Preprints201809 0228 v1
Preprints201809 0228 v1
v1
Atoms in Molecules (AIM), Natural Bond Orbital (NBO), and normal coordinate analysis have been
carried out at the global minimum structures of TH5+ (T = C/Si/Ge). All these analyses lead to a
consistent structure for these three protonated TH4 molecules. The CH5+ has a structure with three
short and two long C-H covalent bonds and no H-H bond. Hence, the popular characterization of
protonated methane as a weakly bound CH3+ and H2 is inconsistent with these results. However,
SiH5+ and GeH5+ are both indeed a complex formed between TH3+ and H2 stabilized by a tetrel bond,
with the H2 being the tetrel bond acceptor. The three-center-two-electron bond (3c-2e) in CH5+ has an
open structure, which can be characterized as a V-type 3c-2e bond and that found in SiH5+ and GeH5+
is a T-type 3c-2e bond. This difference could be understood based on the typical C-H, Si-H, Ge-H and
H-H bond energies. Moreover, this structural difference observed in TH5+ can explain the trend in
proton affinity of TH4. Carbon is selective in forming a ‘tetrel bond’ and when it does, it might be
worthwhile to highlight it as a ‘carbon bond’.
1. Introduction
The structure of a molecule provides valuable insights into its properties. CH5+ though, has
gained notoriety for eluding a formal definition of its structure. Since its discovery in 1952 in mass
spectrometric experiments [1], attempts at assigning this molecule with a definite structure have
proved to be futile. This is because CH5+ is highly fluxional with theoretical investigations revealing
that the potential energy surface (PES) is quite shallow [2,3]. The PES of CH5+ is characterized by 120
equivalent minima which are easily accessible via low lying saddle point structures [4–6]. The
difference in energy between these saddle point structures and the Cs symmetry minima is small and
decreases with increasing levels of theory [7,8]. Though there is a consensus that the molecule is
highly fluxional and the ground state minimum has Cs symmetry, results regarding the Cs symmetry
minimum veer between a structure having a pentacoordinated carbon center with no interaction
between the H atoms [9,10] and a structure where CH5+ is made up of a methyl cation (CH3+)
complexed with H2. [11,4,12] The crux of the matter then is, whether H2 retains its identity in the
molecule or not.
The nature of bonding in CH5+ can also be examined by studying the possible weak
interactions that could lead to its formation. It is known that the hydrogen bonded structures are
intermediates in proton transfer reactions. The hydronium ion (H3O+) and ammonium ion (NH4+) are
formed by the protonation of H2O and NH3, respectively. These are often mediated by a hydrogen
bonded complex such as H2O∙∙∙HX or H3N∙∙∙HX. Rotational spectroscopic investigations on several
CH4∙∙∙HX (X= F, Cl, CN, and OH) dimers [13–15] reveal that the global minimum structure is the one
where HX is the hydrogen bond donor and CH4 is the acceptor. The H atom of HX forms a hydrogen
bond with the carbon center through the tetrahedral face of CH4. It has been pointed out that this
hydrogen bonded structure having a ‘pentacoordinate carbon’ could be a precursor to the formation
of CH5+ [16]. The three O-H bond lengths in H3O+ are equal and the same is true for the four N-H
bonds in NH4+. If one were to guess the geometry of CH5+, without any prior knowledge, a trigonal
bipyramidal would be a reasonable choice. Even in this structure, one would expect two long C-H
bonds in the opposite directions and three short C-H bonds in a plane, such as in Fe(CO)5 which is
also a fluxional molecule. However, the established structure for CH5+ has Cs symmetry with three
short C-H bonds in one side (not in the same plane) and two long C-H bonds in the opposite side. It
appears as though the H2 moiety is separated from the CH3+ tripod. This arrangement then is
reminiscent of a ‘carbon (tetrel) bond’ [17,18], which was proposed recently. The positively charged
central carbon can accept electron density from the sigma electrons of H2 forming a tetrel bonded
complex. If such is the case then, a normal mode analysis on CH5+ could reveal H-H vibration and the
vibrational frequency could show a red-shift. Moreover, Atoms in Molecules (AIM) and Natural Bond
Orbital (NBO) theoretical analysis could yield evidence for the carbon bond.
The infrared spectrum obtained by Oka and coworkers [9] was an important step towards solving the
structure of CH5+. The spectrum is complicated, with nearly 900 lines that depend on CH5+. However,
none of these spectral lines could be assigned. In subsequent years, efforts were made to decipher this
spectrum by obtaining another extended spectrum [12], by computing the spectrum using quantum
calculations [5] and by obtaining the spectrum at low temperatures [19]. However, recent tentative
assignments using combination differences by Asvany et al. [19] from the low temperature spectra
and the calculated rotation- bending energy levels which were compared with this spectra [20] are the
Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 13 September 2018 doi:10.20944/preprints201809.0228.v1
closest one has come to solving the structure. This study could not give any information about H-H
vibration in CH5+.
