Sie sind auf Seite 1von 9

View Online

PAPER

www.rsc.org/pccp | Physical Chemistry Chemical Physics

Divalent carbon atom as the proton acceptor in hydrogen bondingw


Miros$aw Jab$onski*a and Marcin Palusiak*b

Downloaded by STADT UND UNIVERSITAETS on 29 September 2011


Published on 07 May 2009 on http://pubs.rsc.org | doi:10.1039/B901968E

Received 2nd February 2009, Accepted 27th March 2009


First published as an Advance Article on the web 7th May 2009
DOI: 10.1039/b901968e
Proton-accepting properties of the divalent carbon atom in carbodiphosphoranes and their simple
derivatives as well as in carbenes have been investigated. Both these groups of chemical
compounds may be characterized by the formula CL2, where L is a s electron donor. Therefore,
the carbon atom within both these systems, being in its atomic state, can have one or two lone
electron pairs and, as a result, it may form hydrogen bonds of the type DH  CL2, where C acts
as a proton acceptor. Complexes of C(NH3)2, C(PH3)2, C[P(CH3)3]2, CF2, CCl2, and imidazol2-ylidene with such proton donors as H2O, HCF3, HCN and HCCH have been analyzed by
means of high-level quantum chemical methods. Density functional theory (DFT) and secondorder MllerPlesset (MP2) approaches have been applied in conjunction with the aug-cc-pVTZ
basis set. The electron density distribution calculated by means of the atoms in the molecules
procedure has also been analyzed. Proton-accepting properties of the carbon atom are discussed
in detail. It is shown that the divalent carbon atom in the group of chemical systems investigated
should be treated as a normal proton acceptor, similar to the much more electronegative O or N
atoms. Moreover, hydrogen bonds of the type DH  CL2 within the complexes investigated have
been found to be rather strong. The highest proton accepting ability of the carbon(0) atom found
for the (NH3)2C derivative of carbodiphosphorane is explained on the basis of the
LeerHammond postulate. Within the group of carbenes, the strongest hydrogen bonds are
formed by imidazol-2-ylidene. This is attributed to the signicant aromatic character of the
imidazol-2-ylidene ring that increases the proton-accepting properties of the carbene carbon atom.

Introduction
Hydrogen bonding (H-bonding) is one of the most frequently
investigated phenomena in various areas of chemistry, chemical physics, molecular biology and others. This interaction
plays a crucial role in many physical, chemical and biological
processes.1,2,3 As it is well known, H-bonding (DH  Y) is
considered as a type of interaction between a hydrogen
atom of the proton donor system (DH) and some specic
H-accepting centre (Y) which possesses an excess of electron
density. Most often, the proton acceptor (Y) has a lone
electron pair which can be partially shared with the proton
donor (DH) during the H-bond formation. By generalization
of the proton acceptor form to any region of high electron
density distribution, interactions of the form DH  p or
DH  H (dihydrogen bonds) may also be attached to the
group of hydrogen bonds. These so-called non-classical
hydrogen bonds are often considered as unusual, however,
since no specic electron lone pair participates in the H-bonding
formation.4,5 Thus, on the basis of the above, it can be stated
a

Department of Quantum Chemistry, Faculty of Chemistry, Nicolaus


Copernicus University, Gagarina 7, 87-100, Torun, Poland.
E-mail: teojab@chem.uni.torun.pl
b
Department of Crystallography and Crystal Chemistry, University of
odz, Pomorska 149/153, 90-236, odz, Poland.
E-mail: marcinp@uni.lodz.pl
w Electronic supplementary information (ESI) available: Details of
BSSE correction, material supporting discussion, Cartesian coordinates and energies of all H-bonded complexes discussed in the paper.
See DOI: 10.1039/b901968e

This journal is


c

the Owner Societies 2009

that a carbon atom may hardly be considered as a good proton


acceptor in H-bonding. There are, however, compounds in
which a lone electron pair is located on the carbon atom. This
interesting and quite unique electronic state of the carbon
atom exists in carbenes, carbodiphosphoranes and compounds
related to these two chemical classes. Thus, a question arises:
how far such a lone electron pair located on the carbon atom
may contribute to H-bonding, or, in other words, how far the
divalent carbon can serve as the Lewis base in H-bonding?
The rst report regarding the isolation of stable carbenes
dates back to 1985.6 Since 1991, when Arduengo et al.7
successfully synthesized imidazolin-2-ylidene and its derivatives,
the chemistry of carbenes has been greatly extended.8,9 It is
known that the divalent carbon atom in carbenes may exist in
both singlet and triplet states.9 It has been found that the
ground state of the CH2 molecule corresponds to the triplet
state. Nevertheless, the proximity of highly electronegative
atoms (halogens, oxygen or nitrogen), which possess a lone
electron pair, stabilizes the singlet state by delocalization of
the lone electron pair into the empty p orbital of the carbene
carbon atom.911 Thus, carbenes in the singlet state possess a
lone electron pair located on the carbon atom which can
potentially be involved in the process of H-bond formation.
Another class of compounds in which the divalent carbon
atom may serve as a Lewis base is a group of carbodiphosphoranes (CDPs) and their amine analogues. Frenking
and his co-workers12 have recently characterized the bonding
situation in CDPs by means of high quality quantum chemical
calculations. The state of the art bonding analysis performed
Phys. Chem. Chem. Phys., 2009, 11, 57115719 | 5711

Downloaded by STADT UND UNIVERSITAETS on 29 September 2011


Published on 07 May 2009 on http://pubs.rsc.org | doi:10.1039/B901968E

View Online

Fig. 1 An example of H-bonding of the type DH  C investigated in


this paper. The interaction between the solvent molecule (methanol in
this case) and N,N-bismesityl-imidazol-2-ylidene stabilizes the crystal
structure; d(OH) = 0.88 A, d(H  C) = 1.854 A, +(OH  C) =
173.411 (data from CSD, no esds available).

