Other possible sources are hot sur-
to show whether there really was a small difference between the faces or adiabatic compressions of the air–combustible mixture. It
limits in oxygen and air. Nowadays, it is known that the lower limit was found that for a combustible–oxidant (air) mixture, there is a
is practically the same in oxygen and in the air. On the other hand, minimum energy required to cause an ignition when the mixture
the upper limit is much higher in oxygen than in air. This happens is within the flammability interval. This energy is called ignition
because the lower limit is in the oxidant excess zone, and switching energy.
nitrogen for oxygen does not affect the important parameters as Ignition energy depends on the mixture’s composition. The
does flame temperature [40]. minimum ignition energy for hydrocarbons in air was measured
Using central ignition in a globe vessel or with downward prop- for several common substances. Most of the results were around
agation of flame in a tube, the LFL of methane is rather higher in 0.2 mJ, usually for a rich mixture. Some examples of minimum igni-
oxygen than in air, presumably because the mean molecular heat of tion energies are shown in Fig. 7. Based on experimental results,
oxygen is higher than that of nitrogen between room temperature the minimum ignition energy for petroleum derived fuels including
and that of a flame limit, for example 1200–1400 ◦ C. For upward or aviation kerosene is expected to be close to 0.2 mJ [6].
horizontal propagation the LFL of methane is slightly lower in oxy- Flammability limits depend on the ignition source. The most
gen than in air. UFL of ammonia, methane, and propane is much common sources are sparks, generally used with the visual crite-
greater in oxygen than in air [1]. Table 5 shows results for flamma- rion, and fuse wires, generally used with the pressure criterion. The
bility limits in air and in oxygen. spark systems are capacitance sparks, which are developed from
42 C.J.R. Coronado et al. / Journal of Hazardous Materials 241–242 (2012) 32–54

discharged condensers. The duration of these discharges can be as results for mixtures that contain H2 and unsaturated hydrocarbons;
short as 0.01 ␮s or as long as 100 ␮s for larger engines [26]. it is also valid only if the components are chemically similar.
Energy and its distribution in both time and space have an effect Using Le Chatelier’s law, Carvalho and McQuay [40] showed that
on the values obtained for flammability limits. This effect is particu- if there are two different hydrocarbons in individual mixtures with
larly true for difficult to ignite materials that are not very energetic air that are both in the lower flammability limit, when they are
when burning. What is more, if too high an energy ignition source added together, the resulting mixture will also be in the lower
is used, all that can be seen is dissipation of the ignition energy and flammability limit. Similar reasoning is valid for the upper lim-
not the propagation of a flame. The spark is an optimal and accept- its. For the demonstration, consider: (a) a mixture of a moles of
able ignition source [28]. An electric spark was used as an igniter hydrocarbon A and x moles of air, and (b) a mixture of b moles of
mechanism in the experiments in this study. hydrocarbon B and y moles of air, both in the lower flammability
As already mentioned in the previous section, tests conducted limit. Then:
in smaller volumes using larger ignition energies tend to result in 100a
wider flammability range and therefore they are more conservative LFLA = (%), (11)
a+x
could lead to potentially higher prevention and protection costs.
Considering the fact that low ignition energy or small test volume 100b
LFLB = (%), (12)
can mask the flammability of certain fuels, the prudent practice is to b+y
rely on flammability data obtained using sufficiently large ignition Using Le Chatelier’s law:
energies in sufficiently large test enclosures [33].
Minimum ignition energy is a function of electrode spacing. It 100
LFLA+B = , (13)
becomes asymptotic to a very small spacing below which no igni- (CA /LIA ) + (CB /LIB )
tion is possible. Minimum ignition energy decreases as electrode
in which
spacing increases, reaches its lowest value at a certain spacing,
then begins to rise again. For small spacings, the electrode removes 100a
CA = (%), (14)
large amounts of heat from the incipient flame and thus large min- a+b
imum ignition energy is required. As the spacing increases, the 100b
surface area to volume ratio decreases, and consequently the igni- CB = (%). (15)
a+b
tion energy required decreases [26].
Many papers have been published about ignition [58–63]. Thus:
It should be noted that experiments on ignition energies car-
100
ried out by Pasamehmetoglu and Unal [64] in 11.25 L and 400 L LFLA+B =
(100a/(a + b))((a + x)/100a) + (100b/(a + b))((b + y)/100b)
combustion vessels at initial pressures of 100 kPa and tempera-
tures of 295 K. Ignition energy bounds of methane–nitrous oxide, 100 100(a + b)
= = , (16)
((a + x + b + y)/(a + b)) (a + b) + (x + y)
ammonia–nitrous oxide, and ammonia–nitrous oxide–nitrogen
mixtures have been determined. Lower and upper flammability
limits (mixing fan on, turbulent conditions) for ignition energies of which is the concentration of (a + b) moles of fuel in a mixture with
8 J were: H2 –N2 O: 4.5–5.0% H2 (LFL), 76–80% H2 (UFL); CH4 –N2 O: (x + y) moles of air, indicating that the new mixture will be exactly
2.5–3.0% CH4 (LFL), 43–50% CH4 (UFL); NH3 –N2 O: 5.0–5.2% NH3 in the lower flammability limit.
(LFL), 67.5–68% NH3 (UFL).
4. Flammability limits in the aeronautical industry
3.6. Flammability limits of fuel mixtures
4.1. Importance for the aeronautical industry
LFL and UFL for fuel mixtures can be determined by the Le Chate-
lier’s law [3]: Since the introduction of kerosene as a fuel for civilian air trans-
port in the 1950s, aeronautical designers have been careful about
100 the ullage in aircraft fuel tanks because they may contain a mixture
LFLmixture = , (9)
(C1 /LFL1 ) + (C2 /LFL2 ) + . . . + (Ci /LFLi ) of various vapors, fuel, and air which can ignite in the presence of
100 an ignition source when this mixture is within the flammability
UFLmixture = , (10) limits of the fuel in use. After the flight TWA 800 accident, which
(C1 /UFL1 ) + (C2 /UFL2 ) + . . . + (Ci /UFLi )
fell near the coast of New York in 1996, concerns about formation
in which C1 ,C2 ,. . .,Ci (%, volumetric basis) are the proportions of of vapors in aircraft fuel tanks reached a peak. It was determined
each fuel gas in the fuel mixture. According to Kondo et al. [65], the that the crucial reason for the lost plane was an explosion inside
Le Chatelier’s law accurately predicts LFL with for a wide variety of the Fuel-CWT tank (airplane wing). One significant factor for TWA
combustible mixtures. 800 was the very small mass of fuel in the CWT. At low mass-to-
The flammability limits of mixtures with propane were well volume ratios, Jet A vapor pressure decreases due to depletion of the
explained by Le Chatelier’s original formula [66]. Likewise, Kondo more volatile components in the fuel [43]. At the moment of explo-
et al. [67] measured a carbon dioxide dilution effect on flammability sion, the plane had approximately 50 gal of Jet A fuel. The fuel and
limits for various flammable gases. The values obtained were ana- air contained inside the tank were heated by the air conditioning
lyzed using Le Chatelier’s extended formula developed in a previous system (air packs) located directly below the CWT. This increased
study. fuel vaporization in the ullage region, forming a flammable mix-
On the other hand, for most cases, experimental values for UFL ture which later caused the accident. Since then, flammability limit
are lower than those predicted using the law. Von Niepenberg studies for the aeronautical industry have been increasingly taken
et al. [68] used Le Chatelier’s law to predict flammability limits into account in aircraft design [5,6,43].
for mixtures that contained inert gases. Hustad and Sonju [69] had Required conditions for ignition and flame propagation inside
experimental results that were very near those predicted by Le the aircraft fuel tank depend on various parameters that include
Chatelier’s law for fuel mixtures at high temperature and pressures. fuel type, temperature, tank pressure, and oxygen concentration.
According to Glassman [26], the law does not produce accurate It was also noted that the foam and spray of fuel, which may
C.J.R. Coronado et al. / Journal of Hazardous Materials 241–242 (2012) 32–54 43

