You are on page 1of 8

Available online at www.sciencedirect.

com
Proceedings
ScienceDirect of the
Combustion
Institute
Proceedings of the Combustion Institute 35 (2015) 2431–2438
www.elsevier.com/locate/proci

Flame structure and particle-combustion regimes


in premixed methane–iron–air suspensions
Philippe Julien ⇑,1, Sam Whiteley, Samuel Goroshin, Michael J. Soo,
David L. Frost, Jeffrey M. Bergthorson
Department of Mechanical Engineering, McGill University, Montreal, Canada

Available online 14 June 2014

Abstract

Flame structures and particle combustion regimes are studied in hybrid fuel mixtures of methane and
iron using a modified Bunsen burner with two different oxidizing environments. The first is a stoichiometric
methane–air mixture, in which the iron reacts with the hot gaseous combustion products in a kinetically-
controlled regime; the second is a lean methane–oxygen–nitrogen mixture, allowing iron to react with the
excess oxygen in a diffusion-controlled regime. The particle seeding concentration is monitored using laser
attenuation and varies from 0 to 350 g/m3. Burning velocities are obtained from flame photographs and
condensed phase temperatures are determined from particle emission spectra via polychromatic fitting
of the spectra to Planck’s law of blackbody radiation. High-speed imaging is also used to qualitatively
characterize the different flame structures. Results are also compared to a methane flame seeded with inert
silicon carbide particles instead of iron. It is shown that, independent of the combustion regime, a critical
concentration of iron powder is required to form a coupled flame front in the combustion products of the
methane flame. Furthermore, after the metal-powder flame formation, a double front structure separated
by a dark zone is observed in the kinetic regime, whereas the two flames overlap and form a single Bunsen
cone in the case of the diffusion regime.
Ó 2014 The Combustion Institute. Published by Elsevier Inc. All rights reserved.

Keywords: Iron–methane flame; Flame structure; Hybrid; Diffusion; Kinetic

1. Introduction different systems. Two or more reactive fronts


can interact by different physical mechanisms,
Multi-front reactive waves are complex phe- including diffusion and convection of either heat
nomena that are known to occur in a variety of or active species, and through radiative heat trans-
fer. Double detonation fronts have been studied
for reactive particle-laden gases [1,2]. Similar phe-
⇑ Corresponding author. Address: McGill University, nomena have been observed in systems undergo-
Department of Mechanical Engineering, 817 Sherbrooke ing self-propagating high-temperature synthesis
St. W., Montreal, Quebec H3A 0C3, Canada. Fax: +1 [3], metalized propellants [4] and also figure into
(514) 398 7365. proposed novel nuclear reactor concepts [5]. A
E-mail address: philippe.julien@mail.mcgill.ca simple system to study multi-front phenomena
(P. Julien). consists of mixtures containing two fuels that have
1
Presenting author.

http://dx.doi.org/10.1016/j.proci.2014.05.003
1540-7489/Ó 2014 The Combustion Institute. Published by Elsevier Inc. All rights reserved.
2432 P. Julien et al. / Proceedings of the Combustion Institute 35 (2015) 2431–2438