Bader [21] opines that a distinction must be made between the molecular geometry and its structure,
as molecular geometry is a non-generic property of the molecule defined by a set of nuclear
coordinates, whereas, structure is a generic property defined by the network of bonds between atoms
in a molecule. He notes that the “difficulties ascribed to the notion of molecular structure are the
inabilities to assign a single geometrical structure…to a molecule in a ‘floppy’ state wherein the
nuclear excursions cover a wide range of geometrical parameters.” Bader proposed the use of the
charge density topology of a molecule, to determine linkages of the atoms present. This is used to
assign a molecular graph which defines the molecular structure. Therefore, the structure of CH5+
could be defined by performing a topological analysis of the electron density. Okulik et al. have in
fact carried out a topological analysis of the electron density on the CH5+ molecule and concluded that
it is a pentacoordinated carbocation with no interaction present between the H atoms [10] thereby
providing evidence against the structure of CH5+ being a complex between CH3+ and H2. Earlier
studies by Marx and Savin, subjected CH5+ to a similar analysis, using the electron localization
function (ELF), which is a local measure of the Pauli repulsion [22]. This helps locate regions having a
pair of electrons. Their findings lead to four basins, three for the 2c-2e C-H bonds and one for the 3c-
2e bond involving the H2 moiety. This does not answer the question about whether or not there is H-
H bond. Asvany et al. depict the structure of CH5+ with a 3c-2e bond with the H atoms of the H2
moiety connected [19], whereas the molecular graph for CH5+ clearly shows no interaction between
the H atoms (Figure 1). The question we then ask is, what does the molecular graph for the congeners
of CH5+, SiH5+ and GeH5+ look like? Would this comparison help in choosing the right structure for
CH5+?
Figure 1.(a) Molecular graph for CH5+; (b) CH5+as depicted in Reference 19.
Considering the difficulty in defining a structure for CH5+ it is interesting to examine the structures of
SiH5+ and GeH5+. Are they similar, considering they belong to the same group or different? SiH5+ was
first observed in ionized silane-methane mixtures using mass spectrometry by Beggs and Lampe [23].
The optimized geometries and the heats of hydrogenation for SiH3+ led Schleyer et al. to surmise that
SiH5+ is a weakly bound complex of SiH3+ and H2 [24]. The rovibrational spectrum of SiH5+ obtained by
Boo and Lee provides evidence for the H2 moiety rotating freely with respect to the SiH3+frame [25].
Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 13 September 2018 doi:10.20944/preprints201809.0228.v1
GeH5+ on the other hand was difficult to obtain and was first observed in ion beam scattering
experiments by Senzer and coworkers [26]. Kohda-Sudoh et al. described the shape of GeH5+ to be a
loose complex between the germyl cation (GeH3+) and H2 [27]. The comprehensive study of all the
structures of the GeH5+ by Schreiner et al. concluded that the geometry having Cs symmetry was the
minimum on the potential energy surface [28]. It is evident that the ground state minimum structure
for the TH5+ molecule has Cs symmetry.
Theoretical and spectroscopic results for SiH5+ and GeH5+ suggest that these are weakly bound
complexes of the TH3+ (T= Si, and Ge) cation with H2. The topological analysis of the charge density
would then allow us to observe a tetrel bonded interaction between the TH3+ cation and the H2
moiety. This has not been carried out so far to the best of our knowledge. In this work we compare
the electron density topologies of SiH5+ and GeH5+ with CH5+ and investigate if there is any evidence
for a tetrel bond. Natural bond orbital theory and normal mode analysis on these structures have
been carried out as well. This has helped in arriving at unambiguous conclusions about the structure
of CH5+. The results are presented here.
2. Computational methods
The geometries of the TH5+ species (T= C, Si, and Ge) were optimized at the MP2 level of theory with
the aug-cc-pVTZ basis set. Frequency calculations were also performed to ascertain the nature of the
optimized geometries. The absence of an imaginary frequency confirms a structure to be a minimum
and its presence indicates that the structure is a saddle point on the potential energy surface. The
harmonic frequencies were also calculated using the B3LYP functional with the same basis set.
Relaxed potential energy scans were carried out by varying the distance between the TH3+ moiety and
H2. This was computed using the keyword opt=modredundant. This allows all the parameters
defining the molecule to alter during a scan. All calculations were performed using the Gaussian 09
suite of programs [29].
To establish the nature of bonding, the topological analysis of the electron density for the optimized
structures of the TH5+ was done using the AIMAll software [30]. It was also analyzed at select points
of the potential energy scans to study the appearance or disappearance of various weak interactions
and their role in the formation of the TH5+ species. The wave function required for this calculation
was obtained from the Gaussian computation at MP2/aug-cc-pVTZ level. The presence of the 3c-2e
bonds and the nature of the bonding orbitals involved in the TH5+ species were analyzed using
NBO6.0 as implemented in Gaussian 09 [31].
3. Results
3.1.1 CH5+
Methane distorts to accommodate the presence of the added proton to form CH5+. The CH5+
molecule was optimized at MP2/ aug-cc-pVTZ. The three lowest energy stationary structures
obtained are shown below in Figure 2. The geometry with Cs(I) symmetry is a minimum on the
potential energy surface which is confirmed by the vibrational frequency analysis. The CH3 unit of the
Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 13 September 2018 doi:10.20944/preprints201809.0228.v1
molecule is more pyramidal than planar. Thus, CH5+ bears resemblance to a protonated methane
structure rather than to a planar methyl cation (CH3+) interacting with an elongated H2 molecule.
