Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Journal of Analytical and Applied Pyrolysis 123 (2017) 204–215

Contents lists available at ScienceDirect

Journal of Analytical and Applied Pyrolysis


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

Thermal cracking of 1-n-butyldecalin at high pressure (100 bar)


Darwin A. Rakotoalimanana a , Roda Bounaceur a,∗ , Baptiste Sirjean a , Françoise Béhar b ,
Valérie Burklé-Vitzthum a , Paul-Marie Marquaire a
a
Laboratoire Réactions et Génie des Procédés, UMR 7274, CNRS-UL, Nancy, F-54001, France
b
Total Exploration and Production, Paris La Défense, F-92078, France

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

Article history: The pyrolysis of 1-n-butyldecalin at high pressure (100 bar) was studied in gold sealed tube reactors at
Received 24 May 2016 400 ◦ C. It leads to 4 main chemical families of products in order of predominance: (a) normal alkanes
Received in revised form (methane, ethane, propane and n-butane), (b) bicyclics (decalin and 1-methyldecalin), (c) hydroaromat-
10 November 2016
ics (tetralin and 5-methyltetralin) and (d) diaramatics (naphthalene and 1-methylnaphthalene). This
Accepted 8 December 2016
distribution of products suggests that breaking of side alkyl chain and aromatization reactions should
Available online 10 December 2016
prevail at high pressure and low temperature. Reaction pathways are effectively proposed to account for
the formation of main products. Both normal alkanes (except CH4 and C2 H6 ) and bicyclics should mainly
Keywords:
Butyldecalin
come from C C bond breaking reactions on the side alkyl chain of 1-n-butyldecalin. Methane and ethane
Pyrolysis should come from further decomposition of fuel radicals, including ring opening through beta-scissions
Naphthene given the absence of any C12 and C13 compounds in similar quantities. Both hydroaromatics and diaromat-
High pressure ics are secondary products, which result from aromatization reactions of decalin and 1-methyldecalin. In
Reaction pathways our conditions, 1-n-butyldecalin and n-butylcyclohexane have a similar thermal reactivity. A higher yield
in n-butane was observed in the case of 1-n-butyldecalin. In comparison with the pyrolysis of decalin at
low pressure (≤1 atm) and high temperature (>500 ◦ C), the production of naphthalene and the absence
of any BTX (Benzene, Toluene and Xylene) in detectable quantities is here worth mentioning.
© 2016 Elsevier B.V. All rights reserved.

1. Introduction cyclic model compound i.e. n-butylcyclohexane. It completed other


previous results on alkylcyclohexanes e.g. [9,10] at high pres-
An estimation of the thermal evolution during geological time sure (100–400 bar). These studies confirm altogether the inductive
(millions of years) of reservoir oils requires a good knowledge of the effect of the side alkyl chain of the thermal reactivity of an alkylcy-
kinetics of the thermal cracking reactions [1] occurring at high pres- clohexane in these conditions. This is also supported by mechanistic
sure (100–1000 bar) and low temperature (150–220 ◦ C). The first approaches, either via quantum chemical calculations e.g. [12,13]
step consists in simulating those reactions at laboratory scale by or free-radical detailed mechanisms e.g. [14,15]. Breaking of side
studying kinetically the pyrolysis of oils e.g. [2], oil fractions e.g. [3] alkyl chain and aromatization reactions effectively prevail over ring
or model compounds e.g. [4–8] at high pressure (100–700 bar) but opening at reservoir conditions, which leads to a large amount of
at higher temperature (300–450 ◦ C) to obtain acceptable reaction normal alkanes, as main pyrolysis products alongside with unsub-
times (from few hours to 1 month). stistuted and methylated cyclics in lower quantities.
As petroleum constitutes a very complex chemical system Tissot and Welte [1] showed the predominance of monocyclic
with hundreds of compounds, it is necessary to firstly elucidate and also bicyclic compounds among naphthenes in most reservoir
the chemistry of thermal cracking of each main chemical family oils. Some examples of compounds are depicted in Fig. 1.
i.e. alkanes e.g. [4,5], aromatics [7,8] and naphthenes e.g. [9,10]. Yet, until now, the literature on the thermal cracking of bicyclic
Our previous paper [11] highlighted the lack of knowledge of naphthene only concerns the decalin (Fig. 1) at higher tempera-
the thermal cracking of naphthenes and presented experimental ture (≥450 ◦ C) and mostly at atmospheric pressure (oil refining
results of the pyrolysis at 100 bar and around 400 ◦ C of a mono- conditions). Bredael and Rietvele [16] initiated the study of the
pyrolysis of decalin to compare it with tetralin in terms of yields
in benzene, toluene and xylene (BTX) for steam-cracking process.
By operating at 700–950 ◦ C and 0.5 bar in heated quartz reac-
∗ Corresponding author. tor, they showed that the thermal cracking of decalin results in
E-mail address: roda.bounaceur@univ-lorraine.fr (R. Bounaceur).

http://dx.doi.org/10.1016/j.jaap.2016.12.001
0165-2370/© 2016 Elsevier B.V. All rights reserved.
D.A. Rakotoalimanana et al. / Journal of Analytical and Applied Pyrolysis 123 (2017) 204–215 205

Fig. 1. Structure of di- and tricyclic compounds.