very different activation energies and reaction observed when the mixture contains excess
mechanisms. Deflagrations in such mixtures have oxygen. The gases are seeded with micron-sized
been studied theoretically [6,7] and experimentally iron powder with a particle concentration varying
for CH4–NO2–O2 mixtures [8] and hybrid alumi- from 0 to 350 g/m3. Flames are directly observed
num–methane mixtures [9,10]. with high-speed imaging and burning velocities
For aluminum–methane–air mixtures, previ- are obtained using an estimate of the flame surface
ous studies have demonstrated that an aluminum area from the flame photographs. Emission spec-
flame front may form in the products of the meth- troscopy is used to determine the temperature of
ane combustion with micron-sized aluminum par- the condensed-phase species in the flame. The
ticles reacting in a kinetically-controlled regime results are then compared to that of a methane
[9]. In this case, the aluminum flame front forms flame seeded with inert silicon carbide (SiC) parti-
only above a critical particle concentration and cles. In the kinetically-controlled regime, an iron
is believed to be coupled to the methane flame flame front forms above a critical concentration
front. However, due to the intense luminosity of and couples to the methane flame, even though
the aluminum flame it is not possible to directly the particles do not ignite. In the case of the diffu-
observe the methane flame front once the alumi- sion-controlled regime, single particles do ignite,
num flame has formed. In addition, aluminum even at the lowest concentrations, but again form
typically burns in the vapor phase and produces a distinct iron flame front above a critical
a large quantity of sub-oxides which increases concentration.
the complexity of the problem by introducing
other potential limiting factors, such as the rate
of evaporation [10]. For these reasons, a different 2. Experimental methods
particulate fuel is selected to simplify the investi-
gation of multi-front reactive waves. Iron is a suit- 2.1. Iron powder and gaseous mixtures
able candidate as it burns purely heterogeneously
by surface reactions. The iron–air adiabatic flame The iron powder used in the present investiga-
temperature is on the order of the methane flame tion was produced by Alfa Aesar. The Sauter
temperature, which allows the observation of both mean diameter of the particles is d32 = 2.20 lm
flames simultaneously. It has also been shown that with a volume-weighted size distribution obtained
iron can burn in either a diffusion-controlled or a from a Malvern Mastersizer shown in Fig. 1.
kinetically-controlled regime depending on, Figure 2 shows a SEM photograph of the spheri-
among other things, the oxidizing environment cal iron powder.
[11]. Two different gaseous mixtures are used: the
The present study investigates stabilized flames first is a stoichiometric methane–air mixture in
with a double-front structure in hybrid mixtures which, after the methane flame, there is no excess
of iron and methane and focuses on the effect of oxygen; in the second case, all flow rates are kept
the particle combustion mode on this flame struc- identical but the air is replaced with the modified
ture and its propagation speed with different mixture made up of 26% oxygen and the balance
oxidizing mixtures. The kinetically-controlled nitrogen, which has little effect on the flame tem-
regime is observed in the products of methane– perature and the operation of the dust dispersion
air combustion, whereas the diffusion regime is system. The resulting mixture corresponds to an
equivalence ratio of 0.8, which leaves approxi-
mately 5% excess oxygen after the methane has
been fully consumed.

2.2. Dust burner and diagnostics

Experiments are performed with a Bunsen bur-


ner designed to operate with gaseous and dust
fuels. Detailed schematics of the apparatus can
be found in previous publications [9]. The powder
is fed from a reservoir to the burner by a piston
and is dispersed by the gas mixture through an
air-knife. The resulting two-phase flow is lamina-
rized in a 60-cm-long tube and ignited at the exit
of the nozzle.
The dust concentration is monitored by laser
attenuation. A beam passes through the two-
phase mixture through a slit in the nozzle covered
with high-temperature, high-optical-clarity tape.
Fig. 1. Iron particle size distribution. The Beer–Lambert law is used to correlate the
P. Julien et al. / Proceedings of the Combustion Institute 35 (2015) 2431–2438 2433

