You are on page 1of 13

Computational study of the thermal

decomposition and the thermochemistry of


allyl ethers and allyl sulfides

Lucía Aristizabal, Mariana Ángel,


Camila Orozco, Pablo Ruiz, Jairo
Quijano & Rafael Notario

Structural Chemistry
Computational and Experimental
Studies of Chemical and Biological
Systems

ISSN 1040-0400

Struct Chem
DOI 10.1007/s11224-018-1074-8

1 23
Your article is protected by copyright and
all rights are held exclusively by Springer
Science+Business Media, LLC, part of
Springer Nature. This e-offprint is for personal
use only and shall not be self-archived in
electronic repositories. If you wish to self-
archive your article, please use the accepted
manuscript version for posting on your own
website. You may further deposit the accepted
manuscript version in any repository,
provided it is only made publicly available 12
months after official publication or later and
provided acknowledgement is given to the
original source of publication and a link is
inserted to the published article on Springer's
website. The link must be accompanied by
the following text: "The final publication is
available at link.springer.com”.

1 23
Author's personal copy
Structural Chemistry
https://doi.org/10.1007/s11224-018-1074-8

ORIGINAL RESEARCH

Computational study of the thermal decomposition


and the thermochemistry of allyl ethers and allyl sulfides
Lucía Aristizabal 1 & Mariana Ángel 1 & Camila Orozco 1 & Pablo Ruiz 1 & Jairo Quijano 1 & Rafael Notario 2

Received: 8 December 2017 / Accepted: 3 January 2018


# Springer Science+Business Media, LLC, part of Springer Nature 2018

Abstract
In spite of the several experimental and computational studies on the thermal decomposition of allyl ethers and allyl sulfides,
there are still disagreements on aspects of the reaction mechanism, such as the true nature of the transition states and the grade of
synchronicity of the reactions. This work presents a computational study of the gas-phase thermolysis reaction of allyl ethers and
allyl sulfides substituted at α-carbon, at the M05-2X/6-31+G(d,p) level of theory and a temperature range from 586.15 to
673.15 K. The substituent groups were methyl, ethyl, n-propyl, i-propyl, allyl, benzyl and acetonyl. It was found that the sulfides
react faster than the homologous ethers and that the substituent groups with the capacity of delocalize charge increase the reaction
rate. Through natural bond orbital calculations, the transition states were characterized. The synchronicities and atomic charges of
the studied reactions were determined. A computational study at the G3 level of theory on the thermochemistry of allyl ethers and
sulfides was also carried out.

Keywords Allyl ethers . Allyl sulfides . Reaction mechanism . Substituent group . Thermal decomposition . Enthalpy of
formation . M05-2X . G3

Introduction decomposition than allyl ethers [4]. The greater reactivity in


the sulfides was explained by the enhanced acidic character of
The gas-phase thermal decomposition reaction of allyl the Hα involved in the 1,5-hydrogen transfer due to the higher
ethers, RCH 2 OCH 2 CH=CH 2 , and allyl sulfides, polarizability of the sulfur atom respect to the oxygen atom in
RCH2SCH2CH=CH2, has been studied both experimentally ethers. Some experimental studies carried out on the thermal
[1–11] and computationally [12–18]. The experimental stud- decomposition of allyl sulfides and allyl ethers were focused
ies agree to ensure that this reaction implies a retro-ene, in the effect of the substituent groups on the α-carbon of the
unimolecular and concerted mechanism, via a six-membered molecule, in relation to the reaction rate.
cyclic transition state and a first-order rate law. The nature of The most complete experimental study on the thermolysis
the products reported in the decomposition reaction includes of allyl ethers was carried out by Kwart et al. in 1973 [1], who
an olefin and a carbonyl or thiocarbonyl compound. It has studied the thermal decomposition of a series of allyl alkyl
been reported that allyl sulfides have a higher rate of thermal ethers possessing α hydrogen on the alkyl moiety, including
allyl methyl, allyl ethyl, allyl isopropyl, and diallyl ethers, and
Electronic supplementary material The online version of this article several allyl aryl ethers, observing an insensitivity of the acti-
(https://doi.org/10.1007/s11224-018-1074-8) contains supplementary
material, which is available to authorized users.
vation energy respect to changes of the substituent groups.
This fact led to the researchers to conclude that the reaction
* Pablo Ruiz occurred through a concerted transition state and without
paruizr@unal.edu.co charges development. Other experimental studies on thermal
decomposition reactions of allyl ethers are devoted to allyl
1
Laboratorio de Fisicoquímica Orgánica, Facultad de Ciencias,
methyl [10], allyl ethyl [8], allyl cyclohexyl [7], and diallyl
Universidad Nacional de Colombia, Sede Medellín, ethers [9]. To our knowledge, there is only one very recent
050034 Medellín, Colombia theoretical study devoted to the thermolysis of allyl ethers,
2
Instituto de Química Física Rocasolano, CSIC, Serrano 119, carried out by Espinoza et al. in 2016 [18], who studied the
28006 Madrid, Spain mechanism of the elimination kinetics of allyl cyclohexyl and
Author's personal copy
Struct Chem

