PDF hosted at the Radboud Repository of the Radboud University
Nijmegen
The following full text is a publisher's version.
For additional information about this publication click this link.
https://hdl.handle.net/2066/225091
Please be advised that this information was generated on 2021-11-25 and may be subject to
change.
Chemistry—A European Journal
Full Paper
doi.org/10.1002/chem.202000857
& Cycloaddition
Understanding the 1,3-Dipolar Cycloadditions of Allenes
Song Yu,[a] Pascal Vermeeren,[a] Kevin van Dommelen,[a] F. Matthias Bickelhaupt,*[a, b] and
Trevor A. Hamlin*[a]
Abstract: We have quantum chemically studied the reactivity, site-, and regioselectivity of the 1,3-dipolar cycloaddition
between methyl azide and various allenes, including the archetypal allene propadiene, heteroallenes, and cyclic allenes,
by using density functional theory (DFT). The 1,3-dipolar cycloaddition reactivity of linear (hetero)allenes decreases as
the number of heteroatoms in the allene increases, and formation of the 1,5-adduct is, in all cases, favored over the
1,4-adduct. Both effects find their origin in the strength of
the primary orbital interactions. The cycloaddition reactivity
Introduction
Allenes are a class of unsaturated hydrocarbon that contain
two cumulated double bonds and have received significant attention in the past decade due to their privileged role in the
synthesis of natural products through cycloaddition reactions.[1]
The simplest allene, propadiene (CCC), for instance, reacts with
both cyclopentadiene and 1,3-dipoles to form either a substituted norbornene[2] or a heterocycle,[3] respectively (Scheme 1 a
and b), both of which are common motifs in natural products.
Intramolecular Diels–Alder reactions[4] as well as 1,3-dipolar cycloadditions[5] (Scheme 1 c and d) of allenes provide strategies
for the construction of complex polycyclic molecules.[6] In addition, the cycloaddition reactivity of allenes can be broadened
[a] S. Yu, P. Vermeeren, K. van Dommelen, Prof. Dr. F. M. Bickelhaupt,
Dr. T. A. Hamlin
Department of Theoretical Chemistry
Amsterdam Institute of Molecular and Life Sciences (AIMMS)
Amsterdam Center for Multiscale Modeling (ACMM)
Vrije Universiteit Amsterdam
De Boelelaan 1083, 1081 HV Amsterdam (The Netherlands)
E-mail:
[email protected]
[email protected]
[b] Prof. Dr. F. M. Bickelhaupt
Institute for Molecules and Materials (IMM)
Radboud University
Heyendaalseweg 135, 6525 AJ Nijmegen (The Netherlands)
Supporting information and the ORCID identification numbers for the authors of this article can be found under:
https://doi.org/10.1002/chem.202000857.
T 2020 The Authors. Published by Wiley-VCH GmbH. This is an open access
article under the terms of Creative Commons Attribution NonCommercial
License, which permits use, distribution and reproduction in any medium,
provided the original work is properly cited and is not used for commercial
purposes.
Chem. Eur. J. 2020, 26, 11529 – 11539
of cyclic allenes was also investigated, and the increased
predistortion of allenes, that results upon cyclization, leads
to systematically lower activation barriers not due to the expected variations in the strain energy, but instead from the
differences in the interaction energy. The geometric predistortion of cyclic allenes enhances the reactivity compared to
linear allenes through a unique mechanism that involves a
smaller HOMO–LUMO gap, which manifests as more stabilizing orbital interactions.
to heteroallenes, such as ketenimine (CCN),[7] ketene (CCO),[8]
carbodiimide (NCN),[9] isocyanate (NCO),[10] and even to carbon
dioxide (OCO).[11]
In contrast, strained allenes, that is, cyclic allenes, have received less attention in the field likely due to their lower kinetic stabilities.[12] Nevertheless, experimental studies have shown
that strained allenes can be formed in situ and trapped instantaneously by either dienes or 1,3-dipoles.[13] For example, Houk
and co-workers studied the formation and subsequent trapping of 1,2-cyclohexadiene in a Diels–Alder reaction
(Scheme 2 a).[13d] Lofstrand et al. synthesized and subsequently
trapped 1,2-cyclohexadiene, through a 1,3-dipolar cycloaddition under mild conditions (Scheme 2 b).[13e] Houk and Garg recently carried out a systematic study on the synthesis of azacyclic allenes as well as their reactivity towards cycloadditions
(Scheme 2 c).[13g] These examples clearly illustrate that cyclic allenes can serve as prominent building blocks in the construction of polycyclic compounds and may also engage in rapid reactions in analogy with strained alkenes and alkynes.[14]
A number of theoretical studies have shed light on the cycloaddition reactivity of allenes. A concerted asynchronous
mechanism has been proposed to be a more energetically favorable reaction pathway for cycloadditions of allenes.[13g, 15]
Gandolfi and co-workers proposed that the differences in the
extent of structural deformations determine the trends in reaction barrier heights of the cycloadditions of allenes.[15a,b] On the
contrary, an activation strain analysis on transition structures,
by Garg and co-workers, concluded just the opposite, namely,
that the strength of the interaction plays a large role in determining the regioselectivity of the Diels–Alder reactions of azacyclic allenes.[13g] To the best of our knowledge, a thorough investigation into the reactivity, site-, and regioselectivity of cycloadditions of allenes has not yet been reported.
11529
T 2020 The Authors. Published by Wiley-VCH GmbH
Chemistry—A European Journal
Full Paper
doi.org/10.1002/chem.202000857
Scheme 1. Inter- and intramolecular Diels–Alder reactions and 1,3-dipolar cycloadditions of propadiene.
Scheme 2. Diels–Alder reactions and 1,3-dipolar cycloadditions of strained allenes.
We have performed a systematic computational study of the
concerted 1,3-dipolar cycloaddition reactions of allenes, including the linear (hetero)allenes propadiene (CCC; L3), ketenimine
(CCN), ketene (CCO), carbodiimide (NCN), isocyanate (NCO),
and carbon dioxide (OCO) and a series of cyclic allenes 1,2-cyclooctadiene (C8), 1,2-cycloheptadiene (C7), and 1,2-cyclohexadiene (C6). These cyclic allenes have all been synthesized[16]
and might be relevant reactive dienophiles/dipolarophiles in
bioorthogonal chemistries in the future.[17] As azides are
common reactants in 1,3-dipolar cycloadditions,[18] as well as
strain-promoted azide–alkyne cycloadditions (SPAACs),[14a-c]
methyl azide (Az) was chosen as the model 1,3-dipole in this
study. The activation strain model (ASM)[19] in combination
with quantitative Kohn–Sham molecular orbital (KS-MO) theory
and the matching energy decomposition analysis (EDA)[20]
were employed to provide insight into the factor controlling
the reactivity in these cycloaddition reactions. This approach
has proven valuable for understanding of the reactivity of related pericyclic reactions and continues our current research
line into the reactivity of cyclic dienophiles and dipolarophiles.[21]
Computational Methods
All calculations were carried out in ADF2017,[22] using the
BP86[23] functional with the TZ2P basis set.[24] The exchangeChem. Eur. J. 2020, 26, 11529 – 11539
www.chemeurj.org
correlation (XC) functional has been proven to be accurate in
calculating the relative trends in activation and reaction energies for cycloadditions.[25] Additionally, single-point energies
were computed at BP86-D3(BJ)/TZ2P,[26] M06-2X/TZ2P,[27] and
COSMO(toluene)BP86/TZ2P[28] on the fully optimized BP86/
TZ2P geometries in order to assess the effect of a meta-hybrid
functional, dispersion-corrections, and solvation on the computed reactivity trends. Frequency calculations were performed
in order to characterize the nature of the stationary points.
