You are on page 1of 11

DOI: 10.1002/chem.

200902024

Double Group Transfer Reactions: Role of Activation Strain


and Aromaticity in Reaction Barriers

Israel Fernndez,*[a] F. Matthias Bickelhaupt,*[b] and Fernando P. Cosso*[c]

Abstract: Double group transfer typal DGT reactions using density overall reaction barrier DE ¼6 = DE-
(DGT) reactions, such as the bimolecu- functional theory (DFT) at the OLYP/ ACHTUNGRE(z=zTS) is interpreted in terms of the
lar automerization of ethane plus TZ2P level. The main trends in reactiv- strain energy DEstrain(z) associated with
ethene, are known to have high reac- ity are rationalized using the activation deforming the reactants along the reac-
tion barriers despite the fact that their strain model of chemical reactivity. In tion coordinate z plus the interaction
cyclic transition states have a pro- this model, the shape of the reaction energy DEint(z) between these de-
nounced in-plane aromatic character, profile DE(z) and the height of the formed reactants: DE(z) = DEstrain(z) +
as indicated by NMR spectroscopic pa- DEint(z). We also use an alternative
rameters. To arrive at a way of under- fragmentation and a valence bond
Keywords: activation strain model ·
standing this somewhat paradoxical model for analyzing the character of
aromaticity · density functional
and incompletely understood phenom- the transition states.
calculations · double group transfer
enon of high-energy aromatic transi-
reactions · reactivity
tion states, we have explored six arche-

Introduction another in a concerted reaction pathway.[1] This definition


includes textbook reactions like the diimide reduction of
Double group transfer (DGT) reactions are a general class double or triple bonds,[2] the Meerwein–Ponndorf–Verley re-
of pericyclic reactions that occur through the simultaneous duction (MPV) of carbonyl groups,[3] and some type II dyo-
migration of two atoms or groups from one compound to tropic reactions that are characterized by the intramolecular
transfer of the two groups (generally hydrogen atoms).[4]
The archetypal DGT process is the thermally allowed con-
[a] Dr. I. Fernndez
Departamento de Qumica Orgnica, Facultad de Qumica certed and synchronous transfer of two hydrogen atoms
Universidad Complutense, 28040 Madrid (Spain) from ethane to ethylene. The process is suprafacial on both
Fax: (+ 34) 913944310 reaction sites and therefore these [s2s+s2s+p2s] transforma-
E-mail: israel@quim.ucm.es tions may be considered to be a thermally allowed pericyclic
[b] Prof. Dr. F. M. Bickelhaupt reaction according to the Woodward–Hoffmann rules.[5]
Department of Theoretical Chemistry
and Amsterdam Center for Multiscale Modeling
Strikingly, DGT reactions share a common feature, namely,
Scheikundig Laboratorium der Vrije Universiteit they proceed in a concerted and synchronous fashion
De Boelelaan 1083, 1081 HV Amsterdam (The Netherlands) through a transition state featuring a six-membered ring
Fax: (+ 31) 205987629 that is highly in-plane aromatic,[6] as indicated by the high
E-mail: fm.bickelhaupt@few.vu.nl
negative nuclear independent chemical shift (NICS) values[7]
[c] Prof. Dr. F. P. Cosso
computed at the (3,+1) ring critical point of the electron
Departamento de Qumica Orgnica I-Kimika Organikoa I Saila
Facultad de Qumica-Kimika Fakultatea density.[8]
Universidad del Pas Vasco-Euskal Herriko Unibertsitatea Despite the aromatic character of these transition states,
and Donostia International Physics Center (DIPC) DGT reactions are associated with relatively high barriers.
P.K. 1072, 20080 San Sebastin-Donostia (Spain) So far, the origins of the high activation barriers typically
Fax: (+ 34) 943015270
E-mail: fp.cossio@ehu.es
observed for DGT reactions in particular, and pericyclic pro-
Supporting information for this article, which includes Cartesian coor-
cesses in general, have not been fully understood. The rela-
dinates and total energies of all stationary points, is available on the tively recent introduction of the so-called activation strain
WWW under http://dx.doi.org/10.1002/chem.200902024. model has allowed us to gain more insight into the physical