a ring opening in these conditions and ultimately leads to large Table 1


Structure of 1-n-butyldecalin (BD).
yields in BTX with an optimum at 900 ◦ C. Virk et al. [17] compared
the pyrolysis of cyclohexane, decalin and perhydrophenanthrene Compound Formula Structure M (g/mol) Densitya
to propose reaction pathways based on free-radical mechanisms
leading to benzene from decalin and perhydrophenanthrene.
Ondrushka et al. [18] studied the pyrolysis of decalin in a BD C14H26 194 0.849
tubular quartz reactor (600–800 ◦ C, 1 bar) and focused on the
formation of olefins (unsaturated hydrocarbons) and particularly
3-methylenecyclohexene to propose some free-radical mecha- a
Determined at 20 ◦ C and 1 bar.
nisms. They also noticed the absence of naphthalene among the
pyrolysis products in their conditions. In supercritical conditions Table 2
(550–750 ◦ C, 2–100 bar), Stewart et al. [19] compared the pyroly- Mass balances of pyrolysis of 1-BD in gold tubes.
sis of n-decane, decalin and tetralin. At 414 bar and between 460
t Charge Gas BD C6+ DCM total Conversion
and 540 ◦ C, decalin mainly leads to the formation of small normal h mg mg/g %
alkanes (ie. methane, propane, propane and butane) but also to
0 100 0 1000.0 0 0 1000.0 0
methylcyclohexene and methylhexahydroindane. They showed a 9 101.7 22.7 902.6 38.4 0 963.7 9.7
tendency of cycle contraction, which turns out to be even more 24 102.0 48.1 821.6 92.5 0.9 963.0 17.8
favored at higher pressure, which was supported by previous 48 77.0 125.8 698.0 144.0 1.5 969.3 30.2
work of Metzger et al. [20]. Yu and Eser [21] proposed decom- 72 77.0 171.2 584.3 207.4 2.6 965.5 41.6
144 52.4 242.9 310.0 421.9 3.1 978.0 69.0
position pathways of decalin based on free-radical mechanisms
to describe the formation of these products and pointed out that
isomerization reactions are significant at high pressure (450 ◦ C,
23–75 bar). configuration isomers (e.g. 2-n-butyldecalin). The properties of BD
Following works aimed at modeling the thermal cracking of are given in Table 1.
decalin at atmospheric pressure to better understand the thermal
cracking and the oxidation at high temperature (>500 ◦ C) of fuel or 2.1. Pyrolysis of 1-n-butyldecalin (1-BD)
jet fuels. Chae and Violi [22] implemented an ab initio approach to
elucidate main decomposition reaction pathways at atmospheric The recently-developed protocol [11] was implemented to
pressure and in the temperature range of 550–1300 ◦ C leading kinetically study the pyrolysis of BD at high pressure (100 bar) and
to BTX compounds. Dagaut et al. [23] proposed semi-detailed 400 ◦ C. The chosen conditions are provided in Table 2.
free-radical mechanism (11,000 reactions) validated at high tem- Pyrolysis experiments were carried out in gold sealed tubes in an
perature (>500 ◦ C) and atmospheric pressure with the results of isobaric and isothermal regime. Given the chemical inertia, ductil-
Zámostný et al. [24] among others. This model was recently adapted ity and malleability of gold material, these reactors appear to be the
by Zhu et al. [25] to account for the results obtained at higher most convenient to kinetic studies of high pressure pyrolysis e.g.
pressure i.e. 20–50 bar. Another smaller detailed free-radical model [6–8,27,28]. A preliminary step consisted in eliminating any pos-
(1774 reactions) was developed by Zeng et al. [26] at atmospheric sible impurities of the walls of the tubes. The tubes were heated
pressure and higher temperature i.e. 650–1230 ◦ C to show that (700 ◦ C) for 1 h, boiled in hydrochloric acid (1N) and finally washed
the formation of monocyclic aromatic hydrocarbons prevails over with distillate water and acetone. The gold tubes were loaded with
the formation of polycyclic aromatic hydrocarbons at these condi- a charge of 50–100 mg and sealed by an ultrasonic welding sys-
tions. tem (Telsonic) in an anoxic environment (i.e. N2 atmosphere) to
The current kinetic study of the pyrolysis of 1-n-butyldecalin avoid any air contamination which would lead to the formation of
at 100 bar provides supplementary insights in better understand- unwanted compounds such as oxygenated hydrocarbons. This step
ing the thermal cracking of naphthenes at high pressure and was effectively realized in a glove box, whose oxygen concentration
low temperature (400 ◦ C). The possible kinetic effect of a second was controlled below 10 ppm.
cycle is here investigated in comparison with n-butylcyclohexane For each experiment, 2–6 tubes were confined into a N2 -
[11]. Besides, comparison with the thermal cracking of decalin pressurized autoclave to apply to the chosen pyrolysis conditions.
at high pressure is also made in this paper. Some reaction path- This autoclave was then placed in a preheated oven, where temper-
ways are proposed to account for the formation of main pyrolysis ature (confidence interval of 1 ◦ C) and pressure (confidence interval
products. of 0.1 bar) were kept constant: real-time values were measured and
recorded. Once the temperature in the autoclave core was stabi-
lized at the chosen value, pyrolysis time was initiated. At the end
2. Material and methods of the pyrolysis time, the hot autoclave was thermally quenched
in a water bath at room temperature (around 20 ◦ C) and then
BD (C14 H26 ) was synthesized by BOC Sciences Creatives slowly depressurized to avoid any possible rupture of the tubes.
Dynamics Inc. (USA). The purity (≥98%) was certified by gas chro- The extracted tubes were eventually checked for any leakage of
matography – mass spectrometer (GCMS). Impurities were mostly products by weighing.
206 D.A. Rakotoalimanana et al. / Journal of Analytical and Applied Pyrolysis 123 (2017) 204–215

Fig. 2. Chromatogram of C6+ HC pyrolysis products at 400 ◦ C, 100 bar and after 144 h.

2.2. Analysis of products Table 3


Main hydrocarbon products at 400 ◦ C and 100 bar.

The same analytical protocol as for the study of the thermal Alkanes
cracking of n-butylcyclohexane at 100 bar [11] was followed to
Methane (C1 ) Ethane (C2 ) n-Propane (C3 ) n-Butane (C4 )
recover and analyze pyrolysis products. For each pyrolysis condi- n-Pentane Isobutane Isopentane
tion, the analysis of the gaseous and liquid products was separated.
Thus, at least 2 tubes were treated. Alkenes

Ethylene Propylene
2.2.1. Gaseous products
The gold tubes were pierced within a closed system. The gaseous Cyclics

products were recovered through a vacuum line equipped with a


vacuum pump leading to a fixed-volume sealed tank. Thanks to a
pressure gauge, the total quantity of gas generated during pyrolysis
was determined before being injected in GC (HP 6890). GC was
equipped with two detectors: a flame ionization detector (FID) for Decalin (D) 1-Methyldecalin (MD) 1-n-Butyldecalin (BD)
C1 -C5 hydrocarbons (HC) and a thermal conductivity detector (TCD)
for non-HC compounds (H2 and N2 in our study). GC was calibrated Hydroaromatics
with a standard gas.

