You are on page 1of 9

JID: PROCI

ARTICLE IN PRESS [m;July 21, 2016;17:11]

Available online at www.sciencedirect.com

Proceedings of the Combustion Institute 000 (2016) 1–9


www.elsevier.com/locate/proci

Ignition delay times of Jet A-1 fuel: Measurements in a


high-pressure shock tube and a rapid compression
machine
A.R. De Toni a,∗, M. Werler b, R.M. Hartmann a,b, L.R. Cancino a,
R. Schießl b, M. Fikri c, C. Schulz c, A.A.M. Oliveira a, E.J. Oliveira d,
M.I. Rocha d
a Combustion and Thermal Systems Engineering Laboratory – LABCET/UFSC, Mechanical Engineering Department,
Federal University of Santa Catarina, Campus Florianópolis – Trindade, Florianópolis, SC 88.040-900, Brazil
b Institute of Technical Thermodynamics – Karlsruhe Institute of Technology, KIT, Karlsruhe, Germany
c IVG, Institute of Combustion and Gas Dynamics – Reactive Fluids, University of Duisburg-Essen, Duisburg, Germany
d Petróleo Brasileiro S.A. – PETROBRAS, Brazil

Received 3 December 2015; accepted 6 July 2016


Available online xxx

Abstract

Ignition delay time (IDT) measurements for Jet A-1 fuel samples have been performed with a rapid com-
pression machine (RCM) and a high-pressure shock tube (ST). The IDT measurements span a pressure range
from 7 to 30 bar, a temperature range from 670 K to 1200 K, and fuel/air equivalence ratios φ from 0.3 to 1.3.
Expressions fitting the experimental data sets were obtained, with fitting parameters being provided. The
combined RCM/ST data aimed at providing information on the two-stage ignition behavior and on the tran-
sition from NTC chemistry to high-temperature radical chain-branching, which are important and hard to
meet targets in the development of chemical surrogates.
© 2016 by The Combustion Institute. Published by Elsevier Inc.

Keywords: Jet fuel; Rapid compression machine; Shock tube; Ignition delay time; Fuel surrogate

1. Introduction of the resistance of gasoline to autoignition, while


the latter gauges ignition propensity of Diesel fuel.
Autoignition properties of liquid fuels are However, none of these ratings is included in jet
accounted for in terms of octane (ON) and cetane fuel (kerosene) specifications, even though there is
(CN) ratings, with the former being an assessment evidence for a correlation between fuel’s derived
cetane number (DCN) and the occurrence of lean

blow-off [1–3].
Corresponding author. Measurements of autoignition in shock tubes
E-mail address: detoni@labcet.ufsc.br (A.R. De (ST) and rapid compression machines (RCM)
Toni).

http://dx.doi.org/10.1016/j.proci.2016.07.024
1540-7489 © 2016 by The Combustion Institute. Published by Elsevier Inc.

Please cite this article as: A.R. De Toni et al., Ignition delay times of Jet A-1 fuel: Measurements in a high-
pressure shock tube and a rapid compression machine, Proceedings of the Combustion Institute (2016),
http://dx.doi.org/10.1016/j.proci.2016.07.024
JID: PROCI
ARTICLE IN PRESS [m;July 21, 2016;17:11]

