Вы находитесь на странице: 1из 17

j

A
PA
LE
IY
D
CP
AT
L
SS
I
A: GENERAL

ELSEVIER

Applied Catalysis A: General 146 (1996) 207-223

Theoretical study of the mechanism of branching


rearrangement of carbenium ions
M. Boronat a p. Viruela b A. Corma

a,*

a lnstituto de Tecnolog& Qulmica UPV-CSIC, Universidad Polit~cnica de Valencia, c~ Camino de Vera s~ n,


46071 Valencia, Spain
b Departament de Qulmica F&ica, Universitat de Valencia, c~ Dr. Moliner 50, 46100 Burjassot (Valencia),
Spain

Abstract
Owing to the practical interest of the acid catalyzed isomerization reactions of hydrocarbons,
the mechanism of the branching rearrangements of C4H ~- and C5H+1 carbenium ions has been
studied theoretically using ab initio methods which include electron correlation and extended basis
sets. It has been found that the protonated cyclopropane-type species does not appear as a common
intermediate for these reactions, since it is a transition state and not a minimum on the potential
energy surfaces studied. In the case of C4H ~- cation, the protonated methyl-cyclopropane ring is
the transition state for the carbon scrambling reaction in the secondary n-butyl cation, while the
isomerization of n-butyl cation into t-butyl cation occurs via a primary cation. The activation
energies calculated assuming this mechanism are in very good agreement with those obtained
experimentally. For the branching rearrangement of n-pentyl cation two reaction paths have been
considered. In the first one the secondary n-pentyl cation is converted through the 1,2-dimethylcyclopropane ring into the secondary 3-methyl-2-butyl cation, which is converted into the t-pentyl
cation by a 1,2-hydrogen shift. In the second one the secondary n-pentyl cation is directly
converted into the t-pentyl cation through a primary monobranched cation. Comparison of the
calculated activation energies for both mechanisms with the experimental value indicate that this
reaction does not occur via the primary cation as was the case for n-butyl cation, but occurs via
the protonated 1,2-dimethyl-cyclopropane ring.
Keywords." Carbeniumions; Cracking; Isomerization;Mechanism;Rearrangement

1. I n t r o d u c t i o n
C a t a l y t i c t r a n s f o r m a t i o n s o f h y d r o c a r b o n s such as i s o m e r i z a t i o n , a l k y l a t i o n
a n d c r a c k i n g are p r o c e s s e s o f great i m p o r t a n c e for the c h e m i s t r y o f p e t r o l e u m

* Corresponding author.
0926-860X/96/$15.00 Copyright 1996 Elsevier Science B.V. All rights reserved.
PII SO926-860X(96)O0160-3

208

M. Boronat et al. / Applied Catalysis A: General 146 (1996) 207-223

[1-3]. The most active and widely used solid acid catalysts for these reactions
are silica-alumina and zeolites either exchanged with transition metals [4-6] or
in their pure acid forms [7-9]. It is generally accepted that the interaction of
hydrocarbons with solid acids results in the formation of carbenium ions [10,11 ].
Consequently, it has been assumed that the mechanism of heterogeneous
reactions on solid acids is similar to that of homogeneous reactions in superacid
media, although the influence of the solid acid catalyst on the formation and
reactivity of the carbenium ion is not explicitly considered in this formal
mechanism.
The mechanism of homogeneous isomerization reactions in superacid media
involves three steps: formation of the carbenium ion either by protonation of an
alkene or by hydride transfer from an alkane, rearrangement of this carbenium
ion, and deprotonation or hydride ion abstraction to give the rearranged hydrocarbon. The carbenium ion rearrangements involved in the second step may be
classified as branching and non-branching. The classical mechanism for the
non-branching rearrangements, in which the degree of chain branching remains
the same, supposes them to proceed by a succession of 1,2-hydrogen and alkyl
shifts via secondary ions as intermediates. The branching rearrangements, which
are about a thousand times slower, involve a decrease or an increase in the
degree of chain branching. For this type of rearrangements, a mechanism with
only 1,2-hydrogen and alkyl shifts would necessarily include primary carbenium
ions as intermediates. This is not consistent neither with the experimental fact
that n-butane is not isomerized at an observable rate by H F / S b F 5 to isobutane
under conditions where n-pentane and n-hexane are rapidly converted into their
branched isomers [12], nor with the finding that the rate of scrambling or
isomerization of n-butane-l-13C to n-butane-2-13C is comparable to that of
isomerization of n-pentane to isopentane [13]. Consequently, the mechanism
proposed by Brouwer [14,15] which includes a protonated cyclopropane ring as
intermediate has been accepted.
According to this mechanism (see Scheme 1), the positive charge on C 2
carbon atom attacks the C 4 carbon atom and a protonated cyclopropane ring is
formed. The opening of the cyclic intermediate at one of the other two sides of

,.Ici\+

- -

/ c I..
3-

/C2--C4QR

\I

\ /

--C 1\+ 1c3\ /


C2
C4
I
I"R

..I 1\c
--C
~

+/ cl/_

. i I2--C4...R

\1
\ /
Y
q,, ~/c,3,,)/
,,/C2"--~---C4~.R
H+

>cN3
__~Cl [C2__~4j
/
[
~'R ~
Scheme 1.