4.2. Ignition

There is little data regarding ignition energy for aeronautical


fuels, but minimum ignition energy of 0.2 mJ for hydrocarbons is
suggested in CRC [70] as being suitable for aeronautical fuels. Exper-
iments by Nestor [4] and Ott [7] used sparks with energy ranging
between 4 and 20 J. These values are applied to create standard
figures involving flammability limits in most aeronautical fuel man-
uals [36]. Results achieved by Nestor [4] for flammability limits
with ignition energy of 20 J are shown in Fig. 8. In this figure, the
“safe” regions are those in which fuel ignition will not occur and
the medium region indicates flammability. The vertical line indi-
cates the altitude in thousands of feet (e.g., the cruising altitude
in commercial aviation is typically 35,000 ft). Results from other
researchers are shown in Fig. 9. In this figure, the data are plotted
in the form of tank altitude versus fuel temperature for specified
spark ignition energy. Table 6 shows a summary of Jet A fuel ignition
tests applied since the 1960s.
In Table 6, Nestor [4] conducted a test in a 4-ft-long vertical
transparent tube with an ignition source at the bottom. Successful
ignitions at various fuel temperatures were observed as upward
flame propagation to the top of the tube [5]. The ignition source
in Ott’s test was a moving arc that traveled up and down a pair
of 1-ft-long electrodes. Ott [7] used the pressure generated during
combustion as a measure of fuel flammability for both static fuel
and sloshing fuel tests. The 1992 Air Force [71] tests were similar to
Nestor’s [4] flame tube static ignition tests, except that the 1992 Air
Force tests involved higher flash point samples and a broader range
of spark energies [5]. Shepherd et al. [6] conducted Jet A ignitability
tests designed to determine the effect of fuel load and spark ignition
Fig. 8. Flammability limits for Jet A fuel in air [70] based on data by Nestor [4] with energy at various fuel temperatures.
ignition energy of 20 J. According to Shepherd et al. [6], the key factor in determin-
ing ignition energy is the composition and concentration of fuel
vapor. The concentration of fuel vapor in equilibrium with a liq-
form during refueling or during a flight with fuel oxygen libera- uid fuel is proportional to fuel vapor pressure P,fuel and inversely
tion, might increase fuel flammability limits. In this sense, when proportional to ambient pressure Pa :
working with flammability limits of fuels for the aeronautical
P (Tfuel )
industry, the following details must be taken into account: altitude Xfuel = . (17)
Pa
changes, temperature changes, ullage changes, tank ventilation,
state changes of fuel (spray) by agitation due to aircraft movement, The vapor pressure of fuel in turn [43] depends first on liquid
and fuel mixtures during refueling operations. fuel temperature Tfuel and second on the amount of liquid and the

Fig. 9. Jet A flammability data comparison [6].


44 C.J.R. Coronado et al. / Journal of Hazardous Materials 241–242 (2012) 32–54

Table 6
Ignition tests for Jet A fuel [6].

Author Flash point (◦ F) Vessel size (L) Fuel concentration (kg/m3 ) Ignition source Vessel pressure (atm)

Nestor [4] 118, 120, 125 Tubular (9.9) 100 Multiple sparks (5 and 20 J) 1 and 0.58
Nestor [4] Tank (9.5) 100 Spark (16 and 24 J) 1
Ott [7] 188, 123 Tank (303) 100 Electric arc 1 and 0.68
Air Force [5] 130, 140 Tubular (12.4) 100 Spark (3 mJ to 16 J) 1 and 0.69
Shepherd et al. [6] 118 Cylindrical (2) 3 and 200 Spark (1 mJ to 100 J)

Fig. 10. Measured ignition energy for Jet A vapor–air mixtures, 200 kg/m3 mass load,
and an initial pressure 0.585 bar [6].
Fig. 11. Measured ignition energy for Jet A vapor–air mixtures, 3 kg/m3 mass load,
history of handling. To examine temperature dependence, Shep- and initial pressure 0.585 bar [6].

herd et al. [6] studied dependence on fuel load by testing two


situations, one with 200 kg of fuel per m3 of tank fuel correspond- [3] for mist (spray) of Jet A sprayed directly onto electrodes. Fig. 14
ing to a 25% full tank and the other with a nearly empty tank with shows the relation between Jet A ignition energy and altitude.
3 kg/m3 corresponding to a thin layer of residual fuel. The tests had
a tank pressure of 0.585 bar, which corresponds to an altitude of 4.3. Flammability properties of aviation fuels
about 14,000 ft, at which the TWA Flight 800 explosion occurred.
In the first case, Fig. 10, the experiments produced the following There are at least 23 specifications in ASTM D1655 for Jet A or
results in the 1.8-L vessel (about 450 mL of liquid fuel): below 30 ◦ C, Jet A-1 fuels, which set maximum or minimum limits for stated
the ullage mixture cannot be ignited even with spark energies of up properties or measurements. Compare this to nine specifications
to 100 J. Ignition energy decreases rapidly as temperature increases. in the ASTM Diesel fuel requirements or six specifications in the
Above 50 ◦ C, the mixture can be ignited with spark energies on automotive gasoline requirements [72]. According to ASTM D1655,
the order of 1–10 mJ. In the second case, Fig. 11, the experiments there are three grades of aviation turbine fuel: the kerosene-based
produced the following results in the 1.8-L vessel (about 7 mL of Jet A and Jet A-1 (approximate boiling point range of 160–300 ◦ C)
liquid fuel): dependence on temperature was very similar to the and wide-cut or naphtha-based Jet B (approximate boiling range of
200 kg/m3 case. The expected mass-load effect is not obvious [6]. 50–300 ◦ C). The difference between Jet A and Jet A-1 is that freezing
Fig. 12 shows the minimum ignition energy for Jet A and Jet B point of Jet A-1 is −47 ◦ C and Jet A is −40 ◦ C. Jet B is little used
fuel. Fig. 13 shows the values of ignition energy measured by Kuchta today except in some Arctic operations, although it is essentially

Fig. 12. Minimum electric ignition energy for aviation fuels Jet A and Jet B [5].
C.J.R. Coronado et al. / Journal of Hazardous Materials 241–242 (2012) 32–54 45

Table 7
Summary of flammability limits for ethanol–air mixtures at 298 K and 1 atm [1].