2 A.R. De Toni et al. / Proceedings of the Combustion Institute 000 (2016) 1–9

provide insights into the effects of fuel composi- These measurements will be used as one of the
tion, equivalence ratio, pressure, and temperature validation targets for the development of chemical
on autoignition-related phenomena. The assess- surrogates for the Jet A-1 fuel.
ment of these effects is particularly important
when new jet fuel mixtures are proposed, es-
pecially, those containing biofuel components. 2. Jet fuel and surrogates
Dooley et al. [4,5] and Hui et al. [6] reported mea-
surements of ignition delay time for alternative The kerosene obtained from fractional distil-
jet fuels made from hydro-processed esters and lation of crude oil has a typical carbon number
fatty acids (HEFA) based on vegetable oils and distribution between 8 and 16 mostly formed by
animal fats, Fischer–Tropsch fuels made from coal paraffinic, naphthenic (cycloalkane), and aromatic
and natural gas (GtL), and surrogate mixtures hydrocarbons with boiling point in the 200–300 °C
for these alternative fuels. Their results indicated range. Wood et al. [15] developed one of the earlier
that all alternative fuels presented higher reactivity surrogate mixtures, a 14-components blend, in an
than Jet A, with GtL and camelina HEFA being attempt to match JP-4 and JP-5 boiling range,
consistently the more reactive, reinforcing the need atomization quality and combustion behavior.
for the assessment of combustion properties of Since then, surrogate fuels evolved to include from
new fuels, in relation to established fuels, before simpler mixtures of hydrocarbons, adequate to
the new mixtures are widely used. computational fluid dynamic (CFD) studies, to
The comparison of the behavior of the igni- more elaborate blends designed to emulate chem-
tion delay time of a complex fuel with that of ical, physical, or combustion properties in detailed
pure species provides general guidelines for the chemical kinetics and experimental investigations.
development of chemical surrogates amenable to A surrogate fuel is usually defined in terms of
theoretical studies using detailed chemical kinetics the number of species in its composition and its be-
mechanisms. The comparisons are needed in a large havior relative to ‘targets’. Farrell et al. [16] defines
range of temperatures and pressures [2,6], usually three main types of targets, in order of increasing
achieved by a combination of measurements in complexity: property targets, development targets,
shock tubes (ST) and rapid compression machines and application targets. A property target refers
(RCM). However, several autoignition studies for to chemical and physical fuel properties, like the
jet fuels [6–13] showed that even when using the hydrogen-to-carbon ratio and the distillation curve.
same fuel sample the agreement between different Development targets are the chemical kinetics and
sets of data is limited. While measurements of thermophysical behavior typically evaluated in
Vasu et al. [7,8] using ST and Dooley et al. [9] using experiments under well-controlled conditions,
ST and RCM for Jet A (POSF 4658) at 650–1250 K such as ignition delay time (IDT) measurements
and 20 bar presented good agreement, the RCM in shock tubes and rapid compression machines,
results of Kumar and Sung [3,10] at 650–770 K species evolution profiles in flow reactors, or
and 15 bar using the same facility of Dooley et al. multicomponent spray vaporization. Application
[9], compare well with previous ST measurements targets are based on measurements obtained from
at 15 bar but do not present the behavior expected engine tests in transient and steady-state operation,
from the previous RCM measurements obtained such as combustion efficiency, and gaseous and
at 7 bar. The experiments at lower temperature particulate emissions.
require care in properly vaporizing the fuel since Regarding the availability of validated oxi-
kerosene contains species with relatively low vapor dation mechanisms for relevant hydrocarbons,
pressure. Also, cracking of the fuel molecules may Colket et al. [17], Farrell et al. [16], and Pitz et
occur at higher preheating temperatures, an effect al. [18] pointed out that, as late as 2006, only
whose extent is hard to evaluate. While it is possible linear alkanes (n-heptane, n-decane, n-dodecane,
to work with the liquid fuel injected into the test n-hexadecane) and one branched alkane (iso-
chamber of the ST and the RCM [11,14,15], in octane) received sufficient attention to enable
most facilities the fuel and oxidant are premixed in reliable modeling of their oxidation, while mech-
the gas phase before feeding the reactor. anisms for naphthenic and aromatic compounds
Here, we present measurements of ignition lacked validation. Further studies [5,19] have
delay time of a Jet A-1 fuel produced in Brazil in indicated that jet fuel is better approximated
the 660–1250 K temperature range at pressures of by alkanes with one or two branches, instead
7, 15, and 30 bar and at equivalence ratios from 0.3 of highly-branched molecules like iso-cetane
to 1.3, using the high-pressure shock tube at IVG (2,2,4,4,6,8,8-heptamethylnonane). However,
and the rapid compression machine at KIT. The highly-branched alkanes and linear ones are much
measurements aimed at detecting the extent of the more affordable to formulate fuel mixtures for
NTC behavior and the occurrence of two-stage large-scale experiments. Therefore, since the earlier
ignition to aid in the selection of an appropriate studies with surrogate fuels, it has become common
surrogate mixture to emulate ignition delay time. practice to select one or more ‘reference fuels’ from

Please cite this article as: A.R. De Toni et al., Ignition delay times of Jet A-1 fuel: Measurements in a high-
pressure shock tube and a rapid compression machine, Proceedings of the Combustion Institute (2016),
http://dx.doi.org/10.1016/j.proci.2016.07.024
JID: PROCI
ARTICLE IN PRESS [m;July 21, 2016;17:11]