13022  2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eur. J. 2009, 15, 13022 – 13032
FULL PAPER
factors that control how the activation barriers arise in dif- quency. Furthermore, transition states were verified to connect the sup-
ferent fundamental processes.[9] The activation strain model posed reagent and product minima by carrying out intrinsic reaction co-
ordinate (IRC) calculations.[19]
is a fragment approach to understanding chemical reactions
Activation strain analyses of reaction profiles: Insight into how the reac-
in which the height of reaction barriers is described and un- tion barriers arise was obtained through activation strain analyses of the
derstood in terms of the original reactants. Bickelhaupt and various DGT processes.[9, 11] The activation strain model is a systematic
co-workers have pioneered the systematic use of this meth- extension of the fragment approach from equilibrium structures to transi-
odology for elementary reactions in organic chemistry and tion states as well as nonstationary points (e.g., points along a reaction
coordinate). It is a fragment approach to understanding chemical reac-
catalysis.[9] They showed, for instance, how the activation
tions in which the height of reaction barriers is described and understood
barrier of SN2 reactions originates from steric congestion in terms of the original reactants. Thus, the potential-energy surface
around the carbon atom that is nucleophilically attacked, DE(z) is decomposed, along the reaction coordinate z, into the strain
and how electronic factors modulate the barrier height.[9, 10] DEstrain(z) associated with deforming the individual reactants plus the
Not much later, Houk et al.[11] were the first to successful- actual interaction DEint(z) between the deformed reactants [Eq. (1)]:
ly apply the activation strain model (or distortion/interaction
DEðzÞ ¼ DEstrain ðzÞ þ DEint ðzÞ ð1Þ
model) to pericyclic reactions, in particular to [3+2] cycload-
dition reactions. The latter, curiously, are concerted and Here, the reaction coordinate is defined as the projection of the IRC
occur through highly synchronous and in-plane aromatic (vide supra) on the HC distance between the hydrogen atom that is to
be transferred and the carbon atom to which it is transferred. This reac-
transition states, very similar to the situation for DGT reac-
tion coordinate z undergoes a well-defined change in the course of the
tions.[12] Thus, Houk et al.[11a] found that the barrier heights reaction from 1 to the equilibrium HC distance in the reduced prod-
for the cycloadditions of a given 1,3-dipole with ethylene uct.
and acetylene are nearly the same despite very different re- The strain, DEstrain(z), is determined by the rigidity of the reactants and
action thermodynamics and fragment molecular orbital the extent to which groups must reorganize in a particular reaction mech-
(FMO) energy gaps. One of the key findings was that the anism, whereas the interaction, DEint(z), between the reactants depends
on their electronic structure and how they are mutually oriented as they
activation strain, that is, the energy to distort the 1,3-dipole approach each other. It is the interplay between DEstrain(z) and DEint(z)
and dipolarophile to their corresponding transition-state ge- that determines if and at which point along z a barrier arises. The activa-
ometries, is the major factor controlling the reactivity differ- tion energy of a reaction DE ¼6 = DEACHTUNGRE(zTS) consists of the activation strain
ences of 1,3-dipoles. Furthermore, interaction energies be- DE6¼strain = DEstrainACHTUNGRE(zTS) plus the TS interaction DE6¼int = DEintACHTUNGRE(zTS) [Eq. (2);
see Figure 1]:
tween the 1,3-dipole and the dipolarophile differentiate re-
activity for a series of substituted alkenes when the distor- DE° ¼ DE° °
ð2Þ
strain þ DEint
tion energies are nearly constant.
Herein, we aim for a deeper understanding of the origins
of the barrier heights of the fundamental class of DGT reac-
tions. We report the results of a density functional theory
(DFT) study of six archetypal examples of such transforma-
tions based on the activation strain model. It will appear
that, in good agreement with Houks findings for [3+2] cy-
cloadditions, the activation strain associated with the struc-
tural rearrangement of the reactants is also the controlling
factor for the high energy of the cyclic DGT reaction barri-
ers.

Theoretical Methods

Computational details: All calculations were performed with the Amster-


dam density functional (ADF) program developed by Baerends and
others.[13, 14] The molecular orbitals (MOs) were expanded in a large un- Figure 1. Illustration of the activation strain model for a DGT reaction.
contracted set of Slater-type orbitals (STOs) containing diffuse functions
(TZ2P). This basis is of triple-z quality and has been augmented by two
sets of polarization functions, that is, p and d functions for the hydrogen
atom and d and f functions for the other atoms. An auxiliary set of s, p, Molecular orbital theory and energy decomposition analysis: The interac-
d, f, and g STOs was used to fit the molecular density and to represent tion DEint(z) between the strained reactants was further analyzed in the
the Coulomb and exchange potentials accurately in each SCF cycle. Rela- conceptual framework provided by the Kohn–Sham molecular orbital
tivistic effects were accounted for by using the zeroth-order regular ap- (KSMO) model.[20] To this end, it was further decomposed into three
proximation (ZORA).[15] physically meaningful terms [Eq. (3)]:
Equilibrium and transition-state geometries were fully optimized at the
OLYP[16] density functional, which involves Handys optimized exchange DEint ðzÞ ¼ DV elstat þDEPauli þDEorb ð3Þ
(OPTX). All stationary points were confirmed by vibrational analysis:[17]
for equilibrium structures, all normal modes have real frequencies, The term DVelstat corresponds to the classical electrostatic interaction be-
whereas transition states[18] have one normal mode with an imaginary fre- tween the unperturbed charge distributions of the deformed reactants

Chem. Eur. J. 2009, 15, 13022 – 13032  2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemeurj.org 13023
I. Fernndez, F. M. Bickelhaupt, and F. P. Cosso

and is usually attractive. The Pauli repulsion DEPauli comprises the desta- Table 1. Selected reaction parameters (E in kcal mol1, D in ) of DGT
bilizing interactions between occupied orbitals and is responsible for any reactions.[a]
steric repulsion (see ref. [20] for an exhaustive discussion). The orbital in-
Reaction DE° DEr Dshort
teraction DEorb accounts for charge transfer (interaction between occu-
pied orbitals on one moiety with unoccupied orbitals on the other, in-
cluding HOMO–LUMO interactions) and polarization (empty–occupied 1 45.4 0.0 2.732
orbital mixing on one fragment due to the presence of another frag-
ment).
Since the Kohn–Sham MO method of density functional theory (DFT) in 2 41.8 12.1 2.622
principle yields exact energies and, in practice, with the available density
functionals for exchange and correlation, is rather accurate, we have the
special situation that a seemingly one-particle model (an MO method) in
3 44.5 7.0 2.656
principle fully accounts for the bonding energy.[20] Through these energy
decomposition analyses (EDA), it has previously been possible, for exam-
ple, to compare the strength of conjugation, hyperconjugation, and aro-
maticity in p-conjugated systems[21, 22] and to analyze the nature of a 4 43.7 12.1 2.567
chemical bond in terms of electrostatic attraction versus s-, p-, and d-or-
bital (covalent) bonding.[23, 24]
Alternative analyses of transition states: We have also carried out an al- 5 49.0 19.6 2.655
ternative analysis of the transition states. In this approach, the model
system is partitioned, not into the two original reactants (e.g.,
C2H6+C2H4), but instead into the organic scaffold between which double 6 24.5 0.0 2.382
hydrogen transfer takes place plus the two hydrogen atoms (i.e., the two
fragments [C2H4–C2H4]··+ACHTUNGRE[H–H]··, each of which has a triplet-valence
electron configuration). To model the progress of the interactions along [a] Computed at the OLYP/TZ2P level. Dshort = the sum of the shortest
the IRC from TS to reactants in which the biradical character localizes distance of the breaking XH and forming HY bonds in the TS []
towards one of the two C2H4 moieties, a similar analysis has been carried (i.e., upper X-H-Y moiety in Figure 2).
out for the interaction between the organic scaffold of the ethane mole-
cule that is forming plus the same two hydrogen atoms (i.e., the two frag-
ments [C2H4]··+ACHTUNGRE[H–H]··, each of which again has a triplet-valence electron
configuration). The details of this approach are provided later on in the
discussion. The MO and EDA analyses in this alternative fragmentation
provide insight from a different perspective that is complementary to and
augments the activation strain analyses.
NICS values and valence bond analyses: The aromatic character of the
transition states has been confirmed by the computation of the NICS[7]
values computed at the (3,+1) ring critical point of the electron density[8]
(see below). These calculations have been carried out with the Gaussi-
an 03 suite of programs[25] using the gauge invariant atomic orbital
(GIAO) method,[26] at the OLYP level using the triple-zeta plus polariza-
tion basis sets (def2-TZVPP)[27] with the optimized OLYP/TZ2P geome-
tries. This scheme is denoted GIAO-OLYP/def2-TZVPP//OLYP/TZ2P.
For the computational details related to the valence bond analyses, see
the Results and Discussion section.