allyl ethyl ethers, using CBS-QB3 and several density func- sulfides substituted in the α-carbon, with the following sub-
tional theory (DFT) methods. They concluded that the transi- stituents groups: methyl, ethyl, n-propyl, isopropyl, allyl, ben-
tion state is of non-planar geometry and polar nature and found zyl, and acetonyl, with the aim of determining the effect they
that the most advanced process in the transition state is related cause on the reaction rate and to compare with the experimen-
to the migration of the hydrogen towards the terminal sp2 car- tal values, when they are available in the literature.
bon, and the reactions occur in an asynchronous way. Calculations have been made at four different temperatures,
The most complete experimental studies on the thermolysis 586.15, 625.15, 648.15 and 673.15 K. The Wiberg bond in-
of allyl sulfides were carried out by Martin et al., who studied dices [19] were used as a computational tool to quantify the
the thermal decomposition of allyl methyl, diallyl, and allyl synchronicity or asynchronicity of the studied reactions, to
benzyl [4], allyl chloromethyl, allyl cyanomethyl, allyl 1- find the percentage of evolution of each bond involved in
cyanoethyl, and allyl neopentyl [6], allyl n-propyl [3], allyl the reaction, and to conclude about the driving force for the
n-butyl [2], allyl cyclohexyl [7], and allyl acetonyl sulfides reaction.
[5]. The replacement of the methyl group on the α-carbon The knowledge about the thermochemistry of these series
for an acetonyl group that presents a higher attracting- of compounds is very scarce, the only enthalpy of formation
electron character increased in about 50 times the reaction rate of an allyl ether or allyl sulfide available in the literature is that
[5]. The results obtained were interpreted by the authors in the of allyl ethyl sulfide, 18 ± 3 kJ mol−1, measured by Mackle
sense of proposing that the fragmentation of the Cα–H bond is and Mayrick, in 1962 [20]. For this reason, thermochemical
the determinant of the rate [2, 3]. Some substituents were used computational studies at the G3 level have also been carried
[6] with the intention of verifying the inductive and out in order to obtain the enthalpies of formation of the differ-
mesomeric effects. The authors determined that the systems ent allyl sulfides and allyl ethers studied in this work and to
in which there is conjugation with the substituents present a compare between them.
decrease in the activation energy, due to a stabilization of the
molecule by resonance effect. The experimental and compu-
tational investigation about the mechanism and kinetics of the
thermal decomposition in gas phase of diallyl sulfide [11] Methodology
allowed the researchers to affirm that the reaction is a non-
synchronous process. The same authors carried out computa- Calculations were carried out using the Gaussian 09 compu-
tional studies on the thermal decomposition reaction of allyl n- tational package [21]. The geometric parameters of all the
propyl [13] and allyl n-butyl sulfides [12] through DFT species involved in the reactions were fully optimized at
methods with different basis sets. The authors studied the DFT level with the M05-2X functional [22, 23] and the 6-
mechanism by steps in the formation of free radicals and the 31+G(d,p) [24] basis set. The vibrational frequencies were
molecular mechanism through a six-membered cyclic transi- calculated at the same level, scaling the zero point vibrational
tion state. They obtained values for the activation energies of energies (ZPE) by the factor 0.9631 [25]. Thermal corrections
both processes with the radical mechanism much higher than to enthalpy and entropy values were evaluated in the range
with the concerted mechanism. These results allowed con- 586.15–673.15 K. Intrinsic reaction coordinate (IRC) calcula-
cluding that the radical mechanism is unlikely to occur. tions [26, 27] were also performed verifying that the transition
Through an analysis of the natural bond orbital (NBO) charges states connect with reactants and products. Electronic ener-
on each atom involved in the reaction, the researchers con- gies, zero-point vibrational energies, thermal correction to en-
clude that the mechanism is asynchronous and that the transi- thalpies, and entropies, for all the species involved in the ther-
tion state presents a polar character. A computational study at mal decomposition of allyl ethers and allyl sulfides are col-
the G3(MP2) level on the effect of the substituent groups in lected in Table S2 of the Supporting Information.
the reactivity of allyl alkyl sulfides [14] shows how they The bonding characteristics of the species involved in the
change the geometry of the transition state. Contrary to how reactions have been investigated using the NBO analysis of
it occurs with H and CH3 groups, the F and Cl groups whose Reed and Weinhold [28–30] which provides the atomic natu-
electronic effects are –I decrease the activation energy by an ral total charges and the Wiberg bond indices [18] used to
increase in the acidic character of the transferred hydrogen. follow the progress of the reaction. The NBO analysis has
Other theoretical studies on the thermolysis reaction of allyl been performed using the NBO program [31] implemented
sulfides include those on allyl methyl [15, 17], diallyl [16], in the Gaussian 09 computational package [21].
and allyl chloromethyl sulfides [17]. The rate constants have been calculated according to the
With the purpose to systematize and compare the reactivity classical transition state theory [32, 33] through the Eyring-
of allyl ethers and allyl sulfides, a theoretical study has been Polanyi equation:
carried out at the M05-2X/6-31+G(d,p) level on the thermal ≠
decomposition reaction of a series of allyl ethers and allyl k ðT Þ ¼ ðK B T =hÞe−ΔG ðT Þ=RT
ð1Þ
Author's personal copy
Struct Chem

