Sie sind auf Seite 1von 12

Reac Kinet Mech Cat (2015) 116:315326

DOI 10.1007/s11144-015-0913-5

Kinetics of DielsAlder reactions between


1,3-cyclopentadiene and isoprene

Jiri Krupka1

Received: 10 June 2015 / Accepted: 3 August 2015 / Published online: 13 August 2015
 Akademiai Kiado, Budapest, Hungary 2015

Abstract The work deals with a detailed study of the kinetics of the cycloaddition
reactions between isoprene and cyclopentadiene. The laboratory experiments were
carried out in a batch reactor in a cyclohexane solution at different temperatures.
Measured concentrations of the reactants and 1:1 DielsAlder adducts were fitted to
the kinetic model. Kinetic parameters for formation of 5 individual isomeric
codimers of cyclopentadiene and isoprene were determined. Kinetic measurements
gave evidence that 5-methyl-cis-3a,4,7,7a-tetrahydro-1H-indene is formed by two
reaction routes: (i) by the DielsAlder reaction of cyclopentadiene with isoprene in
which cyclopentadiene acts as a dienophile, and (ii) by the Cope rearrangement
from endo-5-isopropenyl-2-norbornene, the dominant codimer. It depends on the
reaction conditions which of them prevails.

Keywords Kinetics  DielsAlder reactions  Cope rearrangement 


Codimerization  Activation energy  Cyclopentadiene  Isoprene

Introduction

At this moment, several mechanisms should be considered regarding to DielsAlder


reactions:

non-polar mechanisms (synchronical mechanism, biradicaloidal one step-


mechanism, stepwise biradical mechanism [14],
polar mechanisms (one step-two stage mechanism, stepwise zwitterionic
mechanism) [57].

& Jiri Krupka


jiri.krupka@vscht.cz
1
Department of Organic Technology, Faculty of Chemical Technology, University of Chemistry
and Technology, Technicka 5, 166 28 Prague 6, Czech Republic

123
316 Reac Kinet Mech Cat (2015) 116:315326

Cyclopentadiene should be considered as moderate electrophile, because its


global electrophilicity, x, is equal to 0.83 eV [8]. On the other hand, the global
electrophilicity of isoprene is equal 0.94 eV [9]. In consequence, the reaction
between title compounds should be interpreted as non-polar cycloaddition [10].
In our preceding papers [1113], we presented the results of a kinetic study of the
thermal dimerizations of 1,3-cyclopentadiene, isoprene (2-methyl-1,3-butadiene),
isomers of methyl-1,3-cyclopentadiene, cis- and trans-1,3-pentadiene and of
codimerizations of cyclopentadiene with cis- and trans-1,3-pentadiene. Kinetic
parameters were determined related to the formation of individual isomeric products
of the reactions mentioned above.
The present paper is focused on a quantification of the kinetics of the thermal
codimerization of cyclopentadiene (CPD) with isoprene (ISP). Six isomeric adducts
[1421], the so called codimers which will be denoted below by symbols CPD-
ISP 16, are the products of this reaction. In all cases they are formed by the
[4 ? 2]-cycloaddition, i.e., the so called DielsAlder reaction. The reaction scheme
is presented in Scheme 1. So far, the kinetics of only two dominant products were
described, namely endo-5-(prop-1-en-2-yl)bicyclo[2.2.1]hept-2-ene [22, 23],
denoted for short as endo-5-isopropenyl-2-norbornene (CPD-ISP 2) and 5-methyl-
cis-3a,7,7a-tetrahydro-1H-indene [16, 22] (CPD-ISP 3). Muja et al. [22] constructed
a kinetic model including three competitive parallel reactions: the dimerization of
CPD and a formation of two above-mentioned dominant codimers CPD-ISP. In
comparison with these reactions, the extent of the isoprene dimerization is
negligible and was not taken into account. Endo-5-isopropenyl-2-norbornene was
denoted by the authors as the codimer A, and 5-methyl-cis-3a,7,7a-tetrahydro-1H-
indene as codimer B. Experimental data from the thermal intervals of 80130 C for
the CPD dimerization and 120150 C for codimerization were used. The results of
their measurements constitute a part of Table 4 (see Results and discussion
section). Iwase et al. [16] found out that, at 145170 C, the Cope rearrangement of
endo-5-isopropenyl-2-norbornene to thermodynamically more stable 5-methyl-cis-
3a,7,7a-tetrahydro-1H-indene takes place. The exo-isomer does not undergo the
Cope rearrangement. On the basis of this finding, the authors assume that the
tetrahydroindene codimer is not formed by the reaction of cyclopentadiene and
isoprene but, instead, by the subsequent rearrangement of endo-5-isopropenyl-2-
norbornene. The kinetic parameters of the rearrangement, determined experimen-
tally by the authors, are included in Table 2 (see Results and discussion section).
Generally, the Cope rearrangement [24] is a thermal [3, 3] sigmatropic shift which
causes the isomerization of the 1,5-diene. Mechanistic aspects of similar [3.3]-
sigmatropic shift in diene/heterodiene reaction systems have been analyzed recently
[25, 26]. Mechanistic aspects of a Lewis acid-catalyzed tandem DielsAlder/either
Cope or Claisen domino reaction are also discussed in the literature [27, 28].
It is the aim of the present study to determine the kinetic parameters of the non-
catalytic thermal codimerization of cyclopentadiene with isoprene related to the
formation of the individual isomeric products given in Scheme 1 and, in the case of
both above mentioned dominant isomers, to compare the results with the published
ones.