Figure 2. Geometries [] of the transition states of our DGT reactions,


computed at the OLYP/TZ2P level. See Table 1 for the position of heter-
Results and Discussion oatoms.

Structures and reaction profiles: In Table 1, we provide an ers of 40–50 kcal mol1. Note that the Meerwein–Ponndorf–
overview of our DGT model reactions together with the re- Verley reaction (reaction 6) has a much lower barrier of
action barriers that we have computed at the OLYP/TZ2P 24.5 kcal mol1. Whereas correlations should of course
level. Structures of the corresponding transition states (TS) always be treated with precaution, one can observe that the
are depicted in Figure 2. We consider the parent reactions of reaction barrier DE° increases if the shortest of the two X-
ethane with ethene and ethyne (reaction numbers 1 and 2), H-Y contacts (i.e., the upper one in the TS structures of
and the corresponding processes involving heteroatoms, that Figure 2) becomes longer, as measured by the Dshort = rACHTUNGRE(X
is, the reduction by ethane of imine and formaldehyde (reac- H)+rACHTUNGRE(HY), which is also provided in Table 1. The correla-
tions 3 and 4) as well as nitrile (reactions 5). In addition, we tion that we find between Dshort and the barrier height
have examined the uncatalyzed Meerwein–Ponndorf–Verley matches nicely with kinetic and structural experiments that
reduction of formaldehyde by methanol (reaction 6).[5] show that changes in Dshort of only 0.1–0.17  translate into
There appears to be a correlation between the barrier a rate spread of 104 s1.[28]
height and the geometries of the TS, in particular, the Figure 2 shows the corresponding transition states of the
breaking XH and forming HY distances. As can be seen above-mentioned transformations. In all cases, the transfor-
in Table 1, all reactions proceed through high energy barri- mations occur through highly symmetric six-membered

13024 www.chemeurj.org  2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eur. J. 2009, 15, 13022 – 13032
Double Group Transfer Reactions
FULL PAPER
cyclic transition states in which the corresponding CC, C energy LUMO of the substrate. This means that for SN2 and
N, and CO bonds present bond lengths that are intermedi- E2 reactions the net interaction DEint is stabilizing along the
ate between double and single bonds. The observed bond- entire reaction coordinate and the destabilization is con-
length equalization and planarity of the transition states tained entirely in the strain DEstrain term.
agree nicely with the highly in-plane aromatic character of If we now proceed further along the DGT reaction coor-
these species, which has been confirmed by the calculation dinate, the trend in DEint inverts at a certain point, after
of the NICS(0), computed at the (3, + 1) ring critical point which this term becomes more and more stabilized as one
of the electron density as defined by Bader,[8] and NICS(1)zz further approaches the transition state. Inspection of
values[29] (the out-of-plane component of the NICS comput- Figure 3 shows that this stabilization in the DEint curve
ed 1  above the ring critical point, Table 2). occurs shortly after the onset of the strain curve. A decom-
position of DEint into the different terms of the interaction
energy furthermore shows that the dominant term causing
Table 2. Computed NICS values [ppm] of transition states TS1–TS6.
this inversion is the orbital interaction energy DEorb (see
TS1 TS2 TS3 TS4 TS5 TS6
Figure 4). Note that the electrostatic attraction DVelstat is cer-
[a]
NICS(0) 27.04 25.76 26.38 27.16 25.41 24.48
tainly far from being negligible, but it is not the dominant
NICS(1)[b] 12.77 12.23 10.97 11.50 12.25 9.27
NICS(1)zz[c] 35.74 34.77 31.90 31.41 34.63 25.46 bonding term. The reason for the onset of sizable orbital in-
teractions at shorter HC distances is that the original CH
[a] NICS(0) computed at GIAO-OLYP/def2-TZVPP//OLYP/TZ2P at the
(3,+1) ring critical point. [b] NICS(1) computed 1  above the (3,+1) bonds are beginning to elongate. This causes the associated
ring critical point. [c] NICS(1)zz denotes the zz tensor component of s*CH orbitals to drop in energy and become more localized
NICS(1). on the transferring hydrogen atoms. As a result, the
HOMO–LUMO interactions between the p-electron system
of the hydrogen acceptor reactants and the s*CH orbitals of
the hydrogen donor become stronger and more stabilizing
Activation strain analyses: Next we address the main pur- (see Figure 4).
pose of our study: to understand the factors that are respon- The reason that the overall energy DE still goes up until
sible for the high activation barriers of DGT reactions, using reaching the TS is, of course, also the increase in the strain
the activation strain model. Figure 3 shows the plots of the energy that goes with the aforementioned elongation of the
computed potential-energy surface along the IRC trajecto- CH bonds in the hydrogen-donor reactant. We also find a
ries, projected onto the distance rHC (see the Theoretical contribution to the strain stemming from the conformational
Methods section) for reactions 1 to 6 in Table 1, together change in the hydrogen donor (ethane or methanol) from
with the change of the energy contributors to DE(z), staggered to eclipsed. But the breaking of the two CH
namely, the strain DEstrain(z) and the instantaneous interac- bonds quickly turns into the dominant contribution of the
tion DEint(z) between the deformed reactants. All our model strain term as the TS is approached.
systems present quite similar energy trends. Thus, we can Interestingly, the uncatalyzed Meerwein–Ponndorf–Verley
see that at the early stages of the process the reaction pro- reaction (reaction 6 in Table 1; Figure 3f) presents the
file DE monotonically becomes more and more destabilized lowest computed energy barrier. This may be explained by
as the reactants approach each other. Interestingly, a sharp the following reasons: 1) the total bond-dissociation energy
increase of DE occurs in the proximity of the transition- required to liberate the two migrating hydrogen atoms,
state region (i.e., at HC distances in the range from 2.0 to thereby breaking the CH and OH bonds, is approximate-
1.6 ), thereby leading to the observed high reaction barri- ly 7 kcal mol1 lower in CH3OH than in CH3CH3 (see
ers. Scheme 1); 2) we found a stabilizing effect when using
At long HC distances, the interaction energy between CH3OH as a “hydrogen donor,” which is not able to occur
the deformed reactants (DEint) becomes destabilizing and using CH3CH3, in the formation of an OH···O hydrogen
causes the net energy DE to go up too, as can be seen in bond between both reactants, which facilitates the hydrogen
Figure 3. The initial increase of DEint can be traced to steric migration if both reactants are in close proximity. The latter
(Pauli) repulsion between the reactants in the early stages of stabilizing contribution is nicely reflected in the higher com-
the reaction. Thus, before anything else happens, the reac- puted DEint value (i.e., DEint  34 kcal mol1 for reaction 6
tants approach and overlap occurs between closed shells, no- and DEint  22 kcal mol1 for the parent reaction 1,
tably between CH bonds of the hydrogen donor and the p Table 1) and in a not-so-high strain energy (i.e., DEstrain
system of the hydrogen acceptor, which causes Pauli repul-  59 kcal mol1 for reaction 6, and DEstrain  65 kcal mol1 for
sion. This behavior of DEint is interesting as it differs from the parent reaction 1). Both effects contribute to the ob-
reactions such as SN2 substitution and E2 elimination.[9a,h] In served decrease of the activation barrier in this system.
the latter, there is of course also Pauli repulsion between Based on the above discussion, it can be concluded that,
closed shells of the reactants as they approach each other. similarly to the pericyclic processes studied so far, the acti-
But in addition, there is in these reactions, from the begin- vation strain associated with the structural rearrangement of
ning, a potent donor–acceptor interaction between the the reactants is also the controlling factor for the high
HOMO of the nucleophile/base and the relatively low- energy of the DGT reaction barriers. This makes sense in