by these authors indicated that the PC bonding should be


considered as a donoracceptor interaction between the
phosphoric lone electron pair and the empty valence p orbital
of carbon.13 As a consequence of this unique bonding situation,
the divalent carbon atom has one s type and one p type lone
pair orbital. The chemistry of CDPs was later characterized in
detail.14 Similarly to carbenes, the divalent carbon atom in
CDPs may also serve as a Lewis base in H-bonding.
A search through the Cambridge structural database15
suggests that the divalent carbon atom may in fact participate
in H-bonding as a proton acceptor (see Fig. 1). A good
example could be the short (O)H  C contact found in the
crystal structure of N,N-bismesityl-imidazolylidene.16 In this
case, a lone electron pair located on the carbon atom in the
imidazol-2-ylidene ring may participate in H-bonding in which
the methanol molecule acts as a proton donor. The geometry
of this potential H-bridge is practically linear, and may suggest
a relatively strong binding between both interacting molecules.
However, crystal data correspond to a chemical situation where
crystal eld eects may play a signicant or even crucial role in
the determination of the structure. Thus, it cannot be excluded
that this unique short contact may be the eect of the packing of
the molecules in the crystal state, particularly as it is impossible to
estimate the exact position of the H atoms by X-ray analysis.
The aim of our studies is to characterize intermolecular
interactions in which a carbon atom having a lone electron
pair acts as a proton acceptor. For this purpose, a detailed
analysis based on high-quality quantum chemical calculations
was performed for several interacting systems in which
C(NH3)2, C(PH3)2, C[P(CH3)3]2, CF2, CCl2, and imidazol2-ylidene were acting as proton acceptors, whereas H2O,
HCF3, HCN and HCCH were playing the role of proton
donors. The rst three proton acceptor molecules represent
CDPs, the next three are carbenes.

of approximation: second order MllerPlesset perturbation


theory (MP2)17 and the DFT/B3LYP18 method, both in
conjunction with the aug-cc-pVTZ basis set19 for all
atoms. Calculations were performed with the use of the
GAUSSIAN03 set of codes.20
To increase the chance of nding distinct local minima on the
potential energy surface, the geometry optimization was performed with several starting geometries. They can be roughly
characterized as having linear or bent DH  C contacts.
Dimers with bent DH  C contacts were designed as being
on one plane or with the proton acceptor and proton
donor molecules being on dierent planes. In most cases, the
optimization procedure led to the same or very similar geometries (mainly with some negligible dierences in the DH  C
angle, in this case the geometry corresponding to the lowest
energy was always taken for further analysis). However, in
some cases, two minima were found with signicantly dierent
geometries and energies. A frequency analysis was used to
verify that the optimized geometries correspond to the ground
state stationary points. No imaginary frequencies were found.
In order to ensure adequate convergence and the reliability of
computed frequencies, the cutos on forces and step size that
are used to determine convergence were tightened in the case of
DFT calculations (tight option in the optimization procedure).
Because of technical restrictions for the largest systems, i.e.
complexes with C[P(CH3)3]2, only calculations with DFT
approximation were performed. However, these systems exhibit
properties that are very similar to those of C(PH3)2 and MP2
approximation should not lead to any new results.
For the optimized geometries, a detailed analysis of the
electron distribution function was made according to the concept
of the atoms in molecules theory proposed by Bader,21 using
the AIM2000 program.22,23 The properties of electron density
calculated at bond critical points (BCP) of the H-bonds under
consideration were characterized. The electron density at BCP,
rBCP, its laplacian, r2rBCP, the density of the total electron
energy, HBCP, and its two components, densities of potential and
kinetic electron energies, VBCP and GBCP, were discussed.
In the supermolecular approach, the interaction energy, DE,
is calculated as the dierence between the total energy of the
complex and the sum of total energies of isolated monomers.
The interaction energy obtained has been corrected using
counterpoise correction (DE(CP) in tables with data), as
proposed by Boys and Bernardi.24 Additionally, counterpoise
correction taking into account deformation energy was also
applied (DE(BSSE) in tables with data). DE(BSSE) converges to
DE as the basis set approaches the complete basis set (CBS)
limit.25 The BSSE-corrected interaction energy obtained on the
DFT level of theory was almost the same as the uncorrected
counterpart (the dierence being less than 0.2 kcal mol1).
(See ESI for details concerning both CP corrections.)w

Results and discussion


H-bonding to CDPs and their derivatives

Methodology
Geometries for both monomers as well as complexes were
optimized without any symmetry constraints using two levels
5712 | Phys. Chem. Chem. Phys., 2009, 11, 57115719

As already mentioned, the C(0) carbon atom in CDPs has a


very unique electronic structure, since it possesses two lone
electron pairs. One of the lone pairs occupies one of the sp2
This journal is


c

the Owner Societies 2009

Downloaded by STADT UND UNIVERSITAETS on 29 September 2011


Published on 07 May 2009 on http://pubs.rsc.org | doi:10.1039/B901968E

View Online

orbitals (while the other two sp2 orbitals are lled with lone
electron pairs donated by ligands coordinating the C(0) atom).
The second lone electron pair is described by a single pp
orbital. Therefore, the two lone electron pairs of the C(0)
atom may lead to H-bonding.
Table 1 contains selected parameters characterizing H-bonds
in which C(0) acts as the proton acceptor. Additionally, the
AIM properties of the electron density function calculated at
critical points are collected in Table 2.
Clearly, there are two factors inuencing the strength of the
interaction under discussion: (i) the proton-donating features
of the proton donor, and (ii) the character of the ligand
coordinating the C(0) atom. The rst factor is evidently
connected with the properties of the donor HD. It is generally
known that the strongest proton donors (D) are usually those
atoms which are characterized by a relatively large electronegativity. Thus, the increasing proton donating abilities in the
set of H-donating molecules investigated in this paper are
HCCH o HCF3 o HCN o H2O. It should be mentioned
that, usually, the sp3 hybridized carbon atoms (HC groups
with sp3 hybridized carbon atoms) are the weakest proton
donors.3 The acidity of the HCF3 molecule (in the sense of
Brnsted acidity) is amplied by the presence of three strongly
electronegative uorines directly attached to the carbon atom.
The second of the factors mentioned is connected directly
with the proton accepting ability of the C(0) atom. From the
data collected in Table 1, it can be seen that the strongest
intermolecular interactions characterize the amino derivative
of CDPs. For those systems for which stable complexes linked
by the DH  C(0) bridge were found this is reected by the
(relative) shortest distances of the H-bridge. Therefore, one
may consider the C(0) atom in the C(NH3)2 molecule as a
stronger proton acceptor than in the case of C(PH3)2 or
C[P(CH3)3]2. This conclusion is supported by the fact that,

for the C(NH3)2 system, proton transfer from the proton


donating molecule to the C(0) atom is observed for the
strongest H-donors investigated in this paper, i.e. for H2O or
HCN. It should be mentioned, however, that in practically
all systems with C(NH3)2 acting as the proton acceptor,
geometries with additional hydrogen bonds were found. This
is a consequence of the fact that the NH group exhibits
relatively strong proton donating properties and, whenever
energetically protable, it forms additional H-bonds. Thus, for
those cases, the energy of complexation also includes the
energy of these additional interactions. Yet, the C(NH3)2
system seems to be a stronger proton acceptor than C(PH3)2
and C[P(CH3)3]2 molecules. For instance, this can be found for
linear complexes with HCCH as the proton donor. In this
case, no additional interaction inuences the geometry of the
H-bridge and the shortest H  C(0) distance of 2.2 A can be
found for the (NH3)2C  HCCH system (DFT/aug-cc-pVTZ).
When the additional interaction is present (bent geometry of
the (NH3)2C  HCCH system) then this distance is even
shorter (2.1 A) due to the cooperation of two H-bonding
interactions. Thus, the additional interaction cannot be
omitted in the analysis of energetics and geometry of a given
complex. The presence of additional stabilizing interactions
has also been discussed recently by Grabowski et al.26
In the case of the (NH3)2C  HCF3 complex, two local
minima were found at the MP2 level of theory. In both these
structures, additional interactions of the NH  F type were
present. Fig. 2 shows the proper graphical representations.
Interestingly, the stronger interaction as measured by the
binding energy (11.30 kcal mol1, see Table 1) is present in
the system in which there is only one NH  F hydrogen bond
(complex (a) in Fig. 2). The total strength of the interaction
in complex (b) is somewhat smaller (9.96 kcal mol1),
even though three individual H-bonds contribute to this

