You are on page 1of 21

Research on Chemical Intermediates (2023) 49:3205–3225

https://doi.org/10.1007/s11164-023-05031-5

MP2, DFT, and IQA study of substituent effect


on the structure, stability, and bonding properties of ­CX2
singlet and triplet carbenes and related carbenoids

Zahrasadat Emami‑Meibodi1 · Hossein Tavakol1   · Kiamars Eskandari1

Received: 20 March 2023 / Accepted: 8 May 2023 / Published online: 22 May 2023
© The Author(s), under exclusive licence to Springer Nature B.V. 2023

Abstract
During this study, six carbenes, consisted of various substitutions (hydrogen, deute-
rium, fluorine, chlorine, bromine, and methyl) and their 24 related carbenoids (using
Li, Na, Be, and Mg metals) were designed. These species were theoretically stud-
ied to obtain the extensive and comprehensive information about their structures,
stabilities, atomic specifications, and bonding properties using MP2/aug-cc-pVTZ
level of theory. Moreover, the PBE1PBE DFT method was used for IQA analyses
using AIMAll program package. The calculated molecular parameters showed that
the electronegativity and the size of ligand are effective on the studied structures.
Moreover, the electronegativity effect is more important than the size. Atomic
hybridizations results showed the p indexes of carbon in triplet carbenes are also
smaller than those in singlet carbenes, but this difference in halogen-containing car-
benes is smaller than the other carbenes. In population analyses, except for sodium-
based carbenoids, all carbenoids have higher Eg values than the carbenes. The ΔG
values for α-elimination reaction, as a method of preparation of these carbenes, were
obtained in order of f < c < b < m < h, which is reversely related to the electronegativ-
ities of the connected ligands. IQA analyses were performed to evaluate the relative
stability of carbenes. It was found that the classical interaction in C–F is attractive
(negative) unlike the other mentioned bonds energy for carbenes. This electrostatic
term in C–F is larger in the singlet state than the triplet state, which leads to the sin-
glet state of ­CF2 being more stable and consequently more favorable than its triplet
state.

* Hossein Tavakol
h_tavakol@iut.ac.ir
Zahrasadat Emami‑Meibodi
Zemami@ch.iut.ac.ir
Kiamars Eskandari
k.eskandari@iut.ac.ir
1
Department of Chemistry, Isfahan University of Technology, Isfahan 84156‑83111, Iran

13
Vol.:(0123456789)
3206 Z. Emami‑Meibodi et al.

Keywords  Carbene · Carbenoids · Singlet · Triplet · IQA · DFT

Introduction

As an important member of reactive intermediates, the chemistry of carbene has


been an interesting topic in many experimental and theoretical studies [1]. This
class of compounds has exclusive properties and chemical behavior, which make
it an interesting structure for synthesizing useful compounds as well as studying
the fundamentals of organic structure and reactivity [2–4]. Singlet and triplet car-
benes, carbenoids, and N-heterocyclic carbenes are different categories of this class
[5–8]. The structure, reactivity, and stability of each category are different from the
others and should be studied separately. Many parts of the fundamental aspects of
carbene (all classes) have been investigated since the first report on the spectro-
scopic observation of a carbene [9]. Some of the most interesting properties of these
structures are the differences between singlet and triplet carbenes, the singlet–triple
energy gap (EST), and the structures and properties of carbenoids [10]. Some car-
benes have small EST values (less than 2  kcal/mol), such as 9-fluorenylidene with
only a 1.1 kcal/mol singlet–triple energy gap [11]. Theoretical studies showed that
the electropositive substituents could stabilize the triplet carbenes [12]. The ES–T
is equal to the electron–electron repulsion minus the energy gap between σ and pπ
orbitals (σ–pπ gap). By increasing the σ–pπ gap, more energy is needed to overcome
the electron–electron repulsion for preparing the triplet state. If any electronic or
spatial effects reduce the energy difference between these orbitals, the triplet state
is the preferred state [13, 14]. Harrison also confirmed that by changing the car-
bene substituent from silicon to fluorine, the stability of singlet carbene is greatly
increased [15]. Moreover, both resonance and the size of the substituent could
change the structure of carbene and EST value. The bigger substituent increases the
central angle of carbene and stabilizes the triplet state [16]. In the course of this
topic, several theoretical or combined theoretical–experimental studies have been
performed to clarify some aspects of this topic [17–22]. In addition, a few reports
on the calculation of the singlet–triplet energy gap have been reported many years
ago [23–26]. The importance of the structure of carbene will be revealed more when
we know its effects on the reactivity and chemical behavior. In a reported study,
the relationship between the nucleophilicity of cyclic amino carbene with its struc-
ture has been examined [27]. Moreover, the potency of carbene as a ligand could be
affected by changing its substituents [28, 29].
Beside carbenes, carbenoids or metal carbenes are special species with similar
properties to carbenes [30]. Despite several works on their reactivities [31–35], the
fundamental works on the structure of carbenoids are limited [36–39]. There are
several reports on the study of simple lithium carbenoids ­(CH2LiX) [40–45] and
the structures and properties of these carbenes studied well. Moreover, the study of
halogenated lithium carbenoids (­CCl3Li) was reported [46, 47] and there are some
more comprehensive studies [48–50] and one review [51] related to these struc-
tures. However, the lack of a more comprehensive work, with considering various
substituents and metal, is obvious. Therefore, based on our experiences in the area

13
MP2, DFT, and IQA study of substituent effect on the structure,… 3207

L
Carbenes (singlet and triplet): C
L
Cl L L
Carbenoids (singlet and triplet): C M' C
M L Cl L

L: H, D, Me, F, Cl, Br

M (group I elements): Li, Na M' (group II elements): Be, Mg

Scheme 1  The general chemical structures of the studied molecules

of computational chemistry, it seems necessary to perform a complete study on


the structure and bonding properties of carbenes and carbenoids. In this line, the
effects of substituents, NBO, QTAIM, and IQA analysis of these properties should
be worked out carefully.

