Sei sulla pagina 1di 8

Pyrolysis: Mathematical Modeling of Hydrocarbon Pyrolysis Reactions

E Ranzi, Politecnico di Milano, Milano, Italy


ã 2015 Elsevier Inc. All rights reserved.

Introduction 1
Kinetic Modeling of Pyrolysis Reactions 1
n-Butane Pyrolysis 3
Successive Pyrolysis Reactions 4
Pyrolysis of Large Hydrocarbons: n-Decane 5
Pyrolysis of Liquid Feedstocks 7
Automatic Generation of Reaction Mechanisms 7
References 8

Introduction

Pyrolysis comes from the Greek word pῦr meaning ‘fire’ and lύo meaning ‘to break.’ Pyrolysis is typically a decomposition process, at
high temperatures in the absence of oxygen. Pyrolysis or steam cracking of hydrocarbons is the main petrochemical process for
producing ethylene and propene. Feedstocks ranging from light alkanes (LPG, ethane, propane, and butane) to complex liquid
feedstocks (naphthas and gas oil) are mixed with steam and preheated in the convection section of a large steam cracking furnace. The
pyrolysis reactions mostly take place in tubular reactors in the radiant section of the furnace at 900–1200 K and pressures slightly
above atmospheric pressure. Depending on the cracking conditions, free-radical reactions lead to the formation of ethylene, propene,
butadiene, and aromatics (BTX: benzene, toluene, and xylenes) with different selectivities. The products obtained depend on feed
composition, hydrocarbon-to-steam ratio, cracking temperature, and residence time. The pyrolysis process also results in a slow
deposition of carbonaceous material (coke) on the reactor walls. Therefore, a steam cracking furnace requires a periodical decoking
phase, where a steam/air mixture is passed through the coils, converting the carbon deposit to carbon monoxide and carbon dioxide.
To be suitably flexible, an overall mathematical model for the simulation of pyrolysis reactors should include the mass, heat,
and momentum balance equations, together with physicochemical properties, firebox patterns, and accurate fluid dynamic
characterization. As it is usual for chemical reactors, the key feature remains the kinetic scheme, that is, reaction mechanisms,
rate expressions, and kinetic parameters. Only a detailed kinetic model can provide an accurate description of the chemical
evolution of the system over a wide range of feed and operating conditions. Thus, the complexity of the modeling of the pyrolysis
process is due both to the large number of species and elementary reactions involved and often to the difficulty of an accurate
characterization of the liquid feedstocks. In fact, a further challenge to model the cracking behavior of heavy liquid feedstocks is
related to a balanced level of molecular detail of the reaction network and of the feedstock characterization. To this aim,
commercial indexes of the feed are needed, at least in terms of specific density, PIONA, and boiling curve.
A large number of species and reactions characterize a wide range and comprehensive kinetic scheme of hydrocarbon pyrolysis.
Nevertheless, they are taking a great advantage from their modular and hierarchical structure, as well as from the mechanistic
analogies among similar reactions. This means that pyrolysis reactions can be studied by starting from the simpler systems and then
progressively extending the kinetic model to more complex cases. Pyrolysis reactions of large hydrocarbon species rapidly give rise
to small radicals and species. Hierarchically, the interactions of small hydrocarbon species (such as hydrogen, methane, ethane and
ethylene, and small radicals) are the key core of the kinetic scheme and they constitute a common ground in all the pyrolysis
systems. Moreover, pyrolysis reactions are also important in the high-temperature combustion of liquid fuels, such as gasoline,
kerosene, diesel, and jet propulsion fuels.
Similar kinetic schemes are useful to explain the formation of polycyclic aromatic hydrocarbons and carbon particles (soot) in
combustion processes, and they also allow to deal with the kinetic modeling of the formation of carbonaceous deposit on pyrolysis
coils (fouling process). The kinetic modeling of visbreaking and delayed coking and the pyrolysis of plastic wastes are final
examples of the interest of pyrolysis reactions.
Catalytic cracking processes are applied extensively in refineries to produce gasoline from crude oil (fluid catalytic cracking) and
by adding hydrogen in order to upgrade particular streams (hydrocracking).