Table 1 Selected numerical data obtained for carbon(0) complexes. Distances in A, angles in 1, frequencies in cm1, intensities in km mol1,
energies in kcal mol1
H-acceptor
C(NH3)2

H-donor Method d(DH) Dd(DH)a d(H  C) +(DH  C) n(DH) Dn(DH)b I(DH) Ic/Imc DE
HCF3

HCCH
C(PH3)2

H2O
HCF3e
HCN
HCCH

C[P(CH3)3]2 H2O
HCF3
HCN
HCCH

MP2
MP2g
B3LYP
MP2
B3LYPh
B3LYPi
MP2
B3LYP
MP2
B3LYP
MP2
B3LYP
MP2
B3LYP
B3LYP
B3LYP
B3LYP
B3LYP

1.094
1.095
1.097
1.091
1.089
1.084
0.983
0.983
1.089
1.094
1.091
1.093
1.077
1.075
0.992
1.097
1.105
1.079

0.0080
0.0090
0.0083
0.0294
0.0273
0.0227
0.0212
0.0212
0.0038
0.0050
0.0261
0.0271
0.0147
0.0134
0.0301
0.0077
0.0397
0.0176

2.122
2.337
2.159
2.051
2.118
2.201
1.963
2.005
2.252
2.310
2.043
2.100
2.212
2.331
1.971
2.287
2.040
2.320

143
129
144
139
144
174
168
179
156
174
168
179
160
179
178
176
178
178

3087
3079
3001
3067
3065
3126
3439
3403
3142
3043
3074
3046
3252
3251
3227
2998
2884
3190

-114
121
123
364
347
286
383
394
59
81
393
399
179
161
571
126
564
222

20.8
125.5
80.0
133.6
584.4
793.8
936.1
984.1
43.2
65.0
1069.7
1031.5
559.2
580.2
1196.4
108.5
1183.6
655.9

DE(CP) DE(BSSE)

0.9 12.45 12.29 11.30


5.7 11.29 11.16 9.96
3.1 9.41 10.11 9.19
1.4 10.03 9.91 9.16
6.5 6.76 7.24 6.62
8.8 6.26 6.36 6.10
168.3 9.37 8.79 8.40
212.0 7.41 7.63 7.29
2.0 7.27 6.48 6.24
2.5 4.96 5.01 4.79
13.9 9.78 9.13 8.72
15.7 7.79 8.01 7.60
5.9 6.22 5.48 5.31
6.4 3.68 3.62 3.51
257.8 9.50 9.95 9.28
4.2 5.94 6.11 5.62
17.9 9.28 9.96 9.01
7.3 4.19 4.21 3.99

a
The change (elongation) of the proton donor bond length. b DH stretching vibration frequency shift. c The ratio of the IR intensity of the DH
stretching mode in the bound and isolated forms of the proton donor molecule. d Proton transfer to the C(0) atom is observed for the system
interacting with H2O or HCN playing the role of H-donor. e The additional interaction of the NH  F or PH  F type is observedsee text for
more details. f The at structure with additional NH  F H-bond. g L-shaped structure with two additional H-bonds of the type NH  F. h The
bent H-bridge geometry. i The linear H-bridge geometry.

This journal is


c

the Owner Societies 2009

Phys. Chem. Chem. Phys., 2009, 11, 57115719 | 5713

View Online

Table 2 Selected AIM electron density parameters calculated at BCP(H  C). Data (DFT/aug-cc-pVTZ) obtained for complexes including
derivatives of CDPs as the proton acceptor. All values in au
H-acceptor
C(NH3)2

C(PH3)2

Downloaded by STADT UND UNIVERSITAETS on 29 September 2011


Published on 07 May 2009 on http://pubs.rsc.org | doi:10.1039/B901968E

C[P(CH3)3]2

H-donor
a

HCF3
additional HB
HCCHb
additional HB
HCCHc
H2O
HCF3
HCN
HCCH
H2O
HCF3
HCN
HCCH

rBCP

r2rBCP

GBCP

VBCP

HBCP  103

0.023
0.012
0.025
0.011
0.021
0.030
0.017
0.025
0.016
0.034
0.019
0.030
0.017

0.056
0.053
0.055
0.039
0.047
0.048
0.040
0.048
0.038
0.046
0.041
0.049
0.038

0.014
0.011
0.015
0.008
0.012
0.017
0.010
0.014
0.009
0.018
0.010
0.016
0.009

0.014
0.009
0.015
0.006
0.012
0.021
0.009
0.016
0.008
0.024
0.010
0.019
0.009

0.172
2.183
0.741
1.747
0.109
4.403
0.391
1.787
0.633
6.162
0.027
3.429
0.327

a
Additional NH  F H-bonds (HB) are observed. Parameters for these interactions are also collected in the table. b Doubly-bonded complex
with additional NH  p interaction present within the structure. c Singly-bonded complex with CH  C(0) H-bridge being close to linear
geometry.

of MP2, as well as in one of DFT geometries, the additional


interaction stabilizes the complex. This additional interaction
may be classied as NH  p, where the p proton-accepting
orbital is from the C2H2 molecule (see Fig. 3). Comparing
singly- and doubly-bonded (NH3)2C  HCCH complexes
(both found at the DFT/aug-cc-pVTZ level of theory), the
interaction energy is only a little larger for the doubly-bonded
system (6.62 kcal mol1 vs. 6.10 kcal mol1 for its
singly-bonded counterpart). This indicates that the additional
interaction does not signicantly contribute to the total energy
of interaction (about 0.5 kcal mol1 from direct comparison).
However, if we compare the interaction energy obtained for
the linear (NH3)2C  HCCH complex with that of
(PH3)2C  HCCH (6.10 kcal mol1 and 3.51 kcal mol1,
respectively, see Table 1), then a signicant dierence occurs,
being close to 40% of the interaction energy of the linear
(NH3)2C  HCCH complex. This conrms the previous