,,I
--C3~+
I/
f 2 --C4~. R

M. Boronat et al./ Applied Catalysis A: General 146 (1996) 207-223

209

the ring results in the formation of a secondary monobranched carbenium ion if


R is an alkyl group. If R is an hydrogen atom, i.e., for C4H ~- cation, the
opening of the cyclic intermediate at side a leads to a n-butyl cation in which a
terminal and a non-terminal carbon atoms have changed positions (C scrambling
reaction), but the opening of the intermediate at side b becomes more difficult
because it would lead to a primary carbenium ion.
The accepted values for the activation energies for the carbon scrambling
process in the n-butyl cation and for the conversion of n-butyl into t-butyl cation
are 7.5 and 18.0 k c a l / m o l respectively [16,17]. The similarity of the last value
to the activation energy for the carbon scrambling in the isopropyl cation, 15.7
kcal/mol [18], is consistent with the supposition that the branching isomerization reaction of n-butyl cation passes through the high-energy primary isobutyl
cation. The low barrier obtained for the scrambling process, however, suggests
that this reaction passes through a protonated methyl-cyclopropane species,
more stable than the primary isobutyl cation. For the rearrangement of the
t-pentyl cation to the n-pentyl cation an experimental activation energy value of
18.3 k c a l / m o l is reported by Brouwer [14,15]. This reversible rearrangement
leads to interchange of the methyl and methylene carbon atoms and, consequently, to scrambling of the methyl and methylene protons in the observable
t-pentyl cation. From PMR spectroscopic measurements on the t-pentyl cation at
high temperatures Saunders and Rosenfeld have obtained a more precise value
of 18.8 k c a l / m o l for the activation energy of this proton scrambling process
[19]. The isomerization of the secondary 3-methyl-2-butyl cation into the tertiary
2-methyl-2-butyl or t-pentyl cation (the last step in Scheme 1) has been found to
have an activation energy value of 2.1 k c a l / m o l [20].
In order to establish the mechanism of the branching rearrangements of
carbenium ions, we present in this paper a complete theoretical study of the
potential energy surfaces of C4H ~- and C5H/1 cations which include geometry
optimization and characterization of the stationary points at correlated levels.

2. Computational details
All ab initio molecular orbital calculations in this work were performed on an
IBM 9021/500-2VF computer and on IBM RS/6000 workstations of the
University of Valencia using the Gaussian 88 [21] and Gaussian 92 [22]
computer programs.
The geometry of the stationary points on C4H ~- and C5H/1 potential energy
surfaces was first fully optimized by using the Hartree-Fock procedure and the
6-31G * basis set [23,24] (HF/6-31G* ) which has polarization functions (d-type)
on non-hydrogen atoms. Afterwards, electron correlation was included by means
of the second-order Moeller-Plesset perturbation theory that takes into account
the core electrons. Two types of calculations were carried out: single point

210

M. Boronat et al./Applied Catalysis A: General 146 (1996) 207-223

calculations on the HF optimized geometries using the MP2 treatment and the
6-31G * basis set ( M P 2 / 6 - 3 1 G * / / H F / 6 - 3 1 G * ), and complete geometry optimization of all stationary points at this correlated theoretical level ( M P 2 / 6 31G * ). The Berny analytical gradient [25] and the eigenvalue following [26,27]
methods were used for the minima and transition states geometry optimizations,
respectively. All H F / 6 - 3 1 G * and M P 2 / 6 - 3 1 G * stationary points were characterized by calculating the Hessian matrix and analyzing the vibrational normal
modes. The relative energies at these two levels were corrected by the zero point
energy (ZPE) obtained from frequency calculations.
3. Results and discussion
3.1.

C 4 H9+

According to Scheme 1, there are four structures involved in the mechanism


of branching rearrangement of C4H ~- cation: the secondary n-butyl cation, the
protonated methyl-cyclopropane ring, the primary isobutyl cation and the tertiary isobutyl cation. The geometry of these four structures has been completely
optimized without any symmetry restriction.
The tertiary isobutyl cation D is the most stable minimum on C4H ~- potential
energy surface both at the H F / 6 - 3 1 G * and M P 2 / 6 - 3 1 G * theoretical levels. As
can be seen in Fig. 1, the conformation adopted minimizes the repulsions
between the hydrogen atoms and also between the C - H and C - C bonds.

.479)
1.491

1.447(1.403),f\2.364(1.923)
\~
A
~1.517)

1.518(1.S11)
)1.494
1.496

"~1.288
B
~1.473(1.459)

Fig. 1. Structures of the C4H ff cation. The HF/6-31G* bond lengths are given first, followed by the
MP2/6-31G* values in parentheses.

M. Boronat et al. / A p p l i e d Catalysis A: General 146 (1996) 207-223

211

The structure of the secondary n-butyl cation has already been studied at
many standard theoretical levels by Schleyer et al. [28]. They found that both the
partially methyl-bridged form A (shown in Fig. 1) and a trans-hydrogen bridged
form (not shown) are minima on the potential energy surface, structure A being
slightly more stable than the hydrogen-bridged one at every considered theoretical level. Since apart from being the most stable, structure A shows a strong
positive charge on a secondary carbon atom, it seems more suitable to induce
the cycle formation process and consequently we have taken it as the starting
point for the carbon scrambling and branching isomerization reactions.
In the present work the geometry of structure A has been completely
optimized at the HF/6-31G* and MP2/631G* levels and, as was previously
reported by Schleyer, both stationary points have been found to be minima on
the potential energy surface. The bond lengths depicted in Fig. 1 show the
changes in geometry due to the inclusion of electron correlation. The degree of
methyl-bridging becomes more important when electron correlation is included,
as can be observed from the !engthening of the C3-C 4 bond from 1.584 ,~ at the
HF/6-31G * level to 1.653 A at the MP2/6-31G * level, and also from the fact
that the C2C3C 4 angle is closed from 102.4 at the HF/6-31G * level to 77.5
at the M P 2 / 6 - 3 1 G * level.o This implies an important ShOortening of the C2-C 4
bond length from 2.364 A at the H F / 6 3 1 G * level to 1.923 A at the M P 2 / 6 3 1 G *
level, while the C 1-C3 bond length remains the same (2.582 and 2.549 ,~). The
calculated energy differences between the secondary A and the tertiary D butyl
cations (or relative energies) summarized in Table 1 are in good agreement with
the experimental values of the enthalpy of rearrangement of n-butyl to t-butyl
cation, which are 15-17 [29-31] and 14-15 [32,33] kcal/mol in gas phase and
in solution, respectively.
Three different structures were proposed by Wiberg and Kass [34] for the
protonated methyl-cyclopropane ring: a corner-protonated species formed by
adding a proton to the methine carbon atom, an edge-protonated species in
which the proton is shared by two neighboring methylene carbon atoms, and a
structure formed by adding a proton to a methylene carbon atom which resulted
to be the open form of the secondary n-butyl cation A. Although at the
Hartree-Fock level the corner-protonated species was found to be more stable
than the edge-protonated one, the inclusion of the electron correlation reversed
Table 1
Calculated relative energies (kcal/mol) of the stationary points found on C4H ~- potential energy surface
Method