Kinetic Modeling of Pyrolysis Reactions

An abundance of groundwork regarding both fundamental and applied chemical kinetics of pyrolysis processes is available in the
literature. Pioneering studies of free-radical pyrolysis reactions mostly refer to Rice, Herzfeld, and Kossiakoff.1–3 The excellent book
of Benson on Thermochemical Kinetics allowed to systematically analyze fundamental reactions of free radicals relevant to
hydrocarbon pyrolysis.4–10 Table 1 summarizes the main reaction classes of chain radical mechanisms for pyrolysis reactions,
that is, initiation, termination, and propagation reactions:

Reference Module in Chemistry, Molecular Sciences and Chemical Engineering http://dx.doi.org/10.1016/B978-0-12-409547-2.11542-2 1


2
Pyrolysis: Mathematical Modeling of Hydrocarbon Pyrolysis Reactions
Table 1 Main reaction classes in pyrolysis mechanism

1. Initiation and termination (or recombination) reactions


(a) Unimolecular: CH3 CH2 CH2 CH3 , C2 H5  + C2 H5 
(b) Bimolecular: CH3 CH3 + CH2 ¼ CH2 , C2 H5  + C2 H5 
2. Propagation reactions
(a) H-abstraction (metathesis) reactions

R + R 0 H , RH + R 0 ðe:g:, CH3  + CH3 CH3 , CH4 + C2 H5  Þ
(b) Addition of radicals to unsaturated molecules and alkyl radical decomposition
00
R + R0 H , R ðe:g:, CH3  + CH2 ¼ CHCH3 , CH3 CH2 CH CH3 Þ
(c) Alkyl radical isomerization reactions via (1–4) and (1,5) and (1,6) H transfers
 
R , R0 e:g:, CH2 CH2 CH2 CH2 CH3 , CH3 CH2 CH2 CH CH3 ;  CH2 ðCH2 Þ3 CH2 CH3 , CH3 ðCH2 Þ3 CH CH3 ;  CH2 ðCH2 Þ4 CH2 CH3 , CH3 ðCH2 Þ4 CH CH3
3. Purely molecular reactions
(a) Four- and six-center concerted reactions
00
RH , R 0 H + R H ðe:g:, C5 H10 , C2 H4 + C3 H6 Þ

RH , R 0 H ðe:g:, cycloC6 H12 , CH2 ¼ CH  CH2  CH2  CH2  CH3 ; CH2 ¼ CH  CH ¼ CH  CH2  CH3 , CH3  CH ¼ CH  CH ¼ CH  CH3 Þ
(b) Diels–Alder cycloaddition reactions
00
RH + R0 H , R H ðe:g:, C2 H4 + C4 H6 , cyclohexeneÞ
Pyrolysis: Mathematical Modeling of Hydrocarbon Pyrolysis Reactions 3

1. Chain initiation and termination reactions:


(a) The splitting of hydrocarbons at their weakest bonds produces free radicals.
(b)Chain ending occurs through recombination and/or disproportionation of radicals.
2. Chain propagation reactions:
(a) Free radicals abstract hydrogen atoms from the hydrocarbon to form saturated molecules and new free radicals.
(b)Free radicals can stabilize themselves by splitting off olefins with b-decompositions.
(c) Radicals with a long carbon skeleton can isomerize through internal H abstractions via 6- or 5-membered ring formation.
Only initiation and termination reactions modify the number of radicals, so these steps control the total concentration of reactive
free-radical intermediates. Usually, b-decomposition and H-abstraction reactions occur together in a chain propagation sequence.
As shown in Table 1, also molecular reactions can play a significant role. Typical examples are Diels–Alder cycloaddition
reactions and olefin isomerization, dehydrogenation, and ‘ene’ decomposition via four- and six-center reactions.
National Institute of Standards and Technology (NIST, formerly the National Bureau of Standards (NBS)) spent several efforts
in establishing a comprehensive database of gas-phase reactions. Initially, the focus was on combustion reactions. In fact,
combustion is largely a competition between oxidation and pyrolysis chemistry.11 Presently, the database includes essentially all
reported gas-phase kinetics studies involving elementary reactions.12 The attention to larger and more unsaturated species with the
ultimate goal of describing soot formation process further highlights the importance of pyrolysis reactions.