Methods

Six carbenes ­(CR2) with different substituents ­(CH2, ­CD2, ­CMe2, ­CF2, ­CCl2, and
­CBr2) have been considered in this study to consider isotope, substituent, and spa-
tial effects for carbenes. In addition, four different carbenoids have been considered
involving the first two metals of groups I and II (Li, Na, Be, and Mg). To com-
plete the capacity of metal in group II carbenoids, one chlorine atom was added. The
choice of these structures was based on the maximum possibilities with the mini-
mum calculations and therefore, the heavy substituents (iodine, large alkyl groups,
…) or metal have not been used. The general chemical structures of the studied mol-
ecules are shown in Scheme 1.
The optimizations of geometries, single point energies, frequency calculations,
NBO calculations, generating wave functions, and population analyses have been
performed using the Gaussian 09 program package [52]. In this line, MP2/aug-cc-
pVTZ basis set level of theory has been employed to obtain the energies, molecu-
lar parameters, charges, and hybridizations. IQA calculations were performed by
PBE1PBE method combined with aug-cc-pVTZ basis set using AIMAll program
package [53]. It is noticeable that since the IQA calculations are available only for a
few methods [53] (The MP2 method is not available in this program), we performed
the IQA analyses on the PBE0 wavefunctions. The compatibility between IQA and
PBE1PBE ensures that the total molecular energy obtained with the IQA terms
is very close to the original energy calculated with Gaussian 09. The difference
between IQA-based and the original total energies was smaller than 1.0 kJ/mol. The
optimizations of all structures were performed without symmetry restriction or pre-
defined conformational structures. The absence of imaginary frequency verified that
a structure was a real minimum at its respective level of theory. All possible isomers
and structures were searched to be sure that each structure is a global minimum.
GaussView 5.0 program was employed for the visualization of the results (optimized

13
3208 Z. Emami‑Meibodi et al.

structures and molecular orbitals) [54]. To calculate chemical potential (µ), chemical
hardness (g), and electrophilicity index (ω), the Koopman’s theorem [55] has been
employed according to Eqs. 1–3 [56]:
( )
ELUMO + EHOMO
𝜇= (1)
2
( )
ELUMO − EHOMO
𝜂= (2)
2

𝜇2
𝜔= (3)
2𝜂
In the IQA calculations, we will use the advantages of the IQA approach to fig-
ure out what makes a carbene stable or unstable. In this part, only the singlet and
triplet structures of CX2 (X = H, F, Cl, and Br) are considered. In the IQA approach
[57–61], the total electronic energy of a molecule can be written as
∑ ∑∑
Emol (IQA) = Eself (A) + Eint (A, B)
A A B>A

In which Eself (A) and Eint (A, B) are, respectively, atomic self-energy of atom A
and interatomic interaction energies between A and B atoms. Eint (A, B) can be par-
titioned into classical electrostatic, Vcl (A, B) , and exchange–correlation, VXC (A, B) ,
components:
Eint (A, B) = Vcl (A, B) + VXC (A, B)

Results and discussion

Optimized structures

The structures of all carbenes and carbenoids (total of 30 structures) were first opti-
mized to find a global minimum in singlet and triplet states. The graphical presenta-
tion of the optimized structures for carbenes is shown in Fig. S1 and for carbenoids
in Figs. 1, 2, 3, and 4 (each figure for one type of carbenoids). The related optimized
parameters for carbenes and carbenoids are listed in Tables  1 and 2, respectively.
The brief names were also provided for all structures. For carbenes, each brief name
includes two parts: a letter representing the carbene elements and a number repre-
senting the multiplicity of carbene (h for C ­ H2, d for C­ D2, m for C­ Me2, f for C­ F2, c
for ­CCl2, b for C­ Br2, 1 refers to the singlet, and 3 refers to the triplet states). In the
studied carbenoids, the symbol of the metal was placed between the letter and num-
bers. For example, HLi1 refers to the singlet state of the lithium carbenoid of C ­ H2
(h).

13
MP2, DFT, and IQA study of substituent effect on the structure,… 3209

Fig. 1  The optimized structures of singlet and triplet lithium carbenoids obtained from MP2/aug-cc-
pVTZ calculations

Fig. 2  The optimized structures of singlet and triplet sodium carbenoids obtained from MP2/aug-cc-
pVTZ calculations

13
3210 Z. Emami‑Meibodi et al.

Fig. 3  The optimized structures of singlet and triplet beryllium carbenoids obtained from MP2/aug-cc-
pVTZ calculations

In all carbenes (singlet and triplet), the value of the C–L bond lengths is
equal to the normal value in similar compounds. In singlet carbenes, the C–L
bond length in four carbenes is 0.00–0.06 Å higher than that in triplet carbenes.
However, in C ­ F2 and C­ Me2 carbenes, both singlet and triplet have similar C–L
bond lengths. These differences are related to the difference in the hybridization
of singlet and triplet carbenes. As we know, singlet carbenes are sp2-hybridized,
which have less s-character than the triplet carbenes (sp-hybridized). The less
s-character of the bond is the reason for the higher bond length. A similar expla-
nation could be said for L–C–L bond angles, which have a higher value in triplet
carbenes than the singlet carbenes. This angle in each triplet carbene is 15–33
degrees higher than the similar angle in singlet carbene. The highest L–C–L
angles (33 degrees) are obtained in ­CH2 and ­CD2 carbenes, which have the least
electronegative atom (hydrogen). The smallest L–C–L angles are obtained in
­CF2 (15 degrees) and C ­ Cl2 (19 degrees), which have the highest electronegative
atoms.
In carbenoids, for L=H, D, and F, the C–L bond lengths are similar or smaller
than the related carbenes, while in the remained carbenoids, the C–L bond lengths

13
MP2, DFT, and IQA study of substituent effect on the structure,… 3211

Fig. 4  The optimized structures of singlet and triplet magnesium carbenoids obtained from MP2/aug-cc-
pVTZ calculations

Table 1  The optimized parameters for singlet and triplet carbenes obtained from MP2/aug-cc-pVTZ cal-
culations
Singlet Triplet S-T difference
C–L (Å) L–C–L C–L (Å) L–C–L C–L (Å) L–C–L (degrees)
(degrees) (degrees)

CH2 (h1, h3) 1.111 102 1.078 135 0.033 −33


CD2 (d1, d3) 1.111 102 1.080 135 0.031 −33
CMe2 (m1, m3) 1.460 113 1.460 134 0.000 −21
CF2 (f1, f3) 1.300 105 1.308 119 −0.008 −15
CCl2 (c1, c3) 1.713 109 1.667 128 0.046 −19
CBr2 (b1, b3) 1.878 110 1.822 130 0.056 −20

The brief names were mentioned in parentheses, where 1 refers to the singlet and 3 refers to the triplet
states

are larger than the related carbenes. The differences between C–L bond lengths in
singlet and triplet carbenoids are small and nearly similar to the related carbenes.
The L–C–L bond angles in both singlet and triplet carbenoids are smaller than
the relative carbenes insomuch carbenoid has connected to four atoms and their
structures are tended to tetrahedron shape. This similarity in the structures of sin-
glet and triplet carbenoids decreases the difference between them (L–C–L bond
angle in singlet and triplet carbenes). The singlet–triplet difference in L–C–L

13
Table 2  The optimized parameters for singlet and triplet carbenoids, obtained from MP2/aug-cc-pVTZ calculations
3212

Singlet Triplet S–T difference


C–L (Å) C–M (Å) L–C–L C–L (Å) C–M (Å) L–C–L C–L (Å) C–M (Å) L–C–L (degrees)

13
(degrees) (degrees)