where KB, h, R, and ΔG≠(T) are Boltzmann’s constant, Table 1 Identification of the reaction products of the thermolysis of
RCH 2 XCH 2 CH=CH 2 compounds (allyl ethers, X = O, and allyl
Planck’s constant, the universal gas constant, and the
sulfides, X = S), according to the mechanism represented in Fig. 1. In
standard-state Gibbs energy of activation, at the absolute tem- all reactions propene and other product (P2) are formed
perature, T, respectively.
The enthalpies of formation in the gas phase of allyl ethers Substrate Reaction product P2
and allyl sulfides have been calculated from the energies ob- RCH2 X
tained at the G3 level of theory [34]. G3-calculated energies at
0 K, and enthalpies at 298 K, are collected in Table S3 of the Me O Formaldehyde
Supporting Information. S Thioformaldehyde
Et O Acetaldehyde
S Thioacetaldehyde
Results and discussion n-Pr O Propionaldehyde
S Thiopropionaldehyde
Thermolysis reactions The thermal decomposition reaction of i-Pr O Acetone
a series of allyl ethers and allyl sulfides was computationally S Thioacetone
simulated at the M05-2X/6-31+G(d,p) level of theory, through Allyl O Acrolein
a molecular mechanism via a six-membered cyclic transition S Thioacrolein
state generating as products propene and a carbonyl or Benzyl O Benzaldehyde
thiocarbonyl compound, depending of the reactant is an ether S Thiobenzaldehyde
or a sulfide, respectively (Fig. 1). In Table 1 are the identified Acetonyl O Acetylformaldehyde
products formed in all the reactions studied. S Acetylthioformaldehyde
The M05-2X/6-31+G(d,p)-optimized structures of the re-
actants of the thermolysis reactions studied are shown in Fig. 2
(allyl ethers) and Fig. 3 (allyl sulfides), and the corresponding decompose easier than the allyl ethers. The results indicate
transition states optimized at the same level of theory are that the reaction is favored when the acetonyl group is the
shown in Fig. 4 (allyl ethers) and Fig. 5 (allyl sulfides). In substituent of the molecule, this occurs in both the allyl ethers
Table S4 of the Supporting Information are the collected di- and allyl sulfides series, with ΔG≠ values of 190.8 and
hedral angles and the imaginary frequencies for the transition 162.8 kJ mol−1, respectively. Contrary to this, when the sub-
states involved in the reactions. The transition states present a stituent is the methyl group, ΔG ≠ are the highest,
non-planar geometry; for allyl ethers, the dihedral angles vary 211.2 kJ mol−1for allyl methyl ether and 182.4 kJ mol−1for
between 64° and 66°, whereas the sulfide atom propitiates a allyl methyl sulfide.
small approach to planarity of the transition state with angles Based on the ΔG≠ values, the increasing order of reactivity
varying between 70° and 75° for the different substituents. obtained is the following: Me, n-Pr, Et, i-Pr, allyl, benzyl, and
The calculation of vibrational frequencies allowed acetonyl, for the allyl ethers, and Me, Et, n-Pr, i-Pr, allyl,
obtaining the kinetic and thermodynamic parameters of the benzyl, and acetonyl, for the allyl sulfides. These results con-
reactions, which are given in Table 2. As it was expected, all firm the stabilization mechanism of the substituent groups at
activation entropies present negative values due to the cyclic the α-carbon in the transition state. Acetonyl, benzyl, and allyl
character of the transition state, which presents less freedom groups, due to their structural configurations, can act as charge
degrees. The allyl sulfides have activation Gibbs free energy delocalizers. The alkyl groups Me, Et, n-Pr, and i-Pr are sta-
values (ΔG≠) smaller than the corresponding to the series of bilized by hyperconjugation, increasing their effectivity in the
allyl ethers. This fact confirms that the allyl sulfides same order. It can be said that the studied reaction is

Fig. 1 Concerted reaction mechanism for the thermal decomposition reaction of allyl ethers (X = O) and allyl sulfides (X = S)
Author's personal copy
Struct Chem

Fig. 2 M05-2X/6-31+G(d,p)-
optimized structures of the
reactants involved in the thermal
decomposition of allyl ethers: I
allyl methyl ether, II allyl ethyl
ether, III allyl n-propyl ether, IV
I II III
allyl isopropyl ether, V diallyl
ether, VI allyl benzyl ether, VII
allyl acetonyl ether

IV V

VI VII

thermodynamically favorable and proceed through an exer- studied are shown in Fig. S1. Allyl ethers present greater ki-
gonic process. In Table 2 it can been observed that Gibbs free netic impediment than allyl sulfides in the thermal decompo-
energies of reaction, ΔrG, values present mostly negative sition reactions; however, the obtained products are more
signs. stable.
The reaction profiles for allyl methyl ether and allyl methyl From the calculated ΔG≠ values and using Eq. (1), the
sulfide are depicted in Fig. 6. Profiles for the other reactions rate constants (k) for each one of the reactions at the four

Fig. 3 M05-2X/6-31+G(d,p)-
optimized structures of the
reactants involved in the thermal
decomposition of allyl sulfides:
VIII allyl methyl sulfide, IX allyl
ethyl sulfide, X allyl n-propyl
sulfide, XI allyl isopropyl sulfide,
XII diallyl sulfide, XIII allyl
benzyl sulfide, and XIV allyl VIII IX X
acetonyl sulfide

XI XII

XIII XIV
Author's personal copy
Struct Chem

Fig. 4 M05-2X/6-31+G(d,p)-
optimized structures of the
transition states involved in the
thermal decomposition of allyl
ethers: I allyl methyl ether, II allyl
ethyl ether, III allyl n-propyl
ether, IV allyl isopropyl ether, V
diallyl ether, VI allyl benzyl ether,
VII allyl acetonyl ether

TS-I TS-II TS-III

TS-IV TS-V

TS-VI TS-VII

studied temperatures have been calculated (Table 3). The whereas lower values are obtained with acetonyl, benzyl, and
computational rate constants have been compared with the allyl groups. Ea values for the thermolysis reaction of allyl
available experimental values. Calculated rate constants ethers vary from 177.3 kJ mol−1 (for allyl acetonyl ether) to
present a reasonable agreement with the experimental rate 200.3 kJ mol−1 (for allyl methyl ether), whereas the activation
constants reported in the literature. In general terms, it is energies are lower for the thermolysis reaction of allyl sul-
observed that the theoretical rate constants are lower than fides, ranging from 153.8 kJ mol−1 (for allyl acetonyl sulfide)
the experimental ones for the allyl ethers and higher for the to 170.9 kJ mol−1 (for allyl methyl sulfide).
allyl sulfides. Population partition calculations (NBO) were carried
The reactions presented an Arrhenius-type dependen- out with the purpose of obtaining the Wiberg bond indi-
cy of the rate constants with the temperature. The plots ces (βi) [19] for each bond and the atomic charges for
of ln(k) vs. 1/T adjust perfectly with straight lines that each atom involved in the reaction. Calculated values for
correspond to the linearized Arrhenius equation repre- all the species involved in the reactions studied are col-
sented as lnk = ln A − Ea/RT. The Arrhenius parameters, lected in Table S5 of the Supporting Information. The
the frequency factor lnA and the activation energy Ea Wiberg bond indices permit to calculate the relative var-
calculated for all the reactions studied are shown in iation of the bond index in the transition state, δβi, de-
Table 4, together with the experimental values available fined as [35]:
in the literature.
R
 P R