Local minima present only real frequencies, whereas transition
state structures have one imaginary frequency. The potential
energy surface (PES) was calculated using the intrinsic reaction
coordinate (IRC) method, which follows the imaginary eigenvector of the transition structure towards the reactant and
product. The resulting PES was analyzed with the aid of the
PyFrag 2019 program.[29] All chemical structures were illustrated using CYLview.[30]
Quantitative analyses of the activation barriers associated
with the studied reactions are obtained by means of the activation strain model (ASM) of reactivity.[19] Herein, the PES,
DE(z), is decomposed into the strain energy, DEstrain(z), and the
interaction energy, DEint(z) [Eq. (1)]. All energy terms are projected onto the reaction coordinate z, the average distance of
newly forming bonds, which undergoes a well-defined change
during the course of the reactions and has been proven to
provide reliable results for cycloaddition reactions.[21a,25b,c, 31]
11530
T 2020 The Authors. Published by Wiley-VCH GmbH
Chemistry—A European Journal
Full Paper
doi.org/10.1002/chem.202000857
DEðzÞ ¼ DE strain ðzÞ þ DE int ðzÞ
ð1Þ
The DEstrain(z) is associated with the rigidity as well as the
structural deformation of the reactants from their equilibrium
geometry to the geometry acquired along the reaction coordinate. The total DEstrain(z) can be further divided into the strain
energy associated with deforming each respective reactant
[Eq. (2)].
DE strain ðzÞ ¼ DE strain,reactant,A ðzÞ þ DE strain,reactant,B ðzÞ
ð2Þ
The DEint(z) is related to the electronic structure of the reactants and their spatial orientation and takes the mutual interaction between the deformed reactants into account. In order
to obtain a deeper insight into the physical mechanism behind
the interaction energy, we employ the canonical energy decomposition analysis (EDA).[20] This analysis method decomposes the interaction energy between the two deformed reactants, within the framework of Kohn–Sham DFT, into three
physically meaningful terms [Eq. (3)].
DE int ðzÞ ¼ DV elstat ðzÞ þ DE Pauli ðzÞ þ DE oi ðzÞ
ð3Þ
The electrostatic interaction, DVelstat(z), corresponds to the
classical electrostatic interaction between the unperturbed
charge distributions of the deformed reactants. The Pauli repulsion, DEPauli(z), comprises the repulsion between closedshell occupied orbitals and is, therefore, destabilizing. The orbital interaction, DEoi(z), accounts for the stabilizing orbital interactions such as electron-pair bonding, charge transfer (interaction between the occupied orbitals of fragment A with the unoccupied orbitals of fragment B, and vice versa), and polarization (e.g., occupied-unoccupied orbital mixing on fragment A
due to the presence of fragment B and vice versa). A detailed
step-by-step protocol on how to perform the activation strain
and energy decomposition analysis can be found in ref. [19a].
The magnitude of the orbital interaction of a 1,3-dipolar cycloaddition mainly comes from two distinct orbital interaction
mechanisms, namely, the normal electron demand (NED) interaction, occurring between occupied orbitals of the dipole and
unoccupied orbitals of the dipolarophile (the allene in this
study), and the inverse electron demand (IED) interaction, originating from the interaction between the unoccupied orbitals
of the dipole with occupied orbitals of the dipolarophile. The
stabilization of a specific orbital interaction mechanism is proportional to the orbital overlap squared divided by their respective orbital energy gap, that is, S2/De.[32] Thus, with the
help of this relation, we can quantify the importance of the individual orbital interaction mechanisms.
The atomic charge distribution was analyzed by using the
Voronoi deformation density (VDD) method.[33] The VDD
method partitions the space into so-called Voronoi cells, which
are non-overlapping regions of space that are closer to nucleus
A than to any other nucleus. The charge distribution is determined by taking a fictitious promolecule as reference point, in
which the electron density is simply the superposition of the
atomic densities. The change in density in the Voronoi cell
Chem. Eur. J. 2020, 26, 11529 – 11539
www.chemeurj.org
when going from this promolecule to the final molecular density of the interacting system is associated with the VDD
atomic charge Q. The VDD atomic charge QA of atom A is calculated according to Equation (4):
Q
VDD
A
¼@
Z
½1ðr Þ @ 1promolecule ðrÞAdr
ð4Þ
Voronoi cell of A
So, instead of computing the amount of charge contained in
an atomic volume, we compute the flow of charge from one
atom to the other upon formation of the molecule. The physical interpretation is therefore straightforward. A positive
atomic charge QA corresponds to the loss of electrons, whereas
a negative atomic charge QA is associated with the gain of
electrons in the Voronoi cell of atom A.
Results and Discussion
1,3-Dipolar cycloadditions of linear allenes
As a starting point, we studied the 1,3-dipolar cycloaddition reaction between methyl azide (Az) and the following linear
(hetero)allenes: propadiene (CCC), ketenimine (CCN), ketene
(CCO), carbodiimide (NCN), isocyanic acid (NCO), and carbon
dioxide (OCO). For each 1,3-dipolar cycloaddition, two regiospecific cycloadducts can be formed, the 1,5-adduct with the
methyl group adjacent to the second double bond and the
1,4-adduct with them on opposite sides (Scheme 3). Additionally, the asymmetric heteroallenes CCN, CCO, and NCO are
able to form two site-specific adducts, i.e., coordinating with
either of the two double bonds.
Scheme 3. The 1,3-dipolar cycloaddition between Az and a linear allene.