A.R. De Toni et al. / Proceedings of the Combustion Institute 000 (2016) 1–9 3

each hydrocarbon class and adjust its proportions real fuel since both MURI surrogates were unable
to match the desired real-fuel properties. For to reproduce the real fuel’s smooth transition from
instance, the six-component Utah/Yale surrogate NTC chemistry to high-temperature radical chain-
[20] (30% n-dodecane, 20% n-tetradecane, 10% branching mechanism. The authors argue that
iso-cetane, 20% methylcyclohexane, 15% o-xylene, none of the a priori evaluated properties (DCN,
5% tetralin by vol.) has two aromatic species to H/C ratio, molecular weight, and TSI) seem to
match jet fuel’s volatility and smoke point; the correlate with the occurrence of this transition.
Drexel/S5 surrogate [20] (26% n-dodecane, 36% In order to select a surrogate mixture for the
iso-cetane, 14% methylcyclohexane, 6% decalin, Jet A-1 fuel of interest here, this work focus on the
18% 1-methylnaphthalene by vol.) was formu- NTC behavior and the occurrence of two-stage ig-
lated to mimic the low-temperature reactivity of nition as detected by ignition delay measurements.
JP-8; and the binary Aachen surrogate [21] (80% Therefore, shock tube and RCM measurements are
n-decane, 20% 1,2,4-trimethylbenzene by weight) combined to span the low to high temperature igni-
focuses on the overall H/C ratio and autoignition tion delay along a few pressures and stoichiometry.
behavior, but has also similar heat-of-combustion
and molecular weight.
The Multi University Research Initiative
3. Materials and methods
(MURI) [4,5,9,19,22–24] extensively tested real
fuels in shock tubes, rapid compression machines,
3.1. High pressure shock tube
flow reactors, counterflow and diffusion flames,
outlined new concepts on the nature of real fuel
The measurements were carried out in a heated
oxidation pathways and proposed approaches to
(125 °C) high-pressure shock tube in the Institute
the formulation of surrogates. One of the impor-
for Combustion and Gas Dynamics (IVG) at the
tant takeaways from these works is the ‘distinct
University of Duisburg-Essen that can be oper-
chemical functionalities’ concept that defines that
ated up to post-combustion pressures of 500 bar.
a surrogate fuel must be able to produce a radical
This facility and the experimental procedures are
pool similar to that of the real fuel. Therefore,
described elsewhere [27–30]. Gas mixtures were
the chemical class composition of surrogate mix-
prepared by injecting liquid Jet A-1 into an evacu-
tures can significantly differ from that of the real
ated, heated (125 °C) stainless-steel mixing vessel.
fuel while the combustion properties of the fully
The total amount of fuel and air were controlled
prevaporized fuel remain remarkably similar. This
manometrically to ensure the desired equivalence
concept was associated with an a priori property
ratio. The mixture was stirred for 120 min to ensure
matching method involving four property tar-
homogeneity before it was filled in the test section
gets: derived cetane number (DCN), H/C ratio,
of the shock tube. The shock speed was mea-
average molecular weight, and threshold sooting
sured over two intervals using three piezo-electric
index (TSI). Based on these principles and exper-
pressure gauges to determine the post-reflected
imental data of the average Jet A sample (POSF
shock temperatures. The estimated uncertainty in
4658), two surrogates mixtures were proposed:
reflected shock temperature is less than ± 25 K.
three-component MURI 1 (42.67% n-decane,
Pressure data were recorded with a time resolution
33.02% iso-octane, 24.31% toluene by mol) and
of 0.1 μs. The ignition delay times were determined
four-component MURI 2 (40.4% n-dodecane,
by extrapolating the steepest increase of the CH∗
29.5% iso-octane, 22.8% n-propylbenzene, 7.3%
chemiluminescence emission (at 431.5 nm) signal
1,3,5-trimethylbenzene by mol) with an associ-
to its zero level on the time axis as shown in Fig. 1.
ated comprehensive reaction mechanism for the
To also determine the delay times of the
former. The overall agreement between the real
first-stage ignition (cool flame), CH2 O∗ chemi-
fuel and both surrogates was significant but some
luminescence was simultaneously detected in the
issues remained. The three-component MURI 1
400–450 nm wavelength range. The experiments
presented a mismatch in average molecular weight
were carried out with synthetic air containing
thus affecting the evaluation of targets that depend
79.5% N2 and 20.5% O2 . As driver gas, tailored
on transport properties, e.g., extinction strain
mixtures of 5–20% Ar in He were used to adjust its
rates. This issue was amended in the formulation
acoustic properties according to [31,32] to ensure
of the MURI 2, which comprises heavier linear
long test times.
alkane and aromatics. As of 2015 the only reaction
mechanisms that include the whole palette of rele-
vant hydrocarbons, i.e., the four species considered 3.2. Rapid compression machine
in MURI 2 surrogate and also naphthenics are
the computationally-generated schemes from the Details of the Rapid Compression Machine fa-
Milano group [25,26]. One of these schemes was cility at the Institute of Technical Thermodynam-
employed by Dooley et al. [24] to study the role ics (ITT) at Karlsruhe Institute of Technology have
of cycloalkanes in the combustion behavior of the been described elsewhere [33]. The fuel-mixture is

Please cite this article as: A.R. De Toni et al., Ignition delay times of Jet A-1 fuel: Measurements in a high-
pressure shock tube and a rapid compression machine, Proceedings of the Combustion Institute (2016),
http://dx.doi.org/10.1016/j.proci.2016.07.024
JID: PROCI
ARTICLE IN PRESS [m;July 21, 2016;17:11]

