Surabhi
Gupta
a,
Charlotte N.
Cummings
b,
Nicholas R.
Walker
b and
Elangannan
Arunan
*a
aDepartment of Inorganic and Physical Chemistry, Indian Institute of Science, Bangalore, 560012, India. E-mail: arunan@iisc.ac.in
bChemistry-School of Natural and Environmental Sciences, Newcastle University, Bedson Building, Newcastle-upon-Tyne NE1 7RU, UK
First published on 4th July 2024
The microwave spectra of five isotopologues of phenylacetylene⋯methanol complex, C6H5CCH⋯CH3OH, C6H5CCH⋯CH3OD, C6H5CCH⋯CD3OD, C6H5CCD⋯CH3OH and C6H5CCH⋯13CH3OH, have been observed through Fourier transform microwave spectroscopy. Rotational spectra unambiguously unveil a specific structural arrangement characterised by dual interactions between the phenylacetylene and methanol. CH3OH serves as a hydrogen bond donor to the acetylenic π-cloud while concurrently accepting a hydrogen bond from the ortho C–H group of the PhAc moiety. The fitted rotational constants align closely with the structural configuration computed at the B3LYP-D3/aug-cc-pVDZ level of theory. The transitions of all isotopologues exhibit doublets owing to the methyl group's internal rotation within the methanol molecule. Comprehensive computational analyses, including natural bond orbital (NBO) analysis, atoms in molecules (AIM) theory, and non-covalent interactions (NCI) index plots, reveal the coexistence of both O–H⋯π and C–H⋯O hydrogen bonds within the complex. Symmetry adapted perturbation theory with density functional theory (SAPT-DFT) calculations performed on the experimentally determined geometry provide an insight into the prominent role of electrostatic interactions in stabilising the overall structural arrangement.
For the PhAc⋯H2O complex, it is evident that there is a clear preference for water to bind with the acetylenic moiety (rather than the phenyl ring) while acting as a hydrogen bond donor, as proposed by IR/UV double resonance experiment and confirmed by microwave experiment.7,10 This interaction is accompanied by a secondary interaction involving the ortho C–H group of PhAc and the oxygen atom of water. In contrast, in the PhAc⋯H2S complex, microwave spectroscopic experiment showed a structure in which, hydrogen sulphide is located above the phenyl ring and donates a hydrogen bond to the π-cloud of the phenyl ring.9 In order to obtain a deeper understanding of the effects of replacing water with alcohols as the donor molecule in the complexes formed with PhAc, an IR-UV double resonance spectroscopic method was utilized.3,4
During their investigations using the IR-UV double resonance spectroscopic technique, Singh et al. reported some intriguing differences in hydrogen bonding tendencies exhibited by H2O and CH3OH when interacting with PhAc.4 H2O showed a preference for forming hydrogen bonds with the acetylenic π-cloud of PhAc, while CH3OH donated hydrogen bonds to the phenyl π-system. These observations were derived from fluorescence dip infrared (FDIR) spectra obtained for H2O and CH3OH complexes with PhAc. To further substantiate the intriguing methyl group-induced hydrogen bond switching phenomenon, they conducted energy decomposition analyses employing symmetry-adapted perturbation theory (SAPT). The SAPT calculations brought to light a disparity in energy contributions between the PhAc⋯H2O and PhAc⋯CH3OH complexes. In contrast to the predominantly electrostatic energy component characterizing the PhAc⋯H2O complex, the PhAc⋯CH3OH complex exhibited dispersion as its primary energy-driving force. Given the similarity in mass between CH3OH and H2S, it was inferred that dispersion plays a dominant role in both PhAc⋯CH3OH and PhAc⋯H2S complexes. A delicate balance of interactions, which encompasses a range of forces such as electrostatic interactions between permanent multipoles, induction, dispersion, and exchange-repulsion, plays an important role in regulating intricate molecular recognition events that occur in chemical and biological systems. Consequently, critical significance lies in the experimental investigation and accurate estimation of the structures of these complexes at their respective global minima.
Surprisingly, nearly a decade later, Karir et al. conducted a study utilizing the slit-jet molecular beam with FTIR detection of the PhAc⋯H2O and PhAc⋯CH3OH complexes, which yielded some conflicting results compared to the earlier findings.2 While the structure of the PhAc⋯H2O complex remained consistent with the previous observations, a discrepancy emerged in the structure of the complex formed between PhAc and CH3OH. Analysis of Fourier-transform infrared (FTIR) spectra obtained from these experiments indicated that methanol also donates hydrogen bonds to the acetylenic π-cloud, a result supported by density functional theory (DFT) calculations, which identified the acetylene-bound structure as the global minimum.
The discrepancy in results for the PhAc⋯CH3OH complex that arises from the combination of molecular beam with IR-UV double resonance spectroscopic and FTIR spectroscopic experimental results can potentially be resolved using microwave spectroscopy. Microwave spectroscopy offers the advantage of providing an unambiguous structure for the complex. A significant amount of research has been carried out on different complexes involving methanol, with microwave spectroscopy being used in many of these studies. Specifically, the barrier height to internal rotation of CH3OH and CH3OH-containing complexes, including the CH3OH dimer, has been a topic of great interest and investigation.14–34 In the case of CH3OH, it has been reported that this barrier height is 373 cm−1, which is the result of the interactions between the internal rotation of the methyl group and the large amplitude motion of the OH group.14 There are reports that suggest that the binding of CH3OH to another molecule can influence (change the barrier height of) the internal rotation of the methyl group in the methanol molecule. Given these findings, it would be of interest to explore the barrier to internal rotation of the methyl group in the PhAc⋯CH3OH complex.