n-Butane Pyrolysis

The kinetic modeling of n-butane pyrolysis constitutes a simple initial example, useful for a better understanding of the pyrolysis
process. Figure 1 shows major and some minor products from n-butane decomposition, under isothermal conditions, at 1100 K
and 1 atm. Ethylene, propene, and methane are the main products, while only minor amounts of butene, ethane, acetylene,
butadiene, benzene, and cyclopentadiene are observed. These model predictions are reliable and validated by several comparisons
with experimental measurements.6 The kinetic scheme available on the CRECK Modeling Web site (http://creckmodeling.chem.
polimi.it/) is used for the simulations reported in Figure 1.
The prevailing initiation reaction is the breaking of a covalent C–C bond with the formation of two ethyl radicals. This initiation
process is highly sensitive to the stability of the formed radicals and its activation energy is equal to the bond dissociation energy
(BDE). In fact, the reverse, radical–radical recombination reaction is highly exothermic and it does not require activation energy.
C–C bonds are usually weaker than the C–H bonds; therefore, the initial formation of H radicals can be disregarded. Radical
initiation and radical recombination or termination reactions control the total radical concentration in the reacting system. Rate
constant of unimolecular decomposition reactions is conveniently estimated through the reverse radical recombination reactions.4
The reference kinetic parameters of unimolecular decomposition reactions of n-alkanes, referring to each single fission of C–C
bonds between secondary C atoms, are as follows:
 
kref ¼ 5:0  1016 expð81000=RT Þ s1

These kinetic parameters for the initiation reaction of n-butane to form two ethyl radicals

nC4 H10 , C2 H5  + C2 H5 

Figure 1 Main products from n-butane decomposition, at 1100 K and 1 atm (model predictions).
4 Pyrolysis: Mathematical Modeling of Hydrocarbon Pyrolysis Reactions

well agree with kinetic parameters suggested in the literature. About 2000 kcal/kmol of extra energy is required to split off a
terminal methyl group. Thus, the kinetic parameters of the second, less favored, initiation reaction of n-butane to form methyl and
1-propyl radical become
 
nC4 H10 , 1-C3 H7  + CH3  k ¼ 1017 expð83 500=RT Þ s1

where the frequency factor accounts for the hydrocarbon symmetry.


Once initiation reactions generate the propagating radicals, the H-abstraction reactions and the subsequent fast decomposition
of 1- and 2-butyl radicals explain the main primary products. Methyl and hydrogen are the effective H-abstracting radicals (R•) with
a favored formation of 2-C4H•9.

H + nC4 H10 , H2 + 1  C4 H9  k ¼ 1:29  104 T 2 expð6500=RT Þ m3 kmol1 s1

H + nC4 H10 , H2 + 2-C4 H9  k ¼ 0:86  104 T 2 expð4000=RT Þ m3 kmol1 s1

CH3  + nC4 H10 , CH4 + 1-C4 H9  k ¼ 2:34  102 T 2 expð7500=RT Þ m3 kmol1 s1

CH3  + nC4 H10 , CH4 + 2-C4 H9  k ¼ 1:56  102 T 2 expð5000=RT Þ m3 kmol1 s1

H abstractions of ethyl and propyl radicals are of limited importance (<10%), in these conditions.
At high temperatures (T > 1000 K), the b-decomposition reactions are the prevailing reaction path of alkyl radicals. Thus, butyl
radicals decompose quickly to form ethylene and propene:

1-C4 H9  , C2 H4 + C2 H5  k ¼ 3  1013 expð30000=RT Þ s1

2-C4 H9  , C3 H6 + CH3  k ¼ 3  1013 expð32000=RT Þ s1

At the usual pyrolysis temperatures, these rate constants are 106.5 and 107 s1, respectively, that is, the lifetime of alkyl radicals is
shorter than 106 s. Similarly, 1-propyl radical can form ethylene and methyl radical:

1-C3 H7  , C2 H4 + CH3  k ¼ 3  1013 expð32000=RT Þ s1

The successive dehydrogenation reaction of ethyl radical forms ethylene and H• radicals:

C2 H5  , C2 H4 + H k ¼ 1014 expð40000=RT Þ s1

These kinetic parameters clearly indicate that the dehydrogenations are less favored than the dealkylation ones. Only 3 wt% of
ethane is observed in the steam cracking products, confirming that ethyl radicals mainly decompose to form ethylene (70%).
Ethane is mostly formed via the H-abstraction reactions, while < 10% is due to the methyl recombination reaction. On the
contrary, propane formation is due both to the H-abstraction reactions of propyl radicals and to the recombination of methyl and
ethyl radicals.
Only at very high temperatures, dehydrogenation reactions of alkyl radicals become relevant, and their rate constants depend on
the type of radical and the number of hydrogen atoms involved:

1-C3 H7  , C3 H6 + H k ¼ 1:0  1014 expð37300=RT Þ s1

1-C4 H9  , 1-C4 H8 + H k ¼ 1:0  1014 expð39940=RT Þ s1

2-C4 H9  , 1-C4 H8 + H k ¼ 1:58  1013 expð39550=RT Þ s1

2-C4 H9  , 2-C4 H8 + H k ¼ 3:16  1012 expð36960=RT Þ s1

Due to the high activation energy of dehydrogenation reactions, butene formation is less favored than the b-decomposition
reactions and it is limited to < 2–3 wt%.
Figure 2 summarizes the pyrolysis mechanism of n-butane, also giving a schematic view of the relative importance of the
different primary reactions, at 1100 K, 1 atm, and 50% butane conversion. The dominant role of H-abstraction reactions, with the
prevailing importance of the 2-C4H•9 radical, is clearly highlighted.

Successive Pyrolysis Reactions


These reactions are the primary chain radical reactions, useful to explain the initial decomposition of n-butane. The maxima of
the profiles of butene and propene fractions in Figure 1 clearly indicate the presence and importance of successive reactions.
After the initial formation, butene and propene undergo successive pyrolysis reactions, mainly constituted by their interactions
with H, methyl, and other reacting species. To give a flavor of the complexity of the reacting mechanism, let us consider the
hydrogen attack on propene. Five different processes, two H-addition and three different H-abstraction reactions, can be
considered11:
Pyrolysis: Mathematical Modeling of Hydrocarbon Pyrolysis Reactions 5

Figure 2 Primary pyrolysis mechanism of n-butane. Arrow thickness indicates the relative importance of the different reaction paths.

H + C3 H6 , 1-C3 H7 
H + C3 H6 , 2-C3 H7 
H + C3 H6 , H2 + CH2 CHCH2  ðallylÞ
H + C3 H6 , H2 +  CHCHCH3 ð1-propenylÞ
H + C3 H6 , H2 + CH2 C CH3 ð2-propenylÞ

At the usual temperatures of steam cracking reactors, addition reactions prevail on the H-abstraction reactions. Among the H
abstractions, the formation of the resonantly stabilized allyl radical (a-C3H•5) is the dominant one, due to the lower BDE of the
allylic H atoms. Similarly, H abstractions on butenes mainly favor 1-methylallyl radical (CH3CH]CHCH•2 or a-C4H•7). The high
concentration of these resonantly stabilized radicals explains the successive formation of heavier species, via addition and
recombination reactions.5 Cyclopentadiene and benzene are typical secondary products, as shown by the inflection in their
mass fraction profiles of Figure 1. Once a significant amount of ethylene is formed, the reactivity of vinyl radical (C2H•3) explains
the formation of 1,3-butadiene (C4H6), via butenyl radicals (a-C4H•7 and p-C4H•7):

C2 H3  + C2 H4 , p-C4 H7 
p-C4 H7  , a-C4 H7 
a-C4 H7  , C4 H6 + H
p-C4 H7  , C4 H6 + H

The allyl and allyl type radicals can add to C2H4 giving rise to cyclopentene (cyC5H8) and methyl-cyclopentene. Successive
dehydrogenation and dealkylation reactions explain the formation of cyclopentadiene. cyC5H6 is also formed by the addition of
vinyl radical to butadiene:

C2 H3  + C4 H6 , CH3  + cyC5 H6

This reaction is very important to explain butadiene disappearance, and it constitutes the first step toward benzene formation
through the successive cycloaddition of ethylene to cyC5H6. Cyclopentadienyl (cyC5H•5) is another important resonantly stabilized
radical, which is mainly formed via H abstractions from cyclopentadiene. Since several years,5 the critical role of the recombination
reaction of cyC5H•5 radicals in the formation of naphthalene and heavy species is well defined:

cyC5 H5  + cyC5 H5  , C10 H8 + H + H

Successive reactions of vinyl and resonantly stabilized radical constitute a very critical and key submechanism for the correct
evaluation of ethylene selectivity. In fact, once the decomposition of the hydrocarbon feed is largely completed, the primary
products and mainly ethylene are involved in successive reactions with detrimental effect on the overall selectivity of the process.
High temperatures and contact times, that is, high severity or cracking extent, favor the production of ethene and benzene, while a
lower severity gives higher amounts of propene and unconverted n-butane.

Pyrolysis of Large Hydrocarbons: n-Decane

Through exploitation of structural or mechanistic analogies among similar reactions, it is possible to predict primary decompo-
sition products of heavy hydrocarbons by means of a limited number of independent kinetic parameters, as reported in Table 2.13,14
6 Pyrolysis: Mathematical Modeling of Hydrocarbon Pyrolysis Reactions

Table 2 Reference kinetic parameters of pyrolysis reactions (units are kmol, m, s, and kcal)

Initiation reactions: unimolecular decomposition of C–C bonds


CH3-Csec Csec-Csec Csec-Cter Csec-Cquat
5  1016 exp(83 500/RT ) 5  1016 exp(81 000/RT ) 5  1016 exp(80 000/RT ) 5  1016 exp(78 000/RT )
H-abstraction reactions of alkyl radicals
Primary H atom Secondary H atom Tertiary H atom
Primary radical 108 exp(13 500/RT ) 108 exp(11 200/RT ) 108 exp(9000/RT )
Secondary radical 108 exp(14 500/RT ) 108 exp(12 200/RT ) 108 exp(10 000/RT )
Tertiary radical 108 exp(15 000/RT ) 108 exp(12 700/RT ) 108 exp(10 500/RT )
Isomerization reactions (transfer of a primary H-atom a)
1–4 H transfer 1–5 H transfer 1–6 H transfer
Primary radical 1011 exp(20 600/RT ) 1.581010 exp(14 500/RT ) 3.16109 exp(19 500/RT )
Secondary radical 1011 exp(21 600/RT ) 1.581010 exp(15 500/RT ) 3.16109 exp(20 500/RT )
Tertiary radical 1011 exp(22 100/RT ) 1.581010 exp(16 000/RT ) 3.16109 exp(21 000/RT )
Alkyl radical decomposition reactions (to form primary radicals)
Primary radical Secondary radical Tertiary radical
1014 exp(30 000/RT ) 1014 exp(31 000/RT ) 1014 exp(31 500/RT )
Corrections of decomposition rates to form:
Methyl radical Secondary radical Tertiary radical Allyl radical
exp(2500/RT ) exp(1500/RT ) exp(2500/RT ) 0.316  exp(8000/RT )
a
Corrections for secondary and tertiary H-atoms are the same as for H-abstractions.

Once the values of the kinetic parameters of a reference reaction have been obtained experimentally and theoretically justified,
analogous reactions can be reasonably predicted without experimental investigation. The theoretical basis of this simplification lies
partially in the assumption that the forces between atoms are very short range,4 that is, in the order of magnitude of the bond
lengths. Thus, each atom contributes constant amounts to the molecule properties. The most widely adopted analogy concerns the
H-abstraction reactions: The rate constants for a single abstracted H atom are assumed to depend only on the abstracting radical
and on the type of H atom, while the frequency factor is assumed to depend only upon the abstracting radical.15
The pyrolysis of n-decane (nC10H22) provides a further example to better show the complexity of the primary pyrolysis reactions
of heavy molecules. There are five different C–C bonds in the n-decane chain, and therefore, there are five different initiation
reactions. For symmetry reasons, it is possible to assume k ¼ 1017 expð81000=RT Þ ðs1 Þ for the following initiation reactions:

nC10 H22 , 1-C8 H17  + C2 H5 


nC10 H22 , 1-C7 H15  + 1-C3 H7 
nC10 H22 , 1-C6 H13  + 1-C4 H9 

while only one half of this value is assumed for the initiation reaction to form two 1-C5H•11 radicals.
Finally, the rate parameters of the initiation reaction to form CH•3 and 1-C9H•19 are 1017 expð83500=RT Þ ðs1 Þ.
Once the radicals are formed in the reacting system, propagation reactions are mostly responsible of the initial feed decompo-
sition. Figure 3 shows the complete reaction path of H-abstraction reactions on n-decane and itself provides proof of the
complexity of the mechanism. All the different H-abstracting radicals (R•) can produce the five n-decyl radical isomers. These
radicals can then isomerize and decompose to form an alkene and a smaller radical.
The H-abstraction reactions on n-decane are estimated taking into consideration the presence of 16 secondary and six primary H
atoms, with the corresponding formation of four different secondary radicals and one primary n-decyl radical. Again, H• and CH•3
are the most effective attacking radicals, and the estimated rate constants for each of the five n-decyl isomers are

H + nC10 H22 , H2 + 1-C10 H21  k ¼ 6:43  103 T 2 expð6500=RT Þ m3 kmol1 s1

H + nC10 H22 , H2 + s-C10 H21  k ¼ 8:60  103 T 2 expð4000=RT Þ m3 kmol1 s1

CH3  + nC10 H22 , CH4 + 1-C10 H21  k ¼ 1:17  102 T 2 expð7500=RT Þ m3 kmol1 s1

CH3  + nC10 H22 , CH4 + s-C10 H21  k ¼ 1:56  102 T 2 expð4900=RT Þ m3 kmol1 s1

Figure 3 shows that decyl radicals can also isomerize. The isomerization reactions of 1-decyl radical to form 5- and 4-decyl
radicals, through 6- and 5-membered ring intermediates, are better depicted in Figure 4.
Due to the difference in activation energy, isomerization reactions of alkyl radicals are competitive at low temperatures, while
b-decomposition reactions prevail at very high temperatures, thus explaining the higher ethylene selectivity when temperature
increases.
A very useful and effective simplification of pyrolysis reactions refers to large alkyl radicals.14,16 In order to reduce the total
number of radicals as well as the total number of elementary reactions, all intermediate alkyl radicals higher than C4 are supposed
Pyrolysis: Mathematical Modeling of Hydrocarbon Pyrolysis Reactions 7

Figure 3 H-abstraction reactions of n-decane and successive isomerization and decomposition reactions of n-decyl radicals.

Figure 4 1-Decyl radical isomerization to form 5- and 4-decyl radicals, via (1,5) and (1,4) H transfers.

to be directly transformed (via isomerization and decomposition reactions) into their final products. The bimolecular
H-abstraction and recombination reactions of heavier alkyl radicals are negligible in comparison with their isomerization
and decomposition reactions, in the usual cracking conditions. This allows describing the fate of these m radicals with the
pseudo-steady-state assumption. The absence of interactions between heavy radicals and the remaining part of the cracking mixture
allows to substitute these radicals directly with their isomerization and decomposition products, without loss of accuracy but with a
significant reduction of the number of intermediate radicals and reactions to be considered in the overall kinetic scheme.

Pyrolysis of Liquid Feedstocks


The main complexity in the modeling of the pyrolysis of liquid hydrocarbon mixtures, such as naphthas, gas oils, and refinery
streams, is that their detailed composition is not available. They are complex mixtures of a huge number of different isomers of
normal and branched alkanes, cyclo-alkanes, and aromatics. PIONA (paraffins, iso-paraffins, naphthenes, olefins, and aromatics),
specific gravity, hydrogen-to-carbon ratio, and boiling curves (TBP curves or ASTM D86) are the commercial indexes commonly
used for their characterization. New analytic systems based on a double gas chromatography (GC GC) complemented by mass
spectrometry and/or ultraviolet ionization techniques further contribute to better identify the isomer structures and to reconstruct
the molecular composition of liquid feedstocks.