2.2.2. Liquid products


The first step consisted in slowly releasing the gaseous phase
Tetralin (T) 5-Methyltetraline (MT)
while avoiding any loss of the volatile products. The gold tubes
were actually pierced in a glass device filled with a “cold” pen- Diaromatics
tane solution, whose temperature was maintained below 15 ◦ C by
an external dry ice bath. The two following steps corresponded
to two extractions of the first solution containing the tube cut up
(successively with pentane and dichloromethane) under reflux for
1 h followed by a filtration step. Both extraction volumes were Naphthalene (N) 1-Methylnaphthalene (MN)
exactly weighed along the whole process, before and after sam-
pling aliquots for composition analyses, i.e. molecular identification
and quantification. Hydrocarbon compounds soluble in pentane not fully quantified, as pentane corresponds to our main solvent.
were firstly identified by a HP 5973 mass spectrometer Mass Selec- However, gas analysis provided an estimation of a part of C5
tive Detector (MSD) coupled to a GC, and then quantified by a gaseous HC. Furthermore, the processing of the gold tube during
HP 6890 GC FID. Complete setting of both analytical apparatus the experiments (piercing, cutting) could account for some mass
was accurately described in our previous paper [11]. No molecu- loss (estimated to 1 or 2 mg/g).
lar identification of compounds soluble in dichloromethane was
performed, as they correspond to a very minor fraction (Table 2). 3. Results

2.3. Mass balances 3.1. Products

Systematic mass balances were calculated for each experiment Fig. 2 shows an overview of the main pyrolysis products in the
to guarantee a quasi-full recovery of the pyrolysis products. Mean “C6+ ” fraction at 400 ◦ C, 100 bar and after 144 h (conversion of about
values corresponding to at least 2 experiments are given in Table 2. 70%). It is worth mentioning the quite low number of pyrolysis
4 fractions were considered in the mass balance: “gas” fraction products: 6 main products can be distinguished. The products (2)
(C1 –C5 HC and H2 ), “C6+ ” fraction (C6+ HC soluble in pentane), “BD” and (3) could not be identified by mass spectrometer with a suffi-
fraction (remaining reactant) and “DCM” fraction (C14+ HC not solu- cient matching percentage. However, structure related compounds
ble in pentane but in dichloromethane). It is worth mentioning that are proposed and are actually quite consistent with decalin related
no black solid residue was observed on the walls of the gold tubes in literature e.g. [16].
our pyrolysis conditions contrary to the case of n-butylcyclohexane The exact structures of the main hydrocarbon products of the
[11]. pyrolysis of 1-BD at 400 ◦ C and 100 bar are provided in Table 3.
We obtained globally good mass balances (above 96%). Some The “gas” fraction is mostly consisting of normal- and iso-alkanes
uncertainties still remained: C5 hydrocarbon compounds were alongside with alkenes and hydrogen (not depicted in Table 3).
D.A. Rakotoalimanana et al. / Journal of Analytical and Applied Pyrolysis 123 (2017) 204–215 207

100 butylcyclohexane. Concerning the higher yield in n-butane in the


case of 1-n-butyldecalin, a possible explanation could be inferred
80 from a mechanistic point of view along BDE values of both fuels
Conversion (%)

(Fig. 8). In a previous paper [15], we showed that normal alka-


60 BCH nes mainly come from breaking reaction of the side alkyl chain
of n-butylcyclohexane (␤-scission reaction of butylcyclohexyl rad-
40 BD ical followed by H-transfer reaction) at 100 bar and around 400 ◦ C.
Thus, an analogy could be made here.
20 Taking into account the double degeneracy of n-
butylcyclohexane due to its structural symmetry, there are
0 two possible pathways for both fuels, which eventually lead to
0 24 48 72 96 120 144 n-butane. Regarding radical sites of fuel radicals, a slight difference
Time (h) could be here highlighted. In the case of 1-n-butyldecalin, one of
Fig. 3. Conversions at 400 ◦ C and 100 bar of 1-n-butyldecalin (1), n-
them corresponds to a radical center on a tertiary carbon, whose
butylcyclohexane (2). formation is de facto favored, as the corresponding hydrogen atom
is more labile (lower C H bond BDE). However, this is not enough
to explain the clear difference in yield in n-butane. All reaction
In the “C6+ ” fraction, 3 chemical families are identified: cyclics
pathways such as further decomposition of fuel radical following
(derivatives of decalin), hydroaromatics (derivatives of tetralin)
ring opening by beta-scission should be also considered (detailed
and diaromatics (derivatives of naphthalene).
in next section).
The distribution of the products suggests two main reaction
Fig. 6 depicts the corresponding curves vs. time of the
pathways: breakings (in alpha and beta position from the cycle)
“C6+ ”fraction (decalin, tetralin, naphthalene, 1-methyldecalin, 5-
of the alkyl side chain and aromatization of the saturated rings. It
methyltetralin and 1-methylnaphthalene).
is worth emphasizing the presence of naphthalene, which is not
Among the “C6+ ” HC products, unsubstituted compounds prevail
detected (or traces) during the pyrolysis of decalin at higher tem-
over methylated compounds. We could here infer that dealky-
perature (>500 ◦ C) and 50 bar. In these conditions, decalin rather
lation reactions are relatively more significant than breaking
leads to BTX compounds by ring opening [19], which are absent in
reactions in ␤-position from the cycle. Concerning the unsubsti-
our distribution of products.
tuted hydrocarbons, we have decalin > tetralin > naphthalene, in
order of predominance. This shows that aromatization reactions
3.2. Conversion (aromatization of one cycle at a time) are also major degradations
pathways of 1-n-butyldecalin at 400 ◦ C. This is also supported by
The conversion of BD was calculated by difference between the the profiles of naphthalene and 1-methylnaphthelene, with a flat
initial load of reactant and its remaining quantity after pyroly- tangent from the diagram’s zero.
sis. The initial load reactant was determined by direct weighing
whereas the remaining quantity was obtained by integration of
the corresponding GC peak. Results (mean values) as time func- 3.4. Trends of products organized by chemical family
tion, at 400 ◦ C and 100 bar, are given in Fig. 3 (lines corresponding
to trends). In the meantime, results are compared with the con- Balances of chemical families are usually done in petroleum geo-
version of n-butylcyclohexane in the same conditions [11]. At chemistry. This is illustrated in Fig. 7 for the 1-n-butylcyclohexane
400 ◦ C and 100 bar, 1-n-butyldecalin is slightly more stable than n- and compared with the case of n-butylcyclohexane. At 400 ◦ C,
butylcyclohexane. This small difference could be explained by the we obtained a large amount of normal alkanes (more than 40
presence of both hydroaromatics and diaromatics, which have an molar% after 144 h). In order of predominance, we had then
inhibiting effect on the thermal cracking of normal alkanes [29,30]. cyclics > hydroaromatics > diaromatics. This is actually similar to
Thus, the second cycle seems to have no significant kinetic effect the case of n-butylcyclohexane at 400 ◦ C, which confirms that both
in comparison with n-butylcyclohexane (BCH) in our experimental model compounds have similar chemistry of thermal cracking at
conditions. high pressure. Subsequently, the second cycle of 1-n-butyldecalin
does not have any significant effect, in comparison with n-
butylcyclohexane.
3.3. Yields of products