The experimental A-factor varies from 8.9 × 10 9 to δβi ¼ β TS
i −β i = β i −β i ð2Þ
4.1 × 10 11 s −1 , while the calculated values vary from
1.0 × 1012 to 4.6 × 1012 s−1. Higher Ea values are obtained δβi gives an idea of the degree of advance of the formation
when RCH2 group is an alkyl group (Me, Et, n-Pr, and i-Pr), or the rupture of the bonds in the transition state.
Author's personal copy
Struct Chem

Fig. 5 M05-2X/6–31 + G(d,p)-


optimized structures of the
transition states involved in the
thermal decomposition of allyl
sulfides: VIII allyl methyl sulfide,
IX allyl ethyl sulfide, X allyl n-
propyl sulfide, XI allyl isopropyl
sulfide, XII diallyl sulfide, XIII
allyl benzyl sulfide, XIV allyl
acetonyl sulfide
TS-VIII TS-IX TS-X

TS-XII TS-XI

TS-XIII TS-XIV

Table 2 Activation parameters


for the thermolysis reaction of RCH2 X ΔS≠ (J mol−1 K−1) ΔH≠ (kJ mol−1) ΔG≠(kJ mol−1) ΔrG(kJ mol−1)
allyl ethers and allyl sulfides,
RCH2XCH2CH=CH2, at Me O −24.8 195.1 211.2 −68.7
648.15 K, obtained from the S −25.9 165.6 182.4 7.4
calculations at the M05-2X/6-31+ Et O −21.3 189.8 203.6 −100.3
G(d,p) level
S −21.2 162.4 176.2 −23.0
n-Pr O −23.5 190.3 205.5 −98.3
S −17.1 161.5 172.6 −23.0
i-Pr O −26.1 181.4 198.3 −129.2
S −17.8 158.7 170.2 −52.4
Allyl O −22.2 181.2 195.7 −105.9
S −18.8 153.9 166.1 −34.2
Benzyl O −29.6 173.8 193.0 −106.2
S −25.2 146.7 163.0 −31.7
Acetonyl O −28.9 172.1 190.8 −75.4
S −21.9 148.6 162.8 −6.8
Author's personal copy
Struct Chem

Fig. 6 Gibbs free energy profiles,


at 648.15 K, calculated at the
M05-2X/6-31+G(d,p) level, for
the thermal decomposition of
allyl methyl ether and allyl methyl
sulfide

Another quantity of interest is the average value, δβav, reaction.


which allows to establish the early or late character of the The synchronicity, Sy, is an indicator that allows to know the
transition state, and it is obtained through the expression concertation degree between the broken and the formed bonds
[35]: in the transition state during the chemical reaction. Sy is calcu-
lated as Sy = 1 − A, where A is defined as the asynchronicity of
δβav ¼ 1=n∑δβi ð3Þ the chemical reaction, given by the expression [35]:

where n is the number of bonds involved in the chemical A ¼ ½1=ð2n−2Þ½∑jδβi −δβ av j=δβ av  ð4Þ

Table 3 Rate constants for the


thermal decomposition of allyl kea (s−1) kcb (s−1) kc/ke
ethers and allyl sulfides,
RCH2XCH2CH=CH2,. RCH2 X T (K)
Experimental values (ke) and 648.15 586.15 625.15 648.15 673.15 648.15
calculated values (kc) Me O 1.16 × 10−3 2.53 × 10−6 3.28 × 10−5 1.29 × 10−4 5.14 × 10−4 0.11
S 2.17 × 10−2 9.33 × 10−4 8.30 × 10−3 2.67 × 10−2 8.68 × 10−2 1.24
−3
Et O 1.71 × 10 1.16 × 10−5 1.40 × 10−4 5.31 × 10−4 2.04 × 10−3 0.31
S – 3.19 × 10−3 2.72 × 10−2 8.56 × 10−2 2.72 × 10−1 –
n-Pr O – 8.02 × 10−6 9.77 × 10−5 3.71 × 10−4 1.43 × 10−3 –
S 7.38 × 10−2 6.29 × 10−3 5.31 × 10−2 1.66 × 10−1 5.24 × 10−1 2.30
i-Pr O 2.60 × 10−3 3.63 × 10−5 3.95 × 10−4 1.42 × 10−3 5.12 × 10−3 0.54
S – 1.03 × 10−2 8.40 × 10−2 2.57 × 10−1 7.98 × 10−1 –
Allyl O 6.60 × 10−3 5.93 × 10−5 6.44 × 10−4 2.30 × 10−3 8.34 × 10−3 0.35
S 7.47 × 10−1 2.45 × 10−2 1.88 × 10−1 5.56 × 10−1 1.67 0.75
Benzyl O 4.35 × 10−3 1.13 × 10−4 1.11 × 10−3 3.78 × 10−3 1.30 × 10−2 0.87
S 3.69 × 10−1 4.96 × 10−2 3.47 × 10−1 9.79 × 10−1 2.80 2.65
Acetonyl O – 1.74 × 10−4 1.68 × 10−3 5.65 × 10−3 1.92 × 10−2 –
S 7.53 × 10−1 5.05 × 10−2 3.61 × 10−1 1.03 2.98 1.33
a
Calculated at 648.15 K from reported Arrhenius equation
b
Calculated at M05-2X/6-31+G(d,p)
Author's personal copy
Struct Chem

