You are on page 1of 7

Combustion and Flame 152 (2008) 293–299

www.elsevier.com/locate/combustflame
Brief Communication

Ignition of syngas/air and hydrogen/air mixtures at low


temperatures and high pressures: Experimental data
interpretation and kinetic modeling implications
Frederick L. Dryer ∗ , Marcos Chaos
Department of Mechanical and Aerospace Engineering, Princeton University, Princeton, NJ 08544-5263, USA
Received 6 May 2007; received in revised form 10 August 2007; accepted 17 August 2007
Available online 25 September 2007

1. Introduction ignition delay, determined by various criteria, were


shown to substantially depart from homogeneous gas
An important aspect of minimizing pollutant emis- phase predictions utilizing any of the recently pub-
sions from industrial gas turbines operating on natural lished H2 /CO detailed kinetic models.
gas has been the implementation of lean, premixed Petersen et al. [1] note “a clear and disturbing
combustion. A similar approach in terms of designing disagreement between experiment and model under
turbines for operating on syngas and pure hydrogen conditions of direct interest to power generation gas
has led to a renewed interest in their combustion dy- turbines operating on syngas.” Through their observa-
namics, and experiments have recently been reported tions that the disparity is similar across a wide range
from a variety of venues [1–10] under conditions rel- of experimental venues and in measurements made
evant to industrial turbine mixing systems (i.e., lower by a number of independent investigators, Petersen
temperatures and higher pressures: T < 1000 K, 10 < et al. [1] infer that missing gas phase kinetic paths
P < 30 atm). and/or uncertainties in kinetic/thermochemical para-
In a recent brief communication in this jour- meters are the most likely sources of the disparity.
nal [1], Petersen et al. reported new data on igni- This inference is further supported by the authors’
tion delay for syngas–air mixtures in a high-pressure position that the disagreement between experiment
shock tube and a flow reactor. Experiments were per- and model “may not be surprising, since few data
formed under lean conditions (φ ∼ 0.5) and for pres- have existed at the temperature and pressure ranges of
sure/temperature ranges of 16–29 atm/940–1150 K the present study to which the (previously) published
(shock tube) and 5 atm/760–785 K (flow reactor). models could be calibrated.” Additional insights as to
These new data were summarized in an Arrhenius- the source(s) of the disagreement are stated to be be-
type plot (Fig. 1) along with other recently published yond the scope of their brief communication [1].
data from the rapid compression studies of Walton We are in complete agreement with Petersen et
et al. [6] and from an earlier high-pressure flow re-
al. [1] that the noted departures of the experimental
actor study described in an Electric Power Research
observations of syngas ignition delay measurements
Institute report [8]. For comparison purposes, all of
from homogeneous gas phase kinetic predictions at
the ignition delay data were normalized to conditions
temperatures lower than about 1050 K are real and
of 20 atm pressure [1]. At temperatures lower than
that the reported measurements impose limitations on
about 1050 K, experimental observations of syngas
gas turbine designs for lean premixing of syngas com-
positions with air. This is likely the most important
* Corresponding author. Fax: +1 (609) 258 6109. message that should be taken from [1] and this brief
E-mail address: fldryer@princeton.edu (F.L. Dryer). communication.
0010-2180/$ – see front matter © 2007 The Combustion Institute. Published by Elsevier Inc. All rights reserved.
doi:10.1016/j.combustflame.2007.08.005
294 F.L. Dryer, M. Chaos / Combustion and Flame 152 (2008) 293–299