lines or molecular bands. By contrast, the top


spectrum, from the methane–lean mixture, exhib-
its molecular bands between 550 and 650 nm on
top of the continuous spectrum. Those bands
are present even at low iron concentration and
are thought to be attributed to an electronic tran-
sition of iron oxide, FeO, identified from the spec-
trum produced by West and Broida [14]. The
temperature is obtained by fitting the continuous
part of the spectra to Planck’s law of blackbody
radiation [13], assuming that iron particles behave
as grey bodies. The accuracy of the temperature
measurements is estimated to be ±120 K. Mea-
surements are taken in the midway between the
base and the tip of the flame and the temperature
can vary over the height of the cone, with the tip
Fig. 2. Scanning electron micrograph of iron powder. being the hottest part. Results are shown in Fig. 4,
with the open squares reflecting the stoichiometric
case and closed squares the enriched-oxygen case.
laser attenuation with the dust concentration. The There is a significant difference in particle tem-
system is calibrated separately for every type of perature depending on the oxidizing environment.
powder [12]. The methane flame temperature without iron
An Ocean Optics USB 4000 spectrometer cou- loading in both cases is approximately 2200 K.
pled with a 100-micron optical fiber is used to The particle temperature in the methane–air flame
obtain emission spectra from the dust flame. Spa- is close to the methane-flame temperature at low
tial resolution is achieved in one dimension by particle concentration and slowly decreases with
coupling the optical fiber to a telescope focused increasing concentration. The particle tempera-
on a rotating mirror reflecting the image of the ture in the oxygen-enriched mixture, however,
flame. The Bunsen cone is scanned with a step has a relatively constant value near 2600 K, even
motor attached to the mirror. The flame front is at low particle concentration, which is about
located by the point of maximum emission inten- 400 K greater than the methane adiabatic-flame
sity [13]. Flame images and videos are acquired temperature.
with a high-resolution digital camera and a high-
speed video camera operating at 300 frames per 3.2. Burning velocities
second. A variable neutral-density filter is used
to reduce the intensity of the light emitted by The burning velocities are obtained from pho-
the flame [9]. The intensity of the filter is changed tographs of the flame. The contour of the Bunsen
according to the luminosity while keeping the cone is traced on a touch-sensitive screen, fitted
exposure time and f-number the same. All optical with a seventh-order polynomial and rotated to
systems are linked to the particle concentration obtain the total surface area, A. The volumetric
monitor to correlate the time of the picture, video
or spectrum with its corresponding dust
concentration.

3. Results and discussion

3.1. Flame spectra and temperature

Flame emission spectra are acquired over the


range of particle concentrations in both types of
methane flames. The integration time varies with
the dust concentration and temperature. Shorter
integration times are required for high particle
loadings due to the greater number of emitters.
The two oxidizing environments produce distinc-
tively different spectra, as illustrated in Fig. 3.
The lower spectrum in Fig. 3 is obtained from
iron particles in the stoichiometric methane–air Fig. 3. Emission spectra for both stoichiometric
flame. The spectrum is composed solely of contin- methane–air and enriched-oxygen–methane cases at an
uous blackbody radiation, without any atomic iron concentration of approximately 200 g/m3.
2434 P. Julien et al. / Proceedings of the Combustion Institute 35 (2015) 2431–2438

Fig. 4. Iron particle temperatures as a function of iron


concentration in both oxidizing environments. Fig. 6. Burning velocities as a function of dust concen-
tration. The gaseous mixture has 5% excess oxygen.

ments. In small quantities, iron particles behave


in a similar fashion and reduce the burning veloc-
ity. In the stoichiometric case, the flame speed
appears to plateau at a particle concentration of
about 250 g/m3, near the maximum value for
which the flame was stable, as discussed below.
The plateau appears to occur sooner for the
excess-oxygen flame, at a concentration of about
100 g/m3. It is suspected that the steepness of
the burning velocity decrease with increasing con-
centration is due to the discrepancy of particle
size. The iron particles range from 1 to 3 lm
whereas the average SiC particle is 6 lm. This dif-
ference in size implies that the SiC particles do not
reach thermal equilibrium within the methane
reaction zone and, thus, have a reduced impact
on the methane flame, compared to iron, for the
Fig. 5. Burning velocities as a function of dust concen- same particle loading.
tration. The gaseous mixture is stoichiometric methane–
air. 3.3. Stoichiometric methane–air flame: No excess
oxygen