Table 4 Experimental and


computational activation Substrate Calculateda Experimental
energies, Ea, ink J mol−1, and lnA
for the thermal decomposition of RCH2 X lnA Ea lnAb Ea Reference
RCH2XCH2CH=CH2
compounds (allyl ethers, X = O, Me O 28.22 ± 0.01 200.3 ± 0.1 25.54 ± 0.05 174.1 ± 2.5 [1]
and allyl sulfides, X = S), in the 27.36 189.0 [10]
temperature range of 586.15 to
S 28.09 ± 0.01 170.9 ± 0.1 25.86 ± 0.58 160 ± 3 [4]
673.15 K
Et O 28.64 ± 0.01 195.0 ± 0.1 25.93 ± 0.39 174.1 ± 2.5 [1]
27.27 ± 0.67 182.3 ± 3.2 [8]
S 28.67 ± 0.01 167.7 ± 0.1 – –
n-Pr O 28.37 ± 0.02 195.5 ± 0.1 – –
S 29.15 ± 0.01 166.8 ± 0.1 26.53 ± 0.37 157 ± 2 [3]
i-Pr O 28.06 ± 0.02 186.6 ± 0.1 26.11 ± 0.21 172.8 ± 1.3 [1]
S 29.06 ± 0.01 163.9 ± 0.1 – –
Allyl O 28.53 ± 0.02 186.5 ± 0.1 26.74 ± 0.23 171.1 ± 0.1 [1]
27.24 ± 0.30 171.1 ± 1.5 [9]
S 28.94 ± 0.01 159.1 ± 0.1 25.36 ± 0.14 138.2 ± 0.7 [4]
22.91 ± 0.28 122.8 ± 1.0 [11]
Benzyl O 27.64 ± 0.01 179.0 ± 0.1 26.55 ± 0.07 172.4 ± 0.4 [1]
S 28.19 ± 0.01 152.0 ± 0.1 25.17 ± 0.41 141 ± 2 [4]
Acetonyl O 27.72 ± 0.01 177.3 ± 0.1 – –
S 28.58 ± 0.01 153.8 ± 0.1 22.91 ± 0.67 125 ± 3 [5]
a
In this work, calculated to M05-2X/6-31+G(d,p)
b
Calculated from reported log A values

In Table 5 are presented the evolution percentages,


%EV = 100 δβi, for each bond involved in the thermal de-
Table 5 M05-2X/6-31+G(d,p)-calculated percentage of evolution
composition of the allyl ethers and allyl sulfides, and the of the bonds (%EV), degree of evolution of bonding at the
δβav and Sy values for each studied reaction. transition states (δBav) and synchronicities (Sy), for the thermal
The most advanced process in all the reactions is the rup- decomposition of RCH2XCH2CH=CH2 compounds (allyl ethers,
ture of the C1–C2bond (66–68% for allyl ethers, and 68–73% X = O, and allyl sulfides, X = S). See Fig. 1 for atom labeling
for allyl sulfides), whereas the slowest process is the formation RCH2= Me Et n-Pr i-Pr Allyl Benzyl Acetonyl
of the X4–C5double bond (40–43% for allyl ethers, and 41–
45% for allyl sulfides). A second advanced process in the Allyl ethers
transition state is the breaking of the C5–H6 bond, with values C1–C2 67.6 66.4 66.6 66.8 67.3 65.6 67.4
above 52% for allyl ethers, and above 60% for allyl sulfides. It C2–C3 46.0 44.5 44.9 42.6 39.4 40.2 40.8
could be thought that the breaking of C1–C2 and C5–H6 bonds C3–O4 53.9 52.3 52.7 50.3 46.7 48.0 48.9
are the processes that initiate the reaction. O4–C5 41.5 43.0 41.6 41.0 39.8 40.6 39.6
The higher reactivity of the allyl sulfides regarding to the C5–H6 53.4 52.5 52.4 54.3 57.1 55.4 58.2
allyl ethers is also reflected in each one of the formation and H6–C1 51.3 50.1 50.0 51.4 53.4 51.6 54.4
breaking processes of the bonds that occur during the chemi- δBav 0.52 0.51 0.51 0.51 0.51 0.50 0.52
cal reaction. In Table 5, it can be verified that all evolution Sy 0.93 0.93 0.93 0.92 0.90 0.91 0.90
percentages for the same bonds are higher for the allyl sulfides Allyl sulfides
than for the allyl ethers; therefore, the δβav values of the allyl C1–C2 72.7 71.3 71.2 71.0 69.1 68.4 70.5
sulfides exceed the values of the allyl ethers. When δβav value C2–C3 50.6 51.2 51.1 49.4 44.0 41.8 42.1
is higher than 0.5, it is said that the transition state is closer to C3–S4 55.0 55.5 55.3 53.7 48.1 46.3 46.4
the products than the reactants, i.e., it is a Blate^ transition S4–C5 44.5 44.2 44.6 44.5 42.7 43.0 40.6
state. Conversely, when that value is lower than 0.5, it is an C5–H6 62.3 60.0 60.0 60.1 61.6 62.0 65.8
Bearly^ transition state, which energetically and geometrically H6–C1 58.8 56.2 56.3 55.9 56.1 56.8 59.4
resembles more to the reactant than to the products. According δBav 0.57 0.56 0.56 0.56 0.54 0.53 0.54
to the obtained results, the transition states for the studied Sy 0.92 0.93 0.93 0.93 0.90 0.89 0.88
reactions with each one of the substituents have a slight Blate^
Author's personal copy
Struct Chem