Figure 2. Optimized geometries and relative energies of three lowest energy stationary structures of
CH5+ calculated at MP2/aug-cc-pVTZ. Number in brackets denotes number of imaginary frequencies.
A quantitative evaluation of the bond lengths reveals the presence of three short C-H bonds
comprising the CH3 tripod and two long C-H bonds which make up the H2 moiety. These two C-H
bonds are about 0.1Å longer than the three other C-H bonds. Therefore, it appears as though the H-
atoms involved in the long C-H bonds are a separate entity with respect to the CH3 tripod. One of the
three short C-H bonds is slightly longer (marked b in Figure 3) and it is with respect to this unique
bond that the H2 unit is eclipsed in the Cs (I) geometry. The distance between the H atoms in the H2
moiety is 0.9748 Å. When compared to the H-H distance in free H2, which is 0.7374 Å, it seems that
this bond is quite elongated with possibly no interaction between the H atoms. The C-H and H-H
bond lengths are summarized in Tables 1 and 2.
The second stationary structure is a saddle point on the potential energy surface, confirmed by the
presence of an imaginary frequency. This structure has Cs (II) symmetry with the H2 moiety staggered
with respect to the unique C-H bond (b). The Cs (II) saddle point is the transition state for the rotation
of the H2 unit about the pseudo C3 axis in the methanium ion as noted in earlier works [4-8]. The
energies computed at MP2/ aug-cc-pVTZ indicate that this structure is only 0.59 kJmol-1 higher in
energy with respect to the minimum Cs(I) structure. Therefore, the rotation of the H2 moiety about the
pseudo C3 axis would be facile.
The third structure is also a saddle point on the potential energy surface. Three of the C-H bonds in
this structure are in a plane and equidistant from each other. This C2v symmetry structure is the
transition state on the reaction coordinate where the H atom flips between the H2 moiety and the CH3
unit. This structure is 2.33 kJ mol-1 higher in energy than the minimum Cs(I) structure. Thus, hydrogen
scrambling via these two saddle point structures is quite facile. CH5+ therefore, easily interconverts
between its 120 equivalent minima, supporting the established fact that it is a highly fluxional
molecule.
The congeners of CH5+, SiH5+and GeH5+ were also optimized at MP2/aug-cc-pVTZ level. The
optimized geometries of TH5+ (T= Si and Ge) resemble a planar TH3+ moiety complexed with an
elongated H2 unit. The Cs (I) structure has the H2 unit eclipsed with the T-H bond (b) and is the
minimum on the potential energy surface. The Cs (II) structure has the H2 unit staggered with respect
to the T-H bond (b). It is energetically quite close to the Cs (I) structure and is a first order saddle point
having one imaginary frequency. This Cs (II) symmetry saddle point is the transition state via which
rotation of the H2 moiety about the TH3+ frame occurs. However, in the case of SiH5+ and GeH5+ the C2v
structure is energetically higher than the dissociation energy of the complex and thus complete
hydrogen scrambling is not facile [32,28].
Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 13 September 2018 doi:10.20944/preprints201809.0228.v1
The TH3+ unit is nearly planar though the T-H bonds are not all equal. The tabulated bond distances
(Tables 1 and 2) show that the T-H bonds in the TH5+ molecule are slightly distorted from those
computed for TH3+. Comparing the H-H bond distances for the H2 moiety in TH5+ with the free H2
molecule, it seems that the H2 unit in TH5+ is slightly elongated. The H-H distance in SiH5+ is 0.7727Å
and in GeH5+ it is 0.7697Å, these are only slightly longer than the H-H distance in the free H2 molecule
which is 0.7374Å. Therefore, the H2 moiety in TH5+ seems to retain its identity in these molecules.
Bader’s ‘Atoms in Molecules’ (AIM) theory examines the topology of the electron density to
understand the nature of bonding in molecules. The interaction between any two atoms is confirmed
by the presence of a bond critical point (BCP) between them. The properties of the electron density at
the BCP provide further insights about the bond. These topological properties for the TH5+ (T= C, Si,
and Ge) and C2H7+ are tabulated in Tables 3 and 4. The molecular graphs obtained from the AIM
analysis are shown in Figures 4 and 5.
3.2.1. CH5+
The molecular graphs for the three lowest lying stationary structures of CH5+ are shown in Figure 4.
The Cs(I) symmetry minimum structure is characterized by the presence of the five bond critical
points (BCPs) which connect each of the H atoms to the C center. The value of the electron density (ρ)
at the five BCPs ranges from 0.2174 to 0.2911 au and the Laplacian of electron density (∇2ρ) from -
0.5017 to -1.2082 au. When ∇2ρ at the BCP is negative and the ρ is large in magnitude, these
interactions are referred to as shared interactions, which is characteristic of a covalent bond. Therefore,
CH5+ is pentacoordinated with five covalent C-H bonds. The most striking observation however, is
the absence of a bond critical point between the H atoms of the H2 moiety. The two long C-H bonds
that make up the H2 moiety have very high ellipticity values (2.1739 and 1.3412) which suggests that
these bonds are unstable. It is quite evident that CH5+ cannot be considered as a weakly bound
complex of CH3+ and H2. These results agree with those obtained previously by Okulik et al. using the
6-311++G** basis set[10].