The yields (in molar fraction) of the main pyrolysis products at 4. Discussion
400 ◦ C as time function are given in Table 4. Those compounds cor-
respond to the two major fractions “Gas” and “C6+ ”. Fig. 4 depicts 4.1. Bond dissociation energies
the corresponding curves vs. time of products of the “gas” frac-
tion (methane, ethane, propane and n-butane). Previous results on Bond dissociation energies (BDE) of 1-n butyldecalin (in the
pyrolysis of n-butylcyclohexane at 400 ◦ C and 100 bar [11] are also trans configuration) were computed using Gaussian09 software
reported in Fig. 4 to draw a parallel between the two compounds. [31]. For the C H and C C bonds located outside the decalin bicycle,
In the “gas” fraction at 400 ◦ C, we obtain propane > n- the high-level composite CBS-QB3 method [32] was applied. The
butane ≈ ethane > methane, in order of predominance. This is bond breaking of C C bonds belonging to the decalin bicycle leads
consistent with the presence of unsubstituted and methy- to the formation of singlet diradicals. To compute the BDE of these
lated compounds in “C6+ ” fraction. In comparison with n- electronic structures, the CBS-QB3’ method, modified for open shell
butylcyclohexane, the yields of normal alkanes (except for n single biradicals [33], was used. It should be noted that the BDE of
butane) are smaller. This difference could suggest that the side acyclic hydrocarbons is equivalent to the activation energy required
alkyl chain is less subjected to breaking reactions in the case of 1- to break the bond because the reverse reaction, i.e., the combi-
n-butyldecalin (Fig. 5). This is actually supported by the calculation nation, is barrierless. In the case of initial C C bond scissions in
of BDE (Fig. 8): the C C bonds on the side alkyl of 1-n-butyldecalin cyclic alkanes, the reverse combination of the biradical into the
are globally stronger than the corresponding C C bonds of n- cycloalkane features a small reaction barrier. Therefore, the energy
208 D.A. Rakotoalimanana et al. / Journal of Analytical and Applied Pyrolysis 123 (2017) 204–215

Table 4
Yields of main products at 400 ◦ C and 100 bar vs. time.

Products

T (◦ C) 400
t (h) 9 24 48 72 144
Conversion (%) 9.7 17.8 30.2 41.6 69.0

“Gas” fraction
H2 3.0 4.9 5.2 4.4 2.6
CH4 1.2 2.6 4.0 5.1 7.0
C2 H6 1.5 3.0 5.8 7.4 10.4
C2 H4 0.0 0.0 0.0 0.0 0.0
C3 H8 2.2 5.2 11.5 13.9 17.1
C3 H6 0.1 0.1 0.1 0.1 0.1
i-C4 H10 0.6 0.2 0.4 1.2 0.4
n-C4 H10 2.1 4.0 8.1 10.2 11.5
i-C5 H12 0.0 0.0 0.1 0.2 0.3
n-C5 H12 0.0 0.1 0.4 0.5 0.6

“C6+ ” fraction
Decalin 1.3 4.1 6.1 8.0 10.8
1-Methyldecalin 0.3 0.8 1.2 1.5 2.0
Tetralin 0.6 1.6 3.0 4.3 8.6
5-Methyltetralin 0.3 0.6 1.6 1.9 4.1
Naphthalene 0.0 0.2 0.7 1.1 3.6
1-Methylnaphthalene 0.0 0.0 0.2 0.4 2.8

N-ALKANES
Methane Ethane
15 15
Mole fracon (%)

Mole fracon (%)

10 10

5 5

0 0
0 24 48 72 96 120 144 0 24 48 72 96 120 144
Time (h) Time (h)

Propane n-Butane
25 15
Mole fracon (%)

Mole fracon (%)

20
10
15
10 5
5
0 0
0 24 48 72 96 120 144 0 24 48 72 96 120 144
Time (h) Time (h)
Fig. 4. Profiles of “gas fraction” at 400 ◦ C and 100 bar: BD (1) and BCH (2) (trends in dotted line).

Fig. 5. ␤-scission reactions on n-butylcyclohexane (1) and 1-n-butyldecaline (2) leading to n-butane.
D.A. Rakotoalimanana et al. / Journal of Analytical and Applied Pyrolysis 123 (2017) 204–215 209

CYCLICS HYDROAROMATICS DIAROMATICS


Decalin Tetralin Naphthalene
15 10 4

Mole fracon (%)

Mole fracon (%)


Mole fracon (%)
8 3
10
6
2
4
5
2 1
0 0 0
0 24 48 72 96 120 144 0 24 48 72 96 120 144 0 24 48 72 96 120 144
Time (h) Time (h) Time (h)

1-Methyldecalin 5-Metyltetralin 1-Methylnaphthalene


2.5 5 3

Mole fracon (%)

Mole fracon (%)


Mole fracon (%)

2.0 4
2
1.5 3
1.0 2
1
0.5 1
0.0 0 0
0 24 48 72 96 120 144 0 24 48 72 96 120 144 0 24 48 72 96 120 144
Time (h) Time (h) Time (h)

Fig. 6. Profiles of “C6+ fraction” at 400 ◦ C and 100 bar (trends in dotted line).

BD BCH
50 70
60
Mole fracon (%)

Mole fracon (%)

40
Alkanes 50
30 40 Alkanes
Cyclics (no BD)
20 30 Cyclics (no BCH)
Hydroaromacs
20
10 Diaromacs Aromacs
10
0 0
0 24 48 72 96 120 144 0 24 48 72 96 120 144
Time (h) Time (h)

Fig. 7. Evolution of chemical families versus time at 400 C and 100 bar (trends in dotted lines).