H2CLiCl 1.092 1.930 107 1.080 3.590 125 0.012 −1.660 −18
D2CLiCl 1.092 1.931 107 1.078 3.590 125 0.014 −1.659 −18
Me2CLiCl 1.503 1.955 111 1.480 3.121 124 0.023 −1.166 −14
F2CLiCl 1.301 2.030 106 1.314 5.230 111 −0.013 −3.200 −5
Cl2CLiCl 1.763 1.960 109 1.705 4.461 117 0.058 −2.501 −8
Br2CLiCl 2.082 1.926 103 1.960 2.160 124 0.122 −0.234 −21
H2CNaCl 1.093 2.300 107 1.080 4.506 125 0.013 −2.206 −18
D2CNaCl 1.093 2.290 107 1.080 4.510 125 0.013 −2.220 −18
Me2CNaCl 1.505 2.320 110 1.480 3.716 123 0.025 −1.396 −13
F2CNaCl 1.311 2.405 105 1.314 5.443 111 −0.003 −3.038 −5
Cl2CNaCl 1.788 2.316 108 1.705 4.635 117 0.083 −2.319 −10
Br2CNaCl 2.088 2.324 103 1.976 2.620 127 0.112 −0.296 −24
H2CBeCl2 1.090 1.680 108 1.080 2.005 136 0.010 −0.325 −27
D2CBeCl2 1.090 1.680 108 1.078 2.000 136 0.012 −0.320 −27
Me2CBeCl2 1.500 1.690 114 1.463 2.011 133 0.037 −0.321 −19
F2CBeCl2 1.276 1.800 109 1.315 1.973 115 −0.039 −0.173 −6
Cl2CBeCl2 1.783 1.700 110 1.674 2.000 126 0.109 −0.300 −16
Br2CBeCl2 1.954 1.690 110 1.846 1.930 118 0.108 −0.240 −9
H2CMgCl2 1.100 2.064 109 1.080 2.442 140 0.020 −0.378 −31
D2CMgCl2 1.090 2.064 109 1.080 2.442 140 0.010 −0.378 −31
Me2CMgCl2 1.501 2.092 112 1.460 2.430 136 0.041 −0.338 −23
F2CMgCl2 1.282 2.245 108 1.340 2.040 108 −0.058 0.205 0
Z. Emami‑Meibodi et al.
Table 2  (continued)
Singlet Triplet S–T difference
C–L (Å) C–M (Å) L–C–L C–L (Å) C–M (Å) L–C–L C–L (Å) C–M (Å) L–C–L (degrees)
(degrees) (degrees)

Cl2CMgCl2 1.792 2.091 109 1.870 2.350 118 −0.078 −0.259 −9


Br2CMgCl2 1.860 2.202 105 1.870 2.350 118 −0.010 −0.148 −13
MP2, DFT, and IQA study of substituent effect on the structure,…
3213

13
3214 Z. Emami‑Meibodi et al.

bond angle in carbenoids is in the range of 0–31 degrees, while in the carbenes
it was between 15 and 33 degrees. The carbon–metal (C–M) distance is also con-
sidered in carbenoids. Interestingly, there are major differences between this dis-
tance in singlet and triples carbenes and mostly, the triplet carbenes have larger
angles (up to 3.2 Å). In fact, in some triplet carbenoids, the structure is similar to
­CL3 anion plus metal cation. The extreme limit of this observation was observed
in the carbenoid involved with fluorine as a ligand and group I elements as a
metal (Li and Na). Moreover, in Br-containing carbenoids, the least difference in
C–D distance of single and triplet carbenes is observed and this value in carbenes
with D, H, and methyl as ligands are nearly the same. By considering all struc-
tural parameters, we could categorize carbenes and carbenoids into three different
groups. First, carbenes and carbenoids contain highly electronegative ligands (Cl
and F), especially fluorine, which have large differences with the other carbenes.
Second, ­CH2 and C ­ D2 carbenes (simple carbenes) and their related carbenoids
have similar structures with each other. Third, ­CMe2 and ­CBr2 carbenes and car-
benoids (large ligands) have larger differences with the first group and smaller
differences with the second group. Therefore, both electronegativity and the size
of the ligand are effective on the structures of carbenes and carbenoids, and the
effect of electronegativity is more than the size effect.

NBO calculations

NBO analysis was employed to find the atomic charges and the hybridizations of
important elements. The related results (atomic charges and hybridization p index)
for all carbenes and carbenoids are listed in Table 3. The reported hybridization p
index is the power of p orbital (i) in spi in bonding between carbene atom and ligand
(L). This value shows the relative shares of p orbitals versus s orbital. Only the
charges of carbonic carbon, ligand (L in C ­ L2 or related carbenoids), and metal are
reported. In singlet carbenes, the charge of carbon atoms in ­CF2 and ­CMe2 are posi-
tive (0.61 and 0.12, respectively), while in the other carbenes, this value is negative.
The high positive charge of carbon in ­CF2 is acceptable since the high electronega-
tivity of the fluorine atom induce the positive charge on carbon. The negative charge
in ­CMe2 carbene is placed on the methyl group, instead of carbene atom. Therefore,
based on the calculated atomic charges, C ­ F2 and C­ Me2 can be considered as elec-
trophilic carbenes and the other carbenes have nucleophilic nature. The charges in
triplet carbenes are nearly similar to those in singlet carbenes with small differences
(0.00–0.05 a.u.), except the charge of carbon atoms in ­CH2 and ­CD2 carbenes. In
these two carbenes, the charge of the carbon atoms in triplet carbene is 0.11 more
negative than that in singlet carbene. The reason of this phenomenon can be under-
stood by comparing p indexes, which in the triplet state of these two carbenes, the
carbon atom has a very smaller p character versus the singlet carbenes (1.99 in tri-
plet and 4.04 in singlet). The smaller p character led to higher electronegativity and
more negative charge. In singlet carbenoids, the charge of carbon atom is tended to
the more negative values and this tendency is higher in singlet carbenes. The largest
effects were observed in carbenoids with D, H, and Me ligands and the least effect

13
Table 3  Atomic charges and hybridization p index in singlet and triplet carbenes and carbenoids obtained, obtained from MP2/aug-cc-pVTZ calculations
Atomic charges in singlet states Atomic charges in triplet states Hybrid. P index
Atom Carbene Li carbe- Na carbe- Be carbe- Mg carbe- Carbene Li carbe- Na carbe- Be carbe- Mg carbe- Singlet Triplet
noid noid noid noid noid noid noid noid carbene carbene