Tube dimensions (cm) Flammability limits (vol%)

Diameter Length LFL UFL

Upward flame propagation


30.6 39 (steel) 3.48 –
15 300 (steel) 4.16 –
10.2 96 (steel) 3.28 –
7.5 150 (glass) 3.56 –
5 150 (glass) 4.24 18.95
5 130 (glass) 3.69 18.00
5 91 (glass) 4.30 –
2.5 150 (glass) 5.02 –
2.5 25 (glass) 5.0 14
Fig. 13. Dependency of ignition energy on fuel temperature for Jet A sprays [3].
Outward flame propagation
15 300 (steel) 4.23 –
7.5 150 (glass) 3.70 –
the same fuel as JP-4, the standard US military fuel in use until 5 150 (glass) 4.32 13.80
about 10 years ago. The current US military aviation fuels are JP-8, 5 91 (glass) 4.40 –
which is nearly equivalent to Jet A, and JP-5, which has a high flash 2.5 150 (glass) 5.18 –
point specification [5]. Downwards flame propagation
Jet A/A-1 and JP-8 fuels are typically mixtures of hundreds of 15 300 (steel) 4.37
compounds, controlled only by the defined boiling point ranges. 7.5 159 (glass) 3.75
Current analyses can identify over 200 chemical species in sample 6.2 33 (glass) 3.70
5 150 (glass) 4.44 11.50
fuels [73]. Jet A type fuels consist of the order of 75–85% paraf- 5 91 (glass) 4.50
fin, both straight chain and cyclic, with the balance almost entirely 5 70 (glass) 3.81
aromatic compounds [74]. 2.5 150 (glass) 5.21
Regarding flammability properties of aviation fuels, the two 1.9 40 (glass) 3.95 13.65
5.0 11.5
most widely used parameters are the fuel’s flash point and flamma-
bility limits given in terms of altitude and temperature data. The
current standardized methods for determining flash point are
5. Tests for flammability limits of ethanol: review
ASTM D56 and ASTM 3828 [46].
The flammability limits of a given fuel provide more information
Coward and Jones [1] also carried out tests using ethanol; they
than the flash-point measurement, especially when presented as
used different vessel sizes and separated the limits by flame prop-
regions of flammability in terms of altitude and fuel temperature.
agation direction. Table 7 summarizes their results. In Table 7, the
At any given altitude (pressure), a fuel will have a lean (or lower)
lower limit of ethyl alcohol in air with upward flame propagation
flammability limit (LFL) and a rich (or upper) flammability limit.
in a 5 cm diameter tube open at the firing end is 4.25 or 4.40%. The
Within these limits, ignition of vapors may occur, while outside
limits in a small vessel have been determined to be 2.6 and 9.0%;
the limits, no combustion phenomenon should be observed. Over
in a 1400 cm3 upright closed tube with 600 cm3 of liquid at the
the years, there have been numerous experiments to determine the
bottom and central ignition, the limits were 3.2% (saturation of the
flammability limits of aviation turbine fuels. A typical plot of these
atmosphere at 10.60 ◦ C) and 18.9% (saturation at 41.2 ◦ C) [1].
limits, including ignition energies, for Jet A and Jet B fuels is shown
Zabetakis [2] also carried out tests using ethanol. The flamma-
in Fig. 9, as taken from Fuel Flammability Task Group [5].
bility limits of ethanol with carbon dioxide–air and ethanol with
Both the flash point and flammability limits given above assume
nitrogen–air mixtures at 298 K and 1 atm are shown in Fig. 15.
isothermal and equilibrium conditions, not considering such things
Fig. 16 shows data for ethanol–water vapor mixtures at 100 ◦ C and
as possible condensation of fuel vapor and variation of ullage com-
position. In order to provide useful data for fuel tank inerting
requirements, data on the reduction of fuel flammability as a func-
tion of oxygen depletion are needed [46].

Fig. 15. Flammability limits of ethanol mixed with inert substances at 298 K and
Fig. 14. Minimum ignition energy with altitude [5]. 1 atm [2].
46 C.J.R. Coronado et al. / Journal of Hazardous Materials 241–242 (2012) 32–54

Table 8
Flammability limits for the most important alcohols [2].

Fuel Formula Molecular weight Stoichiometric Combustion enthalpy LFL (vol%) UFL (vol%)
(g/mol) concentration (vol%) (kcal/mol)

Methyl alcohol CH3 OH 32.04 12.25 159 6.7 36


Ethyl alcohol C2 H5 OH 46.07 6.58 306 3.3 19
n-Propyl alcohol C3 H7 OH 50.09 4.45 448 2.2 14
n-Butyl alcohol C4 H9 OH 74.12 3.37 596 1.7 12
n-Amyl alcohol C5 H11 OH 88.15 2.72 742 1.4 10
n-Hexyl alcohol C6 H13 OH 102.17 2.27 888 1.2 –