4 A.R. De Toni et al. / Proceedings of the Combustion Institute 000 (2016) 1–9

Fig. 1. Determination of the ignition delay time from the CH∗ chemiluminescence signal trace in the ST experiments. Jet
A-1 in air, p5 = 16.3 bar, T5 = 704.3 K, φ = 1.0, τ 1 = 6829 μs, τ = 8176 μs.

created in a mixing vessel heated at 140 °C. Homo- slight gradual increase in pressure (dp/dt = 3%/ms)
geneity of the fuel mixture is ensured by constantly is recorded. Using the maximum pressure increase,
mixing with a magnetically coupled stirrer for at we calculate the resulting variation in temperature
least 120 minutes. The mixture is then filled into the relative to the post-shock temperature T5 based on
combustion chamber up to some pre-defined pres- an adiabatic compression model. The maximum
sure and temperature. An oil bath surrounding the temperature increase was found to be 30 K for our
machine ensures well-defined thermal conditions at measurement conditions (see the column ±T5
the combustion chamber before compression. in Table S1). The determined T5 uncertainties
The compression is achieved by a pneumatic [34] imply an error between ± 11 and ± 16% in the
driver that pushes the piston into the combustion calculated ignition delay times, to be taken into
chamber. A knee-lever mechanism is used to pre- account in chemical kinetics modeling.
vent further piston motion at top dead center po- The measurements are curve-fitted to an equa-
sition. Thus, after compression, constant-volume tion of the form τ = A exp(B/T) p–x φ –y , where
conditions prevail. The RCM allows an adjustment x and y are the pressure and stoichiometry ex-
of the geometric compression ratio to provide a ponents, respectively. Multiple linear regression
wide temperature range. Each condition was mea- analysis using ln(τ ) as the dependent variable and
sured three times and the repeatability was found to (1/T), ln(p) and ln(φ) as independent variables
be very good throughout the whole range of mea- identified the values of x = 0.89 and y = 0.79 using
surement conditions, as detailed below. the data for φ =1.0 and 0.3 listed in Table S1.
For the temperature-and pressure dependence, the
expression
4. Results and discussion  
14510 ± 490
τ /μs = 102.06±0.22 exp
Following Vasu et al. [7], the chemical composi- T/K
tion of Jet A-1 fuel was assumed as C11 H21 with a × ( p/bar )−0.89±0.12 (φ)−0.79±0.06 (1)
density of 0.81 g/cm3 , resulting in a stoichiometric
air fuel ratio of 16.25. Table 1 presents the condi- was determined from a fit for the high-temperature
tions of the measurements in the ST and the RCM. range (925–1195 K).
Numerical simulations of ignition delay times in
4.1. ST ignition delay time measurements and CHEMKIN-Pro [35], using the three-component
simulations mechanism of Dooley et al. [9] which comprises
6633 reactions among 1599 species were per-
The ignition delay times evaluated from the formed. Dooley et al. acknowledge that a detailed
CH∗ emission are listed in the Supplementary assessment and validation of the mechanism was
material, along with the respective pressures p not carried out and that the model should closely
and temperatures T, for stoichiometric and lean reproduce their flow reactor data while predicting
Jet A-1/air mixtures. During the ignition delay, a ignition delays up to a-factor-of-two slower.

Please cite this article as: A.R. De Toni et al., Ignition delay times of Jet A-1 fuel: Measurements in a high-
pressure shock tube and a rapid compression machine, Proceedings of the Combustion Institute (2016),
http://dx.doi.org/10.1016/j.proci.2016.07.024
JID: PROCI
ARTICLE IN PRESS [m;July 21, 2016;17:11]

A.R. De Toni et al. / Proceedings of the Combustion Institute 000 (2016) 1–9 5

Table 1
Investigated conditions (fuel/air equivalence ratio, effective temperature, pressure) for the
ST and RCM measurements.
Equivalence ratio, φ ST RCM
0.3 15 bar, 1000 K/T = 0.83–1.2 15 bar, 1000 K/T = 1.1–1.5
30 bar, 1000 K/T = 0.83–1.25 –
0.7 – 7 bar, 1000 K/T = 1.1–1.5
– 15 bar, 1000 K/T = 1.1–1.5
1.0 –
15 bar, 1000 K/T = 0.8-1.4 7 bar, 1000 K/T = 1.1–1.5
30 bar, 1000 K/T = 0.83-1.11 15 bar, 1000 K/T= 1.45–1.55-

1.3 – 7 bar, 1000 K/T = 1.1–1.5

Table 2
Composition of proposed surrogates for Jet A-1 ignition delay simulations and related a
priori targets.
Surrogate A Surrogate B
n-decane (% mol) 33.5 65.4
iso-octane (% mol) 36.5 14.0
toluene (% mol) 30.0 20.6
Derived cetane number (by regression) 42.0 56.5
CH2 /CH3 x [CH2 + CH3 ] mass fraction 1.304 2.462