Figure 4. Molecular graphs for the three lowest energy stationary structures of CH5+calculated at
MP2/aug-cc-PVTZ
Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 13 September 2018 doi:10.20944/preprints201809.0228.v1
The molecular graphs for SiH5+ and GeH5+ are quite different from that of CH5+ (Figure 5). It is
evident that the slightly elongated H2 moiety retains its identity. A bond critical point (BCP) is present
between the two H atoms. The large magnitude of ρ and the negative sign of ∇2ρ at the BCP (Table 4)
indicate that a covalent bond is present between the two H atoms. The ρ value in GeH5+ is larger than
in SiH5+ which corroborates well with the computed H2 bond lengths (Table 4).
The other prominent difference is the interaction between the planar cation and the H2 moiety. The
molecular graphs for SiH5+ and GeH5+ are best described as conflict structures. This is because the
bond critical point between the two moieties connects the central T atom not with another atom but
with the BCP between the two H atoms. This could be viewed as a bond between the σ bond electrons
of H2 and the positive central T atom. Low values of ρ at BCP and ∇2ρ at BCP being positive indicates
a closed-shell interaction. Such an interaction is usually found in ionic bonds, hydrogen bonds, and van
der Waals molecules.
The BCPs and bond paths clearly indicate the presence of the silyl/germyl cation interacting with an
H2 moiety. This interaction is indicative of a tetrel bonded interaction. The ρ value for SiH5+(0.0465 au)
and GeH5+(0.0527 au) is slightly higher than Koch and Popelier’s criterion for a C-H…O hydrogen
bond [33] (0.002-0.034 au). Clearly, the tetrel bond in these cations are stronger than the strongest C-
H…O hydrogen bond discussed in Koch and Popelier’s work. The Laplacian on the other hand for
SiH5+ and GeH5+ is 0.1194 and 0.1267 respectively. These values are within the range for the criterion
proposed by Koch and Popelier (0.024-0.139). Sosa and coworkers have suggested that the |λ1|/ λ3
Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 13 September 2018 doi:10.20944/preprints201809.0228.v1
ratio must be less than 1 for a closed-shell interaction [34]. This ratio for SiH5+ and GeH5+ satisfy this
criterion. The ellipticity values for these BCPs are quite large which is indicative of an unstable bond
between these two units.
Table 3. Properties of the electron density at the BCP between the central atom and the H2 moiety
Table 4. Properties of the electron density at the BCP between the H2moiety
Would the cations of larger alkane reveal a tetrel bond? In order to address this question, AIM
analysis for C2H7+ was carried out. The result is shown in Figure 6. There is a BCP between the two H
atoms in C2H7+, though the bond path is curved. This BCP appears very close to another BCP,
connecting the H2 to the C atom. A look at the bond ellipticity value at both these BCPs indicates that
they are highly unstable (32.27 and 6.79) and therefore the molecular graph for this structure is highly
unstable. In any case, this result indicates that there could be a tetrel bond between C and H2 in
favorable cases.
TH5+ (T= C, Si, and Ge) shows the presence of a 3c-2e H-T-H bond in the Natural Bond Orbital (NBO)
analysis. In combination with the molecular graphs obtained from the AIM analysis, the 3c-2e bond in
the case of CH5+ can be considered to be ‘open’ or ‘V’ type since there is no interaction between the
Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 13 September 2018 doi:10.20944/preprints201809.0228.v1
two H atoms. Whereas in the case of SiH5+ and GeH5+ we can consider it to be a ‘closed’ or ‘T’ type 3c-
2e interaction since there is an H-H bond. A look at the nature of the orbitals reveals that the long C-H
bonds in CH5+use sp3 orbitals from C and the equivalent bonds in SiH5+ and GeH5+ use an empty p
orbital from the central atom which interacts with the H2 unit. The charge analyses in both AIM and
NBO reveal some interesting observations. The charge on the C atom in CH5+ is negative and the H
atoms carry the positive charge. This therefore reinforces the idea that CH5+ is more like protonated
methane than a CH3+∙∙∙H2 complex. The charge analysis for SiH5+ and GeH5+ on the other hand reveals
that the central T atom is positively charged and the two hydrogen atoms are negatively charged,
indicating that it is TH3+ and H2.
Both AIM and NBO analyses do not give any evidence for the H---H bond in CH5+ and both do give
evidence for the same in TH5+ (T'= Si and Ge). Would the normal mode analysis on these three cations
provide additional evidence about the bonding? Clearly TH5+ has been considered as a weakly bound
complex, with the central T atom tetrel bonded to the H2 moiety. The positively charged central T
atom can therefore accept electron density from the σ electrons of the H2 moiety. This donation of
electron density should lead to the weakening of the H2 bond. The electron density (ρ) obtained from
the AIM analysis shows that the H2 bond is weakened more for SiH5+ than for GeH5+. An elongation of
the H2 bond length is expected as a consequence of the donation of electron density to the T center.
The tabulated bond lengths show that the H2 bond lengths in TH5+ are elongated compared to free H2.