Fig. 8. Bond dissociation energies of 1-n-butyldecalin computed at the CBS-QB3 and CBS-QB3 levels of calculations (see text). (a) C C and (b) C H calculated bond
dissociation energies.
210 D.A. Rakotoalimanana et al. / Journal of Analytical and Applied Pyrolysis 123 (2017) 204–215

required to break the C C bond of the ring is an activation energy compounds have not been detected, their concentrations are very
defined as the energy difference between the transition state (TS) low.
and the reactant [34]. In this study, the BDE leading to biradicals
were calculated as an activation energy (HTS – Hreactant at 298 K). 4.2.2. Formation of cyclics
Fig. 8 presents the results of the BDE calculations. Cyclics should mainly come from C C bond breaking reactions
on the side alkyl chain of 1-n-butyldecalin, as normal alkanes.
4.2. Reaction pathways More precisely, radicals R2 and R3 (Fig. 9) undergo respectively
C C bond ␤-scission to form, on the one hand, propylene and 1-
In a similar way to the case of n-butylcyclohexane [11,15], some methyldecalin, and on the other hand, but-1-ene and decalin after
reaction pathways based on free-radical mechanisms are given metathesis (Fig. 10). As in the case of n-butylcyclohexane [15],
here to explain the formation of normal alkanes and main “C6+ ” alkenes should be quickly consumed by hydrogenation (due to
products i.e. cyclics, hydroaromatics and diaromatics. The reaction the presence of H2 in the reaction medium) and addition reac-
schemes proposed in this subsection should provide guidelines to tions with other radicals (e.g. alkyl radicals). H2 is effectively
an eventual detailed kinetic modeling of the thermal cracking of the major gaseous product at the beginning of the reaction and
1-n-butyldecalin at high pressure (>50 bar), as the development after 18 h, its molar fraction decreases as a function of time
of a detailed kinetic model is not the scope of this work. Aside (Table 4).
[15], two previous works of modeling on similar fuels are worth Another reaction pathway would be the initiations of 1-n-
here being mentioned. Dagaut et al. [23] developed an 11.000 butyldecalin by unimolecular C C bonds homolysis, followed by
reaction model to describe the decalin oxidation and pyrolysis H-transfer reactions. However, it was previously showed [6] that H-
between 480 ◦ C and 1080 ◦ C and at 1 and 10 bar. For similar con- transfer reactions prevail at high pressure over unimolecular bond
ditions (650–1330 ◦ C, 0,04–1 bar), Zheng et al. [26] also recently homolysis.
built a model consisting of 1774 reactions to describe decalin
pyrolysis. Out of structure-related point about decalin and 1-n- 4.2.3. Formation of hydroaromatics
butyldecalin (asymmetry and presence of a side alkyl chain), one The aromatization pathways of decalin and 1-methyldecalin
clear difference remains: conditions of pressure and temperature leading respectively to tetralin and 5-methyltetralin are similar to
as explained in Bounaceur et al. [35] who studied the pyrolysis the aromatization pathways of alkylcyclohexanes during the pyrol-
of n-octane at temperature around 400 ◦ C and pressure ranging ysis of n-butylcyclohexane at high pressure [11]. Different steps are
from 1 to 1000 bar. The authors clearly explained that, depend- depicted in Fig. 11 and include successive reactions of H-transfer
ing on temperature and pressure range, both the chemistry and and C H bond ␤-scission. Given the presence of H• and CH3 • in the
the chemical compounds are different. This is actually illustrated reaction medium, we can also assume an interconversion between
by the difference of product distribution. Zeng et al. observed unsubstituted and methylated compounds by ipso-addition reac-
relatively high yields in monoaromatics and polyaroaromatics tions.
whereas these compounds correspond to the small fraction in our
study. 4.2.4. Formation of diaromatics
The formation of diaromatics from hydroaromatics constitutes
4.2.1. Formation of normal alkanes the step that follows the aromatization of cyclics into hydroaromat-
n-Butane and n-Propane should mainly come from C C bond ics. Thus, the formation of diaromatics follows the same reaction
breaking reactions on the side alkyl chain of 1-n-butyldecalin. pathways as already described in 4.1.2 (Fig. 12).
More precisely, 1-n-butyldecalin is firstly initiated by H-transfer
reactions, leading to the formation of 14 cyclic radicals (Fig. 9). 4.3. Comparison with other studies
Radicals R6 and R5 then undergo respectively C C bond ␤-scission
to form, on the one hand, octalin and n-butane, and on the other 4.3.1. Effect of the number of cycles on reactivity
hand, 1-methylenedecalin and n-propane after metathesis. It is No significant difference was observed between the thermal
worth mentioning that the side products such as octalin and 1- conversion of n-butylcyclohexane and 1-n-butyldecalin at 400 ◦ C
methylenedecalin were not detected among pyrolysis products and 100 bar. Furthermore, no significant difference was found
but constitutes precursors to hydroaromatics and/or diaromatics between the yields of normal alkanes, except for n-butane, which
(described in later subsections). is favored in the case of 1-n-butyldecalin.
Concerning the formation of methane and ethane, such direct However, considering the reaction pathways of both n-
pathways of dealkylation do not seem justified, as no C12 and C13 butylcyclohexane [11,15] and 1-n-butyldecalin, some differences
were observed in significant quantities among pyrolysis products. could be observed, as 1-n-butyldecalin leads to three tertiary
Methane and ethane should come from further decomposition of radicals, whose formation is favored. Two consequences could
fuel radicals, including ring opening (Fig. 9) by beta-scission, which be drawn from this point: firstly, it should favor the formation
results in the formation of very reactive intermediary radicals (with of n-butane (Subsection 3.3) and also free-radical aromati-
an unsaturation). In a similar way to n-butylcyclohexane [15], oth- zation reactions. This brings us to another major difference
ers reactions could occur such as isomerization (intramolecular with n-butylcyclohexane, which quickly forms n-butylbenzene at
relocalization of radical site) or resonance equilibrium (when the 400 ◦ C and 100 bar. In the case of 1-n-butyldecalin, neither 5-n-
radical site is located near the unsaturation). As already showed butyltetralin nor 1-n-butylnaphthalene was observed in the same
in [15], the resulting “network” of reactions corresponds to a sec- conditions.
ondary source of alkanes. We have not detected any compound
with two or more unsaturation as diene or triene. Then, as shown in 4.3.2. Conditions high temperature/low pressure
Fig. 9, these compounds can react via addition with radicals becom- As aforementioned, comparisons of our current results with
ing another source of alkanes or are unresolved compound. For each other works [17,16,19,21,22] on the thermal cracking of decalin
pyrolysis condition, 2–4 tubes were heated simultaneously and at high temperature (>500 ◦ C) and low pressure (≈1 bar) empha-
analyzed for a better consistency of the experimental results. Glob- size two major differences. In these conditions, the pyrolysis
ally, mass balances were mainly were between 96% and 98%, which of decalin does not produce naphthalene but rather BTX
is totally satisfying (Table 2). So we can consider that even if these compounds in high yields. The inverse case is obtained for
D.A. Rakotoalimanana et al. / Journal of Analytical and Applied Pyrolysis 123 (2017) 204–215 211

Fig. 9. Examples of reaction pathways leading to alkanes.

the 1-n-butyldecalin at 400 ◦ C and 100 bar: no BTX com- 4.4. Application to thermal cracking of reservoir oils
pounds but production of naphthalene in the whole conversion
range. In the context of petroleum geochemistry, both monocyclic
and bicyclic compounds are major compounds among naph-
212 D.A. Rakotoalimanana et al. / Journal of Analytical and Applied Pyrolysis 123 (2017) 204–215

Fig. 10. Reaction pathways leading to 1-methyldecalin and decalin.