Fig. 2 Graphical representation of two minima found at the MP2/


aug-cc-pVTZ level of theory for (NH3)2C  HCF3; (a) doubly-bonded
complex, (b) triply-bonded complex. All non-H atoms are labeled.

interaction. The weaker interaction in the case of complex (b)


compared to complex (a) is also reected by larger distances
between the interacting atoms. The NH  F distance in
complex (a) of Fig. 2 is about 2.1 A, whereas distances of
both NH  F hydrogen bonds in the structure (b) are over
2.3 A. Additionally, this is accompanied by a slightly shorter
CH  C(0) distance in the case of the doubly-bonded complex
(see Table 1 and Fig. 2). In the case of DFT calculations, only
one minimum was found in which both interacting molecules
form two H-bonds similarly as shown in Fig. 2(a).
In the case of the (NH3)2C  HCCH complex, only one
minimum was found for MP2 calculations. However, two
minima were found in the DFT approximation. In the case
5714 | Phys. Chem. Chem. Phys., 2009, 11, 57115719

Fig. 3 Molecular graphs of two minima found for the


(NH3)2C  HCCH complex (DFT/aug-cc-pVTZ calculations); big
circles correspond to attractors, small ones to BCPs and ring critical
points (RCPs).

This journal is


c

the Owner Societies 2009

Downloaded by STADT UND UNIVERSITAETS on 29 September 2011


Published on 07 May 2009 on http://pubs.rsc.org | doi:10.1039/B901968E

View Online

conclusion that the carbon atom in (NH3)2C possesses


signicantly higher proton accepting ability than in (PH3)2C.
The energetic and geometrical parameters of individual
H-bonds may result from the presence of multiple interactions,
thus a topological analysis of electron density was applied.
This approach allows one to estimate the strength of the given
individual bond based on the analysis of properties of the
electron density function calculated at the bond critical
point (BCP) of the bond of interest. Table 2 contains
parameters of the electron density function calculated at
the proper BCPs. Data for additional interactions are also
collected in Table 2.
In general, the greater the value of electron density
calculated at the BCP is, rBCP, the stronger the interaction
between two atoms forming this bond is. This feature of the
electron density at BCP applies to any pair of bonded atoms.
In the case of the strongesttypical covalent bondsthe
laplacian of rBCP, r2rBCP, is negative. Positive values of
r2rBCP characterize typical closed-shell interactions. Also,
the values of the total energy density (HBCP) and its components,
kinetic (GBCP, positive by denition) and potential (VBCP,
negative by denition) energy densities, provide valuable
information on the nature of the chemical bond. For the
strongest bonds, mainly covalent in nature, the value of HBCP
is negative and relatively large in its absolute value. For this
kind of interaction the potential energy density, VBCP,
dominates in HBCP. As can be seen from the data collected
in Table 2, in the case of all H-bonds the values of the
laplacian of rBCP are positive. However, values of HBCP
corresponding to DH  C(0) interactions in complexes of
C(PH3)2 and C[P(CH3)3]2 with H2O and HCN as well as in
all complexes of C(NH3)2, in spite of r2rBCP 4 0, are
negative. This interesting situation has already been found
for some metalmetal and metalligand bonding with
characteristics that represent a mix of shared and closed-shell
interactions. In this case, the value of HBCP is usually negative
and close to zero, as found for shared interactions, but with
r2rBCP 4 0, as found for closed-shell interactions.27 Such a
relation between the values of HBCP and r2rBCP is also
characteristic of the strongest H-bonds in which the covalent
contribution to the bonding is signicant.28 Thus, DH  C(0)
bonding in complexes of C(PH3)2 and C[P(CH3)3]2 with H2O
and HCN as well as in all complexes of C(NH3)2 can be
classied as a strong H-bond with characteristics being intermediate between those typical of closed-shell and shared
interactions. This is accompanied by large values of the
interaction energy and topological parameters calculated at
corresponding BCPs. It can also be concluded from the values
of rBCP and r2rBCP that the proton accepting abilities of C(0)
are greater in C(NH3)2 as compared with those of C(PH3)2.
This conclusion is supported by the results of the interaction
energy curves calculated for C(NH3)2 and C(PH3)2 interacting
with H+ and the helium atom. These calculations showed,
that for C(NH3)2, the attractive contribution is more signicant than in the case of C(PH3)2. The two curves for the
interaction with He are close to each other, whereas the
interaction energy for H+ is much lower in the case of
[C(NH3)2  H]+. (See ESI for details and all results of these
calculations.)w
This journal is


c

the Owner Societies 2009

A question then arises: why is C(0) a stronger proton


acceptor in the case of (NH3)2C than in the case of its
phosphine counterpart? In other words, one may ask why it
forms stronger H-bonds when coordinated by NH3 ligands
than when it is coordinated by PH3. In fact, NH3 ligands
should be considered as more electron withdrawing than
PH3 or P(CH3)3, because of the relatively greater electronegativity of nitrogen. Therefore, one would expect that
NH3 ligands, in comparison with PH3 or P(CH3)3 ligands,
should lower the proton-accepting properties of C(0) instead
of increasing them.
Some explanation may follow from the LeerHammond
postulate.29 According to this postulate, in systems for
which two possible ground states related by a proton transfer
reaction can be found, stronger H-bonds usually exist in those
systems which are energetically closer to transition states
corresponding to this proton transfer. More generally, the
closer the geometry (and the total energy) of the given system
is to the transition state which corresponds to the proton
transfer, the stronger is the H-bond related to this proton
transfer. For instance, this postulate is perfectly valid in
the case of tautomeric forms of malonaldehyde and its
derivatives.30 For their two tautomeric structures, the stronger
H-bond was always found for the one which was less stable
thermodynamically, and therefore energetically and geometrically
closer to its corresponding transition state. In the case of the
presently investigated complexes, it is also possible to consider
proton transfer leading to the formation of the CH bond and,
in consequence, resulting in additional bond reorganization.
As already mentioned, in the case of (NH3)2C  H2O, a
two-step proton transfer was found during the process of
geometry optimization and the nal, doubly-H-bonded
complex with the imine derivative instead of (NH3)2C was
found as the stationary ground state (see Fig. 4 for a graphical
representation). The fact that (NH3)2C  H2O undergoes such
a bond reorganization while (PH3)2C  H2O was found as the
ground state stationary point may suggest that NH3 stabilizes
the unique C(0) electronic structure much less eectively than
PH3 or P(CH3)3. The PH3 ligand (with the P atom less
electronegative than N) coordinating the C(0) atom, shares
its lone electron pair more easily relative to the NH3 ligand. In
consequence, the (NH3)2C  HD complexes, if they exist as
stationary points, are relatively less stable thermodynamically.
They are energetically and geometrically closer to the
proper transition state and, according to the LeerHammond

Fig. 4 The structure obtained after the geometry optimization


(MP2/aug-cc-pVTZ level of theory) of the (NH3)2C  H2O complex.
Geometry reorganization is observed with hydrogen atom now
attached to the carbon atom.