The H2 bond in SiH5+ is longer than in GeH5+, this corroborates well with the computed electron
density. It is evident that the H2 unit retains its identity in the molecule. We therefore expect to find a
normal mode in TH5+ which comprises of the H2 stretch. We then anticipate that the weakening and
elongation of the bond would cause this H2 stretch frequency to be red shifted. The computed
harmonic frequencies are tabulated in Table 5. The H2 stretch is red shifted by 466.65 cm-1 in SiH5+ and
by 440.23 cm-1 in GeH5+. Boo and Lee obtained theinfrared spectrum of the H–H stretching mode in
SiH5+ from 3650-3740 cm-1, [25] reasonably close to the predicted value of 3800 cm-1. In contrast there is
no normal mode in CH5+ which can be classified as H2 stretch. There are two normal modes which
involve the two H atoms and the central C atom. These are a symmetric and an antisymmetric stretch
for the H-C-H unit. This provides further evidence that CH5+ has two long and equivalent bonds but it
is not a weak complex between CH3+and H2.
Table 5.Harmonic frequencies of the H-H stretch calculated at MP2/ aug-cc-pVTZ and at B3LYP, also
the scaled frequencies are reported. All values in cm-1.
As discussed earlier, CH5+ is commonly considered protonated methane and H4C∙∙∙HX complex could
be thought of as the hydrogen bonded complex leading to it. However, the structure at the global
minimum shows similarities with a weakly bound complex between CH3+ and H2. All the results
presented above clearly indicate that CH5+ neither has a tetrel bond or a bond between the two H
atoms which are at a longer distance from C compared to the other three H atoms. Would the
reaction path between CH5+ and CH3+ + H2 indicate the presence of a tetrel bond anywhere along the
reaction coordinate? To address this question, the potential energy scan was done at MP2/aug-cc-
pVDZ level by varying the distance between the TH3+ unit and the H2 moiety. The AIM analysis was
carried out at various points along the reaction coordinate. The results for CH5+ are shown in Figure 7
and the results for SiH5+ and GeH5+ are shown in the Supporting Information.
In all these cases the AIM analysis indicates that there is no tetrel bond formation between H2 and
TH3+ as the H2 approaches TH3+ until very close to the minimum. In the CH5+ case the tetrel bond is
not found anywhere along this coordinate and for SiH5+ and GeH5+, only at the minimum a tetrel bond
is seen. In all cases one of the hydrogen atoms is interacting with the central atom along the reaction
coordinate and this interaction is likely the driving force for the reaction. There is a bond path
connecting one hydrogen atom to the central atom which may be considered as a tetrel bond.
Figure 7. Potential energy scan for changing the distance between the CH3+ and H2 moieties
Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 13 September 2018 doi:10.20944/preprints201809.0228.v1
4. Discussions
The Cs (I) symmetry is the global minimum structure for TH5+. The H2 moiety retains its identity in
SiH5+and GeH5+ though the H2 bond is slightly elongated. In the case of CH5+ the H2 moiety is
elongated to such an extent that it can no longer be considered as a separate unit. The two long C-H
bonds that make up the H2 moiety are similar to the other C-H bonds of the CH3+ tripod and the AIM
parameters at all their BCPs clearly show that all five C-H bonds are covalent, though two of them are
slightly weaker than the other three.
The AIM analysis clearly shows that the CH5+ has five covalent C-H bonds and the TH5+(T = Si/Ge)
molecules are a weakly bound complex of the TH3+ cation and H2. The SiH5+and GeH5+ are bound by a
tetrel bond. The NBO analysis shows that all the TH5+ molecules have a 3c-2e bond and the AIM
analysis points out significant differences between CH5+ and SiH5+/GeH5+. We propose that the 3c-2e
bonds in conjunction with the molecular graph obtained from the AIM analysis can be classified as a
‘V’ or ‘T’ type 3c-2e bond. Naturally, some questions arise. Is there any molecular system having a
3c-2e bonds in which every atom is connected by a bond to the other two atoms, such as ‘Δ’? Or is
there a system where the three atoms are arranged linearly with the central atom bonded to the other
two atoms? Though 3c-2e bonds have been discussed extensively [36–38] and it is a text-book
material, we could not find such clear classifications anywhere.
For the sake of completeness, AIM analysis was carried out on well known examples such as C3H3+
and H3+, which are expected to be Δ-type 3c-2e bonds. The results are shown in Figure 8. The C3H3+
does show a Δ-type 3c-2e bond and H3+ is different from all the cases discussed so far and the
molecular graph looks like Y. Interestingly as shown in the previous section C2H7+ shows a Y type 3c-
2e bond. Jensen had drawn a Y type bond for CH5+[37], based on intuition, unlike the V type bond
that has been observed in this work and also in the previous AIM analysis by Okulik et al. Jemmis,
Chandrashekar and Schleyer have investigated linear 3c-2e bonds in CH3Li2+ and other similar
CH3M2+ systems, pointing out the role these structures play in the stereochemistry of the SE2 reactions
[39]. Clearly, describing the bonding in any molecular system as 3c-2e bond does not reveal much
about how the three centres are bonded. The 3c-2e bonds could be V-type, T-type, Δ-type, Y-type or
linear. A more detailed description of 3c-2e bond is beyond the scope of this work and clearly would
be a digression on the main focus of this work i.e. structure and bonding in TH5+.