Chem. Eur. J. 205., 11, 13022 – 13032  205. Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemeurj.org 13025
I. Fernndez, F. M. Bickelhaupt, and F. P. Cosso

Figure 3. Activation strain analysis of the DGT reactions (in kcal mol1, at the OLYP/TZ2P level) along the reaction coordinate projected onto the HC
bond length. a) to f) correspond to reactions 1 to 6, respectively (see Table 1).

view of the geometrically distorted nature of DGT transi- reaction coordinate in the parent ethane plus ethene reac-
tion states in which two CH bonds and one p bond must tion was also computed. As expected, the NICS values
first be partially broken before an equivalent number of become more and more negative (that is, the system be-
new bonds are formed instead. comes more and more aromatic) and reaches the maximum
value for the transition state. Note that the NICS(0) curve is
Aromatic character according to NICS(0): The evolution of a mirror image of the total energy DE (Figure 5a), which
the aromaticity (measured by the NICS(0) values) along the might be an indication of a correlation of both parameters.

13026 www.chemeurj.org  2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eur. J. 2009, 15, 13022 – 13032
Double Group Transfer Reactions
FULL PAPER

Figure 4. Decomposition of the interaction energy between the reactants in model reactions 1–6 (a–f; see Table 1) along the reaction coordinate project-
ed onto the HC bond length, computed at the OLYP/TZ2P level.

In fact, when one plots DE values versus the NICS(0) increases with the concomitant gain in aromaticity. This fact
values, a nice second-order polynomial curve is obtained seems contradictory if we consider that a gain in aromaticity
(correlation coefficient of 0.998), thus confirming the close is usually translated into a gain in stability. Therefore, this
relationship between both parameters (Figure 5b). It is apparent contradiction provides further support to the
clearly shown in both parts of Figure 5 that DE continuously above-mentioned conclusion that the reaction is controlled

Chem. Eur. J. 2009, 15, 13022 – 13032  2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemeurj.org 13027
I. Fernndez, F. M. Bickelhaupt, and F. P. Cosso

energy barrier. Figure 6 shows the valence bond state corre-


lation diagram (VBSCD) for the degenerate double hydro-
gen transfer from eclipsed C2v-symmetric ethane to ethylene.

Scheme 1. Computed (OLYP/TZ2P level) bond-dissociation energies (in


kcal mol1); (d) and (t) indicate doublet and triplet state, respectively.

Figure 6. Valence-bond-state correlation diagram for reaction 1 (see


Table 1). The inset includes the B3LYP/6-31G* spin contributions pro-
jected onto the electron density of eclipsed ethane (C2v symmetry) and
ethylene (D2h symmetry), both in the triplet state and at the ground-state
geometry. Colors range from 0.002 a.u. (blue) to 0.000 a.u. (red).

