You are on page 1of 4

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/319502349

Violation of collision limit in recently published


reaction models

Article in Combustion and Flame · September 2017

CITATIONS READS

0 60

3 authors:

Dongping Chen Kun Wang


Stanford University Stanford University
14 PUBLICATIONS 100 CITATIONS 29 PUBLICATIONS 74 CITATIONS

SEE PROFILE SEE PROFILE

Hai Wang
Stanford University
209 PUBLICATIONS 8,406 CITATIONS

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

morphology of carbon nanoparticles View project

combustion kinetic View project

All content following this page was uploaded by Dongping Chen on 06 September 2017.

The user has requested enhancement of the downloaded file.


Combustion and Flame 186 (2017) 208–210

Contents lists available at ScienceDirect

Combustion and Flame


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

Brief Communications

Violation of collision limit in recently published reaction models


Dongping Chen, Kun Wang, Hai Wang∗
Mechanical Engineering Department, Stanford University, Stanford, CA 94305-3032, USA

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

Article history: Twenty reaction models published in five consecutive issues of Combustion and Flame in 2015 and 2016
Received 12 July 2017 were screened for the occurrence of collision limit violations. It was found that among the 20 models
Revised 10 August 2017
tested, 15 of them contain either considerable numbers of rate coefficients that exceed their respective
Accepted 11 August 2017
collision limits or reactions exceeding the collision limit in a considerable manner. In the worst case, the
rate coefficient exceeds the collision limit by 73 orders of magnitude. It is proposed that computational
tools should be made available for authors to conduct the same rate coefficient screening. A standard
report generated by such screening tools should be submitted as a part of the supplementary material
for a manuscript in which one or more reaction models are proposed; reviewers should consult the report
in forming his/her opinions about the quality of the reaction model during manuscript review.
© 2017 The Combustion Institute. Published by Elsevier Inc. All rights reserved.

The combustion research community has a long-standing his- number of rate coefficients exceeding the collision limit. For exam-
tory in using reaction mechanisms or reaction models1 to explore ple, if a rate rule considers only the forward rate coefficient of an
the underlying physics of complex combustion phenomena [1–3]. endothermic reaction, it can produce an overly large back rate co-
Since the work of Hirschfelder, Curtiss, and others [4–7] in the efficient if the equilibrium constant of the reaction is not properly
1950s and by Dixon-Lewis, Gardiner and others in the 1960s taken into consideration.
[8–11], and of Warnatz, Westbrook, Dryer and others into the To illustrate the problem, we sampled 20 reaction models
1980s (see, e.g., [12,13]), the use of reaction models has been [19–35] published in five consecutive issues of Combustion and
instrumental in advancing our basic understanding of combustion. Flame (Volume 162 Issues 11 and 12, Volume 163, Volume 164 and
In the 1980s, the development of the highly-utilized Chemkin suite Volume 165) for violation of the collision limit in bimolecular re-
of package [14–17] enabled rapid advances of kinetic modeling of action rate coefficients. In all cases, Chemkin compatible reaction
combustion processes. models and their thermochemical and transport data were directly
Recent studies in chemical kinetic modeling typically employ taken from the Supplementary Materials of the respective papers.
large reaction models which in some cases can have 105 reactions For pressure-dependent reactions, only the high-pressure limit bi-
and 104 species [18]. The growth in the size of the reaction mod- molecular reactions were considered. The backward reaction rate
els is unfortunately not matched by an equal magnitude of high- coefficient was computed via the equilibrium constant if no ex-
quality measurements or theoretical calculations of fundamental plicit rate expression is given. Reactions that are declared as irre-
reaction rates. Today, chemical similarity or reaction class rules are versible reactions are not screened for the rate coefficients in the
often used to estimate a vast number of rate coefficients. When reverse direction.
done carefully and properly, the resulting model can yield use- We calculate the collision rate constant kcoll using
ful information about a combustion process, despite the fact that 
the model is largely empirical. When chemical similarity or reac- 8π kB T ∗
kcoll = σ 2 (1,1) Na (1)
tion rules are exercised inappropriately (e.g., mismatched rate co- μ
efficients and thermochemical data), non-physical predictions can
occur. Symptoms of such ill-exercised rules include a substantial where T is temperature, kB is the Boltzmann constant, μ is reduced
mass, σ is the collision diameter of the Lennard–Jones (LJ) 12-6

potential, Na is the Avogadro number, and (1,1 ) is the reduced

Corresponding author. collision integral, which may be calculated in a parameterized form
E-mail address: haiwang@stanford.edu (H. Wang). [36] as
1
The use of the word “mechanism” is for historical reasons. In fact, “reaction
model” is more appropriate here because a mechanism refers to how nature works
∗ =1.16145T ∗−0.14874 + 0.52487e−0.7732T ∗ + 2.16178e−2.437887T ∗
whereas a model is what we use to mimic how nature works. A reaction model
involves assumptions which may or may not be verified by isolated experiments. (2)

http://dx.doi.org/10.1016/j.combustflame.2017.08.005
0010-2180/© 2017 The Combustion Institute. Published by Elsevier Inc. All rights reserved.
D. Chen et al. / Combustion and Flame 186 (2017) 208–210 209