123
Reac Kinet Mech Cat (2015) 116:315326 317

endo-5-methyl-exo-5-
vinyl-2-norbornene

CPD-ISP 5

exo-5-methyl-endo-5-
diene dienophile vinyl-2-norbornene

+ CPD-ISP 6

exo-5-isopropenyl-
2-norbornen

CPD-ISP 1

endo-5-isopropenyl-
2-norbornene
codimer A

CPD-ISP 2

Cope rearrangement

diene dienophile

5-methyl-cis-3a,7,7a-
+ tetrahydro-1H-indene

codimer B
CPD-ISP 3

6-methyl-cis-3a,7,7a-
tetrahydro-1H-indene

CPD-ISP 4

Scheme 1 Chemo-, regio- and stereochemical channels for reaction between cyclopentadiene and
isoprene

Experimental

The reaction kinetics was measured in the liquid phase under isothermal conditions
with cyclohexane present as solvent and in the temperature range of 70140 C,
where the DielsAlder reactions studied can be considered irreversible [29]. CPD
was prepared by thermal decomposition of commercially available dicyclopenta-
diene (Aldrich, 99.5 % endo-dicyclopentadiene) by reflux at 180200 C under a
distillation column with 18 theoretic plates. The distillate of pure CPD was cooled at
the column head using a mixture of acetone and dry ice and was then kept at
-20 C. Quantitative analysis of reaction mixtures was based on gas chromatog-
raphy (SHIMADZU GC-17 A version 3) with a flame ionization detector, a 50-m-
long HP-PONA capillary column and internal normalization method. The details of
analysis, the origin of chemicals and the kinetic procedure were described in our
previous kinetic studies [11, 13]. The identification of the chemical structures of the

123
318 Reac Kinet Mech Cat (2015) 116:315326

isomeric products formed by DielsAlder reactions of cyclopentadiene with


isoprene was the subject of another our study [18].

Regression analysis of kinetic data

The regression analysis of experimental data was performed using the ERA (Easy
Regression Analysis) software [30]. The software has tools for the evaluation of the
statistical significance and reliability of the parameters estimated. The values of the
molar concentrations (in mmol/l) of the reaction components as a function of time
were input parameters; estimated parameters were rate constants at the reference
temperature of 120 C and activation energy Ea. The dependence of rate constant ki
on temperature was expressed by the Arrhenius relation:
 
Eai T  T0
ki ki0 exp 1
RTT0
where T0 is a reference temperature, ki and ki0 are the rate constants of reaction i at
temperature T or T0, Eai is the activation energy of reaction i and R is the universal
gas constant.
A reference temperature was used in order to eliminate the strong correlation
between the pre-exponential factor (frequency factor) Ai0 and the activation energy
Eai in the non-modified Arrhenius equation. Significantly lesser correlation between
the rate constant ki0 and the activation energy Eai is characteristic of the modified
relation (Eq. 1).
The software has a graphical user interface that allows continuous visual control
of the cogency of the computed concentration dependences to the experimental
points during the optimization process. The differential equations in the model were
solved numerically by the Merson modification of the fourth order RungeKutta
method with variable lengths of the integration step. The values of the kinetic
parameters were obtained by a modified optimization algorithm of adaptive random
search. The weighted sum of residual squares was utilized as an objective function
by the algorithm [11].