atmospheric pressure. Table 8 shows flammability limit results for he reports flammability values for pure ethanol of 3.3 vol% for LFL
the most important alcohols. and 19 vol% for UFL, both at 25 ◦ C.
Seaton [75] proposed a mathematical model derived from Le Brooks and Crowl [80] carried out experiments on flamma-
Chatelier’s law to estimate the flammable limits of vapors in air. bility of some alcohols. The experiments were conducted for
Both the lower and the upper flammable limits can be estimated methanol, ethanol, acetonitrile, and toluene. The experimental
from nothing more than the chemical structure of a compound. device consisted of a 20 L volume spherical vessel with a fuse wire
Although there are some outstanding exceptions, the model pre- igniter and ignition energy of 10 J. Fuel was injected using a preci-
dicts LFL values to within 10% of published values. Likewise, the sion syringe. Nitrogen and oxygen were introduced with pressure
model predicts most UFL values to within 20% of published values. gauges. The pressure criterion was applied, considering those above
In the case of LFL for ethanol, he estimated 3.39% (vol) and for the 7% of initial pressure as flammable. The LFL obtained were close to
UFL, he estimated 16.1% (vol). the data in the literature. All the experiments were carried out at
McCormick and Parish [76] prepared a study sponsored by the 298 K and 1 atm.
National Renewable Energy Laboratory (NREL-USA) about the tech- The flammability test device had the following components:
nical barriers to the use of ethanol in diesel fuel. They used ethanol
flammability limits values of 3.3 for LFL and 19% for UFL (vol% at
25 ◦ C). In that work, a report prepared for Growmark, Inc. [77] - A high precision pressure transducer (+/− 0.05% full scale) for low
and Battelle [78] demonstrated that blends of 10%, 15%, and 20% pressures, 0–1.7 bar abs, utilized for the gas mixture.
ethanol in conventional diesel have combustion safety characteris- - A high pressure transducer, 0–13.6 bar abs, utilized to gauge the
tics essentially identical to pure ethanol. These data were acquired pressure during the explosion.
using diesel ethanol blends that contained no emulsifier. - Thermocouples to gauge the temperature at the center of the
Waterland et al. [79] prepared a report sponsored by an agency vessel, at the inner wall, and at the external wall.
of the United States government about the safety and performance
assessment of ethanol/diesel blends (E-Diesel). In this document, The main difference from the other equipment is the liquid
injection port, which is connected to the vessel and protected from
the explosion pressure by a high pressure manual globe valve.
A 1 mL Kloehn #400 precision syringe was used. The liquid was
injected when the vessel pressure was 2.07 kPa (below saturation
pressure of all analyzed fuels). The liquid evaporated as soon as it
entered the vessel due to low pressure inside the vessel. Each test
takes from 30 to 45 min, and they were divided into four stages.
The experimental result from Brooks and Crowl [80] for the LFL
of ethanol was 3.7% (according to Kutcha [3], the LFL is 3.3%). On the
other hand, the experimental result for the OCL (oxygen concentra-
tion limit) was 9.8% (according to Kutcha [3], the OCL is 10.5%). The
saturation pressure of ethanol was taken at 25 ◦ C and 593.5 mmHg
(Perry and Green [81]). The results are shown in Fig. 17.
Brooks and Crowl [82] carried out experiments with the
flammability of vapors above aqueous solutions of ethanol and ace-
tonitrile in a 20 L combustion apparatus. They reported that the
lower flammability limits (LFL) of ethanol decreased from 3.7% for
pure vapor to 3.2% for saturated water vapor.
Astbury [83] presented a brief review of the properties and
hazards of some alternative fuels, including ethanol and methanol.
He also reports some flammability values for the aforementioned
alcohols and other alternative fuels of interest.
Ma [84] reported a methodology based on thermal balance
for estimating the flammability limits of a mixture with ethanol.
Melhem [85] presented a general method for the estimation of
flammability envelopes for chemical mixtures based on chemical
equilibrium. The impact of the mixture’s initial temperature, the
presence of diluents, and elevated system pressures are implicitly
accounted for. Hansen et al. [86] published a review which dis-
cussed the properties and specifications of ethanol blended with
Fig. 16. Flammability limits of ethanol mixed with water vapor–air mixtures at diesel fuel with special emphasis on the factors that were critical
100 ◦ C and 1 atm [2]. to the potential commercial use of these blends. The three previous
C.J.R. Coronado et al. / Journal of Hazardous Materials 241–242 (2012) 32–54 47

Fig. 17. Experimental data for ethanol flammability limits [80].

authors reported that for ethanol at ambient pressure and room liquid injection. The liquid was injected when the pressure in the
temperature, LFL was 3.3 (vol%) and UFL was 19 (vol%). reaction vessel was 2 kPa (0.3 psia), which is below the saturation
Pidol et al. [87] studied the properties of ethanol blended fuels vapor pressure for ethanol used in this research. The liquid flashed
and evaluated their behavior in conventional diesel combustion to vapor upon injection into the reaction test vessel. The evaporator
and advanced combustion with low temperature combustion (LTC). was used only in the experiments to find the UFLs.
These authors reported that the addition of ethanol into Diesel fuel The rubber stopper was attached to four rods connected to four
affects some key properties such as blend stability, cetane num- springs to hold it secure in the vessel port. The 20 L spherical glass
ber or flash point, and the fuel formulation was thus improved. For vessel was equipped with a high precision pressure transducer and
these studies, Pidol et al. used the ethanol flammability limit values two thermocouples, one positioned near the wall (type E), and the
reported by Rakopoulos et al. [88]. other positioned in the inner core of the vessel (type K). The pres-
sure transducer measured initial pressure and pressure variation
during the explosion. The spark gap was 6.4 mm (1/4 in.), and igni-
6. Experimental tests with ethanol tion energy was 90 J. The flammability apparatus was automated
using a Lab-View program.
6.1. Flammability apparatus

A heating chamber flammability apparatus was specially con- 6.2. Description of experimental procedure
structed on the basis of the American standard ASTM E-681 [28]. A
20-L spherical test vessel was used. It should be noted that the vol- The Visual Criterion would be the most appropriate and was
ume specified by the American standard was increased because no chosen for determining the flammability limits of ethanol [89]
rule is provided for use at high temperatures and reduced pressures, (pure and/or mixed) with air. As mentioned, American standard
when the situation requires less air and fuel. A drawing of the appa- E681-04 [28] was chosen as a basis for building a prototype heat-
ratus, which was adapted from Brooks and Crowl [80], is shown in ing chamber that was modified in order to use the visual criterion
Fig. 18. and validate the experimental method for ethanol.
An evaporator was used in the connection line to the vessel. In the experiments it is essential that the vessel have suf-
This guaranteed that the fuel sample entered the vessel as a vapor, ficient ethanol vapor pressure to form flammable mixtures in
even at low temperatures. A 1 mL hypodermic syringe was used for air at atmospheric and negative pressures (below atmospheric

Fig. 18. Flammability apparatus.


Adapted from [80].
48 C.J.R. Coronado et al. / Journal of Hazardous Materials 241–242 (2012) 32–54

irregular flame propagation or insufficient luminescence in the vis-


ible spectrum).
As there is interest in testing ethanol as a fuel for aircraft, the
tests should be bounded by the initial pressure of the local environ-
ment down to 13 kPa (100 mmHg). Considering a typical cruising
altitude of the aircraft of 40,000 ft (approximately 18.82 kPa),
13 kPa is well below the limit of the American standard. The tem-
perature of the heating chamber can vary between 20 and 200 ◦ C.
A uniform mixture of a gas or vapor with air should be ignited in a
closed vessel and propagation of the flame from the ignition source
will be observed. The concentration of ethanol is varied between
each test until the composition that will achieve flame propagation
is determined.
Each run took between 15 and 25 min to complete. First, the
reaction vessel was purged with nitrogen and the pressure was
reduced to less than 2 kPa abs (0.3 psia) to insure removal of all
water and other combustion products from previous runs. Second,
the liquid was injected into the reactor at low pressure to insure its
complete vaporization. Third, the synthetic air was loaded into the
Fig. 19. Visual criterion for flame propagation inside a 20 L vessel [28]. reactor and the concentrations of all the components were deter-
mined from the partial pressures. Fourth, the video camera was
turned on, the mixture was ignited, and the results were collected
pressure) (future experiments) with different temperatures. The and stored.
method is based on electric ignition and the visual observation of All of the experiments were run at atmospheric pressure and
flame propagation in the mixture in the vessel. Users could expe- approximately 25 ◦ C to compare with published data for this pres-
rience some problems if the flames are difficult to observe (e.g., sure and validate the experimental procedure.