CH2 C −0.13 −0.95 −0.93 −0.99 −0.96 −0.24 −0.40 −0.41 −0.32 −0.34 4.04 1.99
H 0.07 0.19 0.18 0.24 0.21 0.12 0.20 0.19 0.19 0.18 0.00 0.00
M 0.79 0.82 0.98 1.34 −0.02 −0.01 0.57 1.24
CD2 C −0.13 −0.95 −0.93 −0.99 −0.96 −0.24 −0.40 −0.41 −0.32 −0.34 4.04 1.99
D 0.07 0.19 0.18 0.24 0.21 0.12 0.20 0.19 0.19 0.18 0.00 0.00
M 0.79 0.82 0.98 1.34 −0.02 −0.01 0.57 1.24
CMe2 C(1) 0.12 −0.41 −0.37 −0.34 −0.44 0.19 0.03 0.02 0.08 0.08 2.47 1.84
C(2Me) −0.77 −0.68 −0.69 −0.66 −0.68 −0.77 −0.70 −0.70 −0.75 −0.77 2.03 2.21
M 0.80 0.81 0.78 1.36 −0.03 −0.01 0.64 1.26
MP2, DFT, and IQA study of substituent effect on the structure,…

CF2 C 0.61 0.45 0.41 0.69 0.54 0.61 0.60 0.60 0.53 0.14 5.04 3.84
F −0.31 −0.29 −0.31 −0.23 −0.25 −0.31 −0.29 −0.29 −0.28 −0.34 2.67 2.90
M 0.73 0.85 0.47 1.21 0.00 0.00 0.55 1.13
CCl2 C −0.18 −0.60 −0.60 −0.73 −0.77 −0.20 −0.29 −0.29 −0.37 −0.41 3.52 1.10
Cl 0.09 0.01 −0.04 0.04 −0.69 0.10 0.10 0.10 0.18 0.15 2.64 5.05
M 0.83 0.89 1.04 1.46 0.00 0.00 0.61 1.27
CBr2 C −0.35 −0.72 −0.69 −0.91 −0.57 −0.40 −0.50 −0.48 −0.53 −0.55 4.31 4.09
Br 0.17 −0.05 0.10 0.12 −0.10 0.20 −0.15 −0.21 0.08 −0.08 2.81 5.36
M 0.76 0.86 1.03 1.23 0.70 0.85 0.54 1.19
3215

13
3216 Z. Emami‑Meibodi et al.

is belong to carbenoids with fluorine ligands. The positive charge of metal in carbe-
noids is mostly in this order: Na > Li > Mg > Be (the charges of two-valence cations
are divided into 2), which are directly related to their electropositivities. However,
in triplet carbenoids, their charges highly tended to the more negative values, espe-
cially for Li and Na carbenoids. In group-I metal triplet carbenoids, the charge of
metal (except for L=Br) is negative (near zero) while in the others, the charges are
still positive. The p indexes of carbon in triplet carbenes are also smaller than those
in singlet carbenes, but this difference in halogen-containing carbenes is smaller
than the other carbenes.

Population analysis

Population analysis is useful to obtain the energies of frontier orbitals, band gaps,
and the relative parameters to get a new insight into the properties of chemical com-
pounds. These data were extracted from the outputs of calculations, which are listed
in Table  4, and the graphical presentation of HOMOs and LUMOs for all singlet
and triplet carbenes and carbenoids is shown in Figs. S2–S6 (in supporting infor-
mation). Based on the presented data, in halogenated carbenes, the HOMO energy
decreases from f1 to b1 (f1 > c1 > b1). The values of these energies are − 0.32, − 0.27,
and − 0.270 eV, respectively, for f1, c1, and b1. The energy of LUMO increases from
f1 to b1 and its energies are f1 =  − 0.08, c1 = −0.12, and b1 =  − 0.12 eV. The energy
difference (Eg) also decreases from f1 to b1 , and the energy difference values are
f1 = 0.23, c1 = 0.12, and b1 = 0.13 eV.
The chemical potentials for all the halogenated carbenes are almost close to each
other and have a slight difference. The chemical potential for all of them is 0.20 eV.
The degree of hardness in f1 to b1 has a decreasing trend. The hardness values are
f1 = 0.11, c1 = 0.07, and b1 = 0.06 eV. Since hardness index has a direct relationship
with stability, we conclude that f1 carbenes are more stable than c1 and b1. The soft-
ness criterion increases from f1 to b1, and the softness values in f1, c1, and b1 are
8.35, 12.98, and 14.14 eV, respectively. The mentioned estimation for the stability of
the compounds is also confirmed by the electrophilicity index. The electrophilicity
indexes (ω) for f1, c1 and b1 are 0.17, 0.26, and 0.29 eV, respectively. Considering
the fact that the electrophilicity index has an inverse relationship with the stability,
the high activity of f1 compared to c1 and b1 is confirmed. In d1 and h1, the energies
of HOMO and LUMO are the same (− 0.27 and − 0.13 eV, respectively). The energy
gap (HOMO–LUMO band gap, Eg) in h1 and d1 has the same value, − 0.13 eV. All
other parameters are also the same for h1 and d1. In lithium-containing halogen
carbenoids, the HOMO’s energy decreases from fli1 to bli1 (fli1 > cli1 > bli1), and
the order of the LUMO’s energy decreases is fli1 > cli1 < bli1. The Eg values also
decrease from fli1 to bli1 and the degree of hardness in fli1 to bli1 has a decreasing
trend. Since the hardness index has a direct relationship with stability, it will be con-
cluded that fli1 carbenoids are the most stable, and bli1 is the least stable.
By comparing the Eg values of singlet carbenes and carbenoids, except for
sodium-based carbenoids, all carbenoids have higher Eg values than the carbenes
in this order: Be-based carbenoids > Mg-based carbenoids > Li-based carbenoids.

13
MP2, DFT, and IQA study of substituent effect on the structure,… 3217

Table 4  The HOMO–LUMO band gap, chemical potential, hardness and electrophilicity indexes for all
structures, obtained from MP2/aug-cc-pVTZ calculations
Singlet Triplet
Eg (eV) μ (eV) η (eV) ω (eV) Eg (eV) μ (eV) η (eV) ω (eV)