Results and discussion

The kinetics of the DielsAlder reactions in the system cyclopentadieneisoprene


was followed at 70, 100, 120 and 140 C. At these temperatures, the reaction
mixtures contained, besides the dimers of cyclopentadiene and isoprene, six
isomeric CPD-ISP codimers. Among them, CPD-ISP 2 and CPD-ISP 3 dominated.
At temperatures up to 120 C, the content of the CPD-ISP 2 was six times higher
than that of the CPD-ISP 3. Experimental data obtained are shown in the top part of
Table 1.
Muja et al. [22] suggest that CPD-ISP 2 and CPD-ISP 3, having the
isopropenylnorbornene and tetrahydroindene structures, respectively, originate

123
Table 1 Experimental data from the kinetic measurements of the reaction system cyclopentadieneisoprene and the rearrangement of CPD-ISP 2 codimer to CPD-ISP 3
Time (min) T (K) Molar concentration of

ISP CPD DCPD CPD-ISP 1 CPD-ISP 2 CPD-ISP 3 CPD-ISP 5 CPD-ISP 6


(mmol/l) (mmol/l) (mmol/l) (mmol/l) (mmol/l) (mmol/l) (mmol/l) (mmol/l)

0 343.15 1123.6 927.7 124.8 0.00 0.63 0.00 0.00 0.00


72 343.15 1148.4 800.8 204.7 0.10 2.32 0.37
150 343.15 1064.1 626.0 224.3 0.21 3.49 0.58
313 343.15 1157.2 512.4 309.3 0.31 5.86 0.90
400 343.15 917.9 370.7 370.1 0.38 7.56 1.16 0.26
1385 343.15 1034.6 167.5 458.5 0.62 14.69 2.22 0.42
Reac Kinet Mech Cat (2015) 116:315326

0 373.15 1036.2 791.8 142.8 0.00 2.43 0.45 0.00 0.00


30 373.15 1137.7 523.6 231.9 0.44 6.97 1.16
70 373.15 1001.3 367.1 313.1 0.58 10.05 1.77
170 373.15 932.0 178.2 399.8 1.12 17.63 3.03 0.68 0.44
295 373.15 992.8 112.2 400.7 1.46 22.38 3.44 0.73 0.53
450 373.15 1160.4 81.3 392.7 1.65 26.22 4.70 0.87 0.63
0 393.15 1052.6 519.6 285.5 0.00 8.82 1.87 0.00 0.00
20 393.15 909.3 287.9 334.9 1.02 14.87 2.91 0.63 0.44
40 393.15 973.7 216.1 342.2 1.26 17.58 3.25 0.78 0.63
70 393.15 902.9 142.4 416.0 1.75 25.41 4.58 0.97 0.68
150 393.15 870.8 74.9 433.8 2.23 31.28 6.16 1.26 0.87
300 393.15 800.6 38.8 435.0 2.52 36.69 7.19 1.46 0.92
0 413.15 847.2 252.2 308.8 1.17 15.12 2.84 0.65 0.42
20 413.15 802.3 118.0 353.8 1.87 24.07 4.74 1.07 0.70
40 413.15 762.6 78.0 368.3 2.24 28.81 6.16 1.26 0.79
85 413.15 720.5 47.7 380.5 2.75 35.45 8.20 1.59 0.99
319

123
Table 1 continued
320

Time (min) T (K) Molar concentration of

ISP CPD DCPD CPD-ISP 1 CPD-ISP 2 CPD-ISP 3 CPD-ISP 5 CPD-ISP 6

123
(mmol/l) (mmol/l) (mmol/l) (mmol/l) (mmol/l) (mmol/l) (mmol/l) (mmol/l)