Table 1 lists the activation energies, DE*, and reaction energies, DErxn, for the studied 1,3-dipolar cycloadditions between
Az and linear (hetero)allenes. Three clear trends can be observed. In the first place, the cycloadditions towards the 1,5-adducts are kinetically and thermodynamically favored over the
formation of the 1,4-adducts. Secondly, for the asymmetric heteroallenes, Az preferentially attacks at the more electropositive
of the two terminal atoms. The only exception, however, is
CCO, which has a slightly lower DE* for the attack at the CO
(19.2 kcal mol@1) than the CC (20.0 kcal mol@1). Thirdly, the cycloaddition reactivity decreases when heteroatoms are introduced in the linear allene, from CCC to CCN and CCO, as well
as from CCO to NCO to OCO and from CCN to NCN. The computed trends in reactivity at BP86/TZ2P agree well with those
11531
T 2020 The Authors. Published by Wiley-VCH GmbH
Full Paper
doi.org/10.1002/chem.202000857
Chemistry—A European Journal
Table 1. Electronic reaction barriers DE* and reaction energies DErxn [kcal mol@1] for the 1,3-dipolar cycloadditions between Az and linear allenes leading
to 1,4- and 1,5-adducts computed at various levels of theory.[d]
Allene
OCO
NCN
NCO
CCO
CCN
CCC
Site
CO
NC
CO
CC
CN
CC
DE*[a]
DE*[b]
1,5-Adduct
DE*[c]
DErxn[a]
DErxn[b]
DErxn[c]
DE*[a]
DE*[b]
1,4-Adduct
DE*[c]
DErxn[a]
DErxn[b]
DErxn[c]
32.1
28.0
27.9
26.4
19.2
20.0
22.6
20.0
19.0
28.8
23.7
24.1
22.7
15.0
15.7
18.0
15.7
14.1
35.8
30.9
28.8
28.9
21.7
26.2
27.1
27.0
24.4
17.2
@19.8
10.7
@18.8
@0.8
@36.4
@25.1
@35.6
@39.1
14.8
@25.2
8.3
@24.0
@2.6
@42.1
@29.1
@41.3
@43.7
53.8
35.8
45.0
42.8
27.1
29.1
26.4
23.5
19.5
50.6
31.6
41.4
39.0
23.4
24.8
22.2
19.0
14.8
62.6
43.5
54.0
50.8
40.2
37.2
34.3
30.8
25.2
34.8
@2.6
29.0
2.9
12.9
@28.6
@10.4
@32.6
@36.2
35.7
@6.6
27.0
@0.1
11.5
@32.2
@13.7
@36.7
@39.7
20.6
@15.3
14.6
@14.9
3.5
@32.2
@20.3
@30.8
@34.0
37.8
1.6
32.4
6.6
16.5
@24.5
@6.1
@28.1
@31.5
[a] Computed at BP86/TZ2P. [b] Computed at BP86-D3(BJ)/TZ2P//BP86/TZ2P. [c] Computed at M06-2X/TZ2P//BP86/TZ2P. [d] See Table S1 for computed enthalpies and Gibbs free energies.
calculated at BP86-D3(BJ)/TZ2P//BP86/TZ2P and M06-2X/
TZ2P//BP86/TZ2P. We note that, when using M06-2X/TZ2P//
BP86/TZ2P, the two site-selective cycloadditions of NCO, forming the 1,5-adduct, have nearly identical reaction barriers. Furthermore, the observed trends in reactivity, site-, and regioselectivity also hold when solvent effects in toluene are included
at COSMO(toluene)BP86/TZ2P//BP86/TZ2P (Table S1 in the Supporting Information).
1,5- versus 1,4-regioselectivity
Next, we turn to the activation strain model (ASM)[19] of reactivity to gain a quantitative insight into the physical factors governing the 1,5- versus 1,4-regioselectivity in the 1,3-dipolar cycloadditions presented herein. In Figure 1, we focus on the
ASM diagram for the 1,5- vs. 1,4-regioselectivity of OCO for
which the difference in DE* is the largest (Table 1). The ASM diagrams of the other linear allenes possess the same characteristics, only less pronounced (Figures S1–S7). The lower DE* for
the formation of the 1,5-adduct originates mainly from a more
stabilizing DEint term, whereas the DEstrain is nearly identical
(Figure 1 a). The canonical energy decomposition analysis
(EDA)[20] reveals that both the more stabilizing DVelstat and DEoi
are the causes of the more favorable DEint term for the 1,5adduct formation compared to the 1,4-adduct (Figure 1 b).
The more stabilizing DEoi for the 1,3-dipolar cycloaddition
yielding the 1,5-adduct can be entirely described to the more
effective orbital overlap of the normal electron demand (NED)
interaction occurring between the HOMO@1Az and LUMOOCO.
Only the lower-lying HOMO@1Az participates in the NED interaction, because its lobes are oriented towards the LUMO of
OCO, while the lobes of HOMOAz are orthogonal to the
LUMOOCO (Scheme S1). As shown in Scheme 4, the HOMO@1Az
has the largest lobe on the nitrogen next to the methyl group,
due to a methyl-induced mix of the p-atomic orbitals of the N3
fragment of Az (Scheme S1e). The LUMOOCO has a larger lobe
on the carbon atom than on the terminal oxygens, due to the
more diffuse nature of the 2p atomic orbital of carbon com-
Figure 1. a) Activation strain and b) energy decomposition analysis of the 1,3-dipolar cycloaddition between Az and OCO, projected onto the average newly
forming C/O···N bond, computed at BP86/TZ2P.
Chem. Eur. J. 2020, 26, 11529 – 11539
www.chemeurj.org
11532
T 2020 The Authors. Published by Wiley-VCH GmbH
Chemistry—A European Journal
Full Paper
doi.org/10.1002/chem.202000857
Scheme 4. Schematic diagrams of the orbital interaction between the
HOMO@1 of Az and the LUMO of OCO for the 1,4- and 1,5-adducts. VDD
charges (red, in electrons) of key atoms in isolated fragment computed at
BP86/TZ2P.
pared to oxygen, which, in turn, leads to a better HOMO@1Az–
LUMOOCO orbital overlap when forming the 1,5-adduct compared to the 1,4-adduct. The computed overlaps of the
HOMO@1Az–LUMOOCO NED interaction for formation of both
adducts on a consistent geometry with an average newly
forming C/O···N bond length of 1.86 a amounts S1,5 = 0.30 and
S1,4 = 0.16. The larger orbital overlap for the formation of the
1,5-adduct is responsible for the more stabilizing DEoi compared to the 1,4-adduct counterpart (Figure 1 b). The NED orbital energy gaps, on the other hand, are identical for the formation of the 1,5- and 1,4-adduct, because the orbital interactions
take place between the same molecular orbitals. In addition,
the cycloaddition resulting in the 1,5-adduct also has a stronger electrostatic attraction between the more negatively
charged nitrogen and the positively charged carbon atom
(Scheme 4) and, therefore, a significantly more stabilizing
DVelstat term (Figure 1 b).
Site-selectivity of asymmetric heteroallenes
After having established that the 1,3-dipolar cycloaddition between Az and (hetero)allene preferentially form the 1,5-adduct,
we have analyzed the site-selectivity of the asymmetric heteroallenes CCN, NCO, and CCO. First, we discuss the site-selectivi-
ty of CCN by applying the ASM analysis. From Figure 2 a we
can clearly see that the attack of Az at the more electropositive CC bond is favored exclusively due to a more stabilizing
DEint term compared to the attack at the CN bond. The more
stabilizing DEint for attack at CC compared to CN compensates
for the destabilizing DEstrain for this pathway. Our EDA indicates
that the more stabilizing DEint term for the attack at CC over
CN originates mainly from a more favorable DEoi supported by
a moderately stronger DVelstat (Figure 2 b).