Fig. 1. Ignition delay times of syngas [1,6,8] and hydrogen mixtures [27]. Conditions: 38.6% H2 + 51.1% CO + 10.3% CO2 + air;
φ = 0.5, 16.5 < P < 28.9 atm [1, shock tube]; φ = 0.5, 11.9 < P < 23 atm [8]; 50% H2 + 50% CO + air, 0.33 < φ < 0.6,
5.0 < P < 5.3 atm [1, flow reactor]; 6.7 < H2 < 13.6% + 4.5 < CO < 9.1% + 16.2 < O2 < 18.7% + 44.1 < N2 < 63.2% in
balance CO2 , 0.3 < φ < 0.7, 12 < P < 23.5 atm [6]; 15% H2 in air, 35 < P < 47 bar [27]. Filled and open circles correspond
to strong and weak ignition events, respectively [27]. All experimental data have been normalized to 20 atm assuming propor-
tionality to P −1 [1]. Lines correspond to ignition delay calculations performed using the mechanism of Li et al. [11] at 20 atm;
the solid line corresponds to the syngas mixture used in the shock tube experiments [1], the dashed line to the conditions of Blu-
menthal et al. [27], and the dash-dotted line to predictions obtained for syngas ignition by catalyzing reactions involved in the
formation and decomposition of H2 O2 (see text). To improve clarity, modeled results are not shown for times greater than 1 s,
although it is noted that predictions can reach values of approximately 1000 s for the lowest temperatures (T ∼ 630 K).

In contrast to the suggestion [1] that kinetic model tions are small in comparison to those produced by
inadequacies may exist since there have been no data the perturbation sources listed above.
to test models in the regime of interest, we argue that The opinions presented herein were first offered,
(1) the principal underlying sources of the dispari- along with supporting materials, at a recent workshop
ties noted for syngas ignition delays predictions in [1] sponsored by the U.S. Department of Energy and the
have been evident in many hydrogen oxidation studies Electric Power Research Institute [12]. The disparities
since the 1960s; (2) the disparities between measured reported in [1] were among several subjects discussed
and predicted syngas ignition delays are mostly a re- that impact development of new industrial gas turbine
sult of the failure of homogeneous gas phase predic- systems for operation on hydrogen. The purpose of
tions to capture perturbations of chemical induction this brief communication is to bring these issues to
processes in the mild ignition regime by one or more the broader audience of readers of [1]. Moreover, we
phenomena. wish to remind the modeling community that ignition
Experiments and analyses of hydrogen–oxygen delay measurements in the mild ignition regime are
ignition delay measurements in the “mild” ignition strongly susceptible to perturbations and that model
regime show high sensitivity to perturbations by the predictions of ignition delays that do not account for
presence of contaminants in the reactants or on ex- these perturbations can be significantly misleading.
perimental surfaces, compressible fluid dynamic ef- Finally, as a result of the fact that these sources of
fects, inhomogeneous mixing, and catalysis from par- perturbations, in any combination, produce within a
ticles or surface materials. In any combination, these factor of 10 or less the same reduction in chemical
perturbations can lead to repeatable reductions of induction times in all of the venues discussed in [1],
the predicted chemical induction time scales in the we show here that a simple zero-dimensional repre-
mild ignition regime by as much as several orders of sentation of their effects as homogeneous catalytic
magnitude. The reactions CO + HO2 = CO2 + OH processes that promote H2 O2 production and decom-
and CO + O + M = CO2 + M also modify syngas position to hydroxyl radicals provides reasonable en-
chemical induction times from those observed for gineering estimates of the perturbed ignition delay
hydrogen–oxygen systems [11], but these modifica- in the mild ignition regime, without modification of
F.L. Dryer, M. Chaos / Combustion and Flame 152 (2008) 293–299 295