flow rate, V_ , is then divided by A to obtain the The flame temperatures indicate that the iron
burning velocity, SL [15]. The burning velocity burns in a kinetically-controlled regime when
as a function of particle concentration is shown introduced into the stoichiometric methane–air
in Figs. 5 and 6. flame. The absence of oxygen after the methane
Figure 5 shows the results in air and Fig. 6 flame means that the iron particles must react with
shows the result in enriched-oxygen. The results the combustion products of methane: water vapor
are also compared with the values obtained with and carbon dioxide. Water molecules diffuse
silicon carbide particles. The stoichiometric meth- quicker than oxygen molecules yet the reaction
ane–air burning velocity is about 35 cm/s, which rate is most likely slower than with oxygen, which
agrees with literature values [16]. The burning would explain why the chemical reaction rates
velocity when calculated with the GRI-30 mecha- control the combustion. The particle temperature
nism [17] in Cantera [18] gives a value 51 cm/s in at low iron concentrations is very close to the adi-
the excess oxygen case, which agrees with the abatic flame temperature of 2220 K. This does not
experimental results. necessarily imply that the particles burn in a kinet-
Silicon carbide particles are used as a heat sink ically-controlled regime since the predicted ther-
in the methane flame, increasing its effective heat modynamic equilibrium iron–air temperature
capacity. This decreases the flame temperature (without methane) is comparable at about
and, in turn, decreases the burning velocity. This 2250 K [11]. In a diffusion-controlled regime, the
trend can be observed in both gaseous environ- particle temperature is independent of the particle
P. Julien et al. / Proceedings of the Combustion Institute 35 (2015) 2431–2438 2435

environment but neither ignition nor flame front


formation is observed. Flame B has an intermedi-
ate iron concentration of about 200 g/m3 and two
distinct Bunsen cones are visible. The thin inner
cone is the methane flame and the thick outer
one is the iron flame. The two cones are separated
by a dark zone, in which there is no flame. After
leaving the first flame, the gas expands causing
the local iron concentration to decrease, which is
why less light is observed in that region. The sec-
ond cone is the iron flame front, which causes the
particles to heat up again and emit more light.
This double-front structure happens only above
a certain critical concentration, which coincides
with the point of the iron flame front formation.
This critical concentration, found from the flame
speed data and high-speed videos, is situated at
Fig. 7. Measured iron-particle temperature and calcu- around 200 g/m3. The dust flame front is sepa-
lated equilibrium temperature for the stoichiometric rated from the methane flame front, which means
methane–air case. that the particles still absorb heat from the meth-
ane reaction zone. However, a leveling off of the
loading and does not follow the gas-phase temper- burning velocity is observed at this point (see
ature. However, as can be seen in Fig. 7, the par- Fig. 5), which implies that enough heat is trans-
ticle temperature follows the thermodynamic ferred from the iron flame to the methane flame
prediction (solid line) for the methane–air mixture to compensate for the heat absorbed and the
including iron particles, further indicating that the flames must be coupled. Furthermore, a methane
particle and bulk gas temperatures are equal and flame loaded with a similar concentration of SiC
that the particles are reacting in the kinetically- particles quenches, which indicates that the meth-
controlled regime. ane flame could not exist at these particle loadings
Figure 8 shows images of iron–methane–air without being supported by the energy released
flames at three different iron concentrations. from the iron flame.
Depending on the concentration, the stabilized Finally, Fig. 8c is taken in the high concentra-
flame has one of three different appearances. tion range of about 325 g/m3. The flame becomes
Flame A has a low concentration of about 80 g/ unstable and cellular due to thermo-diffusive
m3 and only one Bunsen cone is visible. The light instabilities. These instabilities occur when the
observed comes from the blackbody radiation flame is deficient in the more mobile reactant,
emitted by hot iron particles in the flame. This changing the Lewis number of the mixture [15].
alone is not a sign of combustion. At these low Increasing the particle concentration will not vary
concentrations, particles will oxidize in the hot the amount of gaseous reactants since the gas flow