From the VB standpoint,[30] the mechanism of this reaction


can be envisaged as the result of an avoided crossing be-
tween the adiabatic state lines YR !Y*P and Y*R !YP. The
magnitude of the activation energy DEa is given by Equa-
tion (4):

DEa ¼ DEc B ð4Þ

in which B is the resonance energy of the TS, and DEc is the


energy barrier corresponding to the crossing point. This
latter magnitude can be expressed as shown in Equation (5):

DEc ¼ fG ð5Þ

in which G is the promotion energy on going from YR to


YR*. For this kind of pericyclic reaction, G corresponds to
Figure 5. a) Evolution of DE and NICS(0) (computed at the GIAO-
OLYP/def2-TZVPP//OLYP/TZ2P level) along the reaction coordinate. the vertical singlet–triplet gap required to change the elec-
b) Plot of the correlation between DE and NICS(0). tronic state of the reactants to adapt it to that of the prod-
ucts while keeping the singlet-state geometry unchanged.
Therefore, G is calculated as shown in Equation (6):
by the strain in spite of the gain in stability by aromatic de-
localization. G ¼ EST ðC2v C2 H6 ÞþEST ðD2h C2 H4 Þ ð6Þ

Valence bond perspective: Since the dyotropic ring current The two terms of the right side of this expression corre-
present in aromatic structures is closely related to cyclic spond to the vertical singlet–triplet (ST) gap of C2v-symmet-
electron stabilization, we decided to analyze the magnitude ric ethane and D2h-symmetric ethylene, respectively. At the
of the resonance energy of the TS corresponding to reac- B3LYP/6-31G* level, it is found that ESTACHTUNGRE(D2hC2H4) =
tion 1 (see Table 1) to gain a better understanding of the re- 104.1 kcal mol1, which is close to the experimental value of
lationship between its high aromatic character and the 100.5 kcal mol1.[31] Similarly, at the same level we have

13028 www.chemeurj.org  2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eur. J. 2009, 15, 13022 – 13032
Double Group Transfer Reactions
FULL PAPER
found that ESTACHTUNGRE(C2vC2H6) = 241.7 kcal mol1. Therefore, G = equal. The corresponding interactions are also collected in
345.8 kcal mol1. This is a very high value that is caused by Table 3, together with an energy decomposition analysis.
the strongly distorted geometry of the eclipsed geometry of
ethane in the triplet state. Under these conditions, most of
Table 3. Analysis of the interaction between two equivalent HCH2 = CH2· frag-
the spin density of the Y*R state is concentrated in the mi- ments in the circular TS of DGT reaction 1 and the exact same fragments com-
grating hydrogen atoms of ethane and the carbon atoms of bined in a linear fashion.[a]
ethylene (Figure 6), a spin distribution required for pairing Circular (TS of DGT reaction 1) Linear system
the electrons across the novel s bond. The large value of G
can be compared with that computed for the Diels–Alder
reaction between 1,3-butadiene and ethylene. In this case, it
is found that G = 173 kcal mol1 because of the much lower
EST value of 1,3-butadiene.[32] The parameter f reflects the
fraction of G that is responsible for the height of the cross-
ing energy DEc. If we assume a value of f  0.3 for thermally DEint 60.4 31.1
allowed pericyclic reactions,[32] we obtain DEc  103.7 kcal DEPauli 167.6 103.2
mol1 and therefore B  58.6 kcal mol1. This resonance DVelstat[b] 70.7 (31 %) 40.6 (30 %)
energy of the transition state is approximately 20 kcal mol1 DEorb[b] 157.4 (69 %) 93.7 (70 %)
DEs[c] 155.7 (99 %) 92.9 (99 %)
higher than that calculated for the parent Diels–Alder reac-
DEp[c] 1.7 (1 %) 0.8 (1 %)
tion. Therefore, we can conclude that the relatively high ac-
[a] In kcal mol1, computed at the OLYP/TZ2P level. [b] The percentage values
tivation barrier for this reaction is compatible with a highly in parentheses give the contribution to the total attractive interactions
aromatic TS. This apparent paradox is caused by the huge DEelstat+DEorb. [c] The percentage values in parentheses give the contribution to
vertical energy gap required to adapt the electronic state of the total orbital interactions DEorb.
the reactants to that of the products while keeping the geo-
metric features of the reactants in the ground electronic
state. Not unexpectedly, Table 3 shows that interrupting the aro-
matic (i.e., circular) conjugation of the six s electrons leads
Aromaticity and stability of transition states: The above re- to a significant weakening of the interaction, which becomes
sults show that a high energy of the DGT transition state half as strong as in the circular TS. Note that this indeed
can go with a highly aromatic character. It is interesting in comes mainly from a weakening of the orbital interactions,
this context to note that indeed cyclic delocalization causes which for 99 % of those interactions stem from the s-elec-
a stabilization with respect to the corresponding noncyclic tron system in which the aromatic conjugation takes place.
structure. This is well known, for example, for the p-electron Note, too, that the gain in DEorb from the linear to the cyclic
systems of benzene (cyclic delocalization: more stable) as structure (i.e., 63 kcal mol1) is approximately equal to the
compared with 1,3,5-hexatriene (linear, no cyclic delocaliza- resonance B term in VB theory (58.6 kcal mol1, vide
tion: less stable). This is associated with the presence of a supra). In an alternative, more qualitative manner, this may
sixth bond (or overlap or interaction matrix element) in the again be represented by two VB resonance structures that
circular system and the concomitant lowering of the kinetic are in stabilizing, aromatic resonance in the circular DGT
energy in the lowest-energy p level (cf. particle on a ring).[33] transition state (see I, and the corresponding resonance
In terms of simple VB resonance structures, one can illus- structures II and III), whereas such low-energy (nonionic,
trate this with the stabilizing resonance between the two closed-shell) resonance structures are not possible for a
Kekul structures for benzene and, in the case of 1,3,5-hexa- linear arrangement (see IV, which has only resonance struc-
triene, the absence of low-energy resonance structures and ture V, Scheme 2).
thus no stabilizing resonance. There is thus a substantial stabilization due to the aromat-
The same situation occurs for the six s electrons involved ic sextet of electrons in the circular DGT transition state.
in the bond breaking and making in the ring structure of Nevertheless, the DGT transition state has a strong tenden-
DGT transition states. We illustrate that quantitatively by cy to localize two CH bonds and a CC p bond. One
computing the stabilization that occurs by combining the reason is the aforementioned strain and repulsion associated
two equivalent HCH2=CH2C radicals, of which the DGT with pushing the reactants together (e.g., ethane and ethene
transition state of reaction 1 (Table 1) can be constructed in in reaction 1 in Table 1). There is a second aspect that needs
two different ways as shown in Table 2: 1) circularly, thus to be taken into consideration, namely, the question of
yielding the DGT transition state; and 2) in a linear fashion whether the aromatic electrons have an intrinsic propensity
by inflecting one of the two HCH2=CH2C fragments in the to keep the molecule in its regular or circular, “aromatic”
indicated mirror line. The latter system is not a stationary geometry or not. It has been shown in several complementa-
point. It serves merely to estimate the additional aromatic ry studies that the six p electrons in benzene actually favor
stabilization that goes with circular conjugation as compared a localized, cyclohexatriene-type structure.[22a,c–e, 34] Recently,
with a system in which this circular conjugation is interrupt- we have explained this behavior in terms of a quantitative
ed while as many as possible geometry parameters are kept molecular orbital (MO) model.[22c,d] The reason is that three