Two surrogate mixtures were proposed, based presented. As mentioned before, an overestimation
on composition data of the Jet A-1 sample (ASTM of IDT for both surrogates is expected since this
D1319: 79.4% paraffins, 0.8% olefins, 19.8% aro- chemical kinetics model consistently overestimated
matics by vol.) and regarding one of the MURI IDT of the original experiments of Dooley et al.
surrogate targets (ASTM D6890 – DCN 41.9). The [9], in which the a priori target-matching criteria
surrogate ‘A’ (33.5% n-decane, 36.5% iso-octane, was carried out for all four targets. The computed
30.0% toluene by mol) was based on regression IDTs of surrogate ‘A’ presented a significant over-
analysis of derived cetane number, as employed estimation, indicating that a DCN matching us-
by Dooley et al. [9]. Since several mixtures of ing the strategy of adjusted volume of aromatics
these three hydrocarbons can produce the same is not enough to characterize the ignition behavior.
DCN, the mixture chosen as surrogate ‘A’ was Among the likely causes for this lack of agreement
the one with similar volume fraction of aromatics is the limited information provided by the ASTM
(30.0% toluene by mol i.e. 20.5% by vol.). The D1319 composition evaluation which does not
surrogate mixture ‘B’ was formulated consider- distinguish aliphatics (linear and branched) from
ing the experimental evidence [36–38] that linear naphthenics nor does it provide data regarding the
and weakly-branched alkanes are the statistical substituted or unsubstituted nature of the aromat-
‘mode’ of hydrocarbon isomers present in liquid ics. Such information would be helpful to use an-
fuels and this preponderance can be related to a other MURI target, i.e., the overall H/C ratio.
methylene-to-methyl (CH2 /CH3 ) ratio that glob- On the other hand, surrogate ‘B’ computations
ally correlates with the low-temperature reactivity. predicted IDTs very close to the experimental
Therefore, surrogate ‘B’ (65.4% n-decane, 14.0% values for low-to-intermediate temperatures (700–
iso-octane, 20.6% toluene by mol) was formu- 900 K) at 15 bar and stoichiometric mixtures. The
lated regarding the ASTM D1319 composition agreement between experiment and computations
and adding this molecular group composition for other conditions was qualitative, with lean mix-
CH2 /CH3 x [CH2 +CH3 ] metric [19] while relaxing tures having longer IDTs than stoichiometric ones
the DCN matching criteria. Table 2 presents the and an increase in pressure resulting in shorter
composition and target values for the proposed ignition delays. The faster ignition of surrogate ‘B’
surrogates while Fig. 2 shows the comparison compared to ‘A’ is due to the higher fraction of
between experimental and numerical results of n-decane that offers sixteen secondary hydrogen
surrogate ‘B’. sites (methylene) where molecular oxygen can
The numerical IDT prediction based on the promote an H-abstraction to form radicals. The
composition of surrogate ‘A’ is about 60% higher other alkane in the formulation, iso-octane, has
than the IDT values predicted by surrogate ‘B’ lead- only two such sites and they are shielded against O2
ing to a larger deviation compared to the experi- attack by the methyl substituents in both adjacent
ments therefore only modeling with surrogate B is carbons.

Please cite this article as: A.R. De Toni et al., Ignition delay times of Jet A-1 fuel: Measurements in a high-
pressure shock tube and a rapid compression machine, Proceedings of the Combustion Institute (2016),
http://dx.doi.org/10.1016/j.proci.2016.07.024
JID: PROCI
ARTICLE IN PRESS [m;July 21, 2016;17:11]

6 A.R. De Toni et al. / Proceedings of the Combustion Institute 000 (2016) 1–9

Fig. 2. Comparison of measured (shock tube) and simulated ignition delay time for Jet A-1/air mixtures (detailed kinetics
model for Jet A from Dooley et al. [9]); surrogate ‘B’ was used for the numerical calculations.

In the low-to-intermediate temperatures, sur- time below a few milliseconds) to reasonably