Fig. 8. Iron–methane–air flames: (a) low particle concentration (80 g/m3) showing no iron flame front formation; (b)
intermediate particle concentration (150 g/m3) showing a double flame front; (c) high iron concentration (325 g/m3)
showing flame instabilities.
2436 P. Julien et al. / Proceedings of the Combustion Institute 35 (2015) 2431–2438

rates are unchanged and the particles occupy an Figure 10 shows two flame images taken at two
insignificant volume compared to the gases. Thus, different concentrations for the enriched oxygen
an increase in particle concentration will not case. Image A has a low iron concentration and
decrease the concentration of oxidizers, the spe- only the methane flame is visible. It is also possi-
cies actually diffusing in the mixture. An increase ble to see single iron particles burning, after igni-
in particle concentration will increase the overall tion, including the characteristic micro-explosion
heat capacity of the mixture, which in turn affects at the end of the particle track. The heat required
the heat diffusivity and decreases the Lewis num- to bring the particle up to the ignition temperature
ber. When the Lewis number is sufficiently small, comes from the methane flame and the particles
the structure becomes cellular [15]. A similar cellu- may react with the 5% excess oxygen, in addition
lar structure has been observed for pure iron to the water vapor and carbon dioxide. Image B is
flames in microgravity [11]. taken at a concentration of about 150 g/m3. In
this image, only one Bunsen cone is visible in
3.4. Enriched-oxygen–methane flame: 5% excess the diffusion regime, which is the coupled combus-
oxygen tion front consuming both methane and iron. The
formation of the iron flame front coincides with
The experimental results suggest that, with 5% the leveling off of the burning velocity, which
excess oxygen in the methane flame, iron particles occurs at lower concentrations than in the meth-
ignite and burn in a diffusion-controlled regime. ane–air case. Furthermore, the formation of the
The particle temperature, even at low concentra- dust flame front is clearly visible in high-speed
tions, is about 400 K above the methane adiabatic images and is a distinct phenomenon from a mere
flame temperature. Furthermore, the temperature increase of igniting particles in the flow.
remains constant across the concentration range,
further indication that the particles ignite and that 3.5. Flame front formation
the particle and bulk-gas temperatures separate.
The gas-phase temperature cannot be measured These experiments have shown that the forma-
directly by emission spectroscopy, but the iron tion of a dust flame front in the hydrocarbon–air
particles can be replaced with inert SiC particles. flame products is independent of the regime of
The temperature results are shown in Fig. 9. As particle combustion and is analogous to what
expected, the temperature at low SiC concentra- can be observed in a gas phase mixture. A mini-
tions is close to 2220 K and decreases with an mum concentration of iron particles is required
increase in particle loading. This figure also shows for the heat-release rate to be sufficient to sustain
equilibrium thermodynamic calculations per- the propagation of the iron flame. In this respect,
formed for different iron concentrations. The solid single particle models are insufficient to com-
line is the calculation performed for the enriched- pletely describe the observed results, as they
oxygen–methane mixture including iron. The ignore the effect of feedback from the reaction
measured particle temperature is always above of neighboring particles on increasing the local
that of the calculated equilibrium temperature of temperature above the methane product tempera-
the iron–methane–oxygen mixture. ture. This description of flame front formation