HF/6-31G *
HF/6-31G * +ZPE
MP2/6-31G * / / H F / 6 - 3 1 G *
MP2/6-31G *
MP2/6-31G * +ZPE

13.6
14.7
13.9
11.0
12.9

31.5
32.8
20.2
19.9
21.5

32.9
33.3
34.2
33.6
34.1

0.0
0.0
0.0
0.0
0.0

212

M. Boronat et al. / Applied Catalysis A: General 146 (1996) 207-223

the relative stability of these two forms. Besides, the optimized bond lengths
reported by Wiberg and Kass for the corner-protonated species suggest that it
could be best described as a complex between CH3CH ~- cation and CH 2 = C H 2
molecule, without any chemical meaning in the rearrangements studied. Consequently, we have only considered in this work the edge-protonated form of the
methyl-cyclopropane ring (structure B in Fig. 1).
The HF/6-31G* and M P 2 / 6 3 1 G * optimized geometries of B depicted in
Fig. 1 are nearly identical, and exhibit a partially distorted C S symmetry. The
C2-C 3 and C2-C 4 bond lengths in the ring and the C - H bond lengths of the
almost symmetrical hydrogen bridge are equivalent at both theoretical levels.
The only noticeable change that electron correlation introduces in the geometry
of B is a slight shorteningoOf the C3-C 4 bond length from 1.763 A at the
HF/6-31G * level to 1.717 A at the MP2/6-31G * level. Despite this similarity
between the correlated and uncorrelated optimized geometries of B, the calculated relative energies summarized in Table 1 are strongly dependent on the
theoretical level used. At the HF/6-31G * level the relative energy of B is 31.5
kcal/mol, and 32.8 k c a l / m o l with the ZPE correction. The inclusion of the
electron correlation at the M P 2 / 6 - 3 1 G * / / H F / 6 - 3 1 G * level stabilizes this
structure and yields a relative energy value of 20.2 kcal/mol, very similar to the
values obtained at the MP2/6-31G * and MP2/6-31G * + ZPE levels, 19.9 and
21.5 kcal/mol, respectively.
These energetic changes are not surprising if we take into account that the
inclusion of the electron correlation by the Moeller-Plesset perturbation treatment preferentially stabilizes the non-classical bridged structures [35]. Since the
HF/6-31G* optimized geometry of B corresponds to a non-classical bridged
structure, the single point calculation at the MP2/6-31G * / / H F / 6 - 3 1 G * level
is able to introduce all the stabilization due to electron correlation. The
MP2/6-31G * geometry optimization cannot increase the degree of bridging in
B and therefore there is no stabilization with respect to the single point
calculation observed at this level of theory.
What is more surprising is the result obtained from the analysis of the
vibrational normal modes. This analysis indicates that structure B is a transition
state on C4H ~- potential energy surface both at the HF/6-31G* and M P 2 / 6 31G* levels of theory. The negative frequency is clearly associated to the
movement of the bridged hydrogen atom towards one of the methylene carbon
atoms to give the secondary n-butyl cation A. The fact that structure B is not a
minimum but a transition state on the potential energy surface means that the
protonated meth2)l-cyclopropane ring cannot be a true intermediate of the
scrambling and branching isomerization reactions of n-butyl cation as suggested
by Brouwer, but only a species through which the scrambling reaction passes.
Finally, the structure of the primary isobutyl cation C has been calculated.
Taking as a starting point the optimized geometry of the tertiary isobutyl cation
D, the distance between one of the hydrogen atoms attached to C 3 and C2 has

M. Boronat et al. / Applied Catalysis A: General 146 (1996) 207-223

213

been slowly shortened from 2.037 ,~ to 1.075 A, already corresponding to a


primary cation. In each calculation the C 2 - H distance has been fixed and all
other parameters have been allowed to fully optimize. Then, the geometry of the
primary cation obtained has been completely reoptimized using the eigenvector
following transition state search technique. The H F / 6 - 3 1 G * and M P 2 / 6 - 3 1 G *
optimized geometries of C are very similar, and the relative energies summarized in Table 1 are all of them about 33 or 34 kcal/mol. The force constant
calculations indicate that at both theoretical levels structure C is a transition
state with only one imaginary vibration frequency.
Taking into account all these data and the idea that a transition state must be
connecting two minima on a potential energy surface, a new mechanism, shown
in Scheme 2, is proposed for the carbon scrambling and branching isomerization
reactions of n-butyl cation. According to this new mechanism the secondary
n-butyl cation is the starting point for the two reactions that, from the first
moment, follow different reaction paths. When the positive charge on C 2 carbon
atom in minimum A attacks C 4 carbon atom two different processes can take
place: (a) a simultaneous strengthening of the C 2 - C 4 bond and breaking of the
C3-C 4 bond in A directly leads to transition state C and from this, a shift of the
hydrogen atom from C 2 to C 3 leads to minimum D. The primary cation is the
transition state for the conversion of n-butyl into t-butyl cation; (b) strengthening of the C 2 - C 4 bond in A together with an elongation of a C4-H bond length
leads to transition state B. If the C 2 - C 3 bond is then broken and the hydrogen
atom of the bridge moves to C 3, the minimum A' (equivalent to A) is reached
and the carbon scrambling reaction has occurred. The protonated cyclopropane
ring is, at least in the case of C4H ~- cation, not an intermediate but only the
transition state for the carbon scrambling reaction.
Fig. 2 shows the energetic profile for the two reactions studied and Table 2
summarizes the calculated activation energies together with experimental data.
The activation energy for the carbon scrambling process, Eal, has been calcu-