130 413.15 688.6 38.2 385.7 3.03 38.15 10.09 1.73 1.21
0 353.15 0 0 0 0 4793.3 0.0 0 0
840 353.15 0 0 0 0 4784.4 4.6 0 0
6670 353.15 0 0 0 0 4772.6 23.8 0 0
19,840 353.15 0 0 0 0 4752.0 69.1 0 0
0 373.15 0 0 0 0 4640.1 0.0 0 0
900 373.15 0 0 0 0 4582.0 31.2 0 0
9540 373.15 0 0 0 0 4223.2 357.5 0 0
21,060 373.15 0 0 0 0 3812.6 779.7 0 0
0 393.15 0 0 0 0 2356.2 0.0 0 0
490 393.15 0 0 0 0 2244.0 92.7 0 0
2860 393.15 0 0 0 0 1797.7 512.5 0 0
4480 393.15 0 0 0 0 1469.0 749.0 0 0
10,025 393.15 0 0 0 0 836.9 1302.3 0 0
12,850 393.15 0 0 0 0 621.6 1513.4 0 0
0 413.15 0 0 0 0 2281.9 0.0 0 0
195 413.15 0 0 0 0 1918.8 279.7 0 0
360 413.15 0 0 0 0 1747.8 486.3 0 0
500 413.15 0 0 0 0 1591.8 623.9 0 0
1365 413.15 0 0 0 0 722.4 1328.2 0 0
1740 413.15 0 0 0 0 529.4 1512.0 0 0

Composition of the reaction mixtures in mmol/l in relation to the reaction time and reaction temperature
Reac Kinet Mech Cat (2015) 116:315326
Reac Kinet Mech Cat (2015) 116:315326 321

Table 2 Results of the regression analysis of the model of the endo-5-isopropenyl-2-norbornene rear-
rangement and comparison with literature
Reaction/Parameter ki(120 C) 9 106 [min-1] Eai [kJ mol-1]

Value 95 % conf. limits Value 95 % conf. limits

CPD-ISP 2 ? CPD-ISP 3 (exp.) 92.4 87.098.1 138.1 134.8142.1


CPD-ISP 2 ? CPD-ISP 3 [16] 99.1 141.5

First order reaction kinetics

4793 1
2
1 1 1
2

2
c [mmol/l]

2397 3
4 3

4
4 3
4
4 3 3
4 3

3 2
4 3
4 3
4 4 3
2
4
3
0,00 4321 12 1 1

0,00 10530 21060


t [min.]

Fig. 1 Comparison of experimental data of the endo-5-isopropenyl-2-norbornene rearrangement with the


results of the model. Concentration profiles of the reactant (CPD-ISP 2) and the product (CPD-ISP 3)
measured at 4 different temperature levels. Filled circlecCPD-ISP 2, filled trianglecCPD-ISP 3, (1)
80 C, (2)100 C, (3)120 C, (4)140 C (The curves represent concentration dependences yielded
by the model)

independently by parallel reactions of cyclopentadiene with isoprene. In contrast,


according to Iwase et al. [16], only endo-5-isopropenyl-2-norbornene (CPD-ISP 2)
and exo-5-isopropenyl-2-norbornene (CPD-ISP 1) are formed by the reaction of
cyclopentadiene with isoprene, while 5-methyl-3a,7,7a-tetrahydro-1H-indene
(CPD-ISP 3) arises subsequently by the Cope rearrangement from endo-5-
isopropenyl-2-norbornene (CPD-ISP 2). Therefore, in the kinetic model of the
reaction system cyclopentadieneisoprene designed by us, both potential reaction
routes were taken into account. First, however, the kinetic of the Cope
rearrangement of CPD-ISP 2 was studied separately (Table 1, bottom part). The
measurements were performed at 80, 100, 120 and 140 C; undiluted concentrates
of the CPD-ISP 2 codimer were used as the starting mixture. At the beginning of the
measurement, its content was 70 or 36 wt%. The results of the kinetic measure-
ments of the Cope rearrangement are summarized in Table 2. Fig. 1 demonstrates
that the model with kinetic parameters shown in Table 2 fits well the experimental
data. It can be concluded that the identified kinetic parameters are in a reasonable
agreement with the data by Iwase et al. [16].