Chem. Eur. J. 2009, 15, 13022 – 13032  2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemeurj.org 13029
I. Fernndez, F. M. Bickelhaupt, and F. P. Cosso

(i.e., [C2H4–C2H4]··+ACHTUNGRE[H–H]··, both fragments in their triplet


valence states) instead of the interaction between the two
original reactants (i.e., C2H6+C2H4). Figure 7 shows the
schematic orbital-interaction diagram for the DGT transi-
tion state that emerges from our quantitative Kohn–Sham
MO analyses at OLYP/TZ2P. If we move along the IRC to-
wards the reactants, the two C2H4 moieties in the organic
scaffold quickly separate and localize the triplet character
on the p and p* orbitals of one of the two C2H4 fragments.
We have modeled the interactions and overlaps through
analysis of the interacting [C2H4]·· and [H–H]·· fragments
that approach each other along the IRC of the overall reac-
tion system. The results are shown in Figure 8. Clearly, the
SOMO–SOMO overlap and orbital interactions favor locali-
zation of the CH bonds. One can therefore conclude that
the DGT transition state has a strong aromatic character in
the sense that it receives substantial stabilization from circu-
lar conjugation as compared with the less favorable linear
Scheme 2. Resonance and resonance structures in the transition state of
DGT reaction 1 (see Table 1) (I–III) and a linear analogue (IV, V).

localized (i.e., short C=C double bonds) yield effectively


more p overlap than six delocalized (i.e., somewhat elongat-
ed) ones.[22, 34]
A similar behavior can be observed in the present case of
the DGT transition state. Here, the question is whether the
two hydrogen atoms that are transferred have the propensi-
ty to keep the “aromatic geometry” of four delocalized C–
H–C bridges, or if localization towards two regular, short
CH bonds in ethane (and rupture from the remaining
ethene) yields better overlap and more stabilization. To this
end, we have analyzed for DGT reaction 1 (Table 1) the in-
teraction of the organic scaffold between which double hy-
drogen transfer takes place plus the two hydrogen atoms

Figure 8. a) Energy decomposition analysis and b) SOMO–SOMO bond


Figure 7. Schematic orbital interaction diagram between [C2H4–C2H4]·· overlaps for the CH interaction between [C2H4]·· and [H–H]·· fragments
and [H–H]·· in the TS of the DGT reaction 1 (see Table 1). along the IRC of DGT reaction 1 (see Table 1).