/
\1

\i

--CiN+ i c 3 \ /
72 74.

A j
B

[-\1
I+ -It
I -q
~c3/ ~
J ~
L "7<
"

C
Scheme 2.

i<
A'

\1
I/
-q\+/c3--

12
i7'\
D

214

M. Boronat et al. /Applied Catalysis A: General 146 (1996) 207-223


AE

(kcal/mol)

35.0-

I
I

30.0 -

I
I
i I E~a

25.0 -

/--'~\

20.0 15.0-

r.-J
A'

//

~Eal \N

I
I

10.0-

5.0-

0.0-

Fig. 2. Energy profile corresponding to the scrambling and isomerization reactions of the secondary n-butyl

cation. The tertiary ion has been taken as the origin of energies.

lated as the energy difference between the protonated methyl-cyclopropane ring


B and the secondary n-butyl cation A. The H F / 6 - 3 1 G * and H F / 6 - 3 1 G * + ZPE
values, 17.9 and 18.1 kcal/mol respectively, are too high as compared with the
experimental value of 7.5 kcal/mol. As we have seen before, electron correlation stabilizes structure B and the calculated activation energies at the M P 2 / 6 31G * / / H F / 6 - 3 1 G *, M P 2 / 6 - 3 1 G * and MP2/6-31G * + ZPE levels, 6.3, 8.9
and 8.6 kcal/mol respectively, are much closer to the experimental value. These
results indicate that the Hartree-Fock level is not adequate to study mechanisms
in which non-classical species are involved, and that inclusion of electron
correlation is essential to treat these species correctly. The activation energies
for the direct (n-butyl ~ t-butyl), Ea2, and reverse (t-butyl ~ n-butyl), Ea3,
isomerization reactions have been calculated as the energy differences between
the primary cation C and structures A and D, respectively. The calculated
activation energies are quite similar at all theoretical levels, between 18.6 and
22.6 kcal/mol for the direct reaction and between 32.9 and 34.2 kcal/mol for
the reverse reaction, and in good agreement with the experimental values.

Table 2
2 + 13Calculated and experimental activation energies (kcal/mol) for the processes depicted in Fig. 2
Method

gal

Ea2

Ea3

HF/6 -3 1G *
HF/6-31G* + Z P E
MP2/6-31G * / / H F / 6 - 3 1 G *
MP2/6-31G *
MP2/6-31G * + Z P E
Exp.

17.9
18.1
6.3
8.9
8.6
7.5

19.3
18.6
20.3
22.6
21.2
18.0

32.9
33.3
34.2
33.6
34.1
32.0

M. Boronat et al. / Applied Catalysis A: General 146 (1996) 207-223

215

3.2. C5H1~
According to Brouwer's mechanism (see Scheme 1), the following structures
have to be localized and characterized on CsH~-~ potential energy surface: the
secondary n-pentyl cation, the protonated 1,2-dimethyl-cyclopropane ring, the
secondary 3-methyl-2-butyl cation, the tertiary 2-methyl-2-butyl or t-pentyl
cation and the transition state for the hydrogen shift that converts the secondary
branched cation into the tertiary one.
Both at the H F / 6 - 3 1 G * and M P 2 / 6 - 3 1 G * levels the tertiary t-pentyl cation
I is the most stable minimum on CsH~-~ potential energy surface. The optimized
bond lengths depicted in Fig. 3 indicate that the t-pentyl cation is not fully
classical. A partial bridging between the C4-C 5 bond and the positive charge on

I .S53

~.519(1.510)

1.526~'~ 'q
747)

74)

1.59Z

1.462

(I .489)

.836) ~"~0
G

_ 1.471 ~x "474(1.464)

I
Fig. 3. Structures of C5H~1 cation. The HF/6-31G*
MP2/6-31G* values in parentheses.

1.589~ 1"529 (I"516)

J
bond lengths are given first, followed by the

216

M. Boronat et al. /Applied Catalysis A: General 146 (1996) 207-223

Table 3
Calculated relative energies (kcal/mol) of the stationary points found on CsHll + potential energy surface
Method