CH2 0.139 −0.201 0.069 0.292 0.263 −0.133 0.132 0.067


CD2 0.139 −0.201 0.069 0.292 0.263 −0.133 0.132 0.067
CMe2 0.152 −0.137 0.076 0.124 0.194 −0.100 0.097 0.052
CF2 0.239 −0.205 0.120 0.176 0.250 −0.103 0.125 0.043
CCl2 0.154 −0.202 0.077 0.266 0.220 −0.119 0.110 0.064
CBr2 0.139 −0.201 0.069 0.292 0.187 −0.135 0.094 0.097
H2CLiCl 0.177 −0.122 0.089 0.085 0.087 −0.080 0.044 0.073
D2CLiCl 0.177 −0.122 0.089 0.085 0.087 −0.080 0.044 0.073
Me2CLiCl 0.165 −0.116 0.083 0.081 0.091 −0.068 0.045 0.050
F2CLiCl 0.211 −0.168 0.105 0.134 0.110 −0.086 0.055 0.068
Cl2CLiCl 0.191 −0.145 0.095 0.110 0.100 −0.089 0.050 0.080
Br2CLiCl 0.173 −0.150 0.086 0.130 0.194 −0.136 0.097 0.095
H2CNaCl 0.133 −0.116 0.066 0.102 0.106 −0.074 0.053 0.052
D2CNaCl 0.133 −0.116 0.066 0.102 0.106 −0.074 0.053 0.052
Me2CNaCl 0.122 −0.112 0.061 0.103 0.102 −0.069 0.051 0.047
F2CNaCl 0.195 −0.151 0.098 0.116 0.111 −0.077 0.055 0.054
Cl2CNaCl 0.150 −0.139 0.075 0.129 0.092 −0.085 0.046 0.079
Br2CNaCl 0.157 −0.141 0.079 0.126 0.161 −0.135 0.080 0.113
H2CBeCl2 0.268 −0.162 0.134 0.098 0.299 −0.163 0.149 0.089
D2CBeCl2 0.268 −0.162 0.134 0.098 0.299 −0.163 0.149 0.089
Me2CBeCl2 0.238 −0.166 0.119 0.115 0.233 −0.132 0.117 0.075
F2CBeCl2 0.185 −0.228 0.092 0.282 0.249 −0.144 0.125 0.083
Cl2CBeCl2 0.261 −0.177 0.131 0.120 0.236 −0.149 0.118 0.094
Br2CBeCl2 0.225 −0.184 0.112 0.151 0.216 −0.155 0.108 0.112
H2CMgCl2 0.230 −0.161 0.115 0.113 0.279 −0.178 0.140 0.114
D2CMgCl2 0.230 −0.161 0.115 0.113 0.279 −0.178 0.140 0.114
Me2CMgCl2 0.217 −0.151 0.109 0.105 0.226 −0.142 0.113 0.090
F2CMgCl2 0.189 −0.220 0.095 0.255 0.157 −0.122 0.079 0.095
Cl2CMgCl2 0.207 −0.177 0.104 0.151 0.135 −0.146 0.068 0.158
Br2CMgCl2 0.136 −0.228 0.068 0.381 0.209 −0.161 0.104 0.124

Surely, all the other parameters, which are related to the Eg values, have relative
orders accordingly. The Na-based carbenoids have the smallest Eg values, even ver-
sus the carbenes.

Energy values

Since an important energetic aspect of carbenes and carbenoids is the singlet–tri-


ple energy gap (ΔES–T = E(singlet) − E(triplet)), these values were calculated (from

13
3218 Z. Emami‑Meibodi et al.

Table 5  Singlet–triplet energy gaps (in kcal/mol) for all structures, obtained from MP2/aug-cc-pVTZ
calculations. ΔES–T = E(singlet)−E(triplet)
ΔES–T Carbene Li carbenoid Na carbenoid Be carbenoid Mg carbenoid

CH2 (h) 30.8 −47.8 −27.1 −41.1 −28.6


CD2 (d) 17.4 −47.9 −27.0 −41.6 −32.8
CMe2 (m) 4.2 −39.7 −20.3 −29.3 −18.5
CF2 (f) −47.9 −64.8 −46.3 −58.1 −57.9
CCl2 (c) −13.9 −53.1 −36.4 −28.5 −27.9
CBr2 (b) −10.2 −20.1 −19.2 −28.9 −25.2

MP2/aug-cc-pVTZ level of theory) and are listed in Table 5. For carbenes, this gap
obeys from this order: h > d > m > b > c > f, and this value for h, d, and m is positive.
This means that in these carbenes, the triplet state is more stable than the singlet
state. Obviously, all halogen-containing carbenes have a more stable singlet state
and the absolute value of ΔES–T gap increases with the increase of halogen’s elec-
tronegativity or with the decrease of the atomic radius of halogen. Interestingly,
there is a large difference between the ΔES–T gap of ­CH2 and C ­ D2, while there is
no difference in the electron configuration of hydrogen and deuterium. This gap is
highly reduced in their related carbenoids. In all carbenoids, the triplet state is more
stable than the singlet states in high values (18.5–64.8 kcal/mol). The order of ΔES–T
gap’s magnitude in average for carbenoids is f > d≈h≈c > m > b. This order is nearly
the same with hardness, so it can be said that by the increase of hardness, the magni-
tude of the ΔES–T gap increases.
In addition to the ΔES–T gap, the ΔG values for the preparation of each carbene
are also interesting for researchers. To obtain reliable data for this purpose, it was
assumed that all of the six studied carbenes were prepared from the α-elimination
reaction between the related haloalkane as a carbene source and t-butoxide as a base.
This reaction is a common reaction for the preparation of these types of carbenes.
Then, the ΔG values for these reactions (ΔGR) were calculated and this value is con-
sidered as the easiness for the preparation of each carbene. The designed reaction
and the obtained values for singlet carbenes are presented in Table 6. Based on the
listed values, this energy for the preparation of ­CF2 is negative, which shows the high

Table 6  ΔG values for the α-elimination reaction for preparation of carbenes (ΔGR), obtained from MP2/
aug-cc-pVTZ calculations

O L2C OH
L2CHCl Cl

L2C CH2 CD2 CMe2 CF2 CCl2 CBr2

ΔGR (kcal/mol) 69.2 56.6 33.0 −0.2 7.1 8.6

13
MP2, DFT, and IQA study of substituent effect on the structure,… 3219

stability of this carbene. Interestingly, the preparation of C


­ D2 carbene is easier than
­CH2 carbene by 12.6 kcal/mol, while the electronic configurations of these carbenes
are the same. It seems that the preparation of ­CD2 carbenes, since the C–D bond is
stronger than C–H bond, the presence of two C–D bonds, make the existed one C–H
bond weaker and H will be more easily removed by the base and the preparation of
carbene is more facilitated. This observation shows the importance of nuclei dis-
tance on the carbene’s stability, in addition to the electronic parameters. The order
of the ΔG values for the preparation of these carbenes is f < c < b < m < h, which is
exactly the same as the order of electronegativities of the connected ligands. In fact,
by increasing the electronegativity, the acidic nature of α-proton increases and it will
be detached more easily by the base (t-butoxide).