Phys. Chem. Chem. Phys., 2009, 11, 57115719 | 5715

Downloaded by STADT UND UNIVERSITAETS on 29 September 2011


Published on 07 May 2009 on http://pubs.rsc.org | doi:10.1039/B901968E

View Online

postulate, they form stronger H-bonds than their phosphine


counterparts. This explains why the C(0) atom, under some
circumstances, tends to covalently bind the proton with which
it interacts as in the case of the (NH3)2C  H2O complex. In
this way, the less stable L2C(0) complexes may form stronger
H-bonds.
It is also worthwhile to take a closer look at the structure of
the complex in Fig. 4. It may be concluded from the valencies
of C and N atoms, that only one of the two N atoms acts as a
ligand sharing its lone electron pair with the C atom. The other
N atom seems to form a typical covalent bond, additionally
acting as the proton acceptor in the OH  N hydrogen bond.
The molecular graph in Fig. 5 shows the distribution of
critical points and the bonding scheme of this complex. The
characteristics of the CNH2 bond are typical of covalent
bonds, with a relatively large value of rBCP, and large negative
value of r2rBCP (see Fig. 5 for numerical values). Also, the
distance of 1.471 A is typical of single covalent CN bonds. In
the case of the second CN bond, the characteristics clearly
dier from the previous case. The value of rBCP is smaller, and
the (negative) value of r2rBCP is close to zero. Additionally,
the interatomic CN distance is larger than for typical
covalent CN bonds.
Thus, it can be expected that in this specic complex the
carbon atom has a very unique electronic structure, with two
typical covalent bonds (one CN and one CH), one lone
electron pair, and, nally, an empty orbital lled with the lone
electron pair donated by the NH3 group. Additionally, in this
case, sp3 hybridization takes place, which is in contrast with
the C(PH3)2 and C(NH3)2 cases where the carbon atom was
sp2 hybridized and one of the lone electron pairs was located
on the p orbital.
Taking into account the two phosphorus derivatives,
C(PH3)2 and C[P(CH3)3]2, the latter exhibits slightly stronger
proton accepting properties. In the case of both these systems,
no additional interactions with the investigated proton donors
were found (DFT/aug-cc-pVTZ), and the DH  C interaction has the crucial role in the overall bonding eect. A
straightforward comparison of both geometrical and energetic
parameters (see Table 1) as well as the AIM characteristics (see
Table 2) indicate that the C[P(CH3)3]2 molecule can form
stronger H-bonds than its C(PH3)2 counterpart. Nevertheless,

Fig.
5 Molecular
graph
(DFT/aug-cc-pVTZ)
of
the
H3NCHNH2  H2O system from Fig. 4 (MP2/aug-cc-pVTZ level of
theory). Big circles correspond to attractors, small ones are BCPs and
RCP.

5716 | Phys. Chem. Chem. Phys., 2009, 11, 57115719

the dierence in proton accepting properties of C(PH3)2 and


C[P(CH3)3]2 is relatively small. Moreover, since steric eects
become more signicant in the case of C[P(CH3)3]2, the
tendency to form hydrogen bonds of the type DH  C(0)
should also be decreased if the proton donor molecule is of a
larger size.
The changes in spectroscopic parameters support the
general ndings in this section that the DH  C(0) hydrogen
bonds in the complexes investigated are of proper nature, i.e.
they may be referred to as classical hydrogen bonds. The
elongation of the DH proton donor bond is accompanied
by a DH stretching frequency decrease (the so called
red-shift).1,5,31 Also, one can observe an increase of infrared
intensity of the n(DH) band (Ic/Imthe ratio of IR intensity,
I(DH), in the complex and in the isolated molecule is 4 1, see
Table 1). Both these eects are particularly large if H2O or
HCN acts as the proton donor molecule, i.e. in the case of
strong hydrogen bonds. Large values of DH frequency shift
and of the IR intensity increase should help in experimental
identication of hydrogen bonds of the DH  C(0) type in
similar complexes. It should be emphasized that the classical
characteristics of DH  C(0) H-bonds in the investigated
systems indicate that they are similar to the usual DH  Y
hydrogen bonds, where Y = O, N. In other words, the C(0)
atom, although its electronic structure is a little unusual, can
be considered as a typical proton acceptor.
H-bonding to carbenes
In carbenes, the divalent carbon atom has signicantly
dierent electronic structure from that found for the C(0)
atom in CDPs and their derivatives. In this case, the two bonds
formed by the divalent carbon are covalent, so two electrons
of the carbon atom are shared with two attached atoms. The
other two electrons from the valence shell may occupy two
orbitals of similar energy, one being the sp2 hybrid orbital, and
the other being a p orbital perpendicular to the molecular
plane (i.e. to the plane dened by the three sp2 orbitals). Thus,
carbenes may exist in two spin states: the triplet state, as in the
CH2 molecule, if electrons are unpaired and singly occupy
each of these orbitals, or the singlet state, if these electrons are
paired and occupy the same orbital.32 Thus, the singlet state of
carbenes has a lone electron pair. Note that, since the formally
empty p orbital is perpendicular to the plane dened by the
three sp2 orbitals (in CDPs this orbital is lled with the second
lone electron pair), the singlet state may be additionally
stabilized by partial charge transfer from the rest of the system
onto the empty p orbital. Such delocalization of the lone
electron pair of F, Cl or N atoms attached to the carbon atom
onto the formally empty p orbital of carbon may, for example,
take place in CF2, CCl2 or an imidazol-2-ylidene ring. It
should be stressed that the singlet state can exist as a ground
state of the carbene only when extra stabilization through the
donation of electrons onto the empty p orbital takes place.11
Obviously, only the singlet state structure may serve as a good
model for the purposes of our analysis. Diradicals such as
carbenes in the triplet state can hardly be considered as
acceptors of protons in hydrogen bonding. Thus CF2, CCl2
and imidazol-2-ylidene were chosen for further consideration.
This journal is


c

the Owner Societies 2009

View Online

Table 3 Selected numerical data obtained for carbenes. Distances in A, angles in 1, frequencies in cm1, intensities in km mol1, energies in kcal
mol1
H-acceptor