homogeneous ignition delay predictions at other con- anism of Li et al. [11]. Conditions for the Blumen-
ditions. A more detailed discussion of the oxidation thal data [27] (φ = 0.42) are similar to those reported
kinetics of syngas mixtures and hydrogen, including for syngas [1] (φ ∼ 0.5). For comparison purposes,
the points summarized in this brief communication, experimental data are normalized under the same as-
appears elsewhere [13]. sumptions used in [1] to 20 atm; modeling predic-
tions were calculated directly at this pressure. The
near-identical experimental trends of ignition delays
2. Discussion for the syngas–air and hydrogen–air mixtures con-
sidered in [1,27] are striking. To be noted as well is
It is historically well known that carbon monox- that the model predictions for both are also essentially
ide oxidation is strongly influenced by the presence of identical, confirming that the influence of the CO sub-
even small amounts of hydrogen-containing species, mechanism (CO + OH = CO2 + OH, CO + O + M =
including moisture, hydrocarbons, and, most impor- CO2 + M, CO + HO2 = CO2 + OH) on ignition de-
tantly, hydrogen itself (e.g., [14–16]). However, the lay predictions is small in comparison to the influence
discussions in Ref. [1] do not reflect upon the sim- of the hydrogen–oxygen subset. We have shown else-
ilarity in ignition delay behavior for syngas–oxygen where [11] that the reaction CO + HO2 = CO2 + OH,
in shock tubes and other experimental venues to that which has historically had a large uncertainty in its
found in the literature for pure hydrogen–oxygen mix- rate constant, has an influence on chemical induc-
tures. tion times through its competition with H2 + HO2 =
For example, disparities between hydrogen–air H2 O2 + H, but is insignificant in influencing postin-
shock tube ignition delay experiments and homoge- duction chemistry [11]. We agree with Petersen et
neous predictions have been reported, analyzed, and al. [1] that uncertainties in this reaction are not re-
discussed in the literature since the 1960s; see, for ex- sponsible for the disparities in experiments and pre-
ample, [17–32]. Above about 1200 K, the so-called dictions discussed in [1].
“strong” or “sharp” ignition is observed [18], with Blumenthal et al. [27] experimentally identified
reaction initiation starting at a single locus point in the onset of strong and weak ignition experimen-
the reflected region near the end wall and quickly tally using Schlieren imaging. In Fig. 1, the hydrogen
transitioning to a uniform detonation wave. At lower strong ignition cases (filled circles) are modeled rea-
temperatures, a “mild” or “weak” ignition occurs, sonably well by homogeneous kinetic calculations,
characterized by the appearance of random flame ker- whereas weak ignition data (open circles) differ from
nels, with eventual transition to a combusting front. predictions by up to three orders of magnitude, similar
The work of Meyer and Oppenheim [23], later re- to the comparison of observations and predictions for
viewed by Oppenheim [26], also showed that mild syngas mixtures. The onset of weak ignition occurs
ignition is a multidimensional process [25], with ini- for approximately T < 1000 K and it is noted from
tial random flame kernels first appearing close to the Fig. 1 that this point also corresponds to the tempera-
shock tube walls and stagnant corners [28]. Chemi- ture range in which order-of-magnitude discrepancies
cal induction times in the mild ignition regime were are observed for syngas ignition comparisons [1].
also found to be very sensitive to temperature vari- Through combined sensitivity/stability analysis
ations behind reflected shock waves. These results methods, Yetter et al. [36] provide criteria to delin-
have been confirmed by the extensive numerical work eate mild and strong ignition processes kinetically as
of Oran and co-workers [33–35] as well as in the functions of initial reaction conditions. The transition
early work of Gardiner and Wakefield [24]. The recent temperatures noted in [27] are consistent qualitatively
shock tube studies of hydrogen oxidation by Blumen- with those associated with these kinetic definitions.
thal et al. [27,29], Wang et al. [30], and Martynenko et Yetter et al. [36] show sources of the strong sensitivity
al. [31] add further experimental detail. Blumenthal et of mild ignition and chemical induction times to per-
al. [27,29] performed temporal measurements of the turbations and delineate the most important reactions
appearance and transition of the first flame kernel to associated with mild and strong ignition phenomena.
a detonation in the mild ignition regime. Martynenko Blumenthal et al. [27,29] note that inhomogene-
et al. [31] showed that during mild ignition, OH emis- ities in temperature and concentration caused by com-
sion first occurs in the peripheral regions of the shock pressible fluid dynamic effects related to shock re-
tube, close to the walls. flection and bifurcations in the shock tube end-wall
Fig. 1 reproduces the syngas data presented by Pe- corners may not be large enough to entirely explain
tersen et al. [1] along with representative results from the discrepancies in gas phase kinetic predictions and
the H2 /air shock tube study of Blumenthal et al. [27] experimental observations of ignition. Oran et al. [33]
and homogeneous model predictions for both syngas– showed that small temperature perturbations of 10 K
air and hydrogen–air ignition delays using the mech- lead to variations in induction times predicted by their
296 F.L. Dryer, M. Chaos / Combustion and Flame 152 (2008) 293–299