rogate ‘B’ has enough secondary hydrogen sites assign a pressure and temperature to the ignition
to present reactivity similar to the real fuel. event. The data points for first-stage ignition
However, in the transition from the NTC region above 790 K should therefore be considered as
to high temperature, n-decane is mostly gone semi-quantitative. This somewhat fuzzy nature of
while toluene, after an initial H-abstraction at the the points with ignition delay times below ∼5 ms
methyl substituent, forms a resonance-stabilized becomes evident also by the larger scatter in first-
benzylic radical. Therefore, the surrogate ability stage IDT at the nominally-identical measurement
to reproduce the real fuel’s intermediate-to-high points.
temperature reactivity is limited by the absence of First-stage and main-ignition delay times be-
sufficient methylene units in the compounds other come increasingly close as temperature decreases.
than n-decane. Dooley et al. also argue that the Below about 660 K at 7 bar, they practically co-
model shows an elongated NTC behavior, which incide. The temperature where the two begin to
is likely due to an under-emphasis of beta-scission “overlap” is lower for the lean mixture (φ = 0.7)
reactions of the alkyl radicals thus leading to and higher for the rich mixture (φ = 1.3). At 7 bar,
increased low temperature chemistry processes φ = 0.3, no auto-ignition was detected in the RCM
instead of high temperature chain branching. This within the studied temperature range. At a pres-
behavior is also noticeable in our simulations. sure of 15 bar, however, auto-ignition could be de-
tected also for φ = 0.3, and again, two-stage igni-
4.2. RCM ignition delay time measurements tion was observed, as shown in Fig. 4. At 15 bar,
ignition was one order of magnitude faster than at
Figure 3 presents the ignition delay times mea- 7 bar, for both φ = 0.7 and φ = 1.0. Comparison of
sured in the RCM for the Jet A-1 sample at 7 bar Figs. 2 and 4 reveals that RCM and shock tube IDT
for equivalence ratios ranging from slightly lean match well in the temperature and pressure region
(φ = 0.7) to rich (φ = 1.3). The temperatures shown where they overlap i.e. 15 bar, equivalence ratio
in the diagram are effective temperatures represent- 1.0, 1000 K/T=1.4; thus the often described phe-
ing an average of the pre-ignition temperature in nomenon of non-matching ST and RCM ignition
the RCM, as derived from the measured pressure delay times does not appear in our case, see Fig. S1.
curve employing an adiabatic core hypothesis. Within the NTC region, a pronounced pres-
The experimental data and non-reactive pres- sure dependence of the IDT was found, which is
sure histories are available in the Supplementary demonstrated in Fig. 5 by IDTs for φ = 0.7 as a
material. function of temperature. These measurements near
A clear two-stage ignition behavior was ob- 7.2 and 8.1 bar highlight this strong dependence.
served in the range below 760 K. Above 790 K, In the RCM the post-compression pressure at top
the first stage ignition became too fast (delay dead center displays slight shot-to-shot variations

Please cite this article as: A.R. De Toni et al., Ignition delay times of Jet A-1 fuel: Measurements in a high-
pressure shock tube and a rapid compression machine, Proceedings of the Combustion Institute (2016),
http://dx.doi.org/10.1016/j.proci.2016.07.024
JID: PROCI
ARTICLE IN PRESS [m;July 21, 2016;17:11]

A.R. De Toni et al. / Proceedings of the Combustion Institute 000 (2016) 1–9 7

Fig. 3. RCM ignition delay times of Jet A-1/air mixtures at 7 bar for various equivalence ratios. Open symbols denote the
first stage ignition, solid symbols denote the main ignition.

Fig. 4. RCM ignition delay times of Jet A-1/air mixtures at 15 bar for various equivalence ratios. Open symbols denote
the first stage ignition (φ = 0.3 only), solid symbols denote the main ignition.

with an amplitude of around 0.1 bar. Because data curve fitted using expressions in the form
points with post-compression pressures from a log(τ ig ) = A + B/T, such as Eq. (1). These expres-
certain interval are assigned to the same nominal sions, however, cannot describe the NTC region
pressure in the IDT diagrams, the strong pressure with a single set of parameters A and B. In this
dependence of IDT creates an apparent scatter in context Zhou et al. [39] recently proposed fitting
the diagrams. expressions using combinations of different mathe-
The regions of the Arrhenius diagram where matical functions that can reproduce complex igni-
the IDT display a linear behavior are usually tion delay curve shapes. Here, we present a compact

Please cite this article as: A.R. De Toni et al., Ignition delay times of Jet A-1 fuel: Measurements in a high-
pressure shock tube and a rapid compression machine, Proceedings of the Combustion Institute (2016),
http://dx.doi.org/10.1016/j.proci.2016.07.024
JID: PROCI
ARTICLE IN PRESS [m;July 21, 2016;17:11]

8 A.R. De Toni et al. / Proceedings of the Combustion Institute 000 (2016) 1–9

Fig. 5. Strong pressure dependence of IDT (main ignition stage) for φ = 0.7, pressure range between 7 and 8.3 bar. The
color of the scatter symbols represents the pressure, as indicated by the color scale. The dashed line is a nonlinear fit
through data points near 7 bar. (For interpretation of the references to color in this figure legend, the reader is referred to
the web version of this article)