The harmonic frequencies for the TH5+ molecules show that the H2 moiety retains its identity in
SiH5+and GeH5+. The H2 stretching frequency in these molecules is red shifted from the free H2
frequency because of the elongation in the H2 bond. CH5+ is distinctly different with no pure H2
vibration. C2H7+ which is the higher homologue of CH5+ has an H-H interaction as evident from its
molecular graph, but the frequency analysis shows that there is no normal mode which corresponds
to a pure H-H stretch. This indicates that the normal modes of the Y-type 3c-2e bonds could be
different from those of the T-type. Therefore, though it seems by examining the molecular graphs
that only CH5+ is different from its congeners and its homologues, the frequency analysis provides
evidence that the carbocations (CH5+ and C2H7+) are different from their congeners (SiH5+and GeH5+).
Why is the structure of CH5+ distinctly different from that of its congeners, SiH5+ and GeH5+? This
question can be answered by analyzing all the bond energies involved as well as the dissociation
energies for the TH5+ molecules. The relevant bond energies are given in Tables 6 and 7. If we
consider the formation of the CH5+ molecule to be the hydrogenation of the CH3+ species, it is evident
that two C-H bonds are formed at the expense of one H-H bond. The C-H and H-H bond energies are
nearly equal and the two weak C-H bonds can be easily formed by cleaving the H-H bond in H2. This
arrangement is more stable than having a slightly weaker H-H bond and a tetrel bond with the C
atom in CH3+. This is not the case with SiH5+ and GeH5+, as the Si-H and Ge-H bonds are significantly
weaker than H-H bond. The formation of two weak T-H bonds cannot compensate for the cleavage of
the H-H bond and therefore SiH5+ and GeH5+ prefer to form a tetrel bonded complex of the TH3+
species with H2. In general, Si and Ge form a stronger tetrel bond than C [18]. The combination of the
tetrel bond and H-H bond is energetically more preferred than having two weak T-H bonds, when T
= Si/Ge. The difference between the structures of CH5+ and SiH5+ has been rationalized on the basis of
their bond energies by Schleyer et al.[24]. Our analysis includes the bond energies of the T-H bonds
and that of the tetrel bond between T+ and H2, vide infra.
The proton affinities of TH4 are given in Table 6 as well. At first look, it appears counterintuitive to
find the order CH4< SiH4< GeH4 given the reverse order for the T-H bond energy. The proton affinity
of CH4 is about 100 kJ mol-1 less than that of SiH4, which is about 40 kJ mol-1 less than that of GeH4.
This in a way provides further evidence to the structure of TH5+. While the protonation of CH4 results
in the formation of two relatively weaker C-H bonds, protonation of SiH4 and GeH4 results in the
formation of H-H bond and a tetrel bond.
Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 13 September 2018 doi:10.20944/preprints201809.0228.v1
The dissociation energies calculated for the TH5+ molecules going to TH3+ and H2 are given in Table 7.
The CH5+ requires the most energy (-182.99 kJ mol-1) to dissociate into a methyl cation and H2. It is
higher by nearly 133 kJ mol-1 when compared with SiH5+. However, this is not the ‘tetrel bond energy’
between CH3+ and H2 as it does not exist in CH5+. This high energy is due to the fact that the two C-H
bonds need to be broken before forming the H-H bond. The SiH5+ and GeH5+ have nearly equal
dissociation energies (49.7 and 43.8 kJmol-1, respectively) and these can be readily identified as the
tetrel bond energy. These values are larger than the typical tetrel bond energies for neutral molecules,
typically less than 16 kJmol-1 [17,18]. Our results are compared with previous calculated and
experimental values in Table 7 and the trends are similar.
Table 6. T-HBond energies and proton affinities for the TH4 molecule in kJmol-1.
5. Conclusions
The structures of TH5+ (T = C/Si/Ge) have been analyzed based on AIM, NBO and normal coordinate
analysis. All these results give a consistent picture. The CH5+ has a pentacoordinate carbon, having
three shorter C-H bonds and two longer C-H bonds. Both SiH5+ and GeH5+ have a structure in which
the TH3+ is tetrel bonded to H2. Bond energies, proton affinities and hydrogenation energies are all
consistent with these structures. We conclude that the common description of CH5+ as a complex
between CH3+ and H2 is not consistent with all the results presented here. Thus, CH5+ has no hydrogen
bond or carbon bond, and SiH5+ and GeH5+ have a tetrel bond.
Acknowledgements
The authors thank the Inorganic and Physical Chemistry Department, IISc and Prof. Sai G. Ramesh
for the use of computational facilities. S. P. G thanks Council of Scientific and Industrial Research for
a fellowship.
Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 13 September 2018 doi:10.20944/preprints201809.0228.v1
References
1. Tal’roze, V.; Lyubimova, A. Dokl. Akad.Nauk SSSR. In; 1952; Vol. 86, p. 909.
2. McCoy, A. B.; Braams, B. J.; Brown, A.; Huang, X.; Jin, Z.; Bowman, J. M. Ab Initio
Diffusion Monte Carlo Calculations of the Quantum Behavior of CH5+ in Full
Dimensionality. J. Phys. Chem. A2004, 108, 4991–4994, doi:10.1021/jp0487096.