Fig. 9. Emission temperature measurements for iron Fig. 10. Iron–methane-enriched-oxygen flames: (a) low
and silicon carbide particles in enriched-oxygen–meth- concentration (25 g/m3) with no dust flame front; (b)
ane mixture and comparison to thermodynamic equilib- high concentration (140 g/m3) with a coupled flame
rium predictions. and one visible Bunsen cone.
P. Julien et al. / Proceedings of the Combustion Institute 35 (2015) 2431–2438 2437

applies to any reactive particulate system, whether tion refers to a specific phenomenon: the particle
it is a pure dust cloud [12] or for hybrid mixtures transitioning from kinetic oxidation to diffusion-
of gaseous and solid fuels [9]. controlled combustion, causing a separation of
In the kinetically-controlled regime, iron parti- the particle from the bulk-gas temperature. When
cles at low concentration will oxidize in the flame introduced into the excess-oxygen flame, iron par-
but the reaction rate will be too slow to form a ticles heat up and ignite after an induction time,
flame front. The iron flame will form only when which depends on the particle size. At low concen-
there is at least a critical number of iron particles, trations, this ignition occurs after the methane
such that the overall heat released from the dust flame and the particles act merely as heat sinks
cloud is sufficient to propagate a flame at a speed within the reaction zone, decreasing the pure gas
that matches the speed of the methane-flame flame temperature, without contributing to the
products. For the structure to be stable, the two flame propagation. This is represented by
fronts must be coupled, which can only happen Fig. 11b, where the heat release from the ignited
through the exchange of heat between the two iron particles happens far from the flame and is
reaction zones, which is reflected, among other spread over a large distance, reflecting the range
things, in the leveling off of the flame speed [10]. in particle-induction and burn time for the vari-
This situation is illustrated with Fig. 11a, where able particle sizes in the iron powder. As the dust
the two heat release zones are separated from each concentration increases, increased heat release
other and the temperature plateau between reac- from surrounding particles leads to a feedback
tion zones corresponds to the dark region. This mechanism, and locally higher temperatures than
phenomenon has been previously discussed for in the low-concentration case, which causes the
flames in hybrid mixtures of aluminum and meth- particles to ignite closer to the methane flame.
ane in air. This regime has been denoted a control At the critical concentration, the heat released
regime [19], where the second flame front effec- by the iron particles will be enough to form a
tively controls the speed of the first flame, flame front. In such a case, the iron reaction zone
observed by the leveling off of the flame speed. is small and must overlap with the methane reac-
It has also been discussed in the case of binary tion zone; hence only one flame is visible. This is
mixtures of dusts [20]. denoted a merging flame regime [19] and is repre-
It is important to distinguish between particle sented by Fig. 11c.
ignition and flame front formation. Particle igni-

4. Conclusions

The flame structure of methane–iron–air


flames is investigated for particles reacting in the
diffusion-controlled or kinetically-controlled
regimes through the use of different oxidizing mix-
tures. In the kinetic regime at low particle concen-
trations, iron slowly oxidizes in the hot
combustion products of methane and an iron
flame front only forms above a critical concentra-
tion and couples to the methane flame. Finally, at
high concentrations, the flame structure exhibits
instabilities due to thermo-diffusive effects caused
by the increase in the overall heat capacity of
the mixture. The second case, with diffusion-
controlled particle combustion, where single iron
particles ignite in the hot combustion products
and burn with a temperature close to their own
adiabatic flame temperature, an iron flame front
is also seen to form and merge with the leading
methane flame above a critical particle concentra-
tion. Detailed modeling of the phenomena
observed requires further knowledge of high-
temperature iron kinetics with oxygen, water
and carbon dioxide. These rates are currently
unknown and determination of these reaction
Fig. 11. Different combustion regimes observed in this rates should be the focus of future studies.
study: (a) kinetic regime with iron flame front; (b) These results demonstrate that it is possible
diffusion regime with no flame front; (c) diffusion regime to form a stabilized flame front in a reactive-
with flame front. particle suspension whether the particles are
2438 P. Julien et al. / Proceedings of the Combustion Institute 35 (2015) 2431–2438

reacting in a kinetically-controlled or diffusion- tional Congress on Advances in Nuclear Power