numerical scheme of a factor of 2. We estimate that observed and attributed to catalytic activity on the
by themselves, temperature variations on the order stainless steel walls. Additional measurements in sil-
of 150 K would be required to bring homogeneous ica reactor tubes [43] showed similar behavior for
kinetic predictions into line with experimental obser- tube diameters smaller than 10 cm. No ignition de-
vations in [27]. It has also been hypothesized that lay measurements could be extracted in any of these
catalytic effects due to the presence of small parti- experiments, as the reaction actually began during
cles in the gas [27,37,38] could be important in some mixing.
observations. We also add to this hypothesis that cat- Vermeersch [44] was able to perform ignition de-
alytic surface effects may occur in reflected shock lay as well as postinduction kinetic measurements
experiments in the mild ignition regime, since end- for highly diluted H2 /O2 /N2 mixtures in a 10-cm-
wall/corner tube surfaces can be estimated to have fast diameter silica-walled flow reactor tube at pressures
transient heating times [39], in comparison to the ex- from 1 to 9 atm. While the gas phase reaction rates of
perimentally measured chemical ignition delays, for hydrogen–oxygen mixtures after mixing and chemi-
example, under the conditions studied in [27]. cal induction occurred were found to agree with ho-
Further work is clearly necessary to establish mogeneous kinetic predictions, the measured ignition
quantitative contributions of each type of pertur- delays reported in [44] were several orders of mag-
bation that results in near-elimination of chemical nitude shorter than predicted, similar to what is ob-
induction times (in comparison to overall ignition served in Fig. 1. The shortened ignition delays were
delay times) under mild ignition conditions. With- attributed to mixing and catalytic effects occurring
out question, however, these measurements should in the mixing region. It should also be noted that
not be expected to be predicted by zero-dimensional Vermeersch [44] showed that catalytic surface effects
(e.g., SENKIN [40]) modeling calculations, as igni- were absent in the postinduction oxidation as a result
tion takes place inhomogeneously [23–31], especially of the large reactor diameter used (10 cm). This same
in high-energy-density mixtures [31]. Unfortunately, flow reactor was later employed in the postinduction
the computational complexity of modeling the mul- kinetic studies of Mueller et al. [45,47].
tidimensional phenomena and their coupling with Similar results in terms of perturbations of chem-
detailed chemical kinetics has, to date, hampered ical induction times were observed in the Princeton
comprehensive model studies to quantitatively deter- flow reactor studies of Yetter [15,16] and Mueller et
mine the relative importance of the perturbing effects al. [45,47]. These results are consistent with other re-
that might be present in shock tube ignition delay ob- search performed on the moist carbon monoxide oxi-
servations [41]. dation system. In analyzing and modeling this work,
An intriguing point raised in [1] is that data we showed that kinetic perturbations in the mixing
collected in flow reactors [1,8] and in recent rapid region can have significant effects on chemical induc-
compression experiments in the University of Michi- tion processes (chemical “ignition delay times”), but
gan Rapid Compression Facility (UM-RCF) [6] for do not significantly affect the observed postinduction
high-pressure, high-energy-density syngas–oxygen chemistry [48].
mixtures apparently “line up” with the “disparate” It is clear from these results that the differ-
shock tube experimental ignition delay measure- ences in experimental observations and homoge-
ments. Fig. 1 shows that these data also share a par- neous kinetic predictions of chemical induction times
allelism in behavior with those for hydrogen–oxygen are present even at atmospheric pressure and in
in shock tubes. highly diluted mixtures of hydrogen–oxygen and car-
Addressing the flow reactor measurements first, bon monoxide–oxygen systems in the presence of
work not discussed in [1] shows that the shorten- hydrogen-containing species. We have avoided using
ing of ignition delays in flow reactors is not related flow reactor ignition delay measurements for kinetic
solely to the energy density of the reacting mixtures interpretation purposes because of the observed per-
studied. Very dilute flow reactor studies of hydrogen– turbations in the mixing region and the difficulty in in-
oxygen [42–45] and carbon monoxide–hydrogen– terpreting multidimensional mixing and catalytic wall
water–oxygen kinetics [15,16,46,47] at Princeton interactions. It is on the basis of these observations
have shown disparities from ignition delay predic- that our group adopted reaction “time shifting” as
tions similar to those for the high-energy-density flow a means of comparing experimental observations of
reactor results summarized in [1]. Swigart [42] stud- postinduction oxidation kinetics with homogeneous
ied highly diluted hydrogen–oxygen kinetics at at- plug flow kinetic predictions [16].
mospheric pressure and at temperatures of 920–980 K The UM-RCF data [6] reported in [1] also ap-
in stainless steel reactor tubes of diameter 2.54, 5.08, pear to follow the same trends of other data dis-
and 7.62 cm. A distinct increase in the overall rate cussed there, even though the data appear to ex-
of the reaction with decreasing reactor diameter was hibit considerable scatter. The regression analysis of
F.L. Dryer, M. Chaos / Combustion and Flame 152 (2008) 293–299 297