HF/6-31G *
HF/6-31G * +ZPE
MP2/6-31G * / / H F / 6 - 3 1 G *
MP2/6-31G *
MP2/6-31G * +ZPE

13.6
14.2
13.6
10.3
11.7

29.5
30.2
18.5
17.1
17.8

12.9
13.0
11.8
7.2
8.0

15.6
14.5
12.0
14.8
14.4

0.0
0.0
0.0
0.0
0.0

34.2
34.9
35.0
34.5
34.1

C 2 carbon atom can be deduced from the lengthening of the C 4 - C 5 bond to

1.565 ,~ at the HF/6-31G* level and 1.581 A at the MP2/6-31G * level, and
also from the calculated C2C4C 5 angle values of 106.3 and 101.5 at the
HF/6-31G * and MP2/6-31G * levels, respectively. The structure of the t-pentyl
cation has already been studied by comparing the ~3C chemical shifts of the C +
carbon atom calculated by IGLO using ab initio geometries with the experimental values, and our results are in complete agreement with those previously
reported [36].
The secondary n-pentyl cation E has also been found to be a minimum on the
potential energy surface at the HF/6-31G* and MP2/6-31G* theoretical
levels. The optimized bond lengths depicted in Fig. 3 show the same tendencies
that were observed in the secondary n-butyl cation A. Inclusion of electron
correlation at the MP2/6-31G * level lengthens the C3-C 4 bond from 1.599
to 1.656 ~, and closes the C2C3C 4 angle from 102.1 to 78.3 , i.e., increases the
degree of methyl-bridging. This is reflected in the important shortenin~ of the
C a - C 4 bond length from 2.366 ,~ at the HF/6-31G* level to 1.941 A at the
MP2/6-31G* level, while the C3-C 5 bond length remains nearly constant
(2.545 and 2.587 A at the HF/6-31G * and MP2/6-31G * levels respectively).
These geometric changes are reflected in the relative energies summarized in
Table 3. The H F / 6 - 3 1 G * , HF/6-31G* + ZPE and M P 2 / 6 - 3 1 G * / / H F / 6 31G* calculated energies, 13.6, 14.2 and 13.6 kcal/mol respectively, are very
similar. The increase in the degree of methyl-bridging produced by the inclusion
of the electron correlation in the MP2/6-31G * optimization stabilizes structure
E and at the MP2/6-31G* and M P 2 / 6 - 3 1 G * + ZPE levels the calculated
relative energies are 10.3 and 11.7 kcal/mol respectively.
The third minimum found on CsH~] potential energy surface is the secondary
3-methyl-2-butyl cation, structure G in Fig. 3. At the HF/6-31G* level the
C2-C 3 bond length value of 1.592 A and the C3C2C 4 angle value of 98.1 are
indicative of a partial bridging between the C 2-C3 bond and the positive charge
on C 4 carbon atom, similar to that previously reported for the t-pentyl cation. At
the MP2/6-31G* level, however, the C2-C 3 bond is lengthened to 1.733 ~,,
the C3C2C 4 angle is closed to 70.9 and the C 3 - C 4 bond is shortened from
2.295 A at the uncorrelated level to 1.836 .A. The degree of methyl-bridging
becomes so important that this structure could be best described as an unsym-

M. Boronat et al./Applied Catalysis A: General 146 (1996) 207-223

217

metrical comer-protonated 1,2-dimethyl-cyclopropane ring. The relative energies summarized in Table 3 reflect the importance of the geometric changes
produced when electron correlation is included in the calculations. The H F / 6 31G* and HF/6-31G* + Z P E calculated relative energies, 12.9 and 13.0
kcal/mol respectively, are about 2 kcal/mol too high in relation to the
experimental value of 11 kcal/mol reported by Collin and Herman [20] for the
energy difference between the secondary 3-methyl-2-butyl and the tertiary
2-methyl-2-butyl cations. The M P 2 / 6 - 3 1 G * / / H F / 6 - 3 1 G * calculated value,
11.8 kcal/mol, is slightly lower than the two uncorrelated values, probably due
to the fact that there is a partial bridging in the HF/6-31G * optimized geometry
of structure G. But the MP2/6-31G * and MP2/6-31G * + ZPE values are 3.8
and 3 kcal/mol respectively too low in relation to the experimental value. This
overestabilization of structure G at the best levels of theory used in this work
can be explained if we take into account that the inclusion of the electron
correlation using the Moeller-Plesset treatment preferentially stabilizes nonclassical bridged structures [35] and, as can be seen in Fig. 3, the degree of
bridging in G is more important than in any of the other structures.
According to Brouwer's mechanism (Scheme 1), the isomerization of the
linear E to the branched G secondary cations passes through an edge-protonated
cyclopropane ring F. Taking as a starting point the optimized geometry of the
linear cation E, two variables have been simultaneously controlled in order to
obtain structure F: the C2C3C 4 angle and the C3-H bond length. In each
calculation these two variables have been fixed and all other parameters have
been allowed to fully optimize. The geometry of the cyclic structure obtained
has been completely reoptimized at the HF/6-31G * and MP2/6-31G * levels
using the eigenvalue following transition state search technique and the two
stationary points have been characterized by force constant calculations. As in
the case of C4H ~- cation, they have been found to be transition states, with only
one imaginary vibration frequency associated to the movement of the bridged
hydrogen atom.
The two optimized geometries of structure F depicted in Fig. 3 are ne~ly
equivalent, with C - C bond lengths in the ring of 1.450, 1.553 and 1.807 A at
the HF/6-31G * level and 1.443, 1.562 and 1.747 A at the MP2/6-31G * level.
The effect of alkyl substitution on the structure of the protonated cyclopropane
ring can be clearly observed in the different symmetry exhibited by the
hydrogen bridge in B (protonated methyl-cyclopropane ring) and F (protonated
1,2-dimethyl-cyclopropane ring). While the two C - H bond lengths in structure
B are equivalent, the hydrogen bridge in F is markedly unsymmetrical, with
C - H bond lengths of 1.176 and 1.478 A at the HF/6-31G * level and 1.174 and
1.490 ~, at the MP2/6-31G* level.
The relative energies of structure F reproduce the tendencies previously
observed for structure B. The HF/6-31G* and HF/6-31G * + ZPE calculated
values are high, 29.5 and 30.2 kcal/mol respectively. The single point calcula-