controlled regime or, likely, even in a regime Plants 2010, San Diego, CA, 2010.
where both kinetics and diffusion play equal [6] B. Khaikin, A. Filonenko, S. Khudyaev, Combust.
roles. For a flame front to form, the rate of heat Explos. Shock Waves 4 (4) (1968) 343–348.
[7] S. Goroshin, M. Bidabadi, J.H.S. Lee, Combust.
release by the particles must be sufficient to sta- Flame 105 (1–2) (1996) 147–160.
bilize the reaction wave within the dust suspen- [8] M.C. Branch, M.E. Sadequ, A.A. Alfarayedhi, P.J.
sion, which requires that the resulting flame Van Tiggelen, Combust. Flame 83 (3–4) (1991) 228–
speed of the metal front in the local oxidizing 239.
environment be equal to the local flow velocity. [9] M. Soo, P. Julien, S. Goroshin, J.M. Bergthorson,
These metal–hydrocarbon–air flames provide a D.L. Frost, Proc. Combust. Inst. 34 (2) (2013)
relatively clean and simple system in which to 2213–2220.
study the stability and physics of multi-front [10] P. Julien, M. Soo, S. Goroshin, D.L. Frost, J.M.
reactive waves. Bergthorson, N. Glumac, F. Zhang, Combustion of
Aluminum Suspensions in Hydrocarbon Flame Pro-
ducts, J. Propul. Power (2014). http://dx.doi.org/
10.2514/1.B35061.
Acknowledgements [11] F.-D. Tang, S. Goroshin, A.J. Higgins, Proc.
Combust. Inst. 33 (2) (2011) 1975–1982.
Support for this work was provided by the [12] S. Goroshin, I. Fomenko, J.H.S. Lee, Proc. Com-
Natural Sciences and Engineering Research bust. Inst. 26 (2) (1996) 1961–1967.
Council of Canada and Martec, Ltd., under a Col- [13] S. Goroshin, J. Mamen, A. Higgins, T. Bazyn, N.
Glumac, H. Krier, Proc. Combust. Inst. 31 (2)
laborative Research and Development Grant, and (2007) 2011–2019.
the Defense Threat Reduction Agency under con- [14] J.B. West, H.P. Broida, J. Chem. Phys. 62 (7) (1975)
tract HDTRA1-11-1-0014 (program manager 2566–2574.
Suhithi Peiris). [15] C.K. Law, Combustion Physics, Cambrige Univer-
sity Press, New York, 2006.
[16] B. Lewis, G. Von Elbe, Combustion, Flames and
Explosions of Gases, vol. 3, Cambridge University
References Press, 1961.
[17] G.P. Smith, D.M. Golden, M. Frenklach, et al.,
[1] B.A. Khasainov, B. Veyssiere, Shock Waves 6 (1) available at <http://www.me.berkeley.edu/gri_-
(1996) 9–15. mech/>.
[2] F. Zhang, J. Propul. Power 22 (6) (2006) 1289–1309. [18] D.G. Goodwin, An open-source, extensible soft-
[3] A.G. Merzhanov, in: J.B. Munir, J.B. Holt (Eds.), ware suite for CVD process simulation, in: Pro-
Combustion and Plasma Synthesis of High-Temper- ceedings of CVD XVI and EuroCVD Fourteen, 2003.
ature Materials, VCH, 1990, pp. 1–53. pp. 155–162.
[4] R.L. Geisler, R.A. Frederick, M. Giarra, in: Ency- [19] I. Zeldovich, G.I. Barenblatt, V. Librovich, G.
clopedia of Aerospace Engineering, John Wiley & Makhviladze, Mathematical Theory of Combustion
Sons Ltd, 2010. and Explosions, Consultants Bureau, New York,
[5] R.P.T. Ellis, P. Hejzlar, G. Zimmerman, et al., 1985.
Traveling-wave reactors: a truly sustainable and [20] S. Goroshin, M. Kolbe, J.H.S. Lee, Proc. Combust.
full-scale resource for global energy needs, Interna- Inst. 28 (2) (2000) 2811–2817.

You might also like