character for allyl ethers, which is more pronounced in the In Table 6 are the collected values for the NBO charges of
case of allyl sulfides. the atoms involved in the reaction centers for all the studied
With respect to the synchronicity of the reactions, these are reactions. A difference between both series of compounds,
considered as synchronous when the value of the Sy indicator allyl ethers and allyl sulfides, is that in allyl ethers all the
is ≥ 0.9. That is the case for the thermolysis reactions of all the heavy atoms involved in the reaction centers have negative
species studied except for allyl benzyl sulfide (0.89) and allyl charge whereas in allyl sulfides the S atoms have positive
acetonyl sulfide (0.88). In these cases, it can be said that the charge. Another difference is that in allyl ethers, the C1, C2,
reactions are slightly asynchronous. and C3 atoms increase their negative charge from reactants to
An observation of the results shown in Table 5 is the rela- transition states but the charge on the C5 and the O atom
tive asynchronicity between the formation and breaking pro- decreases, whereas in allyl sulfides only the charge on C1
cesses of the bonds in the transition states. If the evolution increases and decreases the charge on the C3, C5, and the S
percentages of the bonds that are forming (C2–C3, X4–C5, atom, being the charge on the C2 atom almost invariant. The
H6–C1) are compared with the values of the bonds that only exception is when the subsituent group is acetonyl where
are breaking (C1–C2, C3–X4, C5–H6), in any of the studied in both cases, ether and sulfide, the charge on the C2 atom
reactions, it is possible to conclude that the breaking pro- decreases.
cesses are more advanced than the forming processes. For
example, for the decomposition reaction of allyl methyl Thermochemistry of allyl ethers and allyl sulfides To our
ether, the average percentage of bonds formation is knowledge, the only enthalpy of formation of an allyl ether
46.3%, while the average breaking percentage is 58.3%. or allyl sulfide available in the literature is that of allyl ethyl
For allyl methyl sulfide, the percentages are 51.3 and sulfide, measured by Mackle and Mayrick, in 1962 [20]. This
63.3%, respectively. It can be said that there is a bond reason has prompted us to carry out a theoretical study deter-
deficiency in the transition states. mining the enthalpies of formation in the gas phase of the

Table 6 M05-2X/6-31+G(d,p)-
calculated NBO atomic charges RCH2= Me Et n-Pr i-Pr Allyl Benzyl Acetonyl
for the atoms involved in the
reaction centers, for the reactants Allyl ethers
(first entry) and transition states C1 −0.456 −0.457 −0.456 −0.459 −0.456 −0.454 −0.445
(second entry) for the thermal −0.580 −0.569 −0.568 −0.569 −0.596 −0.587 −0.604
decomposition reactions of
RCH2XCH2CH=CH2 C2 −0.271 −0.269 −0.270 −0.265 −0.271 −0.273 −0.279
compounds (allyl ethers, X = O, −0.340 −0.350 −0.349 −0.358 −0.339 −0.333 −0.313
and allyl sulfides, X = S) C3 −0.158 −0.157 −0.156 −0.157 −0.165 −0.165 −0.156
−0.275 −0.268 −0.269 −0.262 −0.259 −0.265 −0.267
O4 −0.606 −0.611 −0.612 −0.614 −0.602 −0.602 −0.611
−0.585 −0.596 −0.598 −0.600 −0.574 −0.576 −0.567
C5 −0.345 −0.134 −0.124 0.050 −0.165 −0.137 −0.217
−0.138 0.057 0.065 0.249 0.043 0.057 0.003
H6 0.213 0.221 0.226 0.230 0.222 0.226 0.253
0.203 0.213 0.216 0.222 0.228 0.224 0.226
Allyl sulfides
C1 −0.467 −0.467 −0.468 −0.466 −0.467 −0.466 −0.455
−0.666 −0.645 −0.644 −0.637 −0.658 −0.663 −0.683
C2 −0.248 −0.248 −0.247 −0.248 −0.247 −0.250 −0.263
−0.235 −0.248 −0.248 −0.258 −0.230 −0.230 −0.198
C3 −0.662 −0.664 −0.662 −0.663 −0.665 −0.663 −0.658
−0.507 −0.496 −0.498 −0.496 −0.520 −0.527 −0.531
S4 0.191 0.197 0.200 0.203 0.204 0.207 0.216
0.088 0.073 0.077 0.092 0.118 0.138 0.166
C5 −0.872 −0.648 −0.640 −0.438 −0.665 −0.641 −0.716
−0.811 −0.586 −0.574 −0.383 −0.611 −0.595 −0.667
H6 0.264 0.268 0.267 0.273 0.272 0.289 0.283
0.260 0.266 0.262 0.273 0.280 0.280 0.279
Author's personal copy
Struct Chem