3. Jin, Z.; Braams, B. J.; Bowman, J. M. An ab Initio Based Global Potential Energy Surface
Describing CH5+→ CH3+ + H2+.J. Phys. Chem. A2006, 110, 1569–1574,
doi:10.1021/jp053848o.
4. Kolbuszewski, M.; Bunker, P. R. Potential barriers, tunnelingsplittings, and the
predicted J =1←0 spectrum of CH5+. J. Chem. Phys.1996, 105, 3649–3653,
doi:10.1063/1.472210.
5. Huang, X.; McCoy, A. B.; Bowman, J. M.; Johnson, L. M.; Savage, C.; Dong, F.; Nesbitt,
D. J. Quantum deconstruction of the infrared spectrum of CH5+. Science2006, 311, 60–63.
6. Tian, S. X.; Yang, J. Driving Energies of Hydrogen Scrambling Motions in CH5+. J. Phys.
Chem. A2007, 111, 415–418, doi:10.1021/jp067450j.
7. Schreiner, P. R.; Kim, S.; Schaefer, H. F.; Schleyer, P. von R. CH5+ : The never-ending
story or the final word? J. Chem. Phys.1993, 99, 3716–3720, doi:10.1063/1.466147.
8. Müller, H.; Kutzelnigg, W.; Noga, J.; Klopper, W. CH5+: The story goes on. An explicitly
correlated coupled-cluster study.J. Chem. Phys.1997, 106, 1863–1869,
doi:10.1063/1.473340.
9. White, E. T.; Tang, J.; Oka, T. CH5+: the infrared spectrum observed. Science1999, 284,
135–137.
10. Okulik, N. B.; Peruchena, N. M.; Jubert, A. H. Three-Center−Two-Electron and Four-
Center−Four-Electron Bonds. A Study by Electron Charge Density over the Structure of
Methonium Cations.J. Phys. Chem. A2006, 110, 9974–9982, doi:10.1021/jp063709m.
11. Marx, D.; Parrinello, M. Structural quantum effects and three-centre two-electron
bonding in CH5+. Nature1995, 375, 216.
12. Asvany, O.; Hegemann, I.; Schlemmer, S.; Marx, D. Understanding the Infrared
Spectrum of Bare CH5+. 2005, 309, 5.
13. Legon, A. C.; Roberts, B. P.; Wallwork, A. L. Rotational spectra and geometries of the
gas-phase dimers (CH4,HF) and (CH4,HCl). Chem. Phys. Lett.1990, 173, 107–114,
doi:10.1016/0009-2614(90)85312-Z.
14. Legon, A. C.; Wallwork, A. L. The pairwise interaction of methane with hydrogen
cyanide: a surprising result from rotational spectroscopy. J. Chem. Soc. Chem.
Commun.1989, 588, doi:10.1039/c39890000588.
15. Suenram, R. D.; Fraser, G. T.; Lovas, F. J.; Kawashima, Y. The microwave spectrum of
CH4 – –H2O.J. Chem. Phys.1994, 101, 7230–7240, doi:10.1063/1.468280.
16. Raghavendra, B.; Arunan, E. Hydrogen bonding with a hydrogen bond: The methane–
water complex and the penta-coordinate carbon. Chem. Phys. Lett.2008, 467, 37–40.
17. Mani, D.; Arunan, E. The X–C⋯ Y (X= O/F, Y= O/S/F/Cl/Br/N/P)‘carbonbond’and
hydrophobic interactions. Phys. Chem. Chem. Phys.2013, 15, 14377–14383.
18. Grabowski, S. J. Tetrel bond–σ-hole bond as a preliminary stage of the SN2 reaction.
Phys. Chem. Chem. Phys.2014, 16, 1824–1834.
19. Asvany, O.; Yamada, K. M. T.; Brunken, S.; Potapov, A.; Schlemmer, S. Experimental
ground-state combination differences of CH5+. Science2015, 347, 1346–1349,
doi:10.1126/science.aaa3304.
Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 13 September 2018 doi:10.20944/preprints201809.0228.v1
20. Wang, X.-G.; Carrington, T. Calculated rotation-bending energy levels of CH5+ and a
comparison with experiment. J. Chem. Phys.2016, 144, 204304, doi:10.1063/1.4948549.
21. Bader, R. F. W. Atoms in Molecules: A Quantum Theory; International Series of
Monographs on Chemistry; Oxford University Press: Oxford, New York, 1994; ISBN
978-0-19-855865-1.
22. Marx, D.; Savin, A. Topological Bifurcation Analysis: Electronic Structure of CH5+.
Angew. Chem. Int. Ed. Engl.1997, 36, 2077–2080.
23. Beggs, D.; Lampe, F. SiH5+ Formation in Ionized Silane–Methane Mixtures. J. Chem.
Phys.1968, 49, 4230–4231.
24. Schleyer, P. von R.; Apeloig, Y.; Arad, D.; Luke, B. T.; Pople, J. A. The structure and
energy of SiH5+.comparisons with CH5+ and BH5. Chem. Phys. Lett.1983, 95, 477–482,
doi:10.1016/0009-2614(83)80336-4.
25. Boo, D. W.; Lee, Y. T. Infrared spectroscopy of the siliconium ion, SiH5+. J. Chem.
Phys.1995, 103, 514–519, doi:10.1063/1.470137.