Fig. 20. Minimum volume of ethanol as function of vessel size for different temperatures at 1 atm.

Fig. 21. Maximum volume of ethanol as function of vessel size for different temperatures at 1 atm.
C.J.R. Coronado et al. / Journal of Hazardous Materials 241–242 (2012) 32–54 49

Fig. 22. LFL, initial temperature: 42 ◦ C, sample volume: 1.70 mL, flammable mixture.

The experimental procedure was carried out according to the k) Activate the electrode (for ignition).
following the sequence: l) Observe whether the flame propagates or not, according to the
standard indications.
m) Start the process again in order to obtain values with and with-
a) Double check the equipment.
out flame propagation for the lower and upper flammability
b) Purge of a glass vessel twice using an inert gas (e.g., N2 ).
limits.
c) Purge the vessel once using an oxidant (e.g., synthetic air, 80%
N2 and 20% O2 ).
d) Set a pressure value that guarantees ethanol evaporation dur- Each test was video recorded, and later each video was edited
ing its introduction into the vessel (the ethanol vapor pressure with computer video editing software. This allowed determining
curve should be used as a reference). whether a sample had flame propagation or not at fixed tempera-
e) Set the desired volume of ethanol (a 1 mL scalp 19G hypodermic ture and pressure conditions. Standard indications for determining
syringe could be used; a magnetic stirrer should be operating whether or not a mixture is flammable are of great importance.
when the sample is introduced). A mixture is labeled as flammable if the flame propagates
f) Once the sample has been completely vaporized, slowly allow upward and outward from the vessel walls, forming an arc larger
access of oxidant until the final pressure is reached. than 90◦ measured from the ignition source to the vessel walls
g) Check temperature and pressure inside the vessel at all times. [25,30]. Flame must be continuous along the region of the vessel
h) Allow the magnetic stirrer to operate for another 2 min and then shown in Fig. 19. If flame propagation could not be reproduced,
turn it off. or flame extension was not clear, for example, there was no uni-
i) Turn off the room lamp. form propagation, flame structure was irregular, or the flame did
j) Start video recording and simultaneously record the tempera- not fill at least half the vessel even for the maximum flammable
ture and pressure inside the vessel. concentration in air, an ignition probability of 50% should be used.
50 C.J.R. Coronado et al. / Journal of Hazardous Materials 241–242 (2012) 32–54

Fig. 23. LFL, initial temperature: 40 ◦ C, sample volume: 1.40 mL, non-flammable mixture.

6.3. Mass of ethanol needed to form flammable mixtures with air LFL:

No. of moles of ethanol in the vessel


Minimum and maximum masses of ethanol needed to form a LFLEthanol = × 100
flammable mixture in vessels of 5 and 12 L is calculated for 1 atm Total no. of moles of gas in the vessel
and 25 ◦ C. A minimum and maximum mass of ethanol needed to LV . 1
= × 100 (19)
form a flammable mixture is calculated in vessels of 5 and 12 L MW (V/22.4)(p/p0 )(T0 /T )
(considering the dimensions of the American Standard, E681-04
[28]) mixed with air at 1 atm and 25 ◦ C. in which LV is the sample volume, cm3 ;  is the sample specific
From previous discussions in this article, the following mass, g/cm3 ; MW is the sample substance molar mass, g/mol; p is
flammability limits will be considered: LFLEthanol = 3.3% and the pressure of the test, mmHg; T is the initial temperature of the
UFLEthanol = 19%. For a 5-L vessel at 298 K, the number of moles of test, K; p0 is the standard pressure, 1 atm (760 mmHg = 101.3 kPa);
C2 H5 OH at the LFL is: and T0 is the standard temperature, 273 K. Thus:
5 273 LV .
nLFL = · 0.033 = 0.006784 (18) LFLEthanol = k, (20)
22.4 298 MW
Since the molar mass of ethanol is 46.07 g/mol, the mass of
in which
ethanol to form a flammable mixture with air at the LFL will be
0.006784 × 46.07 = 0.32 g (=0.40 mL). (22.4)(p0 )(100)
If the vessel volume is 12 L, the minimum mass of ethanol is k= . (21)
(V )(T0 )
0.76 g (0.97 mL). For the mixture at the upper limit of flamma-
bility, the calculation procedure is entirely analogous. For the 5-L For example, a test carried out in the 12-L vessel (12,000 cm3 )
vessel, the maximum mass of ethanol is 1.83 g (2.32 mL), and for at 25 ◦ C and 715 mmHg (atmospheric pressure in the town of
the 12-L vessel, the maximum mass of ethanol is 4.39 g (5.56 mL). Guaratinguetá, Brazil) with a sample of ethanol ( = 0.789 g/cm3
Figs. 20 and 21 present the different minimum and maximum vol- and MW = 46.07 g/gmole) produced the first flame spread with 1 mL
umes of ethanol, for different sizes of vessel, for tests at different (L1 ) of ethanol. The last volume that did not produce a flame was
temperatures. This calculation was essential to find the minimum 0.9 mL (L2 ) of ethanol. The sample volume for LFL was then taken
and maximum mass of ethanol for the 20 L vessel that is used in the as the average: 0.95 mL. For this case:
flammability apparatus used in this study.
(22.4)(719)(100)
k= = 519.65
6.4. Description of the procedure to calculate LFL and UFL from (12)(T0 )
experimental data
The value of k remains unchanged for all tests in a 12-L vessel.
Therefore, the LFL obtained was
The experimental results should be optimized, tested, and
eventually modified for volumetric concentration formats in order (0.95)(0.789)(298)
LFLEthanol = k = 3.52% (at 715 mmHg and 25 ◦ C).
to compare with other experimental data from the scientific liter- (46.07)(715)
ature. The procedure is shown for LFL. For UFL, it is the same. For
C.J.R. Coronado et al. / Journal of Hazardous Materials 241–242 (2012) 32–54 51

Fig. 24. LFL, initial temperature: 170 ◦ C, sample volume: 1.10 mL, flammable mixture.