218

M. Boronat et al. / Applied Catalysis A: General 146 (1996) 207-223

tion at the M P 2 / 6 - 3 1 G * / / H F / 6 - 3 1 G * level strongly stabilizes this cyclic


species, yielding a value of 18.5 kcal/mol, while the MP2/6-31G* and
MP2/6-31G* + ZPE calculated values are only slightly lower, 17.1 and 17.8
kcal/mol respectively.
The last step in Brouwer's mechanism is the hydrogen shift that converts the
secondary branched cation into the tertiary one. Starting from the optimized
geometry of I, the distance between C 2 and the hydrogen atom that is going to
migrate has been slowly shortened from 2.12 A to 1.10 A and the geometry of
the obtained structure has been then completely reoptimized at the HF/6-31G *
and MP2/6-31G* levels using the eigenvector following the transition state
search technique. The two stationary points have been characterized by force
constant calculations and they have been found to be transition states on their
respective potential energy surfaces, showing only one imaginary vibration
frequency. At the HF/6-31G* level the relative energy of structure H is 15.6
kcal/mol, and 14.5 kcal/mol with the ZPE correction. The single point
calculation at the MP2/6-31G * / / H F / 6 - 3 1 G * level stabilizes this hydrogenbridged structure and yields a relative energy value of 12.0 kcal/mol. However,
the value obtained from the MP2/6-31G * optimization is higher, 14.8 kcal/mol,
and 14.4 kcal/mol with the ZPE correction. The HF/6-31G * and MP2/6-31G *
optimized bond lengths depicted in Fig. 3 explain this energetic values. At the
HF/6-31G * level, the optimized geometry of H corresponds to an unsymmetrically hydrogen-bridged structure with a C-C bond length of 1.411 A and two
C - H bond lengths of 1.189 and 1.582 A. At the MP2/6-31G* level the
optimized C-C bond length value of 1.437 A and the C - H bond length values
of 1.118 and 1.961 A are indicative of a lesser degree of hydrogen bridging and
consequently the structure is slightly destabilized at this level of theory.
Taking into account all these results, the mechanism depicted in Scheme 3 is
proposed for the branching rearrangement of n-pentyl cation. According to it,
strengthening of the C2-C 4 bond in minimum E together with weakening of the
C3-C 4 bond and migration of one of the hydrogen atoms attached to C 4 to a
bridged position between C 3 and C a leads to transition state F. From this,
braking of the C 3-C4 bond together with migration of the bridged hydrogen
atom to C 3 leads to minimum G. Then, the hydrogen atom attached to C 2
migrates to C 4 through transition state H and minimum I is reached. It is
important to note that this calculated mechanism is not equivalent to that
empirically proposed by Brouwer, the main difference being the nature of the
o

,,t

--c

\c(

c/

\i

[\~1

\]

~'--

It

,,,/

c3

(3

Scheme 3.

,,I/

c,3

lt,.i
|

--c3

M. B o r o n a t et al. / Applied Catalysis A: General 146 (1996) 2 0 7 - 2 2 3

\1

\/

--ctx, + / c 3 \

\+/
c3

I/

/c5-_

7 /k

\
/ CI

219

It

\l

72 C4"~5_.j _
_

~7

"C5--i\

Scheme 4.

protonated 1,2-dimethyl-cyclopropane ring. Our results indicate that this cyclic


species cannot be an intermediate as affirmed by Brouwer, because it is a
transition state and not a minimum on CsH~-1 potential energy surface.
A second mechanism, equivalent to that calculated for the branching isomerization of n-butyl cation, has been considered in this work. As can be seen in
Scheme 4, a simultaneous strengthening of the C2-C 4 bond and breaking of the
C 3 - C 4 bond in the secondary n-pentyl cation leads to a primary monobranched
cation J and from this, a shift of an hydrogen atom from C 2 to C 3 leads to the
t-pentyl cation. The structure of the primary cation has also been calculated.
Starting again from the optimized geometry of the t-pentyl cation I, the distance
between C 2 and one of the hydrogen atoms attached to C 3 has been slowly
shortened from 2.12 ,~ to 1.10 A and then, using the eigenvalue following
method, its geometry has been completely reoptimized at the H F / 6 - 3 1 G * and
M P 2 / 6 - 3 1 G * levels. Characterization of the two stationary points by force
constant calculations indicate that the primary pentyl cation J is a transition state
at both theoretical levels. Since the HF/6-31G* and MP2/6-31G* optimized
geometries of structure .1 depicted in Fig. 3 are very similar and fully classical,
the inclusion of the electron correlation in the calculations produces no stabilization of this structure in relation to the t-pentyl cation and the calculated relative
energies of J summarized in Table 3 are all of them similar, between 34 and 35
kcal/mol.
AE (kcal/mol)

35.0-

/
I

30.0/
/

20.0 -

10.0-

0.0-

\
\
\
\

\
\

I
I

E l/if I ~\

I
I
I

Ea4

5.0-

25.0-

15.0-

\
\

~a6

\\
\

\\ t
I

Fig. 4. Energy profile corresponding to the branching isomerization of the secondary n-pentyl cation. The
tertiary ion has been taken as the origin of energies.