Table 1
Number of reactions having rate coefficient violating the collision limit and the largest ratio of rate coefficient to its collision limit.a

Model no. Total number of bimolecular Applicable T Number of rate coefficients exceeding collision limit
rate coefficients range (K)b
Applicable
T range Tmin = 300 K Tmin = 500 K Tmin = 10 0 0 K

Group 1 (transport database available. kcoll (T) calculated from Eq. (1))
1 2407 298–2381 13 (18) 13 (18) 6 (18) 3 (18)
2 4234 298–1941 50 (1 × 1020 ) 56 (1 × 1020 ) 35 (4 × 104 ) 17 (130)
3 903 313–2050 10 (2 × 1014 ) 10 (2 × 1014 ) 6 (7 × 106 ) 1 (22)
4 2945 298–2200 31 (1 × 104 ) 31 (1 × 104 ) 11 (2 × 102 ) 6 (2 × 102 )
5 1716 403–2300 9 (6) 15 (11) 12 (6) 11 (4)
6 1655 30 0–180 0 39 (2 × 1019 ) 40 (2 × 1019 ) 26 (2 × 1010 ) 9 (3 × 103 )
7 13132 50 0–110 0 9 (8 × 103 ) 20 (3 × 107 ) 10 (8 × 103 ) 6 (9)
8 1178 800–1650 3 (147) 12 (7 × 102 ) 8 (170) 4 (130)
9 807 60 0–180 0 5 (5 × 108 ) 11 (3 × 1015 ) 6 (7 × 1010 ) 4 (2 × 104 )
10 4361 298–2200 60 (1 × 1020 ) 66 (1 × 1020 ) 42 (3 × 106 ) 21 (3 × 102 )
11 2497 500–2319 13 (9 × 103 ) 19 (3 × 109 ) 13 (9 × 103 ) 7 (129)
12 7374 70 0–210 0 0 (<1) 7 (15) 2 (1.5) 0 (<1)
13 3513 30 0–210 0 36 (2 × 1017 ) 38 (2 × 1017 ) 28 (3 × 109 ) 17 (4 × 103 )
Group 2 (transport database unavailable. kcoll = 2.9 × 1015 cm3 /mol s used as the limit)
14 6338 530–1365 26 (2 × 105 ) 41 (3 × 1010 ) 29 (2 × 105 ) 5 (50)
15 41 50 0–110 0 0 (<1) 0 (<1) 0 (<1) 0 (<1)
16 3749 598–1148 4 (2 × 1014 ) 13 (1 × 1031 ) 7 (5 × 1017 ) 2 (2 × 107 )
17 1417 298–1773 56 (2 × 1073 ) 58 (2 × 1073 ) 57 (1 × 1069 ) 55 (5 × 1034 )
18 378 30 0–250 0 6 (2 × 103 ) 6 (2 × 103 ) 5 (90) 4 (13)
19 4482 643–1718 5 (1 × 104 ) 19 (3 × 1010 ) 8 (2 × 105 ) 5 (50)
20 2733 450–1400 17 (5 × 105 ) 31 (2 × 108 ) 12 (1 × 106 ) 4 (1 × 106 )
a
In all cases, Tmax in Eq. (4) was set at 2500 K except for the column showing the results in the applicable range of temperature
of the model. The values in the parentheses refer to the largest ratio of rate coefficient to its collision limit.
b
The range values are extracted from the respective papers and correspond to the minimum and maximum temperature values
of the experiments for which the model was tested against.