6.5. Results and discussion rises to approximately 600–800 ◦ C in less than 0.1 s. This behavior
is similar to that specified by the standard for the case of flame
6.5.1. Behavior of flame propagation propagation. Fig. 23 illustrates the case with no flame propagation.
Fig. 22 shows a sequence of photographs for a test with propaga- When the volume of ethanol increases from the value corre-
tion in a mixture near LFL. Flame propagation was evident and there sponding to the LFL, the flame propagates across the vessel wall,
was an abrupt rise of temperature and pressure. This caused a phe- brightness increases as does the intensity of the detonation noise.
nomenon similar to an explosion. The use of protective glasses and This can be seen in Fig. 24.
headphones is important when carrying out any tests, especially The behavior is similar for UFL, except that the noise is not as
in those to determine the LFL. No vessel, electrode, or other equip- intense as it is for LFL. In most experiments to determine UFL,
ment was lost during the tests. The springs held the cover in place. no detonations were heard, unlike in the experiments for LFL. It
It was noted that for mixtures near the LFL, the flame propagates should be noted that the absence of detonation does not imply
upwards from the ignition spark, reaches the top of the vessel and an absence of flame propagation. The sudden rises in temperature
then extends to the wall and propagates downwards. Temperature registered by the thermocouples and the flame propagation across
52 C.J.R. Coronado et al. / Journal of Hazardous Materials 241–242 (2012) 32–54

Fig. 25. Hydrated ethanol–air mixtures, LFL, and UFL at 101 kPa.

the vessel wall were the factors that led to catalog the mixture as percentage of water (4%) present in the sample (hydrated ethanol
flammable. 96% ◦ INPM). This behavior at the upper limit due to the presence
of inert substances had already been anticipated by Zabetakis [2].
6.5.2. Results for atmospheric pressure
In this stage of the investigation, about 70 experiments were
7. Conclusion
performed; 90% of these tests were video recorded and all of them
were recorded in LabView. Fig. 25 shows flammability results in
A comprehensive literature review of flammability limits
terms of fuel volume as a function of temperature and of tempera-
of combustible mixtures with emphasis on ethanol was con-
ture as a function of the volume percentage of gaseous fuel. Results
ducted. The experimental procedures to determine the limits were
obtained here agreed with data published in the scientific litera-
thoroughly discussed. Flammability limit results at atmospheric
ture on flammability limits of ethanol for behavior of both LFL and
pressure for ethanol were presented.
UFL and in LFL values [2,3,41,80]. This agreement showed that the
equipment was operated correctly and also that the experimental
procedure adopted was carried out appropriately. Acknowledgements
As mentioned, LFLT and the UFLT for any temperature T at 1 atm
pressure can be calculated based on the limits determined for 298 K Funding for this work was provided by the Fundação de Amparo
using Eqs. (7) and (8). Predictions by these equations are also shown à Pesquisa do Estado de São Paulo – FAPESP, Brazil, through projects
in Fig. 25. LFL results agreed well with results from previous inves- 2009/09738-7 (regular sponsorship) and 2009/09008-9 (post doc
tigations. It is specially worth stating that the present work found fellowship to CJRC), and by Empresa Brasileira de Aeronáutica
3.5% for the ethanol LFL at 25 ◦ C, while Kuchta [3] reported 3.3% and (EMBRAER).
Brooks and Crowl [80] 3.7%, for LFL at the same temperature.
Fig. 25 also shows that the behavior of UFL as a function of
temperature is similar to that predicted in the literature (given by References
Eq. (8)). However, UFL values were about 4.5% (absolute) lower in [1] H.F. Coward, G.W. Jones, Limits of Flammability of Gases and Vapors, US Bureau
all cases. Zabetakis [2] reported 19% for UFL at 25 ◦ C. Other articles of Mines Bulletin, United States Government Printing Office, Washington, 1952,
[85,87] report the same value, but they referred to Zabetakis’ data. p. 503.
[2] M.G. Zabetakis, Flammability Characteristics of Combustible Gases and Vapors,
A value of 16.1% for UFL at the same temperature was reported
US Bureau of Mines Bulletin, United States Government Printing Office, Wash-
[75]. The experiments were repeated three times for this pressure ington, 1965, p. 627.
and the data were confirmed. Extreme care was taken to avoid [3] J.M. Kuchta, Investigation of Fire and Explosion Accidents in the Chemical, Min-
condensation of the samples when they were injected into the ing and Fuel-Related Industries – A Manual, US Bureau of Mines Bulletin, United
States Government Printing Office, Washington, 1985, p. 680.
vessel. An evaporator was utilized to guarantee sample injection [4] L.J. Nestor, Investigation of Turbine Fuel Flammability within Aircraft Fuel
in the vapor phase. The difference is probably due to the small Tanks, Federal Aviation Administration, USA, 1967.
C.J.R. Coronado et al. / Journal of Hazardous Materials 241–242 (2012) 32–54 53