The more stabilizing DEoi term for the Az attack at CC can
exclusively be ascribed to its significantly more favorable inverse electron demand (IED) interaction term (Figure 3 a),
which overcomes its less stabilizing NED interaction (Figure 3 b). The IED energy gap for the attack at CC is considerably smaller compared to the attack at CN, 1.7 and 5.0 eV, respectively, while the orbital overlap is also larger for the attack
at CC. This manifest in an orbital stabilization term, that is,103 V
S2/De, of 17.0 and 4.0 for the attack at CC and CN, respectively.
In contrast, the NED interaction is slightly weaker for the
attack at CC than for CN, due to a larger NED energy gap and
a poorer orbital overlap. This, however, can easily be overcome
by the much stronger IED interaction, which leads to a more
stabilizing DEoi and thus a lower reaction barrier for the attack
at the CC bond (Figure 2 b).
In the case of the asymmetric linear heteroallene NCO, the
underlying mechanism behind the preference for the attack at
NC over CO is identical to the above discussed CCN (Figure S8 a). For heteroallene CCO, the DEstrain for the attack at CC
is more destabilizing than for CO, because the terminal carbon
atom needs to deform from a trigonal planar to a tetrahedral
geometry, which overcomes the more favorable DEint, leading
to nearly identical reaction barriers (Figure S8 b). But, the cycloaddition at CO is reversible and goes with a positive reaction
energy (Table S1 for Gibbs free reaction energies), therefore,
the reaction at CC will be preferred thermodynamically. The
finding that allenes prefer to undergo cycloadditions at the
Figure 2. a) Activation strain and b) energy decomposition analysis of site-specific 1,3-dipolar cycloadditions between Az and CCN, projected onto the average newly forming C/N···N bond, computed at BP86/TZ2P. The vertical dotted line indicates the point along the reaction coordinate at which the average
length of newly forming C/N···N bond is 2.19 a.
Chem. Eur. J. 2020, 26, 11529 – 11539
www.chemeurj.org
11533
T 2020 The Authors. Published by Wiley-VCH GmbH
Chemistry—A European Journal
Full Paper
doi.org/10.1002/chem.202000857
Figure 3. FMO diagrams with calculated orbital energy gaps and overlaps of a) the IED (LUMOAz–p-MOCCN) interaction and b) the NED (HOMO@1Az–p*-MOCCN)
interaction for site-specific 1,3-dipolar cycloadditions between Az and CCN at consistent geometries with the average newly forming C/N···N bond of 2.19 a
computed at BP86/TZ2P.
more electropositive terminal atom is in line with several experimental reports.[7, 8, 10]
Influence of heteroatoms on the reactivity
In this section, we discuss the effect of heteroatoms on the reactivity of linear allenes towards the 1,3-dipolar cycloaddition
with Az yielding the 1,5-adduct, by systematically modifying
the nature and number of heteroatoms. Upon going from CCC
to CCN and CCO, while attacking at the kinetically preferred
CC site, the DE* increases from 19.0 to 20.0 kcal mol@1 solely
due to a more destabilizing DEstrain (Figure 4 a). Even though
CCC requires a larger extent of bending over the course of the
reaction compared to CCN and CCO (CCC: 248; CCN: 188;
CCO: 228), the difference in DEstrain can be ascribed to the
more rigid heteroallene CCX (X = N, O) backbone, which can
be reflected by the calculated bending vibrational frequencies
of allenes (CCC: 361 cm@1; CCN: 471 cm@1; CCO: 503 cm@1) as
well as the analysis of the strain energy upon artificially bending of the heteroallene (Figure S9a). The increased rigidity from
CCC to CCN to CCO is due to the increased bond strength between carbon and the heteroatom along the same series.[34]
The DEint term, on the other hand, shows a trend which is opposite to the strain energy, namely, CCO goes with the most
stabilizing interaction energy followed by CCN and CCC. This
trend in interaction energy is exclusively determined by the orbital interactions (see Figure S9b for EDA diagrams), which, in
turn, can be traced to a less stable LUMOallene going from CCO
Figure 4. a) Activation strain analysis for the 1,3-dipolar cycloadditions between Az and CCC, CCN, and CCO attacking at the CC site and b) equilibrium geometries of the allene and respective consistent geometries with the average newly forming C/N···N bond of 2.20 a with internal bending angles [8] computed
at BP86/TZ2P.
Chem. Eur. J. 2020, 26, 11529 – 11539
www.chemeurj.org
11534
T 2020 The Authors. Published by Wiley-VCH GmbH
Chemistry—A European Journal
Full Paper
doi.org/10.1002/chem.202000857
to CCN and CCC (@2.4, @1.5, and @0.9 eV, respectively), and,
therefore, a larger HOMO@1Az–LUMOallene energy gap. This
trend in DEint term, however, is overruled by the larger differences in DEstrain.
For the series CCO, NCO, and OCO, Az attacks at different
sites, namely, CC, NC, and OC, respectively, but the neighboring heteroatom is always oxygen. Along this series, the DE*
systematically increases (Table 1), due to a less stabilizing DEoi
(Figure S10). Similar to the analysis of the site-selectivity of the
asymmetric heteroallenes (Figure 3), going from CC to the
hetero double bonds NC and OC causes a remarkably destabilized IED interaction (LUMOAz–HOMOallene), due to the increased
IED orbital energy gap supported by less efficient orbital overlap (Figure S11). This exact rationale also holds for the comparison of CCN and NCN (Figure S12).
Summarizing, we have analyzed and compared the 1,3-dipolar cycloaddition reactivity of linear (hetero)allenes with Az,
which is all cases prefers to form the 1,5-adduct. The archetypal allene, CCC, is the most reactive. By introducing a heteroatom, the heteroallene becomes less reactive due to the increased rigidity of the CCX (X = N, O) backbone. Additionally, a
second heteroatom diminishes the stabilizing DEoi, making
them even less reactive towards Az.