where T∗ is reduced temperature (T∗ = kB T/ε , where ε is the well temperature range of the model. As Table 1 reveals, this is not the
depth). The collision diameter and well depth are calculated con- case for most of the models tested. In our assessment, five models
ventionally as the arithmetic and geometric means of the self- (1, 5, 8, 12 and 15) are acceptable as they contain no rate coeffi-
collision potential parameters, respectively. For reaction models cient that exceeds the collision limit or the violations are caused
that did not include a transport database we used the colli- by difficulties in fitting some of the rate coefficients into the mod-
sion rate coefficient of H collision with n-dodecane at 2500 K or ified Arrhenius equation. Four models (4, 7, 11 and 18) are found to
2.9 × 1015 cm3 /mol s as an upper limit of the collision rate constant. be on the edge, and the remaining eleven models are quite unsat-
For the 20 reaction models sampled, Table 1 shows the total isfactory in two aspects: the number of rate coefficients exceeding
number of bimolecular-reaction rate coefficients, the number of the collision limit and the extent to which they exceed the colli-
these rate coefficients that exceed their respective collision limits, sion limit. The number of nonphysical rate coefficients is usually
and the largest ratio of the rate coefficient to its collision limit. Ad- not large as compared to the total number of rate coefficients, but
ditional test results can be found in the Supplementary Materials. the extent to which the rate coefficients exceed the collision limit
The models are listed in a random order, though they are separated is often disturbing. In the worst case, the ratio is of the order of
into two groups: Group 1 is comprised of those with transport pa- 1073 due to mismatched thermochemical data assigned to species
rameters provided as a part of the model release/publication, and participating in the reaction and its rate coefficient.
Group 2 contains those without transport parameters provided. The problem usually worsens toward lower temperatures be-
The highest rate ratios are defined as: cause of mismatched assignments of the forward rate coefficient
  and species thermochemistry. The kinetic rate-thermochemistry
κ = max k∗i (3) mismatch is quite worrisome in models 2, 3, 6, 9, 10, 13, 14, 16,
i=1,I
17, 19 and 20. In other cases, a non-physical rate coefficient is
where I is the total number of bimolecular reaction rate coeffi- caused by errors associated with extrapolating a rate expression
cients, and k∗i is the maximum of the ratio of the rate coeffi- to low temperatures when the rate expression was obtained in
cient to its collision limit over a prescribed range of temperature high-temperature experiments. Errors due to typos and other care-
Tmin ≤ T ≤ Tmax , lessness are also more common than expected. Obviously, it is not
straightforward for a reviewer to spot the rate parameter errors
k∗i = max [ki (T )/kcoll,i (T )] (4) during the review process, especially considering the size of many
Tmin ≤T ≤Tmax
of the reaction models shown in Table 1.
Tests were also made in which the Tmin and Tmax values were Models 1, 5 and 8 have k∗ values on the order of 101 or 102 .
taken from the applicable range of temperature in which a model The problem is sometimes unavoidable, as it originates from diffi-
was tested or applied (see Table 1). Additional tests were made culties in fitting rate coefficients into a modified Arrhenius expres-
with Tmax = 2500 K and Tmin = 300, 500 and 1000 K. The resulting sion over an extended range of temperature. The rate coefficients
temperature ranges are sometimes larger than the explicit or im- of many multi-channel, chemically activated reactions do not fol-
plied applicable temperature range of a reaction model. The tests low the Arrhenius expression. In such cases, the range of applica-
were performed considering the fact that a user may apply the bility of a rate expression does not extend below a cut-off temper-
model to temperatures outside the range of applicability. ature, and the authors usually ensure that non-physical rate val-
A physically sound reaction model should have no reaction with ues do not impact model predictions. To this end, we note that
its k∗i value substantially greater than unity over the applicable nonphysical rate coefficients do not always translate into nonphys-
210 D. Chen et al. / Combustion and Flame 186 (2017) 208–210