Automatic Generation of Reaction Mechanisms


The possibility of always applying the same general rules to different reaction classes in a systematic way is very useful for the
automatic generation of the primary reactions of the different isomers of heavy hydrocarbon species.17 Systematic works aiming at
developing an automatic generation of comprehensive kinetic models, both for pyrolysis and for combustion systems, are
extensively present in the literature.16,18,19 Thus, it is feasible to automatically generate large mechanistic schemes involving several
thousands of species and elementary reactions. Nonetheless, it is often convenient to adopt several simplifications and lumping
procedures in order to avoid an excessive number of chemical species and reactions.14
8 Pyrolysis: Mathematical Modeling of Hydrocarbon Pyrolysis Reactions

References
1. Kossiakoff, A.; Rice, F. O. J. Am. Chem. Soc. 1943, 65(4), 590–595.
2. Rice, F. O. J. Am. Chem. Soc. 1931, 53, 1959.
3. Rice, F. O.; Herzfeld, K. F. J. Am. Chem. Soc. 1934, 56(2), 284–289.
4. Benson, S. W. Thermochemical Kinetics, 2nd ed.; Wiley: New York, NY, 1976.
5. Dente, M.; Ranzi, E.; Goossens, A. G. Comput. Chem. Eng. 1979, 3(1), 61–75.
6. Dente, M.; Ranzi, E. Mathematical Modeling of Hydrocarbon Pyrolysis Reactions. In Pyrolysis: Theory and Industrial Practice; Albright, L. F.; Crynes, B. L.; Corcoran, W. H., Eds.;
Academic Press: New York, 1983; p 133.
7. Dente, M.; Bozzano, G.; Faravelli, T.; Marongiu, A.; Pierucci, S.; Ranzi, E. Adv. Chem. Eng. 2007, 32, 51–166.
8. Froment, G. F. Chem. Eng. Sci. 1992, 47(9), 2163–2177.
9. Poutsma, M. L. J. Anal. Appl. Pyrol. 2000, 54(1–2), 5–35.
10. Savage, P. E. J. Anal. Appl. Pyrolysis 2000, 54(1), 109–126.
11. Tsang, W. Ind. Eng. Chem. Res. 1992, 31(1), 3–8.
12. Manion, J. A.; Huie, R. E.; Levin, R. D.; Burgess, D. R., Jr.; Orkin, V. L.; Tsang, W.; McGivern, W. S.; Hudgens, J. W.; Knyazev, V. D.; Atkinson, D. B.; Chai, E.; Tereza, A. M.;
Lin, C.-Y.; Allison, T. C.; Mallard, W. G.; Westley, F.; Herron, J. T.; Hampson, R. F.; Frizzell, D. H. NIST Chemical Kinetics Database, NIST Standard Reference Database
17, Version 7.0 (Web Version), Release 1.6.8, Data version 2013.03; National Institute of Standards and Technology: Gaithersburg, MD, 2013;. 20899–8320 http://kinetics.nist.
gov/.
13. Ranzi, E.; Dente, M.; Pierucci, S.; Biardi, G. Ind. Eng. Chem. Fund. 1983, 22(1), 132–139.
14. Ranzi, E.; Dente, M.; Goldaniga, A.; Bozzano, G.; Faravelli, T. Prog. Energy Combust. Sci. 2001, 27(1), 99–139.
15. Ranzi, E.; Dente, M.; Faravelli, T.; Pennati, G. Combust. Sci. Technol. 1993, 95(1–6), 1–50.
16. Van Geem, K. M.; Reyniers, M. F.; Marin, G. B. Oil Gas Sci. Technol. 2008, 63(1), 79–94.
17. Pierucci, S.; Ranzi, E.; Dente, M.; Barendregt, S. A kinetic generator of hydrocarbon pyrolysis mechanisms. Comput. Aided Chem. Eng. 2005, 20, 241–246.
18. Matheu, D. M.; Dean, A. M.; Grenda, J. M.; Green, W. H. Mechanism Generation with Integrated Pressure dependence: A New Model for Methane Pyrolysis. J. Phys. Chem. A
2003, 107, 8552–8565.
19. Warth, V.; Battin-Leclerc, F.; Fournet, R.; Glaude, P. A.; Come, G. M.; Scacchi, G. Comput. Chem. 2000, 24, 541–560.

Potrebbero piacerti anche