1,3-Dipolar cycloaddition of cyclic allenes
At last, we also analyzed and compared the 1,3-dipolar cycloaddition reactions between methyl azide (Az) and a series of
cyclic allenes, namely, 1,2-cyclooctadiene (C8), 1,2-cycloheptadiene (C7), and 1,2-cyclohexadiene (C6) as well as propadiene
(L3), the most reactive linear allene (vide supra). These cyclic
allenes have all been synthesized and featured in cycloaddition
reactions.[13]
Figure 5 shows the transition state structures of the 1,3-dipolar cycloadditions of Az with the linear allene (Az-L3) and
the cyclic allenes (Az-C8–Az-C6). The transition structures are
concerted asynchronous and become earlier, with regard to
the average forming bond distances, as the ring size of the
cyclic allene decreases. The cycloaddition of the linear L3 is
predicted to proceed with the highest reaction barrier (DE* =
19.0 kcal mol@1) and has the least favorable reaction energy
(DErxn = @34.0 kcal mol@1). The reaction barrier height decreases
along the series L3 > C8 > C7 > C6, and the cycloaddition reaction becomes more exergonic when going from L3 to C6,
which is in line with the Hammond–Leffler postulate[35] (Figure S13). The computed trends at BP86/TZ2P agree well with
those calculated at BP86-D3(BJ)/TZ2P//BP86/TZ2P and M06-2X/
TZ2P//BP86/TZ2P, as well as when solvent effects are included
at COSMO(toluene)BP86/TZ2P//BP86/TZ2P (Table S2).
In order to understand the intrinsic differences in reactivity
between linear and cyclic allenes in the 1,3-dipolar cycloaddition with Az, we performed an ASM analysis. Figure 6 a graphically represents how the DEstrain and DEint components evolve
along the reaction coordinate for 1,3-dipolar cycloadditions of
Az with L3 and C8–C6. Surprisingly, the origin of the increased
reactivity as the ring size of allene decreases can be entirely attributed to the differences in DEint, which becomes more stabilizing from L3 to C6 (Figure 6 a). The total DEstrain for all studied
allenes are nearly identical (Figure 6 a). As expected upon decreasing the size of the ring, the cyclic allene becomes more
predistorted towards the cycloaddition reaction with Az, which
leads to a smaller contribution of the cyclic allene to the total
DEstrain, consistent with the earlier literature.[13f] The contribution of the 1,3-dipole Az to the total DEstrain, however, is more
destabilizing for C6 than for L3 (Figure S14), due to the fact
that the more reactive allenes (vide infra) deform Az to a
larger degree (Figure S15).
The origin of the differences in DEint was uncovered by
means of the EDA method, and the results are shown in Figure 6 b. It is apparent that the DEoi is the major contributor to
the trend in DEint, guided by a smaller contribution of DVelstat.
The DEPauli shows a reverse trend, and, therefore, is not responsible for the trend in DEint. To further probe the key orbital interactions, that cause this difference in DEoi, involved in the
1,3-dipolar cycloadditions of Az with L3 and C8–C6, we analyzed the FMOs participating in these interactions on consistent geometries with an average newly forming C···N bond of
2.48 a (Figure 7).
The FMOs participating in the NED and IED reveal that the
more stabilizing orbital interactions when going from L3 to C8
Figure 5. Transition structures with forming bond lengths [a], computed reaction barriers (DE* [kcal mol@1], blue) with relative reaction rate constants (krel,
black), and reaction energies (DErxn [kcal mol@1], red) for 1,3-dipolar cycloadditions of Az with L3 and C8–C6 computed at BP86/TZ2P.
Chem. Eur. J. 2020, 26, 11529 – 11539
www.chemeurj.org
11535
T 2020 The Authors. Published by Wiley-VCH GmbH
Chemistry—A European Journal
Full Paper
doi.org/10.1002/chem.202000857
Figure 6. a) Activation strain and b) energy decomposition analysis of 1,3-dipolar cycloadditions of Az with L3 and C8–C6, projected onto the average newly
forming C···N bond, computed at BP86/TZ2P. The vertical dotted line indicates the point along the reaction coordinate where the average newly forming C···N
bond is 2.48 a.
Figure 7. FMO diagrams with calculated key orbital energy gaps and overlaps of a) the NED (HOMO@1Az–LUMOallene) interaction and b) the IED (LUMOAz–HOMOallene) interaction for 1,3-dipolar cycloadditions between Az with L3 and C8–C6 at consistent geometries with the average newly forming C···N bond of
2.48 a computed at BP86/TZ2P.
to C6 are exclusively determined by a reduction in orbital
energy gap (Figure 7). The NED interaction between Az and L3
and C8–C6 occurs between the HOMO@1Az and LUMOallene
(Figure 7 a). The least reactive allene L3 has the largest and
least favorable NED orbital energy gap (De = 6.4 eV). As the
ring size decreases from L3 to C8 to C6, the NED orbital
energy gap continuously decreases from 6.4 to 4.8 eV. The orbital overlap in the NED interaction are identical for all reactions (S = 0.15). The IED interaction takes place between the
LUMOAz and HOMOallene (Figure 7 b). Again, L3 has the largest
and, therefore, least favorable IED orbital energy gap (De =
3.7 eV). The IED gap also systematically decreases from 3.7 eV
Chem. Eur. J. 2020, 26, 11529 – 11539
www.chemeurj.org
for L3 to 2.7 eV for C6. The increasingly stabilizing DEoi term
(Figure 6 b), as the ring size of allene decreases, therefore, is a
direct result of the diminishing energy gap for both the NED
and IED interaction, resulted from the continuously stabilizing
LUMO and destabilizing HOMO of allene (Figures 7 and S16).
In order to quantify the effect of allene predistortion on the
HOMO and LUMO, we chose to bent our model system L3. Figure 8 a shows the optimized undistorted structure (top) and
the distorted, bent, structures of L3 (middle and bottom).
Bending of the allene backbone causes a loss in orthogonality
of the two adjacent p systems, because it is accompanied with
a twist in the structure, reducing the dihedral angle from 908,
11536
T 2020 The Authors. Published by Wiley-VCH GmbH
Chemistry—A European Journal
Full Paper
doi.org/10.1002/chem.202000857
Figure 8. a) Front and right-side views of the pristine and constrained optimized structures of L3. b) FMO energies associated with the internal angle computed at BP86/TZ2P.
for the linear allene, to 64.18, for the 1308 bent allene. This observation not only holds for L3, but also for the cyclic allenes
C6–C8 (Figure S17) and is in line with earlier reported literature.[13g] As the backbone of L3 becomes distorted, the LUMO is
stabilized while the HOMO is destabilized (Figure 8 b).
Detailed Kohn–Sham molecular orbital (KS-MO) analysis of
the formation of the HOMO and LUMO of the undistorted
(linear) and distorted (bent to 1308) H2C=C = CH2 (L3), in terms
of an H2C=CCC and a CCCH2 fragment, is shown in Figure 9. For
the archetypal L3 (Figure 9 a), one LUMO (the bold LUMOs in
Figure 9) is solely formed by the p* orbital of H2C=CCC, whereas
the other degenerate LUMO, which is orthogonal to the
former, is a result of the antibonding combination of the p orbitals of two individual fragments. Furthermore, the HOMO
(the bold HOMOs in Figure 9) originates from the antibonding
combination between the p orbital of H2C=CCC and the C@H s
orbital of CCCH2, meanwhile the other degenerate HOMO is the
bonding combination of the p orbitals of both fragments.