ical predictions. For example, for reacting flows that are close to [10] T. Asaba, W. Gardiner, R. Stubbeman, Shock-tube study of the hydrogen–oxy-
or governed by chemical equilibrium or chemistry is not rate lim- gen reaction, Symp. (Int.) Combust. 10 (1965) 295–302.
[11] M. Clyne, B. Thrush, Rates of the reactions of nitrogen atoms with oxygen and
iting in an overall reactive process, nonphysical rate coefficients with nitric oxide, Nature 189 (1961) 56–57.
would not produce an appreciable effect on the prediction. Yet in [12] J. Warnatz, The structure of laminar alkane-, alkene-, and acetylene flames,
this case, the range of applicability of a model containing non- Symp. (Int.) Combust. 18 (1981) 369–384.
[13] C.K. Westbrook, F.L. Dryer, Chemical kinetic modeling of hydrocarbon combus-
physical rate constant must be very limited. Indeed, extending the tion, Prog. Energy Combust. Sci. 10 (1984) 1–57.
model to thermodynamic conditions beyond its intended use un- [14] R.J. Kee, J.A. Miller, T.H. Jefferson, CHEMKIN: a general-purpose, problem-
der any circumstances can be a source of the problem especially independent, transportable, FORTRAN chemical kinetics code package, Report
No. Sandia Report SAND80-8003, Sandia National Laboratories, Albuquerque,
for reaction models that contain nonphysical rate coefficients. We
NM, 1980.
note that the unwritten rule or notion that an experimental kinetic [15] R.J. Kee, J.F. Grcar, M. Smooke, J. Miller, E. Meeks, PREMIX: a Fortran pro-
measurement must accompany itself with a detailed kinetic mod- gram for modeling steady laminar one-dimensional premixed flames, Report
No. Sandia Report SAND85-8240, Sandia National Laboratories, Albuquerque,
eling component in order for the paper to be accepted is rather
NM, 1985.
misinformed and can hinder proper interpretation of the data and [16] A.E. Lutz, R.J. Kee, J.F. Grcar, F.M. Rupley, OPPDIF: a Fortran program for com-
gaining useful insight from the measurement. puting opposed-flow diffusion flames, Report No. Sandia Report SAND96-8243,
From these results, one may conclude that around three quar- Sandia National Laboratories, Albuquerque, NM, 1996.
[17] P. Glarborg, R. Kee, J. Grcar, J. Miller, PSE: a Fortran program for modeling well-
ters of the reaction models published recently contain non-physical stirred reactors, Report No. Sandia Report SAND86-8209, Sandia National Lab-
rate coefficients. To help to resolve this problem, we wish to make oratories, Albuquerque, NM, 1986.
the following specific recommendations. [18] T. Lu, C.K. Law, Toward accommodating realistic fuel chemistry in large-scale
computations, Prog. Energy Combust. Sci. 35 (2009) 192–215.
1. It is crucial for reaction model developers to pay attention to [19] A. Rodriguez, O. Herbinet, F. Battin-Leclerc, A. Frassoldati, T. Faravelli, E. Ranzi,
Experimental and modeling investigation of the effect of the unsaturation de-
collision-limit violations; gree on the gas-phase oxidation of fatty acid methyl esters found in biodiesel
2. When submitting a manuscript for publication, authors have fuels, Combust. Flame 164 (2016) 346–362.
the responsibility to provide explanations in places when and [20] K.B. Brady, X. Hui, C.-J. Sung, Comparative study of the counterflow forced ig-
nition of the butanol isomers at atmospheric and elevated pressures, Combust.
where collision rate violations occur; Flame 165 (2016) 34–49.
3. Computational tools, e.g., those built for Chemkin and Cantera, [21] S. Banerjee, R. Tangko, D.A. Sheen, H. Wang, C.T. Bowman, An experimental
should be made available for authors to conduct the rate coef- and kinetic modeling study of n-dodecane pyrolysis and oxidation, Combust.
Flame 163 (2016) 12–30.
ficient screening. A standard report generated by such screen- [22] N. Lamoureux, H. El Merhubi, L. Pillier, S. de Persis, P. Desgroux, Modeling of
ing tools can be submitted as a supplementary material for NO formation in low pressure premixed flames, Combust. Flame 163 (2016)
a manuscript in which one or more reaction models are pro- 557–575.
[23] Z. Jia, Z. Wang, Z. Cheng, W. Zhou, Experimental and modeling study on
posed;
pyrolysis of n-decane initiated by nitromethane, Combust. Flame 165 (2016)
4. Reviewers should consult the report in forming his/her opin- 246–258.
ions about the quality of the reaction model during manuscript [24] K. Wang, S.M. Villano, A.M. Dean, Fundamentally-based kinetic model for
propene pyrolysis, Combust. Flame 162 (2015) 4456–4470.