Fig. 11. Reaction pathways leading to tetralin and 5-methyltetralin.

thenes in reservoir oils [1]. In Fig. 13, the thermal decomposition with monocyclics at high pressure and low temperature (reservoir
of n-butylcyclohexane (monocyclic model compound) and 1-n- conditions).
butyldecalin (bicyclic model compound) are compared in terms of
chemical families (in wt%). 4.5. Molecular kinetics model at 400 ◦ C and 100 bar
For both compounds, we obtain the same order of predomi-
nance: alkanes > cyclics (either mono- or bi-) > aromatics (either According to Table 4, we have a constant production as func-
mono- or diaromatics). However, in the case of 1-n-butyldecalin, tion of time for all pyrolysis products 400 ◦ C, it is then reasonable
the relative significant production of hydroaromatics is worth men- to develop a one-step global stoichiometric reaction model, which
tioning, they are known to be inhibitors of the thermal cracking is mainly derived from experimental data. Assuming a first-order
of normal alkanes [29,30]. Thus, the presence of a second cycle kinetic law, kinetic parameters and stoichiometric coefficients
seems to induce others products i.e hydroaromatics in comparison were numerically optimized according with our experimental data,
which needed to be expressed in weight fraction in this case. The
D.A. Rakotoalimanana et al. / Journal of Analytical and Applied Pyrolysis 123 (2017) 204–215 213

Fig. 12. Reaction pathways leading to naphthalene and 1-methylnaphthalene.

Fig. 13. Thermal decompositions of n-butylcyclohexane (left) and 1-n-butyldecalin (right).

BD Alkanes
Weight fracon (wt %)

100 16
Weight fracon (wt %)

80 12
Exp
60
Sim 8 C3
40
C1
20 4

0 0
0 24 48 72 96 120 144 0 24 48 72 96 120 144
Time (h) Time (h)

Alkanes Cyclics
Weight fracon (wt %)
Weight fracon (wt %)

16 16

12 12 D
C4
8 8 MD
C2
4 4

0 0
0 24 48 72 96 120 144 0 24 48 72 96 120 144
Time (h) Time (h)

Hydroaromacs Diaromacs
Weight fracon (wt %)

Weight fracon (wt %)

6
10
8 4
T N
6
MT MN
4 2
2
0 0
0 24 48 72 96 120 144 0 24 48 72 96 120 144
Time (h) Time (h)

Fig. 14. Comparison between simulated using stoichiometric model (lines) and experimental (dots) profiles of products at 400 ◦ C.
214 D.A. Rakotoalimanana et al. / Journal of Analytical and Applied Pyrolysis 123 (2017) 204–215

optimization was performed with the IFP GeoKin Compositional ence compared to pyrolysis of n-butylcyclohexane: C C bonds are
software [2], using the method of least squares i.e. minimizing the globally stronger in 1-n-butyldecalin than in n-butylcyclohexane.
function F: Hydroaromatics and diaromatics should be formed by successive
aromatization reactions of both saturated cycles of decalin and

n

F= rk 2 withrk = wk exp − wk sim 1-methyldecalin. Given these free-radical mechanisms, an expla-


nation is provided to account for a higher yield of n-butane in
k=1
comparison with n-butylcyclohexane. In the meantime, a one-
where wk corresponds to the weight fraction of the product k and step global reaction model was developed at 400 ◦ C and optimized
n the total number of products. We obtained for our case the fol- with the experimental results. This model allowed us to esti-
lowing one -step model. The corresponding nomenclature is given mate decomposition rate constant of 1-n-butyldecalin at 400 ◦ C:
in Table 3. 1.83 × 10−6 s−1 .