220

M. Boronat et al. /Applied Catalysis A: General 146 (1996) 207-223

Fig. 4 shows the energetic profile for the branching isomerization of the
n-pentyl cation. The activation energy for the rearrangement of the t-pentyl
cation to the n-pentyl cation, for which experimental data are available, can be
calculated as the energy difference between the primary transition state J and the
tertiary minimum I if the reaction path depicted in Scheme 4 is followed. In the
mechanism of Scheme 3, the rate determining step is the conversion of the
branched G into the linear E secondary pentyl cations, and consequently the
activation energy for the global process is the energy difference between
transition state F and minimum I. The calculated activation energies together
with available experimental data are summarized in Table 4. At the HF/6-31G *
and HF/6-31G * + ZPE levels the calculated activation energies for the two
mechanisms are similar, 29.5 and 30.2 kcal/mol for Ea4 and 34.2 and 34.9
kcal/mol for E,6, and too high as compared with the experimental value of 18.8
kcal/mol. When electron correlation is included the energy of the primary
cation J experiments no changes, and the calculated values for Ea6 are again
between 34 and 35 kcal/mol. However, the transition state F is highly stabilized
by inclusion of electron correlation and consequently the calculated values for
Ea4 at the M P 2 / 6 - 3 1 G * / / H F / 6 - 3 1 G * , MP2/6-31G* and MP2/6-31G* +
ZPE levels are lowered to 18.5, 17.1 and 17.8 kcal/mol respectively. Comparison of the calculated activation energies for the two mechanisms considered
with the experimental value of 18.8 kcal/mol indicates that the branching
isomerization of the n-pentyl cation does not occur through the primary cation,
as was the case for the n-butyl cation, but occurs via the protonated 1,2-dimethyl-cyclopropane ring, following the reaction path shown in Scheme 3.
As already told, this mechanism consists of two steps: the conversion of the
secondary linear cation E into the secondary branched cation G discussed above,
and the conversion of G into the tertiary cation I. The activation energy for this
process, Ea5 in Fig. 4 and Table 4, can be calculated as the energy difference
between transition state H and minimum G. The Ea5 calculated values at the
HF/6-31G * and HF/6-31G * + ZPE levels, 2.7 and 1.5 kcal/mol respectively,
compare well with the experimental value of 2.1 kcal/mol. At the MP2/631G * / / H F / 6 - 3 1 G * level the obtained value is too low, 0.2 kcal/mol, while
the MP2/6-31G* optimization yields a too high barrier for this process, 7.6
kcal/mol, and 6.4 kcal/mol with the ZPE correction. The reported energy
Table 4
Calculated and experimental activation energies (kcal/mol) for the processes depicted in Fig. 4
Method

Ea4

Ea5

Ea6

HF/6-31G *
HF/6-31G * + ZPE
MP2/6-31G * / / H F / 6 - 3 1 G *
MP2/6-31G *
MP2/6-31G * +ZPE
Exp.

29.5
30.2
18.5
17.1
17.8
18.8

2.7
1.5
0.2
7.6
6.4
2.1

34.2
34.9
35.0
34.5
34.1
-

M. Boronat et al. /Applied Catalysis A: General 146 (1996) 207-223

221

difference between the secondary 3-methyl-2-butyl and the t-pentyl cations, 11


kcal/mol, together with the activation energy value for the hydrogen shift that
converts this secondary branched cation into the tertiary one yields an energy
barrier for the conversion of the tertiary I into the secondary G cations of 13.1
kcal/mol, which corresponds to the energy difference between transition state H
and minimum I. Exceptuating the HF/6-31G * result, which is 2.5 kcal/mol
too high, the calculated relative energies of structure H are between 12.0 and
14.8 kcal/mol, i.e., they are basically correct. Consequently, the source of the
difference between the calculated and the experimental values of Ea5 must be in
the calculated energy of the secondary branched cation G. As has been
previously discussed, the Moeller-Plesset perturbation treatment overestabilizes
non-classical bridged structures such as G. The 3 or more kcal/mol of
discrepancy between the calculated and the experimental relative energies of G
reported before added to the 1-2 kcal/mol found for transition state H are the
reason that explains the too high Ea5 values obtained when the MP2 treatment is
used.

4. Conclusions

A theoretical study of the potential energy surfaces of C4H ~- and C5H~1


cations including geometry optimization and characterization of the stationary
points at correlated levels has been carried out in order to establish the
mechanism of branching rearrangement of carbenium ions. The results obtained
suggest alternative mechanisms to that empirically proposed by Brouwer for
these reactions, the main difference being the nature of the protonated cyclopropane-type species. According to Brouwer's mechanism, this cyclic type of
structure is an intermediate, from which two different routes lead to two
different products. However, both the protonated methyl-cyclopropane ring and
the protonated 1,2-dimethyl-cyclopropane ring have been found to be transition
states and not minima on Call ~- and CsH~-1 potential energy surfaces respectively, and consequently they cannot be true intermediates but only species
through which the reactions pass.
The new mechanism proposed for the carbon scrambling and branching
isomerization reactions of the n-butyl cation (Scheme 2) supposes that, from the
first moment, the two reactions follow different reaction paths. The protonated
methyl-cyclopropane ring is the transition state for the carbon scrambling
reaction, and the isomerization of the linear n-butyl cation into the branched
t-butyl cation occurs through a primary cation. However, the branching rearrangement of n-pentyl cation does not occur through a primary cation as shown
in Scheme 4, but it follows the reaction path depicted in Scheme 3. The
secondary n-pentyl cation is converted through the protonated 1,2-dimethylcyclopropane ring into the secondary 3-methyl-2-butyl cation and then, an

222

M. Boronat et al. / Applied Catalysis A: General 146 (1996) 207-223

hydrogen shift converts this branched secondary cation into the tertiary one.
This mechanism can be extrapolated to higher aliphatic carbenium ions because,
as can be observed in Scheme 3, it does not seem probable that addition of a
methyl (or alkyl) group to the C5 carbon atom introduces any important change
in the relative energy or nature of the different species involved.
From the results obtained at the different levels of calculation it can also be
concluded that the inclusion of the electron correlation in the geometry optimizations and in the characterization of the stationary points is essential for the study
of this type of reaction mechanisms, in which non-classical bridged structures
are involved. Thus, the MP2/6-31G* would be the lowest acceptable level of
calculation.