123
322 Reac Kinet Mech Cat (2015) 116:315326

It is incorrect to deal with the kinetics of the codimerization reactions as if they


are isolated. The cycloaddition reactions in a system of two or more dienes (i.e.,
dimerizations and codimerizations) have a competitive-parallel character, and must
be solved together within the kinetic model.
On the basis of the following reaction scheme, a model was constructed enabling
one to calculate kinetic parameters of the codimerization reactions of CPD with
isoprene:
CPD ! 0:5 DCPD 2

CPD ISP ! CPD-ISP 5 3

CPD ISP ! CPD-ISP 6 4

CPD + ISP ! CPD-ISP 1 5

CPD ISP ! CPD-ISP 2 6

CPD ISP ! CPD-ISP 3 7

CPD ISP ! CPD-ISP 3 8

ISP ! 0:5 R DISP 9


The formation rate of the isoprene dimers is much lower than that of the CPD-ISP
codimers and endo-dicyclopentadiene (endo-DCPD) [11]. Therefore, in the model, a
formal overall Eq. 9 was used for the isoprene dimer formation, describing the
conversion of isoprene into all products except for the CPD-ISP codimers. The
formation of exo-dicyclopentadiene was not incorporated into the model and,
instead, it was included in reaction 2. Very low concentrations of 6-methyl-cis-
3a,7,7a-tetrahydro-1H-indene (CPD-ISP 4) in the reaction mixtures did not allow us
to reliably interpret its formation kinetics, so that its formation was not taken into
account in the model.
The regression model consists of the following equations:
r1 k1 c2CPD 10

r2 k2 cCPD  cISP 11

r3 k3 cCPD  cISP 12

r4 k4 cCPD  cISP 13

r5 k5 cCPD  cISP 14

r6 k6 cCPD  cISP 15

123
Reac Kinet Mech Cat (2015) 116:315326 323

r7 k7 cCPDISP3 16

r8 k8 c2ISP 17

dcCPD
 r1  r2  r3  r4  r5  r6 18
dt
dcISP
 r2  r3  r4  r5  r6  r8 19
dt
dcDCPD
0:5 r1 20
dt

(a) 10,0 4 (b) 10,0 4

(4)
4
4

3
3
c [mmol/l]
c [mmol/l]

4 3
4 3

5,00 5,00 4 2
4 2 3
3

(3)
2
2 3
3 2
2 4 3
4 3

3
3 2
2

2 1
2 1
(2) 1
2
1
1

1
2 1 (1) 0,00
1

0,00 1
1

0,00 225 450 0,00 225 450

t [min.] t [min.]

(c) 10,0 4

3
c [mmol/l]

4 3

5,00 4 2
3

2
3
2
4 3

3
2

2 1
1
1
2 1

0,00 1

0,00 225 450


t [min.]

Fig. 2 ac Time dependences of CPD-ISP 3 codimer concentration in the reaction mixture containing
CPD and isoprene. Comparison of experimental data (marks) with the results of regression analysis of
three different kinetic models (lines). (1)70 C, (2)100 C, (3)120 C, (4)140 C (Marks
correspond to the exp. data given in Table 1. For clarity, only the concentrations of key product (CPD-ISP
3 codimer) are shown. The curves represent concentration dependences yielded by the model). The model
considering the CPD-ISP 3 codimer formation a solely by the Cope rearrangement, b solely by the Diels
Alder reaction of CPD and isoprene, c by both reaction routes

123
324 Reac Kinet Mech Cat (2015) 116:315326

Table 3 Results of the regression analysis of the kinetic model including reactions 29
Reaction (no.)/Parameter ki(120 C) 9 103 [l.mol-1 min-1] Eai [kJ mol-1]

Value 95 % conf. limits Value 95 % conf. limits

CPD ? 0.5 endo-DCPD (2) 66.5 64.468.8 66.2 64.6967.67


CPD ? ISP ? CPD-ISP 5 (3) 0.050 0.0470.051 90.0 88.391.7
CPD ? ISP ? CPD-ISP 6 (4) 0.033 0.0320.035 89.3 86.292.4
CPD ? ISP ? CPD-ISP 1 (5) 0.084 0.0810.088 91.3 89.992.6
CPD ? ISP ? CPD-ISP 2 (6) 0.919 0.8780.962 81.2 79.682.8
CPD ? ISP ? CPD-ISP 3 (7) 0.147 0.1390.155 78.4 76.780.0
ki(120 C) 9 103 [min-1]
CPD-ISP 2 ? CPD-ISP 3 (8) 0.093 0.0900.095 138.3 136.9139.7

dcCPDISP 1
r4 21
dt
dcCPDISP 2
r5  r7 22
dt
dcCPDISP 3
r6 r7 23
dt
dcCPDISP 5
r2 24
dt
dcCPDISP 6
r3 25
dt