When L3 is bent to 1308 (Figure 9 b), the p* orbital of H2C=CCC
has an in-phase overlap with the s* orbital of CCCH2 which leads
to a stabilization of the LUMO. In addition, due to the prior
mentioned twisting effect, the fragment p orbitals mix into the
LUMO which results in the additional stabilization. The HOMO,
on the other hand, is stabilized due to the decreased antibonding p–s overlap owing to the bending and twisting of
the backbone, but, at the same time, obtains a slightly stronger destabilization from the mixing of the fragmental p orbitals. This destabilization effect overcomes the stabilizing counterpart, resulting in the overall destabilization of the HOMO.
Figure 9. Diagrams for the Kohn–Sham MO analyses of a) the archetypal and b) 1308 bent L3, where the fragments are H2C=CCC and CCCH2 computed at BP86/
TZ2P.
Chem. Eur. J. 2020, 26, 11529 – 11539
www.chemeurj.org
11537
T 2020 The Authors. Published by Wiley-VCH GmbH
Chemistry—A European Journal
Full Paper
doi.org/10.1002/chem.202000857
These analyses were further verified by investigating both
the pure bending (no twisting) or twisting (no bending) of L3.
Solely bending L3 and maintaining orthogonality of the structure stabilizes the LUMO due to the enhanced p*–s* overlap
and also stabilizes the HOMO because of the decreased p–s
overlap (Figure S18). On the other hand, solely twisting L3 and
maintaining a linear backbone induces a stabilization of the
LUMO, because of an in-phase mixing between the p* and p
orbitals of the fragments, and a significantly destabilization of
the HOMO, due to the mixing between the p and p orbitals
(Figure S19).
Conclusions
1,3-Dipolar cycloadditions of linear allenes and heteroallenes
with methyl azide (Az) favor the formation of the 1,5-adduct
over the 1,4-adduct. In addition, bond formation to the asymmetric heteroallene is preferred at the more electropositive terminal atom. This process becomes less reactive as the number
of heteroatoms in the allene increases. Cyclic allenes experience a significant rate enhancement compared to their linear
allene counterparts. These findings emerge from our quantum
chemical study based on density functional theory calculations.
Our activation strain analyses furthermore identified that the
site-selective preference for the 1,5-adduct compared to the
1,4-adduct is exclusively determined by a more favorable orbital overlap and thus more stabilizing orbital interactions between the reactants. Furthermore, in the case of the asymmetric heteroallenes, the preference for attacking at the more electropositive atoms is caused by a significantly stronger inverse
electron demand (IED) orbital interaction. This is due to the
fact that double bonds involving more electropositive atoms
have lower-lying acceptor orbitals, leading to smaller IED
energy gaps and, thus, more stabilizing orbital interactions
with Az. The archetypal allene, propadiene (CCC) was found to
be the most reactive linear allene. Introducing a heteroatom to
CCC makes the allene less reactive, due to a more destabilizing
DEstrain, originating from a more rigid backbone, as well as less
stabilizing orbital interactions.
The enhanced reactivity of cyclic allenes with respect to
linear ones originates from an enhancement of donor–acceptor orbital interactions, which become more stabilizing as the
ring size of the cyclic allene decreases, and not from a previously reported reduced activation strain. Our activation strain
analyses reveal that, in smaller rings, the allene moiety is more
bent; this goes with a smaller HOMO–LUMO gap in the p-electron system and, hence, with the aforementioned stabilization
of the transition state by stronger donor–acceptor orbital interactions.
Acknowledgements
This work was supported by the Netherlands Organization for
Scientific Research (NWO), the China Scholarship Council (CSC),
Chem. Eur. J. 2020, 26, 11529 – 11539
www.chemeurj.org
and the Dutch Astrochemistry Network (DAN-II). We thank
SURFsara for use of the Cartesius supercomputer.
Conflict of interest
The authors declare no conflict of interest.
Keywords: 1,3-dipolar cycloadditions · activation strain
model · allenes · density functional theory calculations ·
reactivity
[1] a) S. Yu, S. Ma, Angew. Chem. Int. Ed. 2012, 51, 3074 – 3112; Angew.
Chem. 2012, 124, 3128 – 3167; b) A. Hoffmann-Rçder, N. Krause, Angew.
Chem. Int. Ed. 2004, 43, 1196 – 1216; Angew. Chem. 2004, 116, 1216 –
1236; c) N. Krause, A. S. K. Hashmi, Modern Allene Chemistry, Wiley-VCH,
Weinheim, 2004; d) T. M. V. D. Pinho e Melo, Curr. Org. Chem. 2009, 13,
1406 – 1431; e) T. M. V. D. Pinho e Melo, Monatsh. Chem. 2011, 142, 681 –
697; f) A. L. Cardoso, Curr. Org. Chem. 2019, 23, 3064 – 3134.
[2] F. Scheufler, M. E. Maier, Eur. J. Org. Chem. 2000, 3945 – 3948.
[3] A. Padwa, D. N. Kline, K. F. Koehler, M. Matzinger, M. K. Venkatramanan,
J. Org. Chem. 1987, 52, 3909 – 3917.
[4] a) G. Himbert, L. Henn, Angew. Chem. Int. Ed. Engl. 1982, 21, 620; Angew.
Chem. 1982, 94, 631; b) G. Himbert, D. Fink, K. Diehl, P. Rademacher,
A. J. Bittner, Chem. Ber. 1989, 122, 1161 – 1173.
[5] K. S. Feldman, M. R. Iyer, J. Am. Chem. Soc. 2005, 127, 4590 – 4591.
[6] a) L. S. Trifonov, A. S. Orahovats, Helv. Chim. Acta 1989, 72, 59 – 64; b) Y.
Schmidt, J. K. Lam, H. V. Pham, K. N. Houk, C. D. Vanderwal, J. Am. Chem.
Soc. 2013, 135, 7339 – 7348; c) G. Cheng, X. He, L. Tian, J. Chen, C. Li, X.
Jia, J. Li, J. Org. Chem. 2015, 80, 11100 – 11107; d) J. K. Lam, Y. Schmidt,
C. D. Vanderwal, Org. Lett. 2012, 14, 5566 – 5569; e) K. S. Feldman, D. K.
Hester, C. S. Ljpez, O. N. Faza, Org. Lett. 2008, 10, 1665 – 1668; f) K. S.
Feldman, D. K. Hester, M. R. Iyer, P. J. Munson, C. Silva Ljpez, O. N. Faza,
J. Org. Chem. 2009, 74, 4958 – 4974; g) K. S. Feldman, M. R. Iyer, C. Silva
Ljpez, O. N. Faza, J. Org. Chem. 2008, 73, 5090 – 5099; h) K. S. Feldman,
I. Y. Gonzalez, C. M. Glinkerman, J. Am. Chem. Soc. 2014, 136, 15138 –
15141; i) K. S. Feldman, I. Y. Gonzalez, C. M. Glinkerman, J. Org. Chem.
2015, 80, 11849 – 11862.