BD → 2.40C1 + 7.31C2 + 20.28C3 + 20.71C4 + 21.98D + 6.05MD

+10.68T + 6.76MT + 2.51N + 1.02MT References

[1] B.P. Tissot, D.H. Welte, Petroleum Formation and Occurence, Springer-Verlag,
1984.
Simulated results (expressed in wt%) of our optimized one-step
[2] F. Behar, F. Lorant, L. Mazeas, Elaboration of a new compositional kinetic
model are compared with our experimental at 400 ◦ C in Fig. 14. schema for oil cracking, Org. Geochem. 39 (2008) 764–782, http://dx.doi.org/
Globally, a very good agreement is reached with experimen- 10.1016/j.orggeochem.2008.03.007.
tal data for both degradation of 1-n-butyldecalin and profiles [3] T. Al Darouich, F. Behar, C. Largeau, Pressure effect on the thermal cracking of
the light aromatic fraction of Safaniya crude oil – Implications for deep
of products, except the case of hydroaromatics and diaromatics, prospects, Org. Geochem. 37 (2006) 1155–1169, http://dx.doi.org/10.1016/j.
where the model starts to diverge after 72 h by underestimat- orggeochem.2006.04.004.
ing experimental values. This was expected as these compounds [4] F. Behar, M. Vandenbroucke, Experimental determination of the rate constants
of the n-C25 thermal cracking at 120, 400, and 800 bar: implications for
correspond to secondary products (c.f. Subsection 4.1). Thus, a one- high-pressure/high-temperature prospects, Energy Fuels 10 (1996) 932–940.
step reaction could not take into account this aspect. Given this [5] V. Burklé-Vitzthum, R. Bounaceur, P.-M. Marquaire, F. Montel, L. Fusetti,
stoichiometric reaction, a rate constant at 400 ◦ C could be cal- Thermal evolution of n- and iso-alkanes in oils. Part 1: pyrolysis model for a
mixture of 78 alkanes (C1–C32) including 13,206 free radical reactions, Org.
culated: at 400 ◦ C, we obtained a value of kBD = 1.83 × 10−6 s−1 . Geochem. 42 (2011) 439–450, http://dx.doi.org/10.1016/j.orggeochem.2011.
This value can be compared to the case of n-butylcyclohexane 03.017.
(kBCH = 2.44 × 10−6 s−1 ) [11]. Thus, it confirms that 1-n-butyldecalin [6] F. Lannuzel, R. Bounaceur, R. Michels, G. Scacchi, P.-M. Marquaire, An
extended mechanism including high pressure conditions (700 bar) for
and n-butylcyclohexane have a similar reactivity at 400 ◦ C and
toluene pyrolysis, J. Anal. Appl. Pyrolysis 87 (2010) 236–247, http://dx.doi.
100 bar, while 1-n-butyldecalin still remains slightly less reactive org/10.1016/j.jaap.2010.01.001.
than n-butylcyclohexane. By assuming the same global activa- [7] L. Fusetti, F. Behar, R. Bounaceur, P.-M. Marquaire, K. Grice, S. Derenne, New
insights into secondary gas generation from the thermal cracking of oil:
tion energy, we obtained a frequency factor for 1-n-butyldecalin:
methylated monoaromatics. A kinetic approach using
A = 5.0 × 1016 s−1 (vs 6.7 × 1016 s−1 for n-butylcyclohexane). 1,2,4-trimethylbenzene. Part I: a mechanistic kinetic model, Org. Geochem.
41 (2010) 146–167, http://dx.doi.org/10.1016/j.orggeochem.2009.10.013.
[8] V. Burklé-Vitzthum, R. Michels, G. Scacchi, P.-M. Marquaire, Mechanistic
5. Conclusions modeling of the thermal cracking of decylbenzene. Application to the
prediction of its thermal stability at geological temperatures, Ind. Eng. Chem.
In this paper, we investigated the thermal cracking of an Res. 42 (2003) 5791–5808, http://dx.doi.org/10.1021/ie030086f.
[9] W.-C. Lai, C. Song, Pyrolysis of alkylcyclohexanes in or near the supercritical
alkylated decalin at high pressure (100 bar) and 400 ◦ C. The exper- phase. Product distribution and reaction pathways, Fuel Process. Technol. 48
imental and analytical protocols are the same as described in our (1996) 1–27, http://dx.doi.org/10.1016/0378-3820(96)01030-2.
previous paper [11], to study the pyrolysis of 1-n-butyldecalin at [10] J.A. Widegren, T.J. Bruno, Thermal decomposition kinetics of
propylcyclohexane, Ind. Eng. Chem. Res. 48 (2009) 654–659, http://dx.doi.
400 ◦ C and 100 bar in gold tube reactors in isothermal and isobaric
org/10.1021/ie8008988.
regime. [11] D.A. Rakotoalimanana, F. Behar, R. Bounaceur, V. Burklé-Vitzthum, P.-M.
At 400 ◦ C and 100 bar, the pyrolysis of 1-n-butyldecalin leads Marquaire, Thermal cracking of n-butylcyclohexane at high pressure (100
bar) – part 1: experimental study, J. Anal. Appl. Pyrolysis 117 (2016) 1–16,
in order of predominance: small normal alkanes (from C1 to C4 ),
http://dx.doi.org/10.1016/j.jaap.2015.11.009.
cyclics (decalin and 1-methyldecalin), hydroaromatics (tetralin [12] F. Zhang, Z. Wang, Z. Wang, L. Zhang, Y. Li, F. Qi, Kinetics of decomposition and
and 5-methyltetralin) and diaromatics (naphthalene and 1- isomerization of methylcyclohexane: starting point for studying
methylnaphthalene). At a conversion of 40% we obtain 51 wt% of monoalkylated cyclohexanes combustion, Energy Fuels 27 (2013) 1679–1687,
http://dx.doi.org/10.1021/ef302097s.
alkanes, 28 wt% of bicyclics, 17 wt% of hydroaromatics and 4 wt% [13] M. Akbar Ali, V.T. Dillstrom, J.Y.W. Lai, A. Violi, Ab initio investigation of the
of diaromatics. In comparison with the pyrolysis of decalin at thermal decomposition of n-butylcyclohexane, J. Phys. Chem. A 118 (2014)
atmospheric pressure and high temperature, it is worth mention- 1067–1076, http://dx.doi.org/10.1021/jp4062384.
[14] R. Bounaceur, V. Burklé-Vitzthum, P.-M. Marquaire, L. Fusetti, Mechanistic
ing both the formation of naphthalene and the absence of any modeling of the thermal cracking of methylcyclohexane near atmospheric
BTX compounds. Our experimental results showed no significant pressure, from 523 to 1273 K: identification of aromatization pathways, J.
difference of thermal reactivity between 1-n-butyldecalin and n- Anal. Appl. Pyrolysis 103 (2013) 240–254, http://dx.doi.org/10.1016/j.jaap.
2013.02.012.
butylcyclohexane at 400 ◦ C and 100 bar. Thus, a second cycle seems [15] D.A. Rakotoalimanana, R. Bounaceur, F. Béhar, V. Burklé-Vitzthum, P.-M.
to have no effect in comparison with n-butylcyclohexane in these Marquaire, Thermal cracking of n-butylcyclohexane at high pressure (100
conditions, except for the formation of hydroaromatics. bar) – part 2: mechanistic modeling, J. Anal. Appl. Pyrolysis 120 (2016)
174–185, http://dx.doi.org/10.1016/j.jaap.2016.05.003.
Reaction pathways based on free-radical mechanisms are pro-
[16] P. Bredael, D. Rietvelde, Pyrolysis of hydronapthalenes. 2. Pyrolysis of
posed to explain the formation of main pyrolysis products. Both cis-decalin, Fuel 58 (1979) 215–218, http://dx.doi.org/10.1016/0016-
normal alkanes (except methane and ethane) and cyclics should 2361(79)90121-2.
[17] P.S. Virk, A. Korosi, H.N. Woebcke, Pyrolysis of unsubstituted mono- di-, and
result from breaking reactions of the side alkyl chain of 1-n-
tricycloalkanes, Adv. Chem. Ser. 183 (1979) 67–76.
butyldecalin. The reaction pathway to methane and ethane consists [18] B. Ondruschka, G. Zimmermann, M. Remmler, M. Sedlackova, J. Pola, Thermal
in decomposition of fuel radical, including ring opening by beta- reactions of decalin I. A comparative study of conventional and laser-driven
scissions, isomerization and resonance equilibrium. The proposed pyrolysis, J. Anal. Appl. Pyrolysis 18 (1990) 19–32, http://dx.doi.org/10.1016/
0165-2370(90)85002-5.
mechanism are supported by the calculation of BDE of C H and C C [19] J. Stewart, K. Brezinsky, I. Glassman, Supercritical pyrolysis of decalin, tetralin,
bonds of 1-n-butyldecalin, which could partly explain some differ- and N-Decane at 700–800 K. Product distribution and reaction mechanism,
D.A. Rakotoalimanana et al. / Journal of Analytical and Applied Pyrolysis 123 (2017) 204–215 215