Walton et al. [6] (considering pressure, temperature, are assumed to be sufficiently rapid in comparison
equivalence ratio, and oxygen mole fraction as key to the observed ignition delay times in the exper-
parameters) yields a goodness of fit R 2 value of imental configurations considered, surface catalysis
only 0.57. This statistical value, however, is consistent may remove the rate-limiting behavior of these gas
with other syngas flow reactor and shock tube studies phase reactions, significantly shortening chemical in-
shown in Fig. 1 at low temperatures. The low corre- duction times and, therefore, measured ignition de-
lation of these data is a strong indication that other lays. We approximated this effect here as homoge-
dependencies exist which are not represented by the neous gas phase processes with modified rate con-
chosen correlation parameters. stants. Modeling of the syngas conditions shown in
Drawing upon the similarities between hydrogen Fig. 1 using the mechanism of Li et al. [11] includ-
and syngas systems discussed above, the UM-RCF ing catalytically accelerated rates for only the above
data [6] can also be compared against hydrogen ig- reactions (using rates adopted from Deutschmann et
nition data collected in rapid compression machines al. [51]; more details can be found in [13]) results
(RCMs) [3,49]. Walton et al. [6] point out that their in the dash-dotted curve shown in the figure, in rea-
measurements agree (on the order of magnitude of de- sonable agreement with observations. Localized radi-
lay times measured) with the RCM data of Lee and cal production from catalytically accelerated peroxide
Hochgreb [49] (at high temperatures) and indicate formation/decomposition between the third and ex-
that a more quantitative comparison is complicated tended second limits leads to an overall straight-chain
since different mixtures were used (i.e., hydrogen– kinetic conversion to H atoms and OH radicals, and
oxygen–argon [49]). However, when properly scaled, the exothermic reaction of OH with hydrogen to pro-
a quantitative comparison can be performed be- duce water and H rapidly drives the reacting systems
tween the two studies [13]. Whereas the two sets of across the extended second limit.
data [6,49] are very similar at temperatures above Clearly, additional effort is needed to experimen-
1000 K, the UM-RCF experiments result in signifi- tally confirm the relevance of catalytic processes as
cantly shorter ignition delays at lower temperatures. a perturbing factor in each of the above experimental
The overall activation energy (temperature depen- venues. Noting the sensitivity of chemical induction
dence) of the data of Walton et al. [6] is considerably times to the hydrogen peroxide intermediate chem-
different from RCM hydrogen ignition studies [3,49] istry in the mild ignition regime, the hypothesis that
(∼12.5 kcal/mol versus ∼75 kcal/mol), which more surfaces can promote chemical induction processes,
closely approximate that exhibited by homogeneous leading to further reaction and crossing of the ex-
gas phase kinetic predictions [13]. The temperature tended second explosion limit, is a hypothesis en-
dependence of [6] is heavily weighted by the short ig- tirely consistent with all of the experimental obser-
nition delays measured at lower temperatures. Walton vations. An equally important point stressed above
et al. [6] attribute the difference in activation energies is that the combination of several types of perturba-
between model and data to the uncertainty of the re- tions can lead to the same order of magnitude re-
action CO + HO2 = CO2 + OH; this has since been duction in the chemical induction time scales. The
clearly shown not to be the case [1,13]. simple approximating assumption that the perturba-
As discussed above, chemical induction processes tion is derived solely by homogeneous catalysis of the
are strongly influenced by various experimental per- above two reactions leads to an empirical modeling
turbations and they appear to be quite different in result that can be fitted to the experimental obser-
the UM-RCF and other RCM devices. Types of per- vations reported in Petersen et al. [1]. Moreover, at
turbations on ignition delay measurements for rapid temperatures above the mild ignition regime, as well
compression experiments discussed in the literature as at very low reaction temperatures, predictions re-
include surface contamination, particles as catalytic turn to those observed without homogeneous catalysis
centers, and impurities [37,38]. On the other hand, of these reactions. The empirically fitted approxima-
Walton et al. [6,50] report the presence of preigni- tion can serve as an important engineering tool for
tion reaction fronts as well as particles in the reac- conservative computational predictions in designing
tion chamber. All can have significant influence in the gas turbine mixing systems.
mild ignition regime.
A sensitivity analysis of what controls ignition
delay times under mild ignition conditions at high 3. Conclusion
pressures [13,36] shows that the most important, con-
trolling reactions are those that lead to generation of In summary, we note that the magnitude of ig-
reactive radical species from hydroperoxyl radicals, nition delay observations and homogeneous kinetic
namely H2 + HO2 = H2 O2 + H and H2 O2 + M = calculations discussed in Petersen et al. [1] is a re-
OH + OH. If radial diffusion time scales to/from walls sult in large measure of departures of the experimen-
298 F.L. Dryer, M. Chaos / Combustion and Flame 152 (2008) 293–299