IQA analysis of CX2 carbenes

As mentioned in previous sections, carbenes can exist in singlet and triplet states.
Generally, the triplet form of a carbene is more stable, but this is not always the
case. For instance, several reports show that in carbenes of the CX2 form, where X
is a halogen atom, the more stable state is the singlet. First, let us look at the triplet
and singlet states of C
­ H2. In this case, the most stable state is the triplet one. Indeed,
the triplet state is more stable by about 17 kcal/mol than the singlet (Table 7). As
stated, since atomic IQA energies are additive, the sum of atomic energies gives the

Table 7  Molecular energies (in a.u.) of C


­ X2, obtained directly from Gaussian09, Emol (PBE1PBE) , and
IQA approach, Emol (IQA)
CH2 Emol (PBE1PBE) Emol (IQA)

Triplet −39.1156 −39.1156


Singlet −39.0880 −39.0880
Difference (kcal/mol) 17.3 17.3
CF2 Emol (PBE1PBE) Emol (IQA)

Triplet −237.4823 −237.4822


Singlet −237.5598 −237.5598
Difference (kcal/mol) −48.6 −48.7
CCl2 Emol (PBE1PBE) Emol (IQA)

Triplet −958.0782 −958.0782


Singlet −958.1008 −958.1008
Difference (kcal/mol) −14.2 −14.2
CBr2 Emol (PBE1PBE) Emol (IQA)

Triplet −5185.7489 −5185.7490


Singlet −5185.7657 −5185.7657
Difference (kcal/mol) −10.5 −10.5

Difference between singlet and triplet states (defined as Esinglet − ETriplet ) is in kcal/mol

13
3220 Z. Emami‑Meibodi et al.

total energy of the molecule. The total molecular IQA-based energies, Emol (IQA),
reported in Table 7, are in excellent agreement with the energy obtained from SCF
calculations at the same level, indicating the reliability of IQA results. To find the
source of singlet–triplet energy difference, let us take a look at the IQA energy
terms, including Eint (A, B) , Vcl (A, B) , VXC (A, B) , as well as the atomic self-energies,
Eself (A, B).
Interatomic interaction energies and the corresponding classical and
exchange–correlation components of singlet and triplet states of ­CH2 are listed in
Table 8. The carbon–hydrogen interactions, Eint (C, H) , are attractive (negative val-
ues) in both singlet and triplet states. However, the C–H interactions are stronger in
the triplet than the singlet state. The values of VXC (C, H) show that the C–H bond
is predominantly covalent in both the singlet and triplet states, with the covalent
character being slightly more evident in the triplet state. Interestingly, the classical
electrostatic term destabilizes the C–H bond in both states, even though the atomic
charges of carbon and hydrogen atoms are opposite. In fact, the polarization term
causes the electrostatic term to be repulsive. However, the destabilizing role of elec-
trostatics is not large enough to overcome the exchange–correlation term, and hence,
the carbon–hydrogen interatomic interaction energy, Eint (C, H) is stabilizing in both
singlet and triplet states. On the other hand, the hydrogen–hydrogen interatomic
interaction, Eint (H, H) , in the triplet state is positive (repulsive), while it is negative
(attractive) in the singlet form (Table 8). In summary, the C–H interactions tend to
stabilize the triplet state, while the H–H interaction is more favorable in the singlet
state. Nevertheless, the impact of C–H interactions is larger, and the total intera-
tomic interaction is more negative in the triplet state, leading to a stability of about
5.5 kcal/mol compared to the singlet state.
However, this is only part of the picture and additional factors need to be consid-
ered; the atomic self-energies also play a significant role in the energy of molecules.
The self-energies of atoms in both triplet and singlet states are presented in Table 9.
The results show the singlet state is more favorable for hydrogen atoms, while the
triplet state is preferable for carbon atom. With a difference of about 6.0 kcal/mol,
hydrogen atoms are more stable in the singlet state. In contrast, the carbon atom is
more stable in the triplet state, with a stability difference of about 24.0  kcal/mol.
Be that as it may, the total self-energy (the sum of self-energies of both carbon and
hydrogen atoms) favors the triplet state.
As stated, when the hydrogens of ­CH2 are replaced with halogen atoms, the sin-
glet state becomes more favorable (Table 7). The results of IQA partitioning for sin-
glet and triplet states of ­CF2, ­CCl2, and ­CBr2 are presented in Table  8. Similar to
­CH2, in the ­CBr2 and ­CCl2 molecules, the interatomic interaction energies are more
favorable in the triplet state. Compared to their singlet states, the triplet states of
­CBr2 and ­CCl2 exhibit a stronger interatomic interaction (that is, more negative val-
ues) by approximately 26.7 and 4.3  kcal/mol, respectively. Unlike C ­ H2, however,
the atomic self-energies favor the singlet states of C ­ Br2 and ­CCl2. Both carbon and
halogen atoms are more stable in the singlet state, so that the total self-energy favors
the singlet state by around 37.1 and 18.4 kcal/mol in C ­ Br2 and ­CCl2, respectively.
Even though the interatomic interactions in C ­ Br2 and C ­ Cl2 make the triplet state
more stable than the singlet state, the resulting stability does not compensate for the

13
MP2, DFT, and IQA study of substituent effect on the structure,… 3221

Table 8  Interatomic interaction energies (in a.u.) of ­CX2 molecules. Difference between energy compo-
nents of singlet and triplet states (defined as YSinglet − YTriplet , in which Y is VXC , Vcl , or Eint ) are in kcal/
mol
CH2 AB VXC (A, B) Vcl (A, B) Eint (A, B)

Triplet H2 C1 −0.2943 0.0401 −0.2541


H3 C1 −0.2943 0.0401 −0.2541
H3 H2 −0.0050 0.0065 0.0015
Singlet H2 C1 −0.2861 0.0391 −0.2470
H3 C1 −0.2861 0.0391 −0.2470
H3 H2 −0.0071 0.0032 −0.0039
Difference(kcal/mol) 8.9 −3.2 5.5
CF2 AB VXC (A, B) Vcl (A, B) Eint (A, B)

Triplet F2 C1 −0.2816 −0.5501 −0.8317


F3 C1 −0.2816 −0.5501 −0.8317
F3 H2 −0.0316 0.1247 0.0931
Singlet F2 C1 −0.2816 −0.5501 −0.8317
F3 C1 −0.2816 −0.5501 −0.8317
F3 F2 −0.0316 0.1247 0.0931
Difference(kcal/mol) 29.9 −169.8 −139.9
CCl2 AB VXC (A, B) Vcl (A, B) Eint (A, B)

Triplet Cl2 C1 −0.3677 0.0508 −0.3169


Cl3 C1 −0.3677 0.0508 −0.3169
H3 H2 −0.0194 0.0038 −0.0156
Singlet Cl2 C1 −0.3358 0.0301 −0.3057
Cl3 C1 −0.3358 0.0301 −0.3058
Cl3 Cl2 −0.0358 0.0047 −0.0311
Difference(kcal/mol) 29.8 −25.4 4.3
CBr2 AB VXC (A, B) Vcl (A, B) Eint (A, B)

Triplet Br2 C1 −0.3438 0.0281 −0.3157


Br3 C1 −0.3438 0.0281 −0.3157
Br3 Br2 −0.0169 0.0130 −0.0039
Singlet Br2 C1 −0.3200 0.0361 −0.2839
Br3 C1 −0.3200 0.0361 −0.2839
Br3 Br2 −0.0310 0.0062 −0.0249
Difference(kcal/mol) 21.0 5.8 26.7

These data were obtained from PBE1PBE1/aug-cc-pVTZ calculations

instability of self-energies. As a result, in C


­ Br2 and ­CCl2, the singlet state will be the
preferred state.
The singlet state is the most stable state in the C ­ F2 molecule (Table 7), albeit it
has a somewhat different IQA pattern compared to the C ­ Br2 and C ­ Cl2 molecules.
In this molecule, the self-energy of the atoms favors the triplet state; all three atoms

13
3222 Z. Emami‑Meibodi et al.