Table 7 G3-calculated enthalpies of formation of RCH2XCH2CH=CH2 not planar, although the sulfur atom in allyl sulfide induces an
compounds. Δ is the difference between the enthalpy of formation of the
approach to the planarity with respect to the oxygen atom in
allyl sulfide and the allyl ether
allyl ethers.
ΔfH0m(g)/kJ mol−1 The activation energies obtained for the thermal decomposi-
tion reaction of the allyl sulfides are lower than the obtained ones
RCH2 Allyl ethers Allyl sulfides Δ/kJ mol−1
X=O X=S for the counterpart allyl ethers. The rate constants, calculated at
the M05-2X/6-31+G(d,p) level of theory, for the different reac-
Me −108.4 40.1 148.5 tions studied present a reasonable agreement with the experimen-
Et −143.7 14.6 (18 ± 3)a 158.3 tal values reported in the literature for the same reactions, and it
n-Pr −164.5 −8.0 156.5 was found that the substituents groups with the ability of delo-
i-Pr −178.6 −16.7 161.9 calize charge (benzyl, allyl and acetonyl) increase the reaction
Allyl −33.3 114.9 148.2 rate with respect to the alkyl groups.
Benzyl −6.5 142.2 148.7 Calculations of population partition of NBO type were
Acetonyl −267.1 −122.5 144.6 made with the purpose of obtaining the Wiberg bond indexes
and the charges in the atoms involved in the reaction. The
a
Experimental value taken from [20] mathematical treatment of the Wiberg bond indices allowed
to find major reaction parameters such as %EV, δβav and Sy. It
studied allyl ethers and allyl sulfides. They have been calcu- was found that in the transition state, the breaking processes of
lated at the G3 level of theory using atomization reactions [36] the C1–C2 and C5–H6 bonds are the most advanced ones and
that is the standard procedure to obtain enthalpies of formation the formation of the X4–C5 double bond is the slowest pro-
in Gaussian-n theories. The calculated values are collected in cess. The transition states are of Blate^ character, more similar
Table 7. to the products than to the reactants. The studied reactions
As it can be observed, the allyl ethers are more stable than occur through synchronous processes.
the allyl sulfides in all cases, the difference being ca. 150– The enthalpies of formation of allyl ethers and allyl sulfides
160 kJ mol−1. There is a very good correlation between the have been calculated at the G3 level. The allyl ethers are more
enthalpies of formation of allyl sulfides and allyl ethers stable than the allyl sulfides, ca. 150–160 kJ mol−1.
(Fig. 7).
Acknowledgments The authors thank the financial support provided by
Universidad Nacional de Colombia–Medellín under the project number
201010020362.
Conclusions
Compliance with ethical standards
The thermal decomposition reaction of a series of allyl ethers
and allyl sulfides with different substituents in the α-carbon Conflict of interest The authors declare that they have no conflict of
interest.
has been studied. From the measurement of the dihedral an-
gles of the modeled transition states, it was found that they are
References
150
y = (0.988 ± 0.032) x + (150.9 ± 4.9) 1. Kwart H, Sarner SF, Slutsky J (1973) Mechanisms of thermolytic
-1

100 fragmentation of allyl ethers I. J Am Chem Soc 95:5234–5242


-1

2. Martin G, Drayer A, Ropero M, Alonso M (1982) Gas-phase


thermolysis of sulfur compounds. Part III: n-butyl 2-propenyl sul-
50
fide. In J Chem Kinet 14:131–142
3. Martin G, Ropero M (1982) Gas-phase thermolysis of sulfur com-
0
pounds. Part IV. n-propyl allyl sulfide. In J Chem Kinet 14:605–612
4. Martin G, Ropero M, Avila R (1982) Gas-phase thermolysis of
-50 sulfur compounds. Part V. Methyl allyl, diallyl and benzyl allyl
m

sulfides. Phosphorus and Sulfur 13:213–220


0
H f

-100 5. Martin G, Ropero M (1983) Gas phase thermolysis of acetonyl allyl


sulfide. React Kinet Catal Lett 22:171–174
-150
6. Martin G, Martinez H, Suhr H, Suhr U (1986) Gas-phase
-300 -250 -200 -150 -100 -50 0 thermolysis of sulfur compounds. Part VIII. Chloromethyl allyl,
0 -1 cyanomethyl allyl, 1-cyanoethyl allyl and neopentyl allyl sulfides.
H
f m In J Chem Kinet 18:355–362
Fig. 7 Plot of the enthalpies of formation calculated for allyl sulfides vs. 7. Martin G, Ascanio J (1989) Gas phase thermolysis of alkyl allyl
those calculated for allyl ethers ethers, thioethers and amines. React Kinet Catal Lett 38:153–158
Author's personal copy
Struct Chem