tal configurations from behavior dominated solely References


by homogeneous gas phase kinetics. In the regime
of interest (higher pressures, lower temperatures), [1] E.L. Petersen, D.M. Kalitan, A.B. Barrett, S.C. Reehal,
the hydrogen–oxygen chemical induction processes J.D. Mertens, D.J. Beerer, R.L. Hack, V.G. McDonell,
can be significantly perturbed by several nonhomo- Combust. Flame 149 (2007) 244–247.
[2] C.G. Fotache, Y. Tan, C.-J. Sung, C.K. Law, Combust.
geneous effects, which include catalytic aberrations.
Flame 120 (2000) 417–426.
The multiple perturbations that can significantly af- [3] G. Mittal, C.-J. Sung, R.A. Yetter, Int. J. Chem.
fect induction chemistry are very difficult to remove Kinet. 38 (2006) 516–529.
in research experiments and nearly impossible to [4] D.M. Kalitan, E.L. Petersen, J.D. Mertens, M.W.
control in engineering applications. The implications Crofton, ASME Paper GT2006-90488, 2006.
for developing lean premixing schemes for advanced [5] H. Sun, S.I. Yang, G. Jomaas, C.K. Law, Proc. Com-
syngas gas turbine applications are that designs must bust. Inst. 31 (2007) 439–446.
consider the inherent presence of these perturbations [6] S.M. Walton, X. He, B.T. Zigler, M.H. Wooldridge,
Proc. Combust. Inst. 31 (2007) 3147–3154.
on ignition delay as well as those that might occur
[7] M.P. Burke, X. Qin, Y. Ju, F.L. Dryer, in: Fifth U.S.
from potential particle contamination of the air stream Combustion Meeting, San Diego, CA, 2007, paper
exiting the compressor, if stimulated flashback into A16.
the mixing region is to be precluded. [8] W.T. Peschke, L.J. Spadaccini, Electric Power Re-
We acknowledge that improvements in kinetic pa- search Institute, Report EPRI AP-4291, 1985.
rameters for gas phase reactions in the H2 /CO system [9] J.H. Chen, V.J. Jermakian, G.S. Samuelsen, V.G. Mc-
can be made that may influence predicted ignition de- Donell, University Turbine Systems Research, Re-
lays by factors, but not by orders of magnitude in port SR084, 2003; http://www.clemson.edu/scies/utsr/
FinalSR084.pdf.
the mild ignition regime. Compressible flow, physi-
[10] D.J. Beerer, M.U. Greene, G.S. Samuelsen, V.G. Mc-
cal mixing, and/or catalytic surface processes occur- Donell, University Turbine Systems Research, Report
ring in the mild ignition regime are the main sources SR112, 2006.
of the departure of experimentally observed ignition [11] J. Li, Z. Zhao, A. Kazakov, M. Chaos, F.L. Dryer, J.J.
delays from homogeneous gas phase kinetic predic- Scire, Int. J. Chem. Kinet. 39 (2007) 109–136.