The temperature dependence of the rate constants was expressed by Eq. 1. Both
potential reaction routes of the CPD-ISP 3 codimer formation, i.e., the Cope
rearrangement of the CPD-ISP 2 codimer (8) and the DielsAlder reaction (7) of
isoprene with cyclopentadiene, were included into the model. Correspondingly,
experimental data of the thermal stability measurements of the CPD-ISP 2 codimer
alone were also included into the assessment of the model.
Besides the above combined model, the kinetic data measured were assessed also
by the models considering the CPD-ISP 3 codimer formation (1) solely by the
DielsAlder reaction of the starting dienes and (2) solely by the Cope rearrange-
ment. Thus, the data were assessed by three different kinetic models. The plots in
Fig. 2 show the temporal dependences of the CPD-ISP 3 concentration for the
reaction mixtures containing cyclopentadiene and isoprene at four temperatures.
The data marks correspond to the experiment, the curves represent the concentration
dependences assessed by the three kinetic models under discussion. It is evident
from the figure that only that model which includes both reaction routes (plot c) fits
well the experimental data. This is a proof that the CPD-ISP 3 formation can
proceed by both reaction routes and that the reaction conditions determine which

123
Reac Kinet Mech Cat (2015) 116:315326 325

Table 4 Comparison of the measured kinetic parameters for the formation of CPD-ISP 2 and 3 codimers
with literature values
Reaction (no)/Parameter ki(120 C) 9 103 [l.mol-1 min-1] Eai [kJ mol-1]

Exp. [22] [23] Exp. [22] [23]

CPD ? ISP ? CPD-ISP 2 (6) 0.919 1.85 0.671 81.2 71.67 78.7
CPD ? ISP ? CPD-ISP 3 (7) 0.147 0.324 78.4 82.73

The values obtained in this study are bold

route is predominant. If the reaction mixture contains cyclopentadiene and isoprene


and, simultaneously, the CPD-ISP 2 concentration is low, then the DielsAlder
reaction of cyclopentadiene with isoprene leading to the CPD-ISP 3 codimer is the
prevailing reaction route in which cyclopentadiene acts as a dienophile. The other
reaction route, the Cope rearrangement, is preferred at a higher CPD-ISP 2 codimer
concentration and at higher reaction temperatures. The claim of Iwase et al. [16]
that, in the reaction system cyclopentadieneisoprene, the CPD-ISP 3 codimer is
produced exclusively by the subsequent Cope rearrangement from the CPD-ISP 2
codimer, is incorrect.
The estimates of the kinetic parameter values obtained by an optimization of the
regression model which included the reactions 29 are presented in Table 3. In the
temperature range studied (70140 C), the parameters describe well the temporal
dependences of the concentrations of the codimers being produced; the reliability of
the estimates of ki0 a Eai is high and enables one to extrapolate even beyond the
temperature range under study.
Comparing these results with those of our previous kinetic studies [11, 13] we
can claim that, in the temperature range up to 100 C, the overall reaction rate of
cyclopentadieneisoprene codimerizations is by two orders of magnitude lower than
the cyclopentadiene dimerization. The activation energy Ea of the formation of the
two dominant codimers CPD-ISP 2 and 3 (by the cycloaddition reactions) is by ca.
10 kJ/mol higher than Ea of the CPD dimerization yielding endo-DCPD. Ea of the
formation of the other (minority) CPD-ISP codimers is by ca. 20 kJ/mol higher than
that of the reaction of CPD yielding endo-DCPD. Table 4 compares kinetic
parameters of the CPD-ISP 2 and 3 codimers formation with published data. In case
of the rate constants the deviations assume 70200 % of the value, while in the case
of activation energies the differences in the determination of the values do not
exceed 10 kJ/mol. The kinetic parameters of the formation of CPD-ISP 1, 5 and 6
have not been published so far.