H-donor method d(DH) Dd(DH)a d(H  C) +(DH  C) n(DH) Dn(DH)b I(DH) Ic/Imc DE

CF2

H2O
HCF3
H2O

CCl2

HCF3
C(imidazol2yliden) H2Of

Downloaded by STADT UND UNIVERSITAETS on 29 September 2011


Published on 07 May 2009 on http://pubs.rsc.org | doi:10.1039/B901968E

HCF3f

MP2
B3LYP
MP2
B3LYP
MP2
B3LYP
MP2d
MP2e
B3LYP
MP2
B3LYP
MP2
B3LYP

0.966
0.967
1.085
1.089
0.971
0.971
1.086
1.085
1.089
0.981
0.983
1.087
1.093

0.0051
0.0056
0.0003
0.0003
0.0094
0.0089
0.0009
0.0009
0.0005
0.0201
0.0210
0.0016
0.0039

2.233
2.250
2.612
2.685
2.123
2.167
2.477
2.553
2.579
2.002
2.005
2.368
2.382

168
169
176
180
168
173
179
130
177
141
146
139
176

3753
3712
3210
3131
3669
3646
3192
3215
3118
3472
3415
3178
3062

69
84
10
6
153
151
9
14
6
350
382
23
62

197.9
233.2
0.1
0.2
459.5
450.9
4.5
5.2
2.0
557.1
742.2
27.4
85.9

DE(CP) DE(BSSE)

35.6 3.42 2.97 2.89


50.2 2.75 2.74 2.66
0.0 2.30 1.85 1.80
0.0 1.52 1.44 1.41
82.6 4.98 4.35 4.19
97.1 3.57 3.61 3.47
0.2 3.56 2.88 2.79
0.2 3.64 2.97 2.89
0.1 2.10 2.05 1.98
100.2 10.64 10.25 9.84
159.9 8.65 8.92 8.54
1.2 7.76 7.12 6.77
3.3 5.49 5.54 5.35

a
The change (elongationpositive value, shorteningnegative value) of the proton donor bond length. b DH stretching vibration frequency
shift. c The ratio of the IR intensity of the DH stretching mode in the bound and isolated form of the proton donor molecule. d One of two
minima corresponding to linear H-bridge geometry. e One of two minima corresponding to bent H-bridge geometry. f Additional interactions of
the NH  O or NH  F type are observedsee text for more details.

From the set of proton donors we excluded HCN and HCCH


molecules since the addition reaction is a typical one which
carbenes readily participate in. Therefore, only the complexes
of CF2, CCl2 and imidazol-2-ylidene with H2O and HCF3
as proton donors were considered. Selected parameters
characterizing the H-bonding in complexes of carbenes with
H2O and HCF3 are listed in Table 3.
The water molecule is expected to form stronger H-bonds
than HCF3. This is because of the stronger proton-donating
properties of the oxygen atom as compared with the carbon
atom, even if it is additionally modied by three directly
attached uorine atoms. In fact, all data collected in Table 3
conrm these expectations. This is particularly well illustrated
by the data obtained for those complexes where no signicant
additional interactions are present, e.g. in all cases of linear
complexes found for DFT calculations.
Interestingly, the strongest H-bonds are observed for
complexes with imidazol-2-ylidene acting as the protonaccepting molecule, whereas the weakest H-bonds are found
for complexes formed with the participation of the CF2
molecule. Thus, it can be said that the carbene carbon in
imidazol-2-ylidene exhibits the strongest relative protonaccepting properties, whereas the carbon in CF2 is the weakest
proton accepting centre. The carbon atom in CCl2 represents
an intermediate case between those two. So a question may
arise: where do such dierences in the properties of the
carbene systems under discussion come from? On considering
only CF2 and its chlorine counterpart, the answer is relatively
simple; chlorine, being less electronegative, may share its
lone electron pairs more easily, stabilizing the (singlet state)
electronic structure of the carbon atom more eectively. In
this way, the lone electron pair located on a divalent carbon
atom may be involved in H-bond formation more easily in
CCl2 than in CF2. This results in a lower proton-accepting
strength of the carbon atom in CF2 than in CCl2.
The case of imidazol-2-ylidene requires some attention.
In this case, the carbon atom is directly connected to two
nitrogen atoms. Taking into account the previous explanation
of bonding strength of complexes formed by CF2 and CCl2
This journal is


c

the Owner Societies 2009

one could expect that the electronegativity of N must be lower


than that of Cl. Is the nitrogen atom really less electronegative
than chlorine one? Not really, thus some additional stabilization
of the electronic structure of divalent carbon atom in imidazol2-ylidene has to take place. The explanation may be obtained
from the analysis of electron structure of the imidazol-2ylidene ring. There are six p-electrons within this ring; two
of them are involved in a CQC p-bond, and four formally
belong to lone pairs located on N atoms. Therefore, the
imidazol-2-ylidene ring is a cyclic system with 4n + 2
p-electrons and can be considered as aromatic, since all atoms
forming this ring contribute to the p-electron delocalization.
In fact, HOMA (harmonic oscillator model of aromaticity)33,34
values (see Table 4) indicate the signicant aromatic character
of the imidazol-2-ylidene ring.
Therefore, the eect of the p-electron delocalization involving
the empty p orbital of the carbon atom stabilizes the singlet
structure of the carbene. In this way, a lone electron pair
located on the carbon atom may contribute more eectively to
the H-bond formation. On the other hand, the values of
HOMA estimated for H-bonded complexes show that the
aromatic character of the imidazol-2-ylidene ring increases
slightly with increasing value of the interaction energy. This
suggests that both the eect of delocalization within the ring
and the eect of H-bond formation are cooperating. It is easy
to imagine that the hydrogen bond involving the lone electron
Table 4 Values of the HOMA index estimated for the imidazol-2ylidene ring in the isolated imidazol-2-ylidene and in its complexes
with HCF3 and H2O. Data obtained for fully optimized structures at
the level of theory given in the table in conjunction with the aug-ccpVTZ basis set. For comparison, HOMA values obtained at the same
level of theory for pyrrole and imidazole are also given
Imidazol-2-ylidene
Complex
Method

Monomer

HCF3

H2O

Pyrrole

Imidazole

MP2
B3LYP

0.85
0.78

0.86
0.79

0.87
0.81

0.91
0.91

0.92
0.92

Phys. Chem. Chem. Phys., 2009, 11, 57115719 | 5717

View Online

Conclusions

Downloaded by STADT UND UNIVERSITAETS on 29 September 2011


Published on 07 May 2009 on http://pubs.rsc.org | doi:10.1039/B901968E

Fig. 6 The fully optimized structure obtained at the MP2/aug-ccpVTZ level of theory for the complex of imidazol-2-ylidene and HCF3.
Two H-bonds are linking the interacting molecules.

pair of the divalent carbon partially stabilizes the singlet state.