[5] Fuel Flammability Task Group, A review of the flammability hazard of jet a fuel [36] A. Takahashi, Y. Urano, K. Tokuhashi, S. Kondo, Effect of vessel size and shape
vapor in civil transport aircraft fuel tanks, Final Report No. DOT/FAA/AR-98/26, on experimental flammability limits of gases, J. Hazard. Mater. A 105 (2003)
Federal Aviation Administration, USA, 1998. 27–37.
[6] J.E. Shepherd, J.C. Krok, J.J. Lee, Spark ignition energy measurements in Jet A, [37] ASTM E918-83, Standard Practice for Determining Limits of Flammability of
Report FM97-9, Explosion Dynamics Laboratory, 2000. Chemicals at Elevated Temperature and Pressure, American Society for Testing
[7] E. Ott, Effects of fuel slosh and vibration on the flammability hazards of and Materials (ASTM) Washington Office, 2011.
hydrocarbon turbine fuels within aircraft fuel tanks, Technical Report AFAPL- [38] ASTM-E2079, Standard Test Methods for Limiting Oxygen (Oxidant) Concen-
TR-70-65, Fire Protection Branch of the Fuels and Lubrication Division, tration in Gases and Vapors, American Society for Testing and Materials (ASTM)
Wright-Patterson Air Force Base, Ohio, 1970. Washington Office, 2011.
[8] T. Kosvic, L. Zung, M. Gerstein, Analysis of aircraft fuel tank fire and explosion [39] J.C. Jones, Combustion Science: Principles and Practice, Millenium Books, New-
hazards, Technical Report AFAPL-TR-71-7, Air Force Aero Propulsion Labora- town, Australia, 1993.
tory, Wright-Patterson Air Force Base, Ohio, 1971. [40] J.A. Carvalho, M.Q. McQuay, Principles of Applied Combustion, UNESP, São
[9] F. Van den Schoor, F. Verplaetsen, The upper explosion limit of lower alkanes Paulo, Brazil, 2007 (in Portuguese).
and alkenes in air at elevated pressures and temperatures, J. Hazard. Mater. [41] D. Drysdale, An Introduction to Fire Dynamics, 3rd ed., John Wiley & Sons,
128 (2006) 1–9. University of Edinburgh, UK, 2011, ISBN 978-0-470-31903-1.
[10] Physical & thermodynamic properties (methanol, ethanol, propanol, and [42] S. Kondo, K. Takizawa, A. Takahashi, K. Tokuhashi, On the temperature depend-
butanol), Chem. Eng. J. 1 (June) (1976) 10. ence of flammability limits of gases, J. Hazard. Mater. 187 (2011) 585–590.
[11] S.Y. Liao, Q. Cheng, D.M. Jiang, J. Gao, Experimental study of flammability limits [43] J.E. Shepherd, J.C. Krok, J.J. Lee, Jet A Explosion Experiments: Laboratory Testing,
of natural gas–air mixture, J. Hazard. Mater. B 119 (2005) 81–84. Report FM97-5, Explosion Dynamics Laboratory, 1997.
[12] C.J.R. Coronado, Technical-economical analysis of an 100-kg/hr biomass gasifier [44] J. Arnaldos, J. Casal, E. Planas-Cuchi, Prediction of flammability limits at reduced
for powering of an internal combustion engine, Ph.D. Thesis, in Portuguese, pressures, Chem. Eng. Sci. 56 (2001) 3829–3843.
São Paulo State University, Campus of Guaratinguetá, Post Graduate Program [45] D.R. Stull, Fundamentals of fire and explosion, A.I.Ch.E. Monograph Series No.
in Mechanical Engineering, 195 pp., 2007. 10, New York, 1977.
[13] D. Naegeli, W.D. Weatherford Jr., Practical ignition limits for low molecular [46] S.M. Summer, Technical Note Limiting Oxygen Concentration Required to
weight alcohols, Fuel 68 (1989) 45–48. Inert Jet Fuel Vapors Existing at Reduced Fuel Tank Pressures, Report No.
[14] Y. Shebeko, W. Fan, I.A. Bolodian, V.Y. Navzenya, An analytical evaluation of DOT/FAA/ARTN-02/79, U.S. Department of Transportation, Federal Aviation
flammability limits of gaseous mixtures of combustible–oxidizer–diluent, Fire Administration, Office of Aviation Research, Washington, DC, 2003.
Saf. J. 37 (2002) 549–568. [47] J. Chomiak, J. Jarosinski, Flame quenching by turbulence, Combust. Flame 48
[15] E. Brandes, M. Mitu, D. Pawel, The lower explosion point: a good measure for (1982) 241–249.
explosion prevention: experiment and calculation for pure compounds and [48] A.S. Sokolik, W.P. Karpov, E.S. Semenov, Phys. Comb. Expl. 3 (61) (1967).
some mixtures, J. Loss Prevent. Process Ind. 20 (2007) 536–540. [49] W.P. Karpov, E.S. Severin, Phys. Comb. Expl. 14 (33) (1978).
[16] F. Van den Schoor, F. Verplaetsen, J. Berghmans, Calculation of the upper [50] F.C. Lockwood, I.E. Megahed, Combust. Sci. Technol. 19 (1978) 77.
flammability limit of methane/air mixtures at elevated pressures and temper- [51] ASTM E2079, Standard test method for Limiting Oxygen (Oxidant) Concentra-
atures, J. Hazard. Mater. 153 (2008) 1301–1307. tion in Gases and Vapors, American Society for Testing and Materials (ASTM)
[17] F. Van den Schoor, R.T.E. Hermanns, J.A. van Oijen, F. Verplaetsen, L.P.H. de Goey, Washington Office, 2010.
Comparison and evaluation of methods for the determination of flammability [52] I.A. Zlochower, M.G. Gregory, M. Green, The limiting oxygen concentration and
limits, applied to methane/hydrogen/air mixtures, J. Hazard. Mater. 150 (2008) flammability limits of gases and gas mixtures, J. Loss Prevent. Process Ind. 22
573–581. (2009) 499–505.
[18] C.C. Chen, T. Wang, H. Liaw, H. Chen, Nitrogen dilution effect on the flammability [53] L.G. Britton, Flammability hazards of lower aliphatic aldehydes at elevated
limits for hydrocarbons, J. Hazard. Mater. 166 (2009) 880–890. pressure and temperature, Process Saf. Prog. 17 (1998) 138–148.
[19] F. Gharagheizi, Prediction of upper flammability limit percent of pure [54] L.G. Britton, Operating atmospheric vent collection headers using methane gas
compounds from their molecular structures, J. Hazard. Mater. 167 (2009) enrichment, Process Saf. Prog. 15 (1996).
507–510. [55] F.T. Bodurtha, Industrial Explosion Prevention and Protection, McGraw-Hill,
[20] Y. Koshiba, T. Takigawa, Y. Matsuoka, H. Ohtani, Explosion characteristics of New York, USA, 1980, pp. 13–16.
flammable organic vapors in nitrous oxide atmosphere, J. Hazard. Mater. 183 [56] K.L. Cashdollar, I.A. Zlochower, G.M. Green, R.A. Thomas, M. Hertzberg, Flamma-
(2010) 746–753. bility of methane, propane and hydrogen gases, J. Loss Prevent. Process Ind. 13
[21] Y.L. Shoshin, L.P.H de Goey, Experimental study of lean flammability limits of (2000) 327–340.
methane/hydrogen/air mixtures in tubes of different diameters, Exp. Therm. [57] M. Hertzberg, K. Cashdollar, A.I. Zlochower, Flammability limit measurements
Fluid Sci. 34 (2010) 373–380. for dusts and gases: ignition energy requirements and pressure dependences,
[22] H. Miao, L. Lu, Z. Huang, Flammability limits of hydrogen-enriched natural gas, in: Twenty-first Symposium (International) on Combustion/The Combustion
Int. J. Hydrogen Energy 36 (2011) 6937–6947. Institute, 1986, pp. 303–313.
[23] C.M. Piqueras, J. García-Serna, M.J. Cocero, Estimation of lower flammability [58] J. Bond, Sources of ignition: flammability characteristics of chemicals and prod-
limits in high-pressure systems. Application to the direct synthesis of hydro- ucts, J. Hazard. Mater. 31 (1992) 190.
gen peroxide using supercritical and near-critical CO2 and air as diluents, J. [59] F. Norman, F. Van den Schoor, F. Verplaetsen, Auto-ignition and upper explosion
Supercrit. Fluids 56 (2011) 33–40. limit of rich propane–air mixtures at elevated pressures, J. Hazard. Mater. 137
[24] J.R. Rowley, R.L. Rowley, W.V. Wilding, Estimation of the lower flammability (2006) 666–671.
limit of organic compounds as a function of temperature, J. Hazard. Mater. 186 [60] R.K. Eckhoff, M. Ngo, W. Olsen, On the minimum ignition energy (MIE) for
(2011) 551–557. propane/air, J. Hazard. Mater. 175 (2010) 293–297.
[25] S. Kondo, A. Takahashi, K. Takizawa, K. Tokuhashi, On The pressure dependence [61] S. Kondo, A. Takahashi, K. Tokuhashi, Calculation of minimum ignition energy
of flammability limits of CH2 CFCF3 , CH2 F2 and methane, Fire Saf. J. 46 (2011) of premixed gases, J. Hazard. Mater. 103 (2003) 11–23.
289–293. [62] M. Caron, M. Goethals, G. De Smedt, J. Berghmans, S. Vliegen, E. Van’t Oost,
[26] I. Glassman, Combustion, 2nd ed., Academic Press, San Diego, CA, 1996. A. van den Aarssen, Pressure dependence of the auto-ignition temperature of
[27] A.A. Tareq, Flammability characteristics of pure hydrocarbons, Chem. Eng. Sci. methane/air mixtures, J. Hazard. Mater. 65 (1999) 233–244.
58 (2003) 3629–3641. [63] V. Tihay, P. Gillard, D. Blanc, Ignition study of acetone/air mixtures by using
[28] ASTM E681, Standard Test Method for Concentration Limits of Flammability laser-induced spark, J. Hazard. Mater. 209–210 (2012) 372–378.
of Chemicals (Vapors and Gases), American Society for Testing and Materials [64] U.J. Pfahl, M.C. Ross, J.E. Shepherd, K.O. Pasamehmetoglu, C. Unal, Flamma-
(ASTM), W. Conshohocken, PA, 2010. bility limits, ignition energy, and flame speeds in H2 –CH4 –NH3 –N2 O–O2 –N2
[29] Safety information sheet of the chemical product – anhydrous ethyl alcohol mixtures, Combust. Flame 123 (2000) 140–158.
fuel, Brazilian Oil Inc., PETROBRAS S.A., No. FISPQ: Pb0004 P, 23.01.09. [65] S. Kondo, K. Takizawa, A. Takahashi, K. Tokuashi, A. Sekiya, A study on flamma-
[30] M. Vidal, W. Wong, W.J. Roger, M.S. Mannan, Evaluation of lower flammability bility limits of fuel mixtures, J. Hazard. Mater. 155 (2008) 440–448.
limits of fuel–air–diluent mixtures using calculated adiabatic flame tempera- [66] S. Kondo, K. Takizawa, A. Takahashi, K. Tokuhashi, A. Sekiya, Flammability limits
tures, J. Hazard. Mater. 130 (2006) 21–27. of isobutane and its mixtures with various gases, J. Hazard. Mater. 148 (2007)
[31] B.S. Fuman Zhao, Experimental measurements and modeling prediction of 640–647.
flammability limits of binary hydrocarbon mixtures, Master of Science Thesis, [67] S. Kondo, K. Takizawa, A. Takahashi, K. Tokuhashi, Extended Le Chatelier’s for-
University of Tianjin, PR China, 108 pp., 2008. mula for carbon dioxide dilution effect on flammability limits, J. Hazard. Mater.
[32] V. Schroder, M. Molnarne, Flammability of gas mixtures. Part 1. Fire potential, A 138 (2006) 1–8.
J. Hazard. Mater. A 121 (2005) 37–44. [68] H.P. Von Niepenberg, R. Oppenberg, F. Rick, Zusammenstellung Sicher-
[33] E. Brandes, E.A. Ural, Towards a global standard for flammability determination, heitstechnischer Daten zur Vermeidung van Explosionen in Feurräumen,
in: Loss Preventions Symposium, American Institute of Chemical Engineers, Industriefeurung-Essen 11 (1978) 17–31.
Spring National Meeting, New Orleans, 2008. [69] J.E. Hustad, K.O. Sonju, Experimental studies of lower flammability limits of
[34] E.A. Ural, R.G. Zalosh, A mathematical model for lean hydrogen air–steam mix- gases and mixtures of gases at elevated temperatures, Combust. Flame 71
ture combustion in closed vessels, in: Twentieth Symposium (International) on (1988) 283–294.
Combustion, The Combustion Institute, 1984, pp. 1727–1734. [70] CRC, Handbook of Aviation Fuel Properties, CRC Report 530, Society of Auto-
[35] O. Fuss, M. Molnarne, A. Schonbucher, V. Schroder, Sauertoffgrenzkonzen- motive Engineers, Warrendale, PA, 1983.
tration von brennbaren Gasen and Dampfen, Chem. Ing. Tech. 75 (2003) [71] Fuel characterization test, Minimum energy for ignition, Wright Patterson Air
101–104. Force Base, Dayton OH, Classified Air Force report, Attachment D, 1992.
54 C.J.R. Coronado et al. / Journal of Hazardous Materials 241–242 (2012) 32–54