formula that can also fit the NTC region, given by (φ=1, 15 bar), are included in the supplementary
  material.
  θ − a5
log τig = a1 + a2 θ + (a3 + a4 θ ) erf (2)
a6
where erf is the error function, θ is the inverse 5. Summary and conclusions
temperature and ai are curve-fitted parameters.
This function has some properties that make it Ignition delay times of Jet A-1/air mixtures were
attractive for fitting ignition delay time curves. measured in a high-pressure shock tube and a rapid
For instance, in the limit of small (large) values compression machine spanning a pressure range
of θ, erf attains the value –1 (+1), and the function from 7 to 30 bar, temperatures of 670–1200 K and
approaches the forms: fuel/air equivalence ratios from 0.3 to 1.3. In the
  700–870 K range, IDTs show distinct NTC be-
log τig ≈ a1 − a3 + (a2 − a4 ) θ , when θ is small,
havior at both facilities (7 bar at RCM and 15 bar
  for ST) while at temperatures outside this range
log τig ≈ a1 + a3 + (a2 + a4 ) θ , when θ is large. the IDTs showed an Arrhenius-like behavior. The
experiments in rapid compression machine show
Both are of the “standard” Arrhenius form
a two-stage ignition behavior for compressed tem-
A + Bθ with high- and low-temperature values for
peratures below 760 K. Expressions fitting the ex-
A and B. Therefore, the function is “asymptotically
perimental data sets were obtained, with fitting pa-
Arrhenius-like” for small and large θ and can
rameters provided in the supplementary material.
represent the normal branches of ignition delay to
Based on limited information about the Jet
the left and right sides of the NTC region. Between
A-1 sample hydrocarbon makeup (ASTM D1319:
these regions, it can fit both negative and positive
79.4% paraffins, 0.8% olefins, 19.8% aromatics by
slopes. A sample fit with this function is included
vol.), computations of IDT, using the mechanism
in Fig. 5, for a pressure of 7 bar. The parameters
of Dooley et al. [9] were conducted for two different
for this fit are
surrogate compositions. The surrogate ‘B’ (65.4%
a1 = -17.484, a2 = 16.169, a3 = 6.5411, n-decane, 14.0% iso-octane, 20.6% toluene by mol),
a4 = -7.8843, a5 = 1.1706, a6 = 0.22398, which was proposed regarding ASTM D1319 com-
position and a low temperature molecular group
using θ = 1000 K/T, and expressing τ ig in millisec- metric [19] presented good agreement for stoichio-
onds. Fitting parameters for the remaining con- metric data at 15 bar, i.e., reproducing the reactivity
ditions, including the matching shock tube data plateau in the low to intermediate temperatures.

Please cite this article as: A.R. De Toni et al., Ignition delay times of Jet A-1 fuel: Measurements in a high-
pressure shock tube and a rapid compression machine, Proceedings of the Combustion Institute (2016),
http://dx.doi.org/10.1016/j.proci.2016.07.024
JID: PROCI
ARTICLE IN PRESS [m;July 21, 2016;17:11]