The cooperation of these two electronic eects may explain
why the H-bonded complexes of imidazol-2-ylidene are
relatively more stable than complexes of CCl2 and CF2.
It is worthwhile to mention that some of the imidazol-2ylidene complexes are doubly-bonded. See Fig. 6 for the
exemplary structure with two H-bonds linking the interacting
molecules. Therefore, in some cases the total interaction
energy given in Table 3 includes contributions from both
H-bonds.
The AIM-based parameters of the H-bonds in question are
collected in Table 5. Analysis of the electron density properties
of the appropriate hydrogen bonds calculated at BCPs (see
Table 5) indicates the signicant dierence with respect to
r2rBCP and HBCP in complexes of carbenes and H2O or
HCF3, respectively. For systems with HCF3 both the values
of r2rBCP and of HBCP are positive as for typical closed-shell
interactions. In the case of complexes with water molecules,
the values of HBCP are negative, although the values of r2rBCP
are simultaneously positive. Thus, a mix of shared and
closed-shell interaction is present in complexes with water.
Finally, frequency analysis shows, that in the case of
F2C  HCF3 and Cl2C  HCF3 complexes, the stretching
vibration frequency shifts of the proton-donating bond to
higher values (the so-called blue-shift) are observed. This eect
is accompanied by a signicant decrease of infrared intensity
of the stretching vibration band and by a very small decrease
in the length of the DH bond. Thus, hydrogen bonds of the
DH  C type in these systems seem to belong to a group of
the so-called improper, blue-shifting hydrogen bonds.35 For
all other carbenes investigated in this study, red shift of the
n(DH) stretching vibration frequency and an increase of IR
intensity accompany H-bond formation as is the case of
typical strong and intermediate H-bonds.
Table 5 Selected AIM electron density parameters calculated at
BCP(H  C). Data (DFT/aug-cc-pVTZ) obtained for systems with
carbene derivatives as proton acceptors. All values in au
H-acceptor

H-donor rBCP r2rBCP GBCP VBCP

CF2

H2O
HCF3
H2O
HCF3
H2O
HCF3

CCl2
C(imidazol-2-ylidene)

0.018
0.008
0.022
0.010
0.031
0.017

0.045
0.023
0.048
0.027
0.055
0.039

0.011
0.005
0.013
0.006
0.018
0.009

HBCP  103

0.011 0.028
0.004 1.089
0.014 1.138
0.005 1.032
0.021 3.720
0.008 0.714

5718 | Phys. Chem. Chem. Phys., 2009, 11, 57115719

Because of its unique electronic structure, the carbon atom in


CDPs and their derivatives as well as in carbenes may act as a
typical proton acceptor and may form typical hydrogen
bonds, in a manner similar to O and N atoms. Moreover,
the proton-accepting abilities of the carbon atom having lone
electron pair(s) are sucient to allow the formation of strong,
partially covalent H-bonds. In the case of CDPs and their
derivatives, the possibility of stabilization of the electronic
structure of C(0) atom has direct impact on the strength of the
H  C(0) bond. Since, under some conditions, the C(0) atom
tends to covalently bind the proton (and, as a result, change its
electronic structure) the less stable L2C(0) complexes form
stronger H-bonds.
In the case of carbenes, the strongest proton accepting
properties are exhibited by systems in which the singlet ground
state structure is the most stable. Imidazol-2-ylidene, in which
the empty p orbital of the divalent carbon is involved in
p-electron delocalization, acts as the strongest proton accepting
system in the investigated series. This is due to the fact that the
divalent carbon atom in carbenes may act as an eective
proton-accepting centre only in the singlet ground state. The
molecule in the triplet state, being a diradical, cannot form
stable H-bridged complexes.

Acknowledgements
The authors wish to thank Prof. Gernot Frenking for his
inspiring lecture given during the IUPAC Conference on
Physical Organic Chemistry (Santiago de Compostela, Spain,
July 2008). The authors also thank Prof. A. J. Sadlej for
helpful comments and discussions. M.J. acknowledges the
nancial support from Nicolaus Copernicus University,
Torun, Poland, by the internal research Grant No. 366-Ch.
Calculations have been carried out on the multiprocessor
cluster at the Information and Communication Technology
Center of Nicolaus Copernicus University, Torun
(http://www.uci.umk.pl), in Wroc"aw Centre for Networking
and
Supercomputing
(http://www.wcss.wroc.pl),
and
Academic Computer Centre Cyfronet AGH in Krakow
(http://www.cyf-kr.edu.pl).

References
1 G. A. Jerey, in An introduction to hydrogen bonding. Topics in
Physical Chemistry, Oxford University Press, New York, 1997.
2 G. A. Jerey and W. Saenger, in Hydrogen Bonding in Biological
Structures, Springer-Verlag, Berlin, 1991.
3 L. Sobczyk, S. J. Grabowski and T. M. Krygowski, Chem. Rev.,
2005, 105, 3513.
4 G. R. Desiraju and T. Steiner, in The Weak Hydrogen Bond in
Structural Chemistry and Biology, Oxford University Press,
New York, 1999.
5 S. J. Grabowski, in Hydrogen BondingNew Insights, ed.
J. Leszczynski, Challenges and Advances in Computational
Chemistry and Physics, Springer, 2006, vol. 3.
6 A. Baceiredo, G. Bertrand and G. Sicard, J. Am. Chem. Soc., 1985,
107, 4781.
7 A. J. Arduengo III, R. L. Harlow and M. Kline, J. Am. Chem.
Soc., 1991, 113, 2801.