review.
[25] J. Bugler, B. Marks, O. Mathieu, R. Archuleta, A. Camou, C. Grégoire, K.A. Heufer,
E.L. Petersen, H.J. Curran, An ignition delay time and chemical kinetic modeling
Editorial policies of a similar nature exist in other fields. Ex-
study of the pentane isomers, Combust. Flame 163 (2016) 138–156.
amples include the demonstration of grid-independent or grid- [26] L.-S. Tran, R. De Bruycker, H.-H. Carstensen, P.-A. Glaude, F. Monge,
convergent results in computational fluid dynamics, which has M.U. Alzueta, R.C. Martin, F. Battin-Leclerc, K.M. Van Geem, G.B. Marin, Pyroly-
sis and combustion chemistry of tetrahydropyran: experimental and modeling
been implemented and utilized with success for many years.
study, Combust. Flame 162 (2015) 4283–4303.
[27] A. Stagni, A. Frassoldati, A. Cuoci, T. Faravelli, E. Ranzi, Skeletal mechanism
Supplementary materials reduction through species-targeted sensitivity analysis, Combust. Flame 163
(2016) 382–393.
[28] X. Liu, H. Wang, L. Wei, J. Liu, R.D. Reitz, M. Yao, Development of a reduced
Supplementary material associated with this article can be toluene reference fuel (TRF)-2, 5-dimethylfuran-polycyclic aromatic hydrocar-
found, in the online version, at doi:10.1016/j.combustflame.2017.08. bon (PAH) mechanism for engine applications, Combust. Flame 165 (2016)
005. 453–465.
[29] C. Brackmann, V.A. Alekseev, B. Zhou, E. Nordström, P.-E. Bengtsson, Z. Li,
M. Aldén, A.A. Konnov, Structure of premixed ammonia + air flames at at-
References mospheric pressure: laser diagnostics and kinetic modeling, Combust. Flame
163 (2016) 370–381.
[1] J.A. Miller, R.J. Kee, C.K. Westbrook, Chemical kinetics and combustion model- [30] K. Zhang, C. Banyon, C. Togbé, P. Dagaut, J. Bugler, H.J. Curran, An experimental
ing, Ann. Rev. Phys. Chem. 41 (1990) 345–387. and kinetic modeling study of n-hexane oxidation, Combust. Flame 162 (2015)
[2] J.A. Miller, M.J. Pilling, J. Troe, Unravelling combustion mechanisms through 4194–4207.
a quantitative understanding of elementary reactions, Proc. Combust. Inst. 30 [31] H. Ning, C. Gong, N. Tan, Z. Li, X. Li, Low- and intermediate-temperature ox-
(2005) 43–88. idation of ethylcyclohexane: a theoretical study, Combust. Flame 162 (2015)
[3] J.M. Simmie, Detailed chemical kinetic models for the combustion of hydrocar- 4167–4182.
bon fuels, Prog. Energy Combust. Sci. 29 (2003) 599–634. [32] X. Liu, M. Yao, Y. Wang, Z. Wang, H. Jin, L. Wei, Experimental and kinetic mod-
[4] J. Hirschfelder, C. Curtiss, D.E. Campbell, The theory of flame propagation. IV, eling study of a rich and a stoichiometric low-pressure premixed laminar 2,
J. Phys. Chem. 57 (1953) 403–414. 5-dimethylfuran/oxygen/argon flames, Combust. Flame 162 (2015) 4586–4597.
[5] C. Curtiss, J.O. Hirschfelder, Integration of stiff equations, Proc. Natl. Acad. Sci. [33] W. Sun, B. Yang, N. Hansen, C.K. Westbrook, F. Zhang, G. Wang, K. Moshammer,
38 (1952) 235–243. C.K. Law, An experimental and kinetic modeling study on dimethyl carbonate
[6] R.E. Duff, Calculation of reaction profiles behind steady state shock waves. I. (DMC) pyrolysis and combustion, Combust. Flame 164 (2016) 224–238.
Application to detonation waves, J. Chem. Phys. 28 (1958) 1193–1197. [34] K. Narayanaswamy, H. Pitsch, P. Pepiot, A component library framework for
[7] H. Glick, J. Klein, W. Squire, Single pulse shock tube studies of the kinetics of deriving kinetic mechanisms for multi-component fuel surrogates: application
the reaction N2 + O2  2NO between 20 0 0–30 0 0 K, J. Chem. Phys. 27 (1957) for jet fuel surrogates, Combust. Flame 165 (2016) 288–309.
850–857. [35] U. Burke, W.K. Metcalfe, S.M. Burke, K.A. Heufer, P. Dagaut, H.J. Curran, A de-
[8] G. Dixon-Lewis, Flame structure and flame reaction kinetics. V. Investigation of tailed chemical kinetic modeling, ignition delay time and jet-stirred reactor
reaction mechanism in a rich hydrogen + nitrogen + oxygen flame by solution study of methanol oxidation, Combust. Flame 165 (2016) 125–136.
of conservation equations, Proc. R. Soc. A 317 (1970) 235–263. [36] P.D. Neufeld, A. Janzen, R. Aziz, Empirical equations to calculate 16 of the
[9] G. Dixon-Lewis, Flame structure and flame reaction kinetics. I. Solution of con- transport collision integrals  (l,s)∗ for the Lennard–Jones (12–6) potential, J.
servation equations and application to rich hydrogen–oxygen flames, Proc. R. Chem. Phys. 57 (1972) 1100–1102.
Soc. A 298 (1967) 495–513.

View publication stats

You might also like