[72] E.M. Goodger, Liquid fuels for transport, Prog. Energy Combust. Sci. 8 (1982) [81] R.H. Perry, D.W. Green, Perry’s Chemical Engineers Handbook, 7th ed., McGraw-
233–260. Hill, NY, 1997.
[73] H.T. Mayfield, JP-8 Composition and Variability, USAF Report AL/EQ-TR-1996- [82] M.R. Brooks, D. Crowl, Vapor flammability above aqueous solutions of
0006 (AD-A317177), May 1996. flammable liquids, J. Loss Prevent. Process Ind. 20 (2007) 477–485.
[74] W.G. Dukek, Aviation and other gas turbine fuels, in: M. Howe-Grant (Ed.), Kirk- [83] G.R. Astbury, A review of the properties and hazards of some alternative fuels,
Othmer Encyclopedia of Chemical Technology, vol. 3, John Wiley & Sons, 1992, Process Saf. Environ. Prot. 86 (2008) 397–414.
pp. 788–812. [84] T. Ma, A thermal theory for estimating the flammability limits of a mixture, Fire
[75] W. Seaton, Group contribution method for predicting the lower and the upper Saf. J. 46 (2011) 558–567.
flammability limits of vapors in air, J. Hazard. Mater. 27 (1991) 169–185. [85] G.A. Melhem, A detailed method for estimating mixture flammability limits
[76] R. Mc Cormick, R. Parish, Advanced Petroleum Based Fuels Program and Renew- using chemical equilibrium, Process Saf. Prog. 16 (1997) 203–218.
able Diesel Program Milestone Report: Technical Barriers to the Use of Ethanol [86] A. Hansen, Q. Zhang, P. Lyne, Ethanol–diesel fuel blends, a review, Bioresour.
in Diesel Fuel, November 2001, NREL/MP-540-32674. Technol. 96 (2005) 277–285.
[77] Growmark, Inc., Flammability Limits for Ethanol/Diesel Blends Final Report [87] L. Pidol, B. Lecointe, L. Starck, N. Jeuland, Ethanol–biodiesel–diesel fuel blends:
from Battelle Agreement No. CP037328R, September 1998. performances and emissions in conventional diesel and advanced low temper-
[78] Battelle, Flammability limits for ethanol/diesel blends, Final Report prepared ature combustions, Fuel 93 (March) (2012) 329–338.
by Battelle, Columbus, OH, 1998. [88] D.C. Rakopoulos, C.D. Rakopoulos, R.G. Papagiannakis, D.C. Kyritsis, Combustion
[79] R. Waterland, S. Venkatesh, S. Unnasch, Safety and performance assessment heat release analysis of ethanol or n-butanol diesel fuel blends in heavy-duty
of ethanol–diesel blends, National Renewable Energy Laboratory, September DI diesel engine, Fuel 90 (2011) 1855–1867.
2003, NREL/SR-540-34817. [89] S. Kumar, N. Singh, R. Prasad, Anhydrous ethanol: a renewable source of energy,
[80] M.R. Brooks, D. Crowl, Flammability envelopes for methanol, ethanol, acetoni- Renew. Sustain. Energy Rev. 14 (2010) 1830–1844.
trile and toluene, J. Loss Prevent. Process Ind. 20 (2007) 144–150.

Vous aimerez peut-être aussi