This journal is


c

the Owner Societies 2009

Downloaded by STADT UND UNIVERSITAETS on 29 September 2011


Published on 07 May 2009 on http://pubs.rsc.org | doi:10.1039/B901968E

View Online

8 N-Heterocyclic Carbenes in Synthesis, ed. S. P. Nolan, Wiley-VCH,


New York, 2006; M. Dekker, in Carbene Chemistry, ed.
G. Bertrande, Springer, New York, 2002.
9 F. A. Carey and R. J. Sundberg, Carbenes, Part B: Reactions and
Synthesis, Advanced Organic Chemistry, Springer, New York,
2007.
10 C. Boehme and G. Frenking, J. Am. Chem. Soc., 1996, 118, 2039;
J. F. Harrison, R. C. Liedtke and J. F. Liebman, J. Am. Chem.
Soc., 1979, 101, 7162; P. H. Mueller, N. G. Rondan, K. N. Houk,
J. F. Harrison, D. Hooper, B. H. Willen and J. F. Liebman, J. Am.
Chem. Soc., 1981, 103, 5049.
11 N. C. Baird and K. F. Taylor, J. Am. Chem. Soc., 1978, 100, 1333.
12 R. Tonner, F. Oxler, B. Neumuller, W. Petz and G. Frenking,
Angew. Chem., Int. Ed., 2006, 45, 8038.
13 R. Tonner and G. Frenking, Chem.Eur. J., 2008, 14, 3260.
14 R. Tonner and G. Frenking, Chem.Eur. J., 2008, 14, 3273.
15 F. H. Allen, Acta Crystallogr., Sect. B: Struct. Sci., 2002, 58, 380.
CSD release; November 2007.
16 M. Movassaghi and M. A. Schmidt, Org. Lett., 2005, 7, 2453.
17 C. Mller and M. S. Plesset, Phys. Rev., 1934, 46, 618; M. HeadGordon, J. A. Pople and M. J. Frisch, Chem. Phys. Lett., 1988,
153, 503; M. J. Frisch, M. Head-Gordon and J. A. Pople, Chem.
Phys. Lett., 1990, 166, 275; M. J. Frisch, M. Head-Gordon and
J. A. Pople, Chem. Phys. Lett., 1990, 166, 281; J. Almlo,
K. Korsell and K. Faegri Jr, J. Comput. Chem., 1982, 3, 385;
M. Head-Gordon and T. Head-Gordon, Chem. Phys. Lett., 1994,
220, 122; M. Head-Gordon and J. A. Pople, J. Chem. Phys., 1988,
89, 5777.
18 A. D. Becke, J. Chem. Phys., 1993, 98, 5648; C. Lee, W. Yang and
R. G. Parr, Phys. Rev. B, 1988, 37, 785; P. J. Stephens, F. J. Devlin,
C. F. Chabalowski and M. J. Frisch, J. Phys. Chem., 1994, 98,
11623; B. Miehlich, A. Savin, H. Stoll and H. Preuss, Chem. Phys.
Lett., 1989, 157, 200.
19 R. A. Kendall, T. H. Dunning Jr. and R. J. Harrison, J. Chem.
Phys., 1992, 96, 6796; E. R. Davidson, Chem. Phys. Lett., 1996,
260, 514.
20 M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria,
M. A. Robb, J. R. Cheeseman, J. A. Montgomery, Jr.,
T. Vreven, K. N. Kudin, J. C. Burant, J. M. Millam,
S. S. Iyengar, J. Tomasi, V. Barone, B. Mennucci, M. Cossi,
G. Scalmani, N. Rega, G. A. Petersson, H. Nakatsuji, M. Hada,
M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida,
T. Nakajima, Y. Honda, O. Kitao, H. Nakai, M. Klene, X. Li,
J. E. Knox, H. P. Hratchian, J. B. Cross, V. Bakken, C. Adamo,
J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev,
A. J. Austin, R. Cammi, C. Pomelli, J. Ochterski, P. Y. Ayala,
K. Morokuma, G. A. Voth, P. Salvador, J. J. Dannenberg,
V. G. Zakrzewski, S. Dapprich, A. D. Daniels, M. C. Strain,
O. Farkas, D. K. Malick, A. D. Rabuck, K. Raghavachari,
J. B. Foresman, J. V. Ortiz, Q. Cui, A. G. Baboul, S. Cliord,
J. Cioslowski, B. B. Stefanov, G. Liu, A. Liashenko, P. Piskorz,
I. Komaromi, R. L. Martin, D. J. Fox, T. Keith, M. A. Al-Laham,

This journal is


c

the Owner Societies 2009

21
22
23
24
25
26
27
28
29
30
31

32
33
34

C. Y. Peng, A. Nanayakkara, M. Challacombe, P. M. W. Gill,


B. G. Johnson, W. Chen, M. W. Wong, C. Gonzalez and
J. A. Pople, GAUSSIAN 03 (Revision D.01), Gaussian, Inc.,
Wallingford, CT, 2004.
R. F. W. Bader, in Atoms in Molecules: A Quantum Theory, Oxford
University Press, New York, 1990.
F. W. Biegler-Konig, R. F. W. Bader and T. H. Tang, J. Comput.
Chem., 1982, 3, 317.
AIM2000 program, F. Biegler-Konig, University of Applied
Sciences, Bielefeld, Germany.
S. F. Boys and F. Bernardi, Mol. Phys., 1970, 19, 553.
S. S. Xantheas, J. Chem. Phys., 1996, 104, 8821.
S. J. Grabowski, A. J. Sadlej, W. A. Sokalski and J. Leszczynski,
Chem. Phys., 2006, 327, 151.
R. W. F. Bader and C. F. Matta, Inorg. Chem., 2001, 40, 5603.
E. Espinosa, I. Alkorta, J. Elguero and E. Molins, J. Chem. Phys.,
2002, 117, 5529; S. J. Grabowski, Annu. Rep. Prog. Chem., Sect. C:
Phys. Chem., 2006, 102, 131.
J. E. Leer, Science, 1953, 117, 340; G. S. Hammond, J. Am.
Chem. Soc., 1955, 77, 334.
M. Palusiak, S. Simon and M. Sola, Chem. Phys., 2007, 342, 43.
G. C. Pimentel and A. L. McClellan, in The Hydrogen Bond, W.H.
Freeman & Co, San Francisco, 1960; S. N. Vinogradov and
R. H. Linnell, in Hydrogen Bonding, ed. Van Nostrand-Reinhold,
Princeton, 1971; in The Hydrogen Bond, Recent Developments in
Theory and Experiments, ed. P. Schuster, S. Zundel and
C. Sandorfy, North Holland Amsterdam, vol. IIII, 1976.
We neglect singlet states that result from single occupations.
T. M. Krygowski, J. Inf. Comput. Sci., 1993, 33, 70.
The geometry-based index of aromaticity. The formula for HOMA
reads
HOMA 1 

1X
aj Ropt;j  Rj 2
n j1

where: n is the number of bonds taken into the summation; aj is a


normalization constant (for CC and CN bonds aCC = 257.7 and
aCN = 93.52) xed to give HOMA = 0 for a model non-aromatic
system and HOMA = 1 for the system with all bonds equal to the
optimal value, Ropt,j assumed to be realized for fully aromatic
systems, i.e. benzene ring (for CC and CN bonds Ropt,CC = 1.388
A and Ropt,CN = 1.334 A); Rj stands for a running bond length. In
general, the closer HOMA to value 1, the more aromatic the ring
under consideration.
35 P. Hobza and Z. Havlas, Chem. Rev., 2000, 100, 4253; P. Hobza
and Z. Havlas, Theor. Chem. Acc., 2002, 108, 325, and references
thereinA. J. Barnes, J. Mol. Struct., 2004, 704, 3, and references
therein; E. S. Kryachko, in Hydrogen Bonding: New Insights
(Challenges and Advances in Computational Chemistry and Physics),
ed. S. J. Grabowski and J. Leszczynski, , Springer, 2006, vol. 3;
J. Joseph and E. D. Jemmis, J. Am. Chem. Soc., 2007, 129, 4620.

Phys. Chem. Chem. Phys., 2009, 11, 57115719 | 5719

Das könnte Ihnen auch gefallen