tions. The departures first described decades ago for [12] F.L. Dryer, M. Chaos, in: Workshop on Hydrogen Com-
hydrogen–air systems are evident in syngas–air ob- bustion in Gas Turbines, EPRI, Washington, March 22,
servations recently reported [1]. 2007.
[13] M. Chaos, F.L. Dryer, Special Issue on Syngas Com-
Finally, we wish to point out that two-stage ig-
bustion, Combust. Sci. Technol., in press.
nition phenomena, observed in many hydrocarbon– [14] T.A. Brabbs, F.E. Belles, R.S. Brokaw, Proc. Combust.
air systems, display many multidimensional char- Inst. 13 (1971) 129–136.
acteristics similar to those found for hydrogen–air [15] R.A. Yetter, F.L. Dryer, H. Rabitz, Combust. Sci. Tech-
(i.e., local ignition kernels, coalescence into reac- nol. 79 (1991) 97–128.
tion fronts; single ignition kernels, development into [16] R.A. Yetter, F.L. Dryer, H. Rabitz, Combust. Sci. Tech-
detonations, e.g., see [52]). The “hot ignition” phe- nol. 79 (1991) 129–140.
nomena observed in these systems are controlled by [17] G.L. Schott, J.L. Kinsey, J. Chem. Phys. 29 (1958)
1177–1182.
the decomposition of hydrogen peroxide [53]. Cat-
[18] R.A. Strehlow, A. Cohen, Phys. Fluids 5 (1962) 97–
alytic processes that might potentially yield radicals 101.
from hydrogen peroxide at lower temperatures need [19] H. Miyama, T. Takeyama, J. Chem. Phys. 41 (1964)
to be carefully considered in ignition delay mea- 2287–2290.
surements of hydrocarbon–air systems at high pres- [20] G.B. Skinner, G.H. Ringrose, J. Chem. Phys. 42 (1965)
sures. 2190–2192.
[21] V.V. Voevodsky, R.I. Soloukhin, Proc. Combust.
Inst. 10 (1965) 279–283.
[22] C.B. Wakefield, D.L. Ripley, W.C. Gardiner, J. Chem.
Acknowledgments Phys. 50 (1969) 325–332.
[23] J.W. Meyer, A.K. Oppenheim, Proc. Combust.
We acknowledge helpful discussions with Pro- Symp. 13 (1970) 1153–1164.
fessor R.A. Yetter of Pennsylvania State University. [24] W.C. Gardiner, C.B. Wakefield, Astronaut. Acta 15
This work was supported by the U.S. Department of (1970) 399–409.
[25] K. Terao, Jpn. J. Appl. Phys. 16 (1977) 29–38.
Energy under Grant DE-FG02-86ER13503 from the
[26] A.K. Oppenheim, Phil. Trans. R. Soc. London Ser.
Chemical Sciences, Geosciences and Biosciences Di- A 315 (1985) 471–508.
vision, Office of Basic Energy Sciences, Office of Sci- [27] R. Blumenthal, K. Fieweger, K.H. Komp, G. Adomeit,
ence and by the Pittsburgh–National Energy Technol- B.E. Gelfand, in: Proc. 20th Int. Symp. on Shock
ogy Laboratories under Grant DE-FG26-05NT42544. Waves, 1995, pp. 935–940.
F.L. Dryer, M. Chaos / Combustion and Flame 152 (2008) 293–299 299