Table 9  Atomic self-energies (in a.u.) of C ­ X2 molecules. Difference between total self-energies of sin-
glet and triplet states (defined as ESinglet
self self
− ETriplet ) are in kcal/mol
CH2 Atom Eself (A)

Triplet C1 −37.7296
H2 −0.4396
H3 −0.4396
Singlet C1 −37.6914
H2 −0.4493
H3 −0.4493
Difference (kcal/mol) 11.8
CF2 Eself (A)

Triplet C1 −36.8578
F2 −99.5269
F3 −99.5269
Singlet C1 −36.7801
F2 −99.4930
F3 −99.4930
Difference (kcal/mol) 91.3
CCl2 Eself (A)

Triplet C1 −37.5504
Cl2 −459.9390
Cl3 −459.9390
Singlet C1 −37.5690
Cl2 −459.9444
Cl3 −459.9444
Difference (kcal/mol) −18.4
CBr2 Eself (A)

Triplet C1 −37.6133
Br2 −2573.7500
Br3 −2573.7501
Singlet C1 −37.6353
Br2 −2573.7686
Br3 −2573.7686
Difference (kcal/mol) −37.1

These data were obtained from PBE1PBE1/aug-cc-pVTZ calculations

(carbon and fluorines) are more stable in the triplet state, and in total, the self-
energy of the triplet state is about 91.3 kcal/mol lower (more negative) than the sin-
glet state (Table  9). It should be noted that this behavior of ­CF2 is quite different
from that of ­CBr2 and ­CCl2 and can be compared to ­CH2. On the other hand, the
interatomic interaction in ­CF2 favors the singlet state. In fact, in this part, the singlet
state is about 140 kcal/mol more stable than the triplet state (Table 8). The stronger

13
MP2, DFT, and IQA study of substituent effect on the structure,… 3223

interatomic interaction in the singlet state can overcome the instability caused by the
self-energy and stabilize the singlet state of ­CF2.

Conclusion

MP2/aug-cc-pVTZ level of theory was employed for geometry optimizations, popu-


lation analyses, and energy calculations. The obtained molecular parameters showed
both electronegativity and size of the ligand are effective on the structures of car-
benes and carbenoids, and the electronegativity effect is more intense than the size.
Atomic hybridizations and charges were also obtained from NBO calculations. The
results showed the p indexes of carbon in triplet carbenes are also smaller than those
in singlet carbenes, but this difference in halogen-containing carbenes is smaller
than the other carbenes. The results of population analyses showed that the except
for sodium-based carbenoids, all carbenoids have higher Eg values than the carbenes
in this order: Be-based carbenoids > Mg-based carbenoids > Li-based carbenoids.
By considering the α-elimination as a method of preparation of the studied carbenes,
the order of the ΔG values for the preparation of these carbenes is f < c < b < m < h,
which shows the easier preparation of the halogenated carbenes. IQA analyses
showed that the interatomic exchange–correlation interaction in the C–F bond, like
other C–X bonds examined in this section, is more favorable in the triplet state.
What makes the C–F bond different from C–H, C–Cl, and C–Br bonds is the clas-
sical electrostatic contribution. In fact, the classical interaction in C–F is attractive
(negative) unlike the other mentioned bonds. Additionally, the electrostatic contri-
bution in C–F is much greater than the exchange–correlation contribution, so that
this bond can be predominantly referred to as an electrostatic bond. The electrostatic
term in C–F is larger in the singlet state than the triplet state, which leads to the sin-
glet state of ­CF2 being more stable and consequently more favorable than its triplet
state.
Supplementary Information  The online version contains supplementary material available at https://​doi.​
org/​10.​1007/​s11164-​023-​05031-5.

Authors’ contributions ZE performed MP2 and DFT calculations and extracted the related data. HT
designed the study, analyzed MP2 and DFT data, prepared related figures, tables, and scheme, and wrote
the all parts of the manuscript except IQA part. KS performed IQA calculations and wrote related part in
the manuscript.

Funding  No fund is received for this work.

Availability of data and materials  All data are available and could be sent on request from the corre-
sponding author.

Declarations 
Conflict of interest  The authors declare that there is no financial or personal interest.

Ethical approval  This work does not include any human and/or animal studies.

13
3224 Z. Emami‑Meibodi et al.