13030 www.chemeurj.org  2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eur. J. 2009, 15, 13022 – 13032
Double Group Transfer Reactions
FULL PAPER
situation. But, at the same time, repulsion between the reac- c) M. T. Reetz, Adv. Organomet. Chem. 1977, 16, 33; d) K. N. Houk,
tants causes an overlap behavior in which the six aromatic s Y. Li, M .A. McAllister, G. A. ODoherty, L. A. Paquette, W. Sie-
brand, Z. K. Smedarchina, J. Am. Chem. Soc. 1994, 116, 10895; e) G.
electrons achieve best overlap and thus an even more stable Frenking, F. P. Cosso, M. A. Sierra, I. Fernndez, Eur. J. Org. Chem.
geometrical arrangement if the CH bonds are localized 2007, 5410.
under formation of ethane and ethene. The aromatic ring [5] R. B. Woodward, R. Hoffmann, The Conservation of Orbital Sym-
currents, indicated by the NICS values (vide supra), simply metry, VCH, Weinheim, 1970, p. 141.
reflect the small HOMO–LUMO gap in the DGT transition [6] I. Fernndez, M. A. Sierra, F. P. Cosso, J. Org. Chem. 2007, 72,
1488.
state that goes with the distortions (i.e., the occurrence of [7] Z. Chen, C. S. Wannere, C. Corminboeuf, R. Puchta, P. von R.
partial instead of full s CH and p CC bonds) relative to Schleyer, Chem. Rev. 2005, 105, 3842.
the stable reactants. [8] R. F. W. Bader, Atoms in Molecules -A Quantum Theory, Clarendon
Press, Oxford, 1990.
[9] a) F. M. Bickelhaupt, J. Comput. Chem. 1999, 20, 114; b) A. Diefen-
bach, F. M. Bickelhaupt, J. Chem. Phys. 2001, 115, 4030; c) A. Die-
Conclusion fenbach, F. M. Bickelhaupt, J. Phys. Chem. A 2004, 108, 8460; d) A.
Diefenbach, G. T. de Jong, F. M. Bickelhaupt, J. Chem. Theory
From the computational study reported in this paper the fol- Comput. 2005, 1, 286; e) J. N. P. van Stralen, F. M. Bickelhaupt, Or-
lowing conclusions can be drawn. 1) Similarly to other re- ganometallics 2006, 25, 4260; f) G. T. de Jong, F. M. Bickelhaupt,
ChemPhysChem 2007, 8, 1170; g) G. T. de Jong, F. M. Bickelhaupt,
ported pericyclic reactions, the activation strain associated
J. Chem. Theory Comput. 2007, 3, 514; h) A. P. Bento, F. M. Bickel-
with the structural rearrangement of the reactants is the haupt, J. Org. Chem. 2008, 73, 7290. An early study in which the re-
controlling factor for the high activation barrier of double actant strain or distortion energy has been separated from the inter-
group transfer reactions. 2) The gain in stability by reso- action energy for understanding SN2 reactions is: i) D. J. Mitchell,
nance and in-plane aromaticity of the transition structures H. B. Schlegel, S. S. Shaik, S. Wolfe, Can. J. Chem. 1985, 63, 1642.
[10] I. Fernndez, G. Frenking, E. Uggerud, Chem. Eur. J. 2009, 15, 2166.
cannot compensate for the strong destabilizing effect of the [11] a) D. H. Ess, K. N. Houk, J. Am. Chem. Soc. 2008, 130, 10187. For a
strain; 3) The in-plane aromatic character of the six s elec- different activation strain or distortion/interaction study by the
trons involved in the DGT bond rearrangement is confirmed Houk group, see, for example: b) C. Y. Legault, Y. Garcia, C. A.
by the strong stabilization we compute for circular conjuga- Merlic, K. N. Houk, J. Am. Chem. Soc. 2007, 129, 12664.
[12] a) I. Morao, B. Lecea, F. P. Cosso, J. Org. Chem. 1997, 62, 7033;
tion as compared with a fictitious linear system. 4) Never-
b) F. P. Cosso, I. Morao, H. Jiao, P. von R. Schleyer, J. Am. Chem.
theless, the six in-plane aromatic s electrons have a propen- Soc. 1999, 121, 6737; c) I. Fernndez, M. A. Sierra, F. P. Cosso, J.
sity to localize the C–H–C bridges towards two full CH Org. Chem. 2006, 71, 6178.
bonds (under formation of either reactants or products) be- [13] a) Computer code ADF 2007.01, E. J. Baerends, J. Autschbach, A.
cause in this way even more favorable bond overlap is Brces, J. A. Berger, F. M. Bickelhaupt, C. Bo, P. L. de Boeij, P. M.
Boerrigter, L. Cavallo, D. P. Chong, L. Deng, R. M. Dickson, D. E.
ACHTUNGREachieved (this is reminiscent of the aromatic p electrons of
Ellis, M. van Faassen, L. Fan, T. H. Fischer, C. Fonseca Guerra,
benzene, which have a propensity to localize the double S. J. A. van Gisbergen, J. A. Groeneveld, O. V. Gritsenko, M. Gr n-
bonds but are overruled by the s-electron system).[22, 34] ing, F. E. Harris, P. van den Hoek, C. R. Jacob, H. Jacobsen, L.
5) The activation strain model is confirmed to be useful and Jensen, E. S. Kadantsev, G. van Kessel, R. Klooster, F. Kootstra, E.
fully applicable for understanding the intricacies also of per- van Lenthe, D. A. McCormack, A. Michalak, J. Neugebauer, V. P.
Nicu, V. P. Osinga, S. Patchkovskii, P. H. T. Philipsen, D. Post, C. C.
icyclic reactions. Pye, W. Ravenek, P. Romaniello, P. Ros, P. R. T. Schipper, G.
Schreckenbach, J. Snijders, M. Sol
, M. Swart, D. Swerhone, G. te
Velde, P. Vernooijs, L. Versluis, L. Visscher, O. Visser, F. Wang,
T. A. Wesolowski, E. M. van Wezenbeek, G. Wiesenekker, S. K.
Acknowledgements Wolff, T. K. Woo, A. L. Yakovlev, T. Ziegler, Scientific Computing
& Modeling N.V., Amsterdam, 2007; b) C. Fonseca Guerra, J. G.
We thank the Spanish Ministerio de Ciencia e Innovacin (MCINN; Snijders, G. te Velde, E. J. Baerends, Theor. Chem. Acc. 1998, 99,
CTQ2007-67528), Consolider-Ingenio 2010 (CSD2007-0006), and the 391.
Netherlands Organization for Scientific Research (NWO-CW) for finan- [14] a) E. J. Baerends, D. E. Ellis, P. Ros, Chem. Phys. 1973, 2, 41; b) G.
cial support. I.F. is a Ramn y Cajal fellow. te Velde, F. M. Bickelhaupt, E. J. Baerends, C. Fonseca Guerra,
S. J. A. van Gisbergen, J. G. Snijders, T. Ziegler, J. Comput. Chem.
2001, 22, 931.
[1] S. Sankararaman, Pericyclic Reactions—A Textbook: Reactions, Ap- [15] E. van Lenthe, E. J. Baerends, J. G. Snijders, J. Chem. Phys. 1994,
plications and Theory, Wiley, Weinheim, 2005; pp. 326 – 329, and ref- 101, 9783.
erences therein. [16] a) N. C. Handy, A. J. Cohen, Mol. Phys. 2001, 99, 403; b) C. Lee, W.
[2] a) S. H nig, H. R. M ller, W. Thier, Angew. Chem. 1965, 77, 368; Yang, R. G. Parr, Phys. Rev. B 1988, 37, 785.
Angew. Chem. Int. Ed. Engl. 1965, 4, 271; b) M. Franck-Neumann, [17] L. Fan, L. Versluis, T. Ziegler, E. J. Baerends, W. Ravenek, Int. J.
C. Dietrich-Buchecker, Tetrahedron Lett. 1980, 21, 671; c) E. W. Gar- Quantum Chem. Quantum Chem. Symp. 1988, S22, 173.
bisch, Jr., S. M. Schildcrout, D. B. Patterson, C. M. Sprecher, J. Am. [18] L. Fan, T. Ziegler, J. Chem. Phys. 1990, 92, 3645.
Chem. Soc. 1965, 87, 2932. [19] K. Fukui, Acc. Chem. Res. 1981, 14, 363.
[3] This latter process was recently proposed to evolve through a cyclic [20] a) F. M. Bickelhaupt, E. J. Baerends in Reviews in Computational
six-membered transition state at elevated temperatures. See: L. So- Chemistry, Vol. 15 (Eds.: K. B. Lipkowitz, D. B. Boyd), Wiley-VCH,
minsky, E. Rozental, H. Gottlieb, A. Gedanken, S. Hoz, J. Org. Weinheim, 2000, pp. 1 – 86; b) M. Lein, G. Frenking in Theory and
Chem. 2004, 69, 1492. Applications of Computational Chemistry: The First 40 Years (Eds.:
[4] a) M. T. Reetz, Angew. Chem. 1972, 84, 163; Angew. Chem. Int. Ed. C. E. Dykstra, G. Frenking, K. S. Kim, G. E. Scuseria), Elsevier,
Engl. 1972, 11, 130; b) M. T. Reetz, Tetrahedron 1973, 29, 2189; Amsterdam, 2005, p. 291.