[28] R. Blumenthal, G. Adomeit, in: 3rd Workshop on Mod- [41] E.S. Oran, personal communication.
elling of Chemical Reaction Systems, Heidelberg, Ger- [42] R. Swigart, M.S.E. thesis, Aerospace and Mechanical
many, 1996; http://www.iwr.uni-heidelberg.de/groups/ Sciences, Princeton University, Princeton, NJ, 1958;
reaflow/user/crs96/Program/Contrib/a37ad.htm. Report 432.
[29] R. Blumenthal, K. Fieweger, K.H. Komp, G. Adomeit, [43] R.F. Sawyer, Ph.D. dissertation, Aerospace and Me-
Combust. Sci. Technol. 113–114 (1996) 137–166. chanical Sciences, Princeton University, Princeton, NJ,
[30] B.L. Wang, H. Olivier, H. Grönig, Combust. Flame 133 1965; Report 761.
(2003) 93–106. [44] M.L. Vermeersch, Ph.D. dissertation, Department of
[31] V.V. Martynenko, O.G. Penyaz’kov, K.A. Ragotner, Mechanical and Aerospace Engineering, Princeton
S.I. Shabunya, J. Eng. Phys. Thermophys. 77 (2004) University, Princeton, NJ, 1991; Report 1916-T.
785–793. [45] M.A. Mueller, T.J. Kim, R.A. Yetter, F.L. Dryer, Int. J.
[32] P. Sabia, E. Schießwohl, M.R. de Joannon, A. Cav- Chem. Kinet. 31 (1999) 113–125.
aliere, Turk. J. Eng. Environ. Sci. 30 (2006) 127– [46] R.A. Yetter, Ph.D. dissertation, Department of Mechan-
134. ical and Aerospace Engineering, Princeton University,
[33] E.S. Oran, T.R. Young, J.P. Boris, A. Cohen, Combust. Princeton, NJ, 1985; MAE Report 1703-T.
Flame 48 (1982) 135–148. [47] M.A. Mueller, R.A. Yetter, F.L. Dryer, Int. J. Chem.
[34] E.S. Oran, J.P. Boris, Combust. Flame 48 (1982) 149– Kinet. 31 (1999) 705–724.
162. [48] R.A. Yetter, F.L. Dryer, H. Rabitz, Combust. Flame 59
[35] E.S. Oran, V.M. Gamezo, Combust. Flame 148 (2007) (1985) 107–133.
4–47. [49] D. Lee, S. Hochgreb, Int. J. Chem. Kinet. 30 (1998)
[36] R.A. Yetter, H. Rabitz, R.M. Hedges, Int. J. Chem. 385–406.
Kinet. 23 (1991) 251–278. [50] S.M. Walton, X. He, B.T. Zigler, M.S. Wooldridge,
[37] J.E. Elsworth, W.W. Haskell, I.A. Read, Combust. Combust. Flame 150 (2007) 246–262.
Flame 13 (1969) 437–438. [51] O. Deutschmann, R. Schmidt, F. Behrendt, J. Warnatz,
[38] W.W. Haskell, SAE Paper 700059, 1970. Proc. Combust. Inst. 26 (1998) 1747–1754.
[39] P.P. Andreev, Y.M. Tsirkunov, J. Eng. Phys. Thermo- [52] K. Fieweger, R. Blumenthal, G. Adomeit, Combust.
phys. 51 (1986) 909–914. Flame 109 (1997) 599–619.
[40] A.E. Lutz, R.J. Kee, J.A. Miller, Technical Report [53] A. Kazakov, M. Chaos, Z. Zhao, F.L. Dryer, J. Phys.
SAND87-8248, Sandia National Laboratories, 1987. Chem. A 110 (2006) 7003–7009.

You might also like