Acknowledgements
The authors thank the Centre de Informhtica and Departament de Qufmica
Ffsica of the University of Valencia for computing facilities. They thank
C.I.C.Y.T. (Project MAT 94-0359) and Conselleria de Cultura, Educaci6 i
Cibncia de la Generalitat Valenciana for financial support. MB thanks the
Conselleria de Cultura, Educaci6 i Ci~ncia de la Generalitat Valenciana for a
personal grant.
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]

H. Pines, The Chemistry of Catalytic Hydrocarbon Conversion, Academic Press, New York, 1981.
J.H. Jenkins and T.W. Stephens, Hydrocarbon Proc., 60 (1980) 163.
F.E. Condon, in P.H. Emmet (Editor), Catalysis, Reinhold, New York, 1958, Vol. VI, Chap. 2.
P.A. Jacobs, Carboniogenic Activity of Zeolites, Elsevier, Amsterdam, 1977, Chap. 4.
K. Tanabe, Solid Acids and Bases, Academic Press, New York, 1970, p, 73.
J.W. Ward and R.C. Hansford, J. Catal., 13 (1979) 364.
A. Corma and A. Lopez Agudo, React. Kinet. Catal. Lett., 15 (1981) 253.
K.M. Minachev, Acta Phys. Chem., 24 (1978) 5.
M. Daage and J. Fajula, J. Catal., 81 (1983) 394, 405.
B.C. Gates, J.R. Katzer and G.C.A. Schuit, in Chemistry of Catalytic Processes, McGraw-Hill, New
York, 1979, Chap. 1.
H.H. Vogue, in P.H. Emmet (Editor), Catalysis, Reinhold, New York, 1958, Vol. VI, Chap. 5.
D.M. Brouwer and J.M. Oelderik, Rec. Trav. Chim., 87 (1968) 721.
D.M. Brouwer, Rec. Trav. Chim., 87 (1968) 1435.
D.M. Brouwer and H. Hogeveen, Progr. Phys. Org. Chem., 9 (1972) 179.
D.M. Brouwer, in R. Prins and G.C.A. Schuit (Editors), Chemistry and Chemical Engineering of
Catalytic Processes, Sijthoff and Noordhoff, Alphen aan den Rijn, The Netherlands, 1980, p. 137.
M. Saunders, E.L. Hagen and J. Rosenfeld, J. Am. Chem. Soe., 90 (1968) 6882.
P.C. Myhre and C.S. Yannoni, J. Am. Chem. Soc., 103 (1981) 230.
G.A. Olah and M. White, J. Am. Chem. Soc., 91 (1969) 5801.
M. Saunders and J. Rosenfeld, J. Am. Chem. Soc., 91 (1969) 7756.
G.C. Collin and J.A. Herman, J. Chem. Soc. Faraday Trans., 74 (1978) 1939.
M.J. Frisch, M. Head-Gordon, H.B. Schlegel, K. Raghavachari, J.S. Binkley, C. Gonzalez, D.J. DeFrees,
D.J. Fox, R.A. Whiteside, R. Seeger, C.F. Melius, J. Baker, R.L. Martin, R.L. Kahn, J.J.P. Stewart, E.M.
Fluder, S. Topiol and J.A. Pople, Gaussian 88, Gaussian, Pittsburgh, PA, 1988.

M. Boronat et al. / Applied Catalysis A: General 146 (1996) 207-223

223

[22] M.J. Frisch, G.W. Trucks, M. Head-Gordon, P.M.W. Gill, M.W. Wong, J.B. Foresman, B.G. Johnson,
H.B. Schlegel, M.A. Robb, E.S. Replogle, R. Gomperts, J.L. Anches, K. Raghavachari, J.S. Binkley, C.
Gonzalez, R.L. Martin, D.J. Fox, D.J. DeFrees, J. Baker, J.J.P. Stewart and J.A. Pople, Gaussian 92,
Gaussian, Pittsburgh, PA, 1992.
[23] W.J. Hehre, L. Radom, P.v.R. Schleyer and J.A. Pople, Ab initio Molecular Orbital Theory, Wiley-Interscience, New York, 1986.
[24] P.C. Hariharan and J.A. Pople, Chem. Phys. Lett., 16 (1972) 217.
[25] H.B. Schlegel, J. Comp. Chem., 3 (1982) 214.
[26] J. Baker, J. Comp. Chem., 7 (1986) 385.
[27] J. Baker, J. Comp. Chem., 8 (19871 536.
[28] J.W. de M. Carneiro, P.v.R. Schleyer, W. Koch and K. Raghavachari, J. Am. Chem. Soc., 112 (199(/)
4064.
[29] J.L. Franklin in G.A. Olah and P.v.R. Schleyer (Editors), Carbonium Ions, Interscience, Vol. 1, New
York, 1968.
[30] F.P. Lossing and G.P. Semeluk, Can. J. Chem., 48 (19701 955.
[31] J.J. Solomon and F.H. Field, J. Am. Chem. Soc., 97 (1975) 2625.
[32] E.W. Bittner, E.M. Arnett and M. Saunders, J. Am. Chem. Soc., 98 (1976) 3734.
[33] E.M. Arnett and C. Petro, J. Am. Chem. Soc., 100 (1978) 5408.
[34] K.B. Wiberg and S.R. Kass, J. Am. Chem. Soc., 107 (19851 988.
[35] K. Raghavachari, R.A. Whiteside, J.A. Pople and P.v.R. Schleyer, J. Am. Chem. Soc., 103 (1981) 5649.
[36] P.v.R. Schleyer, J.W. de M. Carneiro, W. Koch and D.A. Forsyth, J. Am. Chem. Soc., 113 (1991) 3990.

Вам также может понравиться