References
1. R. A. Moss, M. S. Platz, M. Jones Jr, (Eds.). John Wiley & Sons (2004)
2. G.T. Gurmessa, G.S. Singh, Res. Chem. Intermed. 43, 6447 (2017)
3. F.E. Hahn, Chem. Rev. 118(19), 9455 (2018)
4. M.S. Lee, J.E. Jackson, Res. Chem. Intermed. 20, 223 (1994)
5. J. Vignolle, X. Cattoen, D. Bourissou, Chem. Rev. 109(8), 3333 (2009)
6. H. Tomioka, E. Iwamoto, H. Itakura, K. Hirai, Nature 412(6847), 626 (2001)
7. A. Pawar, S. Gajare, A. Patil, R. Kurane, G. Rashinkar, S. Patil, Res. Chem. Intermed. 47, 2801
(2021)
8. J.C. Lin, R.T. Huang, C.S. Lee, A. Bhattacharyya, W.S. Hwang, I.J. Lin, Chem. Rev. 109(8), 3561
(2009)
9. J. J. Zupancic, G. B. Schuster J. Am. Chem. Soc. 102(18), 5958 (1980).
10. D. Bourissou, O. Guerret, F.P. Gabbaï, G. Bertrand, Chem. Rev. 100(1), 39 (2000)
11. P.B. Grasse, B.E. Brauer, J.J. Zupancic, K.J. Kaufmann, G.B. Schuster, J. Am. Chem. Soc. 105(23),
6833 (1983)
1 2. A. Nemirowski, P.R. Schreiner, J. Org. Chem. 72(25), 9533 (2007)
13. D.R. Myers, V. P. Senthilnathan, M. S. Platz, M. Jones, J. Am. Chem. Soc. 108(14) 4232 (1986)
14. G. Bertrand, Carbene chemistry: from fleeting intermediates to powerful reagents. Boca Raton:
CRC Press (2002).
1 5. P.H. Mueller et al., J. Am. Chem. Soc. 103(17), 5049 (1981)
16. K. Hirai, T. Itoh, H. Tomioka, Chem. Rev. 109(8), 3275 (2009)
17. C. Boehme, G. Frenking, J. Am. Chem. Soc. 118(8), 2039 (1996)
18. Y. Mizuhata, T. Sasamori, N. Tokitoh, Chem. Rev. 109(8), 3479 (2009)
19. A.J. Arduengo, Acc. Chem. Res. 32(11), 913 (1999)
20. C. Hill, F. Bosold, K. Harms, J.C. Lohrenz, M. Marsch, M. Schmieczek, G. Boche, Chem. Ber.
130(9), 1201 (1997)
2 1. P. Bazinet, G.P. Yap, D.S. Richeson, J. Am. Chem. Soc. 125(44), 13314 (2003)
22. A.J. Arduengo III., H.R. Dias, D.A. Dixon, R.L. Harlow, W.T. Klooster, T.F. Koetzle, J. Am. Chem.
Soc. 116(15), 6812 (1994)
23. P.H. Mueller, N.G. Rondan, K.N. Houk, J.F. Harrison, D. Hooper, B.H. Willen, J.F. Liebman, J.
Am. Chem. Soc. 103(17), 5049 (1981)
2 4. E.A. Carter, W.A. Goddard, J. Phys. Chem. 91(18), 4651 (1987)
25. F. Mendez, M.A. Garcia-Garibay, J. Org. Chem. 64(19), 7061 (1999)
26. C.M. Geise, Y. Wang, O. Mykhaylova, B.T. Frink, J.P. Toscano, C.M. Hadad, J. Org. Chem. 67(9),
3079 (2002)
2 7. M. Soleilhavoup, G. Bertrand, Acc. Chem. Res. 48(2), 256 (2015)
28. D. Munz, Organometallics 37(3), 275 (2018)
29. M.S. Sanford, M. Ulman, R.H. Grubbs, J. Am. Chem. Soc. 123(4), 749 (2001)
30. A. Caballero, P.J. Pérez, Chem. Eur. J. 23(58), 14389 (2017)
31. L. Friedman, H. Shechter, J. Am. Chem. Soc. 82(4), 1002 (1960)
32. P.S. Skell, A.Y. Garner, J. Am. Chem. Soc. 78(20), 5430 (1956)
33. R.A. Moss, Acc. Chem. Res. 13(2), 58 (1980)
34. A. Padwa, M.D. Weingarten, Chem. Rev. 96(1), 223 (1996)
35. T. Satoh, Chem. Soc. Rev. 36(10), 1561 (2007)
36. G. Costantino, R. Rovito, A. Macchiarulo, R. Pellicciari, THEOCHEM 581(1–3), 111 (2002)
37. L.R. Domingo, M. Ríos-Gutiérrez, M. Duque-Noreña, E. Chamorro, P. Pérez, Theor. Chem. Acc.
135(7), 1 (2016)
3 8. J. Messelberger, A. Grünwald, P. Pinter, M.M. Hansmann, D. Munz, Chem. Sci. 9(28), 6107 (2018)
39. J. Vaitla, A. Bayer, K.H. Hopmann, Angew. Chem. Int. Ed. 57(49), 16180 (2018)
40. H. Qiu, C. Deng, Chem. Phys. Lett. 249(3–4), 279 (1996)
41. H. Hermann, J.C. Lohrenz, A. Kühn, G. Boche, Tetrahedron 56(25), 4109 (2000)
42. T. Clark, P. V. R. Schleyer, J. Chem. Soc. Chem. Comm. (20), 883 (1979)
43. M.A. Vincent, H.F. Schaefer, J. Chem. Phys. 77(12), 6103 (1982)
44. T. Clark, P.V.R. Schleyer, Tetrahedron Lett. 20(51), 4963 (1979)
45. B.T. Luke, J.A. Pople, P.V.R. Schleyer, T. Clark, Chem. Phys. Lett. 102(2–3), 148 (1983)
46. T. Clark, P.V.R. Schleyer, J. Am. Chem. Soc. 101(26), 7747 (1979)

13
MP2, DFT, and IQA study of substituent effect on the structure,… 3225

4 7. T. Koizumi, O. Kikuchi, Bull. Chem. Soc. Japan 68(1), 120 (1995)


48. R.H. Nishimura, V.E. Murie, R.A. Soldi, J.L. Lopes, G.C. Clososki, J. Braz.Chem. Soc. 26, 2175
(2015)
49. K.G. Taylor, Tetrahedron 38(18), 2751 (1982)
50. S. Molitor, V.H. Gessner, Angew. Chem. Int. Ed. 55(27), 7712 (2016)
51. V.H. Gessner, Chem. Comm. 52(81), 12011 (2016)
52. Gaussian 09, Revision B.01, M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb,
J. R. Cheeseman, G. Scalmani, V. Barone, G.A. Petersson, H. Nakatsuji, X. Li, M. Caricato, A.V.
Marenich, J. Bloino, B. G. Janesko, R. Gomperts, B. Mennucci, H.P. Hratchian, J.V. Ortiz, A.F.
Izmaylov, J.L. Sonnenberg, D. Williams-Young, F.Ding, F. Lipparini, F. Egidi, J. Goings, B. Peng,
A. Petrone, T. Henderson, D. Ranasinghe, V.G. Zakrzewski, J. Gao, N. Rega, G. Zheng, W. Liang,
M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O.
Kitao, H. Nakai, T. Vreven, K. Throssell, J. A. Montgomery Jr, J. E. Peralta, F. Ogliaro, M. J. Bear-
park, J. J. Heyd, E. N. Brothers, K. N. Kudin, V. N. Staroverov, V. N., T. A. Keith, R. Kobayashi,
J. Normand, K. Raghavachari, A. P. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, J. M.
Millam, M. Klene, C. Adamo, R. Cammi, J. W. Ochterski, R. L. Martin, K. Morokuma, O. Farkas,
J.B. Foresman, D. J. Fox, Gaussian, Inc., Wallingford CT, (2016).
53. AIMAll (Version 19.10.12), Todd A. Keith, TK Gristmill Software, Overland Park KS, USA, 2019
(aim.tkgristmill.com)
54. V. GaussView, 6, Dennington, Roy (Todd A.; Millam, John M. Semichem Inc., Shawnee Mission,
KS, Keith, 2016)
55. T. Koopmans, Physica 1, 104 (1933)
56. R.G. Parr, R.G. Pearson, J. Am. Chem. Soc. 105(26), 7512 (1983)
57. A.M. Pendás, M. Blanco, E. Francisco. J. Chem. Phys. 120, 4581 (2004)
58. M. Menéndez, R. Álvarez Boto, E. Francisco, Á. Martín Pendás J. Comput. Chem. 36, 833 (2015)
59. M. A. Blanco, A. Martín Pendás, E. Francisco, J. Chem. Theory Comput. 1(6), 1096 (2005)
60. A. Martín Pendás, M. Blanco, E. Francisco, J. Chem. Phys. 125, 184112 (2006).
61. J.L. Casals-Sainz, F. Jiménez-Grávalos, A. Costales, E. Francisco, Á.M. Pendás, J. Phys. Chem. A
122, 849 (2018)

Publisher’s Note  Springer Nature remains neutral with regard to jurisdictional claims in published maps
and institutional affiliations.

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under
a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted
manuscript version of this article is solely governed by the terms of such publishing agreement and
applicable law.

13

You might also like