8. Egger K, Vitins P (1974) The thermochemical kinetics of the retro- Bearpark M, Heyd JJ, Brothers E, Kudin KN, Staroverov VN,
‘ene’ reactions of molecules with the general structure (allyl)XYH Kobayashi R, Normand J, Raghavachari K, Rendell A, Burant
in the gas phase. IX. The thermal unimolecular decomposition of JC, Iyengar SS, Tomasi J, Cossi M, Rega N, Millam JM, Klene
ethylallylether in the gas phase. In J Chem Kinet 6:429–435 M, Knox JE, Cross JB, Bakken V, Adamo C, Jaramillo J, Gomperts
9. Vitins P, Egger K (1974) The thermochemical kinetics of the retro- R, Stratmann RE, Yazyev O, Austin AJ, Cammi R, Pomelli C,
‘ene’ reactions of molecules with the general structure (allyl)XYH Ochterski JW, Martin RL, Morokuma K, Zakrzewski VG, Voth
in the gas phase. Part X. Unimolecular thermal decomposition of GA, Salvador P, Dannenberg JJ, Dapprich S, Daniels AD, Farkas
diallyl ether. J Chem Soc Perkin Trans 2:1292–1293 O, Foresman JB, Ortiz JV, Cioslowski J, Fox DJ (2010) Gaussian
10. Gutman D, Braun W, Tsang W (1977) Comparison of the thermal 09, Revision B.01. Gaussian Inc., Wallingford
and infrared laser induced unimolecular decompositions of 22. Zhao Y, Schultz NE, Truhlar DG (2006) Design of density func-
allylmethylether, ethylacetate, and isopropylbromide. J Chem tionals by combining the method of constraint satisfaction with
Phys 67:4291–4296 parametrization for thermochemistry, thermochemical kinetics,
11. Gholami MR, Izadyar M (2003) Gas-phase kinetics and mechanism and noncovalent interactions. J Chem Theory Comp 2:364–382
of diallyl sulfide thermal decomposition. J Phys Org Chem 16:153– 23. Zhao Y, Truhlar DG (2008) Density functionals with broad appli-
157 cability in chemistry. Acc Chem Res 41:157–167
12. Izadyar M, Jahangir AH, Gholami M (2004) DFT calculations on 24. Ditchfield R, Hehre WJ, Pople JA (1971) Self-consistent molecu-
the retro-ene reactions, part I: allyl n-butyl sulfide pyrolysis in the lar-orbital methods. IX. An extended gaussian-type basis for
gas phase. J Chem Res 2004:585–588 molecular-orbital studies of organic molecules. J Chem Phys 54:
13. Izadyar M, Gholami M, Haghgu M (2004) DFT calculations on the 724–728
retro-ene reactions, part II: allyl n-propyl sulfide pyrolysis in the gas 25. Merrick JP, Moran D, Radom L (2007) An evaluation of harmonic
phase. J Mol Struct (THEOCHEM) 686:37–42 vibrational frequency scale factors. J Phys Chem A 111:11683–
14. Izadyar M, Gholami M (2006) Substituent effects on the gas phase 11700
reactivity of alkyl allyl sulfides, a theoretical study. J Mol Struct 26. Kenichi F (1970) A formulation of the reaction coordinate. J Phys
(THEOCHEM) 759:11–15 Chem 74:4161–4163
15. Rodríguez LJ, Añez R, Ocando-Mavarez E (2001) Transition struc- 27. Hratchian H, Schlegel H (2005) Finding minima, transition states,
ture for the retro-ene type elimination reaction of propene from and following reaction pathways on ab initio potential energy sur-
allylmethylsulfide. J Mol Struct (THEOCHEM) 536:53–57 faces. Theory and applications of computational chemistry: the first
16. Rodríguez LJ, Fermín J, Añez R, Ocando-Mavarez E (2002) 40 years. Chapter 10. Elsevier, Amsterdam
Transition-state geometry and activation energy calculations for 28. Reed AE, Curtiss LA, Weinhold F (1988) Intermolecular interac-
the retro-ene elimination reaction of propene from diallyl sulfide. tions from a natural bond orbital, donor-acceptor viewpoint. Chem
J Phys Org Chem 15:826–830 Rev 88:899–926
17. Hahn DK, Raghu Veer KS, Ortiz JV (2011) Computational tests of 29. Reed AE, Weinhold F (1983) Natural bond orbital analysis of near-
models for kinetic parameters of unimolecular reactions of organo- Hartree–Fock water dimer. J Chem Phys 78:4066–4073
phosphorus and organosulfur compounds. J Phys Chem A 115: 30. Weinhold F, Landis C (2005) Valency and bonding. A natural bond
14143–14152 orbital donor-acceptor perspective. Cambridge University Press,
18. Espinoza S, Lezama J, Mora JR, Cordova T, Chuchani G (2016) Cambridge
Theoretical calculations on the mechanism of the elimination kinet- 31. Glendening ED, Badenhoop JK, Reed AE, Carpenter JE, Bohmann
ics of allyl cyclohexyl-, amine-, sulfide, -ether, and allyl ethyl ether JA, Morales CM, Weinhold F (2001) NBO 5.0. Theoretical
in the gas phase. Comput Theor Chem 1090:6–16 Chemistry Institute, University of Wisconsin, Madison
19. Wiberg KB (1968) Application of the Pople-Santry-Segal CNDO 32. Glasstone S, Laidler KJ, Eyring H (1941) The theory of rate
method to the ciclopropylcarbinyl and cyclobutyl cation and to processes1st edn. McGraw Hill, New York
bicyclobutane. Tetrahedron 24:1083–1096 33. Benson SW (1969) The foundations of chemical kinetics. McGraw
20. Mackle H, Mayrick RG (1962) Studies in the thermochemistry of Hill, New York
organic sulphides. Part 3.—the gas-phase heats of formation of 34. Curtiss LA, Raghavachari K, Redfern PC, Rassolov V, Pople JA
diphenyl sulphide, dibenzyl sulphide, diphenyl disulphide, 4-thia- (1998) Gaussian-3 (G3) theory for molecules containing first and
1-hexene and 4-thia-5,5-dimethyl-1-hexene. Trans Faraday Soc 58: second-row atoms. J Chem Phys 109:7764–7776
238–243 35. Moyano A, Pericas MA, Valentí E (1989) A theoretical study on the
21. Frisch MJ, Trucks GW, Schlegel HB, Scuseria GE, Robb MA, mechanism of the thermal and the acid-catalyzed decarboxylation
Cheeseman JR, Scalmani G, Barone V, Mennucci B, Petersson of 2-oxetanones (β-lactones). J Org Chem 54:573–582
GA, Nakatsuji H, Caricato M, Li X, Hratchian HP, Izmaylov AF, 36. Notario R, Castaño O, Gomperts R, Frutos LM, Palmeiro R (2000)
Bloino J, Zheng G, Sonnenberg JL, Hada M, Ehara M, Toyota K, Organic thermochemistry at high ab initio levels. 3. A G3 study of
Fukuda R, Hasegawa J, Ishida M, Nakajima T, Honda Y, Kitao O, cyclic saturated and unsaturated hydrocarbons (including aro-
Nakai H, Vreven T, Montgomery Jr JA, Peralta JE, Ogliaro F, matics). J Org Chem 65:4298–4302

You might also like