Combust. Sci. Technol. 136 (1998) 373–390, http://dx.doi.org/10.1080/ [30] V. Burklé-Vitzthum, R. Michels, R. Bounaceur, P.-M. Marquaire, G. Scacchi,
00102209808924178. Experimental study and modeling of the role of hydronaphthalenics on the
[20] J.O. Metzger, J. Hartmanna, D. Malwitz, P. Koell, Thermal organic reactions in thermal stability of hydrocarbons under laboratory and geological conditions,
supercritical fluids., in: Chem Eng Supercrit Fluid Cond, Ann Arbor Sci. (1983) Ind. Eng. Chem. Res. 44 (2005) 8972–8987, http://dx.doi.org/10.1021/
515–533. ie050381v.
[21] J. Yu, S. Eser, Thermal decomposition of jet fuel model compounds under [31] M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R.
near-critical and supercritical conditions. 2. Decalin and tetralin, Ind. Eng. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G.A. Petersson, H. Nakatsuji,
Chem. Res. 37 (1998) 4601–4608, http://dx.doi.org/10.1021/ie980302y. M. Caricato, X. Li, H.P. Hratchian, A.F. Izmaylov, J. Bloino, G. Zheng, J.L.
[22] K. Chae, A. Violi, Thermal decomposition of decalin: an ab initio study, J. Org. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida,
Chem. 72 (2007) 3179–3185, http://dx.doi.org/10.1021/jo062324x. T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J.A. Montgomery Jr., J.E.
[23] P. Dagaut, A. Ristori, A. Frassoldati, T. Faravelli, G. Dayma, E. Ranzi, Peralta, F. Ogliaro, M. Bearpark, J.J. Heyd, E. Brothers, K.N. Kudin, V.N.
Experimental and semi-detailed kinetic modeling study of decalin oxidation Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J.C. Burant,
and pyrolysis over a wide range of conditions, Proc. Combust. Inst. 34 (2013) S.S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J.M. Millam, M. Klene, J.E. Knox, J.B.
289–296, http://dx.doi.org/10.1016/j.proci.2012.05.099. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R.E. Stratmann, O.
[24] P. Zámostný, Z. Bělohlav, L. Starkbaumová, J. Patera, Experimental study of Yazyev, A.J. Austin, R. Cammi, C. Pomelli, J.W. Ochterski, R.L. Martin, K.
hydrocarbon structure effects on the composition of its pyrolysis products, J. Morokuma, V.G. Zakrzewski, G.A. Voth, P. Salvador, J.J. Dannenberg, S.
Anal. Appl. Pyrolysis 87 (2010) 207–216, http://dx.doi.org/10.1016/j.jaap. Dapprich, A.D. Daniels, Ö. Farkas, J.B. Foresman, J.V. Ortiz, J. Cioslowski, D.J.
2009.12.006. Fox, Gaussian 09, Revision D.01, Gaussian, Inc, Wallingford C.T, 2009.
[25] Y. Zhu, D.F. Davidson, R.K. Hanson, Pyrolysis and oxidation of decalin at [32] J.A. Montgomery Jr, M.J. Frisch, J.W. Ochterski, G.A. Petersson, A complete
elevated pressures: a shock-tube study, Combust. Flame 161 (2014) 371–383, basis set model chemistry. VI. Use of density functional geometries and
http://dx.doi.org/10.1016/j.combustflame.2013.09.005. frequencies, J. Chem. Phys. 110 (6) (2009) 2822–2827.
[26] M. Zeng, Y. Li, W. Yuan, Z. Zhou, Y. Wang, L. Zhang, F. Qi, Experimental and [33] B. Sirjean, R. Fournet, P.A. Glaude, M.F. Ruiz-López, Extension of the composite
kinetic modeling investigation on decalin pyrolysis at low to atmospheric CBS-QB3 method to singlet diradical calculations, Chem. Phys. Lett. 435 (1)
pressures, J. Anal. Appl. Pyrolysis 167 (2016) 228–237, http://dx.doi.org/10. (2007) 152–156.
1016/j.combustflame.2016.02.009. [34] B. Sirjean, P.A. Glaude, M.F. Ruiz-Lopez, R. Fournet, Detailed kinetic study of
[27] P. Ungerer, F. Behar, M. Villalba, O.R. Heum, A. Audibert, Kinetic modelling of the ring opening of cycloalkanes by CBS-QB3 calculations, J. Phys. Chem. A
oil cracking, Org. Geochem. 13 (1988) 857–868. 110 (46) (2006) 12693–12704.
[28] F. Behar, H. Budzinski, M. Vandenbroucke, Y. Tang, Methane generation from [35] R. Bounaceur, F. Lannuzel, R. Michels, G. Scacchi, P.-M. Marquaire, V.
oil cracking: kinetics of 9-methylphenanthrene cracking and comparison Burklé-Vitzthum, Influence of pressure (100 Pa–100 Mpa) on the pyrolysis of
with other pure compounds and oil fractions, Energy Fuels 13 (1999) an alkane at moderate temperature (603 K–723 K): Experiments and kinetic
471–481, http://dx.doi.org/10.1021/ef980164p. modeling, J. Anal. Appl. Pyrolysis (2016) s1, http://dx.doi.org/10.1016/j.jaap.
[29] R. Bounaceur, G. Scacchi, P.-M. Marquaire, F. Dominé, O. Brévart, D. Dessort, B. 2016.10.022.
Pradier, Inhibiting effect of tetralin on the pyrolytic decomposition of
hexadecane. Comparison with toluene, Ind. Eng. Chem. Res. 41 (2002)
4689–4701, http://dx.doi.org/10.1021/ie0108853.

You might also like