A.R. De Toni et al. / Proceedings of the Combustion Institute 000 (2016) 1–9 9

Acknowledgments
[19] S.H. Won, S. Dooley, P.S. Veloo, et al., Combust.
The authors gratefully acknowledge the sup- Flame 161 (2014) 826–834.
port for this work from Petróleo Brasileiro – S.A. [20] S. Humer, a. Frassoldati, S. Granata, et al., Proc.
Combust. Inst. 31 (I) (2007) 393–400.
PETROBRAS (Agreement # 0050.0070218.11.9).
[21] S. Honnet, K. Seshadri, U. Niemann, N. Peters, Proc.
The help of Natascha Schlösser in supporting the Combust. Inst. 32 (I) (2009) 485–492.
experiments at IVG is greatly appreciated. [22] S.H. Won, S. Dooley, F.L. Dryer, Y. Ju, Combust.
Flame 159 (2012) 541–551.
Supplementary materials [23] F.L. Dryer, S. Jahangirian, S. Dooley, et al., Energy
Fuels 28 (2014) 3474–3485.
[24] S. Dooley, J. Heyne, S.H. Won, P. Dievart, Y. Ju,
Supplementary material associated with this ar- F.L. Dryer, Energy Fuels 28 (2014) 7649–7661.
ticle can be found, in the online version, at doi: [25] E. Ranzi, A. Frassoldati, A. Stagni, M. Pelucchi,
10.1016/j.proci.2016.07.024. A. Cuoci, T. Faravelli, Int. J. Chem. Kinet. 46 (2014)
512–542.
[26] A. Cuoci, A. Frassoldati, T. Faravelli, E. Ranzi, Proc.
References Combust. Inst. 35 (2014) 1621–1627.
[27] M. Fikri, J. Herzler, R. Starke, C. Schulz, P. Roth,
[1] L.Q. Maurice, H. Lander, T. Edwards, W.E. Harri- G.T. Kalghatgi, Combust. Flame 152 (2008) 276–281.
son, Fuel 80 (2001) 747–756. [28] L.R. Cancino, M. Fikri, A.A.M. Oliveira, C. Schulz,
[2] P. Dagaut, M. Cathonnet, Prog. Energy Combust. Proc. Combust. Inst. 32 (2009) 501–508.
Sci. 32 (2006) 48–92. [29] L.R. Cancino, Development and Application of De-
[3] K. Kumar, C.J. Sung, Combust. Flame 157 (2010) tailed Chemical Kinetics Mechanisms for Ethanol and
676–685. Ethanol-Containing Hydrocarbon Fuels, Federal Uni-
[4] S. Dooley, S.H. Won, J. Heyne, et al., Combust. versity of Santa, Catarina, 2009.
Flame 159 (2012) 1444–1466. [30] L.R. Cancino, M. Fikri, A.A.M. Oliveira, C. Schulz,
[5] S. Dooley, S.H. Won, S. Jahangirian, et al., Combust. in: Proceedings of the 20th International Congress
Flame 159 (2012) 3014–3020. on Mechanical Engineering – COBEM 2009,
[6] X. Hui, K. Kumar, C.J. Sung, T. Edwards, D. Gard- ABCM, Gramado, 2009, pp. 15–20.
ner, Fuel 98 (2012) 176–182. [31] H. Oertel, Stossrohre, Springer Verlag,
[7] S.S. Vasu, D.F. Davidson, R.K. Hanson, Combust. Wien-NewYork, 1966.
Flame 152 (2008) 125–143. [32] H.B. Palmer, B.E. Knox, ARS J. 31 (1961) 826–828.
[8] S.S. Vasu, D.F. Davidson, R.K. Hanson, Shock [33] M. Werler, L.R. Cancino, R. Schiessl, U. Maas,
Waves, Springer, Berlin Heidelberg, Göttingen, 2009, C. Schulz, M. Fikri, Proc. Combust. Inst. 35 (2015)
pp. 293–298. 259–266.
[9] S. Dooley, S.H. Won, M. Chaos, et al., Combust. [34] L.R. Cancino, M. Fikri, A.A.M. Oliveira, C. Schulz,
Flame 157 (2010) 2333–2339. Energy Fuels 24 (2010) 2830–2840.
[10] K. Kumar, C.J. Sung, Fuel 89 (2010) 2853–2863. [35] CHEMKIN-Pro 15083, Reaction Design, (2009) Re-
[11] C. Allen, E. Toulson, T. Edwards, T. Lee, Combust. action Design, San Diego.
Flame 159 (2012) 2780–2788. [36] L.M. Shafer, R.C. Striebich, J. Gomach, T. Edwards,
[12] T. Malewicki, S. Gudiyella, K. Brezinsky, Combust. in: Proceedings of the 14th AIAA/AHI Space Planes
Flame 160 (2013) 17–30. Hypersonic Systems and Technologies Conference,
[13] P. Dagaut, F. Karsenty, G. Dayma, et al., Combust. AIAA, Canberra, 2006.
Flame 161 (2014) 835–847. [37] C.A. Moses, Comparative Evaluation of Semi-
[14] D.F. Davidson, D.R. Haylett, R.K. Hanson, Com- Synthetic Jet Fuels. Contract 33415.02-D, Dayton,
bust. Flame 155 (2008) 108–117. 2008.
[15] C. Wood, V. McDonnell, R. Smith, G. Samuelson, J. [38] F.L. Dryer, Y. Ju, K. Brezinsky, R.J. Santoro, T.A.
Propuls. Power 5 (1989) 399–405. Litzinger, C.J. Sung, Generation of Comprehensive
[16] J.T. Farrell, N.P. Cernansky, F.L. Dryer, et al., in: Surrogate Kinetic Models and Validation Databases
Proceedings of SAE World Congress, SAE Interna- for Simulating Large Molecular Weight Hydrocar-
tional, Detroit, 2007, 2007. bon Fuels. Contract FA9550-07-1-0515, Arlington,
[17] M. Colket, T. Edwards, S. Williams, et al., in: Pro- 2012.
ceedings of the 45th AIAA Aerospace Sciences [39] A. Zhou, T. Dong, B. Akih-Kumgeh, in: Proceedings
Meeting and Exhibit, AIAA, Reno, 2007, pp. 1–21. of the 25th International Colloquim on the Dynam-
[18] W.J. Pitz, N.P. Cernansky, F.L. Dryer, et al., in: ics of Explosions and Reactive Systems, Leeds, 2015.
Proceedings of SAE World Congress, SAE Interna-
tional, Detroit, 2007, 2007.

Please cite this article as: A.R. De Toni et al., Ignition delay times of Jet A-1 fuel: Measurements in a high-
pressure shock tube and a rapid compression machine, Proceedings of the Combustion Institute (2016),
http://dx.doi.org/10.1016/j.proci.2016.07.024

You might also like