Conclusions

It was verified that, in the temperature range of 70140 C, the codimerization of
cyclopentadiene and isoprene yields 6 isomeric products of the DielsAlder type.
Two of them prevail: endo-5-isopropenyl-2-norbornene, referred to in the literature
as codimer A, and 5-methyl-cis-3a,4,7,7a-tetrahydro-1H-indene, denoted as

123
326 Reac Kinet Mech Cat (2015) 116:315326

codimer B. At 120 C, the reaction mixture contains 6 times more codimer A than
codimer B. Two conflicting mechanisms of the formation of codimers A and B are
presented in the literature: (i) codimers are formed by independent parallel reactions
of cyclopentadiene and isoprene and (ii) only endo-5-isopropenyl-2-norbornene
(codimer A) and exo-5-isopropenyl-2-norbornene are formed, while 5-methyl-cis-
3a,7,7a-tetrahydro-1H-indene (codimer B) originates exclusively by the Cope
rearrangement from endo-5-isopropenyl-2-norbornene. To find out which of the
mechanisms is correct, we performed kinetic measurements that gave evidence that
codimer B can be formed by both reaction routes and it depends on the reaction
conditions which of them prevails. Kinetic parameters for the formation of five
isomeric products were determined.

Acknowledgments Financial support from specific university research (MSMT Czech Republic No
20/2015).

References

1. Singleton DA, Schulmeier BE, Hang C, Thomas AA, Leung SW, Merrigan SR (2001) Tetrahedron
57:5149
2. Firestone RA (1996) Tetrahedron 52:14459
3. Dewar MJ, Olivella S, Stewart JJ (1986) J Am Chem Soc 108:5771
4. Fringuelli F, Taticchi A (1990) Dienes in the DielsAlder reaction. Wiley, New York
5. Jasinski R, Kubik M, Lapczuk-Krygier A, Kacka A, Dresler E, Buguszewska-Czubara A (2014) Reac
Kinet Mech Cat 113:333
6. Jasinski R (2014) Comput Theor Chem 1046:93
7. Jasinski R, Kwiatkowska M, Baranski A (2007) Wiad Chem 61:485
8. Jasinski R, Kwiatkowska M, Baranski A (2011) J Phys Org Chem 24:843
9. Domingo LR, Aurell MJ, Perez P, Contreras R (2002) Tetrahedron 58:4417
10. Domingo LR, Saez JA (2009) Org Biomol Chem 7:3576
11. Krupka J, Pasek J, Lederer J, Bilkova D (2014) Pet Coal 56:428
12. Krupka J (2011) Pet Coal 53:233
13. Krupka J (2010) Pet Coal 52:290
14. Duschek CH, Pritzkow W (1970) J Prakt Chem 312:15
15. Tomi P, Fratiloiu R, Toma C, Pirvutoiu T (1977) Rev Chim 28:682
16. Iwase S, Nakata M, Hamanaka S, Ogawa M (1976) Bull Chem Soc Jpn 49:2017
17. Ishii Y, Nakagawa K, Yuki H, Iwase S, Hamanaka S, Ogawa M (1982) J Japan Pet Inst 25:58
18. Krupka J, Kolena J (2012) Pet Coal 54:385
19. Bell A, Jablonski D, Koronich E, Knapp B, Amoroso D (2014) (Promerus, LLC, US) Patent US
8704022
20. Belikova NA, Lermontov SA, Skornyakova TG, Pehk T, Lippmaa E, Plate AF (1979) J Org Chem
USSR 15:436
21. Nigmatova VB, Andreev VA, Pekhk TI, Belikova NA, Bobyleva AA, Anfilogova SI, Dubitskaya NF,
Koroza GA (1990) J Org Chem USSR 26:2213
22. Muja I, Andreescu G, Corciovei M, Fratiloiu R (1975) Rev Chim 26:981
23. Bulicev P, Lavrovskij KP, Rumyancev AN (1966) Neftekhimiya 6:690
24. Cope AC, Hardy EM (1940) J Am Chem Soc 62:441
25. Jasinski R, Kwiatkowska M, Sharnin V (2013) Monatsh Chem 144:327
26. Jasinski R, Baranski A (2010) J Mol Struct THEOCHEM 949:8
27. Davies HML, Dai X (2005) J Org Chem 70:6680
28. Chapuis C, Skuy D, de Saint Laumer JY, Brauchli R (2014) Chem Biodiv 11:1470
29. Muja I, Andreescu G, Corciovei M, Fratiloiu R (1975) Rev Chim 26:545
30. Zamostny P, Belohlav Z (1999) Comput Chem 23:479

123

Das könnte Ihnen auch gefallen