[7] a) C. Larksarp, O. Sellier, H. Alper, J. Org. Chem. 2001, 66, 3502 – 3506;
b) M. Alajarin, B. Bonillo, M.-M. Ortin, R.-A. Orenes, A. Vidal, Org. Biomol.
Chem. 2011, 9, 6741 – 6749; c) W. J. Kauffman, J. Org. Chem. 1970, 35,
4244 – 4245.
[8] a) X.-L. Huang, L. He, P.-L. Shao, S. Ye, Angew. Chem. Int. Ed. 2009, 48,
192 – 195; Angew. Chem. 2009, 121, 198 – 201; b) T.-Y. Jian, X.-Y. Chen, L.H. Sun, S. Ye, Org. Biomol. Chem. 2013, 11, 158 – 163; c) M. Schmittel, H.
von Seggern, J. Am. Chem. Soc. 1993, 115, 2165 – 2177; d) P.-L. Shao, X.Y. Chen, S. Ye, Angew. Chem. Int. Ed. 2010, 49, 8412 – 8416; Angew.
Chem. 2010, 122, 8590 – 8594; e) M. Presset, K. Mohana, M. Hamann, Y.
Coquerel, J. Rodriguez, Org. Lett. 2011, 13, 4124 – 4127.
[9] a) J. Světl&k, J. Leško, A. Martvoň, Monatsh. Chem. 1980, 111, 635 – 642;
b) W.-P. Yen, F.-C. Kung, F. F. Wong, Eur. J. Org. Chem. 2016, 2328 – 2335.
[10] a) K. M. Oberg, T. Rovis, J. Am. Chem. Soc. 2011, 133, 4785 – 4787; b) Z.
Zheng, Y. Cao, D. Zhu, Z. Wang, K. Ding, Chem. Eur. J. 2019, 25, 9491 –
9497.
[11] a) L.-L. Zhao, S.-Y. Wang, X.-P. Xu, S.-J. Ji, Chem. Commun. 2013, 49,
2569 – 2571; b) C.-X. Guo, W.-Z. Zhang, N. Zhang, X.-B. Lu, J. Org. Chem.
2017, 82, 7637 – 7642; c) F. Murillo, J. Barroso, M. G. de los Santos, G.
]vila, S. Pan, M. A. Fern#ndez-Herrera, G. Merino, J. Org. Chem. 2018, 83,
13045 – 13050.
[12] a) R. O. Angus, M. W. Schmidt, R. P. Johnson, J. Am. Chem. Soc. 1985,
107, 532 – 537; b) R. P. Johnson, Chem. Rev. 1989, 89, 1111 – 1124; c) K. J.
Daoust, S. M. Hernandez, K. M. Konrad, I. D. Mackie, J. Winstanley, R. P.
Johnson, J. Org. Chem. 2006, 71, 5708 – 5714.
[13] a) P. Mohanakrishnan, S. R. Tayal, R. Vaidyanathaswamy, D. A. Devaprabhakara, Tetrahedron Lett. 1972, 13, 2871 – 2872; b) A. T. Bottini, L. L.
Hilton, Tetrahedron 1975, 31, 2003 – 2004; c) M. Balci, W. M. Jones, J. Am.
Chem. Soc. 1980, 102, 7607 – 7608; d) M. Nendel, L. M. Tolbert, L. E. Herring, M. N. Islam, K. N. Houk, J. Org. Chem. 1999, 64, 976 – 983; e) V. A.
11538
T 2020 The Authors. Published by Wiley-VCH GmbH
Chemistry—A European Journal
[14]
[15]
[16]
[17]
[18]
[19]
[20]
Full Paper
doi.org/10.1002/chem.202000857
Lofstrand, F. G. West, Chem. Eur. J. 2016, 22, 10763 – 10767; f) J. S.
Barber, E. D. Styduhar, H. V. Pham, T. C. McMahon, K. N. Houk, N. K. Garg,
J. Am. Chem. Soc. 2016, 138, 2512 – 2515; g) J. S. Barber, M. M. Yamano,
M. Ramirez, E. R. Darzi, R. R. Knapp, F. Liu, K. N. Houk, N. K. Garg, Nat.
Chem. 2018, 10, 953 – 960.
a) R. Manova, T. A. van Beek, H. Zuilhof, Angew. Chem. Int. Ed. 2011, 50,
5428 – 5430; Angew. Chem. 2011, 123, 5540 – 5542; b) S. S. van Berkel, S.
Brauch, L. Gabriel, M. Henze, S. Stark, D. Vasilev, L. A. Wessjohann, M.
Abbas, B. Westermann, Angew. Chem. Int. Ed. 2012, 51, 5343 – 5346;
Angew. Chem. 2012, 124, 5437 – 5441; c) C. Wendeln, I. Singh, S. Rinnen,
C. Schulz, H. F. Arlinghaus, G. A. Burley, B. J. Ravoo, Chem. Sci. 2012, 3,
2479 – 2484; d) S. H. Lee, O. K. Park, J. Kim, K. Shin, C. G. Pack, K. Kim, G.
Ko, N. Lee, S.-H. Kwon, T. Hyeon, J. Am. Chem. Soc. 2019, 141, 13829 –
13840; e) J. Escorihuela, A. Das, W. J. E. Looijen, F. L. van Delft, A. J. A.
Aquino, H. Lischka, H. Zuilhof, J. Org. Chem. 2018, 83, 244 – 252.
a) A. Rastelli, M. Bagatti, R. Gandolfi, Tetrahedron 1994, 50, 5561 – 5568;
b) A. Rastelli, M. Bagatti, R. Gandolfi, J. Am. Chem. Soc. 1995, 117, 4965 –
4975; c) M. Manoharan, P. Venuvanalingam, J. Chem. Soc. Perkin Trans. 2
1996, 1423 – 1427; d) A. B. G. P8rez, A. B. Gonz#lez P8rez, J. A. Souto,
C. S. Ljpez, ]. R. de Lera, Eur. J. Org. Chem. 2011, 2933 – 2939; e) C. S.
Ljpez, O. N. Faza, K. S. Feldman, M. R. Iyer, D. K. Hester, J. Am. Chem.
Soc. 2007, 129, 7638 – 7646.
a) C. Wentrup, G. Gross, A. Maquestiau, R. Flammang, Angew. Chem. Int.
Ed. Engl. 1983, 22, 542 – 543; Angew. Chem. 1983, 95, 551 – 551; b) W. R.
Moore, W. R. Moser, J. Am. Chem. Soc. 1970, 92, 5469 – 5474; c) I. Quintana, D. PeÇa, D. P8rez, E. Guiti#n, Eur. J. Org. Chem. 2009, 5519 – 5524;
d) G. Wittig, P. Fritze, Angew. Chem. Int. Ed. Engl. 1966, 5, 846; Angew.