Chem. Eur. J. 2009, 15, 13022 – 13032  2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemeurj.org 13031
I. Fernndez, F. M. Bickelhaupt, and F. P. Cosso

[21] Conjugation and Hyperconjugation: a) D. Cappel, S. T llmann, A. Keith, M. A. Al-Laham, C. Y. Peng, A. Nanayakkara, M. Challa-
Krapp, G. Frenking, Angew. Chem. 2005, 117, 3683; Angew. Chem. combe, P. M. W. Gill, B. Johnson, W. Chen, M. W. Wong, C. Gonza-
Int. Ed. 2005, 44, 3617; b) I. Fernndez, G. Frenking, Chem. Eur. J. lez, and J. A. Pople, Gaussian, Inc., Wallingford CT, 2004.
2006, 12, 3617; c) I. Fernndez, G. Frenking, J. Org. Chem. 2006, 71, [26] K. Wolinski, J. F. Hilton, P. Pulay, J. Am. Chem. Soc. 1990, 112,
2251; d) I. Fernndez, G. Frenking, Chem. Commun. 2006, 5030; 8251.
e) I. Fernndez, G. Frenking, J. Phys. Chem. A 2007, 111, 8028; f) I. [27] F. Weigend, R. Alhrichs, Phys. Chem. Chem. Phys. 2005, 7, 3297.
Fernndez, G. Frenking, J. Org. Chem. 2007, 72, 7367. [28] a) K. Mackenzie, J. A. K. Howard, S. Mason, E. C. Gravett, K. B.
[22] Aromaticity: a) I. Fernndez, G. Frenking, Faraday Discuss. 2007, Astin, S. X. Liu, A. S. Batsanov, D. Vlaovic, J. P. Maher, J. Chem.
135, 403; b) I. Fernndez, G. Frenking, Chem. Eur. J. 2007, 13, 5873; Soc. Perkin Trans. 2 1993, 1211; b) K. N. Houk, J. Y. Li, M. A.
c) S. C. A. H. Pierrefixe, F. M. Bickelhaupt, Chem. Eur. J. 2007, 13, McAllister, G. A. ODoherty, L. A. Paquette, W. Siebrand, Z. K.
6321; d) S. C. A. H. Pierrefixe, F. M. Bickelhaupt, J. Phys. Chem. A Smedarchinal, J. Am. Chem. Soc. 1994, 116, 10895.
2008, 112, 12816; e) S. C. A. H. Pierrefixe, F. M. Bickelhaupt, Aust. [29] H. Fallah-Bagher-Shaidaei, C. S. Wannere, C. Corminboeuf, R.
J. Chem. 2008, 61, 209. Puchta, P. von R. Schleyer, Org. Lett. 2006, 8, 863.
[23] G. Frenking, K. Wichmann, N. Frçhlich, C. Loschen, M. Lein, J. [30] S. S. Shaik, P. C. Hiberty, A Chemists Guide to Valence Bond
Frunzke, V. M. Rayn, Coord. Chem. Rev. 2003, 238–239, 55. Theory; Wiley, New York, 2008, Chapter 6 and references therein.
[24] A. Krapp, F. M. Bickelhaupt, G. Frenking, Chem. Eur. J. 2006, 12, [31] a) A. Vzquez-Mayagoitia, R. Vargas, J. A. Nichols, P. Fuentealba, J.
9196. Garza, Chem. Phys. Lett. 2006, 419, 207; b) E. H. van Veen, Chem.
[25] These calculations were carried out using the Gaussian 03, Revision Phys. Lett. 1976, 41, 540.
E.01, M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, [32] A. Ioffe, S. Shaik, J. Chem. Soc. Perkin Trans. 2 1992, 2101.
M. A. Robb, J. R. Cheeseman, J. A. MontgomACHTUNGREery, Jr., T. Vreven, [33] P. W. Atkins, R. S. Friedman, Molecular Quantum Mechanics, 3rd
K. N. Kudin, J. C. Burant, J. M. Millam, S. S. Iyengar, J. Tomasi, V. ed., Oxford University Press, Oxford, 1997, Chapter 3.
Barone, B. Mennucci, M. Cossi, G. Scalmani, N. Rega, G. A. Peters- [34] a) P. C. Hiberty, S. S. Shaik, J. M. Lefour, G. Ohanessian, J. Org.
son, H. Nakatsuji, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Ha- Chem. 1985, 50, 4657; b) K. Jug, A. M. Koster, J. Am. Chem. Soc.
segawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, M. 1990, 112, 6772; c) A. Gobbi, Y. Yamaguchi, G. Frenking, H. F.
Klene, X. Li, J. E. Knox, H. P. Hratchian, J. B. Cross, V. Bakken, C. Schaefer III, Chem. Phys. Lett. 1995, 244, 27; d) S. Shaik, A. Shurki,
Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. D. Danovich, P. C. Hiberty, Chem. Rev. 2001, 101, 1501; e) B. Kova-
Austin, R. Cammi, C. Pomelli, J. W. Ochterski, P. Y. Ayala, K. Moro- cevic, D. Baric, Z. B. Maksic, T. M ller, ChemPhysChem 2004, 5,
kuma, G. A. Voth, P. Salvador, J. J. Dannenberg, V. G. Zakrzewski, 1352; f) A. Rehaman, A. Datta, S. S. Mallajosyula, S. K. Pati, J.
S. Dapprich, A. D. Daniels, M. C. Strain, O. Farkas, D. K. Malick, Chem. Theory Comput. 2006, 2, 30.
A. D. Rabuck, K. Raghavachari, J. B. Foresman, J. V. Ortiz, Q. Cui,
A. G. Baboul, S. Clifford, J. Cioslowski, B. B. Stefanov, G. Liu, A. Received: July 21, 2009
Liashenko, P. Piskorz, I. Komaromi, R. L. Martin, D. J. Fox, T. Published online: October 22, 2009

13032 www.chemeurj.org  2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eur. J. 2009, 15, 13022 – 13032

You might also like