26. Senzer, S. N.; Abernathy, R. N.; Lampe, F. W. GeH5+ and the proton affinity of
monogermane. J. Phys. Chem.1980, 84, 3066–3067, doi:10.1021/j100460a018.
27. Kohda-Sudoh, S.; Ikuta, S.; Nomura, O.; Katagiri, S.; Imamura, M. Proton affinity of
GeH4 and the shape of GeH5+. J. Phys. B At. Mol. Phys.1983, 16, L529–L531,
doi:10.1088/0022-3700/16/17/005.
28. Schreiner, P. R.; Schaefer, H. F.; Schleyer, P. von R. The structures, energies, vibrational,
and rotational frequencies, and dissociation energy of GeH5+.J. Chem. Phys.1994, 101,
2141–2147, doi:10.1063/1.467720.
29. Frisch, M.; Trucks, G.; Schlegel, H.; Scuseria, G.; Robb, M.; Cheeseman, J.; Scalmani, G.;
Barone, V.; Mennucci, B.; Petersson,. Nakatsuji, G. H; Li, X.; Caricato, M.; Marenich, A.;
Bloino, J.; Janesko, B. G.; Gomperts, R.; Mennucci, B. ; Hratchian, H. P.; Ortiz, J. V. ;
Izmaylov, A. F.; Sonnenberg, J. L.; Williams-Young, D.; Ding, F.; Lipparini, F.; Egidi, F.;
Goings, J. ; Peng, B.; Petrone, A.; Henderson, T.; Ranasinghe, D.; Zakrzewski, V. G.;
Gao, J.; Rega, N.; Zheng, G.; Liang, W.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.;
Hasegawa, J.; Ishida, M. ; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.;
Throssell, K.; Montgomery, Jr., J. A.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.;
Brothers, E. ; Kudin, K. N. ; Staroverov, V. N.; Keith, T.; Kobayashi, R.; Normand, J.;
Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Millam,
J. M.; Klene, M.; Adamo, C.; Cammi, R. ; Ochterski, J. W.; Martin, R. L.; Morokuma, K. ;
Farkas, O.; Foresman,J. B. ; Fox,D. J. Gaussian 09 Revision D. 01, 2009. Gaussian Inc
Wallingford CT2009.
30. Keith, T. A. AIMAll (Version 16.05.18),TK Gristmill Software. Overland Park KS USA
(aim.tkgristmill.com)2016.
31. Glendening, E. D.; Badenhoop, J. K.; Reed, A. E.; Carpenter, J. E.; Bohmann, J. A.;
Morales, C. M.; Landis, C. R.; Weinhold, F. NBO6. 0.Theoretical Chemistry Institute,
University of Wisconsin, Madison, WT (nbo6.chem.wisc.edu)2013.
32. Hu, C.-H.; Shen, M.; Schaefer, H. F. Toward the infrared spectroscopic observation of
SiH5+: the silanium ion. Chem. Phys. Lett.1992, 190, 543–550, doi:10.1016/0009-
2614(92)85189-H.
33. Koch, U.; Popelier, P. L. Characterization of CHO hydrogen bonds on the basis of the
charge density. J. Phys. Chem.1995, 99, 9747–9754.
Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 13 September 2018 doi:10.20944/preprints201809.0228.v1
34. Amezaga, N. J. M.; Pamies, S. C.; Peruchena, N. M.; Sosa, G. L. Halogen Bonding: A
Study based on the Electronic Charge Density. J. Phys. Chem. A2010, 114, 552–562,
doi:10.1021/jp907550k.
35. Hydrogen Available online:
https://webbook.nist.gov/cgi/cbook.cgi?ID=C1333740&Mask=1000 (accessed on Sep 10,
2018).
36. Pitzer, K. S. Electron deficient molecules. i. the principles of hydroboron structures. J.
Am. Chem. Soc.1945, 67, 1126–1132.
37. Jensen, W. B. Extending ball and stick models by using three-center, two-electron
bonding components. J. Chem. Educ.1980, 57, 637.
38. DeKock, R. L.; Bosma, W. B. The three-center, two-electron chemical bond.J. Chem.
Educ.1988, 65, 194.
39. Jemmis, E. D.; Chandrasekhar, J.; Schleyer, P. von R. Stabilization of
D3hpentacoordinatecarbonium ions. Linear three-center-two-electron
bonds.Implications for aliphatic electrophilic substitution reactions.J. Am. Chem.
Soc.1979, 101, 527–533, doi:10.1021/ja00497a004.
40. Darwent, B. deB. Bond dissociation energies in simple molecules. Nat. Stand. Ref. Data
Ser., Nat. Bur. Stand. (U.S.), 31, 1970
41. Bohme, D. K.; Mackay, G. I.; Schiff, H. I. Determination of proton affinities from the
kinetics of proton transfer reactions. VII. The proton affinities of O2, H2, Kr, O, N2, Xe,
CO2, CH4, N2O, and CO. J. Chem. Phys.1980, 73, 4976–4986, doi:10.1063/1.439975.
42. Cheng, T. M. H.; Lampe, F. W. SiH5+ and the proton affinity of monosilane. Chem. Phys.
Lett.1973, 19, 532–534, doi:10.1016/0009-2614(73)85141-3.