Chem. 1966, 78, 905; e) J. D. Price, R. P. Johnson, Tetrahedron Lett. 1986,
27, 4679 – 4682; f) E. T. Marquis, P. D. Gardner, Tetrahedron Lett. 1966, 7,
2793 – 2798.
a) N. J. Agard, J. A. Prescher, C. R. Bertozzi, J. Am. Chem. Soc. 2004, 126,
15046 – 15047; b) S. T. Laughlin, J. M. Baskin, S. L. Amacher, C. R. Bertozzi,
Science 2008, 320, 664 – 667; c) M. F. Debets, C. W. J. van der Doelen,
F. P. J. T. Rutjes, F. L. van Delft, ChemBioChem 2010, 11, 1168 – 1184.
a) R. Huisgen, Angew. Chem. Int. Ed. Engl. 1963, 2, 565 – 598; Angew.
Chem. 1963, 75, 604 – 637; b) F. Himo, T. Lovell, R. Hilgraf, V. V. Rostovtsev, L. Noodleman, K. B. Sharpless, V. V. Fokin, J. Am. Chem. Soc. 2005,
127, 210 – 216; c) V. V. Rostovtsev, L. G. Green, V. V. Fokin, K. B. Sharpless,
Angew. Chem. Int. Ed. 2002, 41, 2596 – 2599; Angew. Chem. 2002, 114,
2708 – 2711; d) C. W. Tornøe, C. Christensen, M. Meldal, J. Org. Chem.
2002, 67, 3057 – 3064.
a) P. Vermeeren, S. C. C. van der Lubbe, C. Fonseca Guerra, F. M. Bickelhaupt, T. A. Hamlin, Nat. Protoc. 2020, 15, 649 – 667; b) F. M. Bickelhaupt,
K. N. Houk, Angew. Chem. Int. Ed. 2017, 56, 10070 – 10086; Angew. Chem.
2017, 129, 10204 – 10221; c) D. H. Ess, K. N. Houk, J. Am. Chem. Soc.
2008, 130, 10187 – 10198; d) L. P. Wolters, F. M. Bickelhaupt, Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2015, 5, 324 – 343; e) I. Fern#ndez, F. M.
Bickelhaupt, Chem. Soc. Rev. 2014, 43, 4953 – 4967; f) W.-J. van Zeist,
F. M. Bickelhaupt, Org. Biomol. Chem. 2010, 8, 3118 – 3127.
a) F. M. Bickelhaupt, E. J. Baerends in Reviews in Computational Chemistry
(Eds.: K. B. Lipkowitz, D. B. Boyd), Wiley, Hoboken, 2000, pp. 1 – 86; b) R.
Chem. Eur. J. 2020, 26, 11529 – 11539
www.chemeurj.org
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
van Meer, O. V. Gritsenko, E. J. Baerends, J. Chem. Theory Comput. 2014,
10, 4432 – 4441; c) L. Zhao, M. von Hopffgarten, D. M. Andrada, G. Frenking, WIREs Comput. Mol. Sci. 2018, 8, e1345.
a) B. J. Levandowski, T. A. Hamlin, F. M. Bickelhaupt, K. N. Houk, J. Org.
Chem. 2017, 82, 8668 – 8675; b) T. A. Hamlin, B. J. Levandowski, A. K.
Narsaria, K. N. Houk, F. M. Bickelhaupt, Chem. Eur. J. 2019, 25, 6342 –
6348.
a) G. 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 –
967; b) C. Fonseca Guerra, J. G. Snijders, G. te Velde, E. J. Baerends,
Theor. Chem. Acc. 1998, 99, 391 – 403; c) ADF, SCM Theoretical Chemistry; Vrije Universiteit, Amsterdam (The Netherlands), 2017; http://
www.scm.com.
a) A. D. Becke, Phys. Rev. A 1988, 38, 3098 – 3100; b) J. P. Perdew, Phys.
Rev. B 1986, 33, 8822 – 8824.
E. van Lenthe, E. J. Baerends, J. Comput. Chem. 2003, 24, 1142 – 1156.
a) A. Talbot, D. Devarajan, S. J. Gustafson, I. Fern#ndez, F. M. Bickelhaupt,
D. H. Ess, J. Org. Chem. 2015, 80, 548 – 558; b) T. A. Hamlin, D. Svatunek,
S. Yu, L. Ridder, I. Infante, L. Visscher, F. M. Bickelhaupt, Eur. J. Org.
Chem. 2019, 378 – 386; c) S. Yu, H. M. de Bruijn, D. Svatunek, T. A.
Hamlin, F. M. Bickelhaupt, ChemistryOpen 2018, 7, 995 – 1004.
a) S. Grimme, S. Ehrlich, L. Goerigk, J. Comput. Chem. 2011, 32, 1456 –
1465; b) J. G. Brandenburg, J. E. Bates, J. Sun, J. P. Perdew, Phys. Rev. B
2016, 94, 115144.
a) Y. Zhao, D. G. Truhlar, J. Chem. Phys. 2006, 125, 194101; b) Y. Zhao,
D. G. Truhlar, Theor. Chem. Acc. 2008, 120, 215 – 241.
a) A. Klamt, G. Scheermann, J. Chem. Soc. Perkin Trans. 2 1993, 799 –
805; b) A. Klamt, J. Phys. Chem. 1995, 99, 2224 – 2235; c) A. Klamt, V.
Jonas, J. Chem. Phys. 1996, 105, 9972; d) C. C. Pye, T. Ziegler, Theor.
Chem. Acc. 1999, 101, 396 – 408.
a) W.-J. van Zeist, C. Fonseca Guerra, F. M. Bickelhaupt, J. Comput. Chem.
2008, 29, 312 – 315; b) X. Sun, T. M. Soini, J. Poater, T. A. Hamlin, F. M.
Bickelhaupt, J. Comput. Chem. 2019, 40, 2227 – 2233.
C. Y. Legault, CYLview, 1.0b; Universit8 de Sherbrooke, Sherbrooke (QC,
Canada), 2009; http://www.cylview.org.
D. Svatunek, N. Houszka, T. A. Hamlin, F. M. Bickelhaupt, H. Mikula,
Chem. Eur. J. 2019, 25, 754 – 758.
T. A. Albright, J. K. Burdett, M.-H. Whangbo, Orbital Interactions in
Chemistry, 2nd ed., Wiley, Hoboken, 2013.
C. Fonseca Guerra, J.-W. Handgraaf, E. J. Baerends, F. M. Bickelhaupt, J.
Comput. Chem. 2004, 25, 189 – 210.
Y.-R. Luo, Comprehensive Handbook of Chemical Bond Energies, CRC
Press, Boca Raton, 2007.
G. S. Hammond, J. Am. Chem. Soc. 1955, 77, 334 – 338.
Manuscript received: February 17, 2020
Revised manuscript received: March 25, 2020
Accepted manuscript online: March 27, 2020
Version of record online: August 7, 2020
11539
T 2020 The Authors. Published by Wiley-VCH GmbH