This study presents a comprehensive analysis of the microwave spectrum exhibited by five isotopologues of the PhAc⋯CH3OH complex. In order to delve deeper into the intricate nature of the interactions within the complex, the study employs techniques such as atoms in molecules (AIM) theory, natural bond orbital (NBO) analysis and non-covalent interactions (NCI) index analysis. Through these methods, the molecular interactions present in the complex are explored. The results of this analysis provide valuable insights into the intricate and complex nature of the PhAc⋯CH3OH compound.
For the experiments conducted at Newcastle University, the introduction of PhAc into a neon flow was executed under a backing pressure of 5 bar. Owing to the low vapour pressure, the PhAc was heated to 50 °C, facilitated by a modified general valve,54 in order to produce the mixture with the carrier gas. A second reservoir containing CH3OH (or isotopic variants) was positioned downstream of the modified general valve allowing the introduction of low concentrations of CH3OH into the neon carrier gas. A detailed description of the spectrometer can be found elsewhere.55,56 To capture the broadband spectrum, a chirped pulse was generated by an Arbitrary Waveform Generator (Tektronix AWG7102) with a frequency range from 2.0 GHz to 8.0 GHz and a duration of 1 μs, was sent directly to a 450 W traveling-wave tube amplifier (TWTA) (Applied Systems Engineering), before transmission to the vacuum chamber. The polarization of the molecules within the sample was achieved following the introduction of the chirp from the horn antenna to the vacuum chamber. The microwave radiation and the expanding gas jet exhibit a mutually perpendicular orientation. Subsequently, the free induction decay (FID) of a 20 μs duration was recorded using a second horn antenna after the microwave pulse interaction. The FID's were then digitised by employing a 100 GS s−1 oscilloscope (Tektronix DPO72304XS). The utilization of a gas pulse lasting approximately 200 μs facilitates the acquisition of eight FID's per gas pulse, which was achieved through the implementation of the ‘fast frame’ operational mode of the oscilloscope. Subsequently, the FID's were co-added in the time domain and then Fourier transformed using either a Kaiser-Bessel or High Resolution Window function. Individual lines in the frequency domain spectrum have a full width at half-maximum of approximately 100 kHz allowing line centres to be measured with an estimated precision of 10 kHz. The arbitrary waveform generator (AWG) and the oscilloscope were referenced to a 10 MHz local oscillator obtained from an Analog Signal Generator (Agilent MXG N5183A) or an Rb-clock (SRS FS725) to guarantee phase coherence within the temporal domain and precision in the transition frequencies.
The pulsed nozzle Fourier transform microwave spectrometer (PN-FTMW) was utilized to acquire experimental data at the Indian Institute of Science Bangalore. A comprehensive description of the spectrometer can be found elsewhere.57 The PhAc and CH3OH both were kept in two different bubblers at ambient conditions. The most optimal signal intensity was achieved by allowing helium to flow through the PhAc and CH3OH bubblers at a rate of approximately 20 SCCM (standard cubic centimetre per minute) and 2 SCCM, respectively. At a four-way junction, the samples were mixed with a bath of helium (about 200 SCCM), and the mixture was then sent to the nozzle. The mixture underwent expansion via a 0.8 mm nozzle after being subjected to a backing pressure of 1.5–2.0 bar. A microwave pulse duration of 0.4 μs was found to be optimal for inducing a- and c-dipole transitions, while a duration of 1.0 μs was found to yield the highest intensity signals for b-dipole transitions. The Doppler doubling phenomenon observed in the spectra is a consequence of the molecular beam's colinear arrangement relative to the cavity axis. The spectra were recorded utilizing a sampling frequency of 5 MHz, and during the scanning for the transitions, 256 points were collected. To obtain the high-resolution spectra, the observed rotational lines were subsequently averaged at 512 points.
![]() | ||
Fig. 1 Structures of the PhAc⋯CH3OH complex and their relative energies (kJ mol−1) optimized at MP2/aug-cc-pVDZ (top panel) and B3LYP-D3/aug-cc-pVDZ (bottom panel) level of theories. |
The energy landscape reveals interesting details. At the MP2 level of theory, the (M–P)Ph structure is the lowest energy configuration, with a marginal 0.4 kJ mol−1 difference from the (M–P)Ac structure. Furthermore, their rotational constants exhibit a high degree of similarity. Discrimination between these two configurations based on their rotational spectra becomes feasible by considering their c-dipole moments. The (M–P)Ph configuration possesses a zero c-dipole moment, rendering it devoid of c-type transitions in its microwave spectrum. In contrast, at the B3LYP-D3/aug-cc-pVDZ level of theory, the (M–P)Ac is the global minimum, exhibiting a 3.0 kJ mol−1 energy difference from the (M–P)Ph configuration. The structures optimized using the B3LYP-D3/aug-cc-pVDZ level of theory exhibit markedly distinct rotational constants and dipole moment components, facilitating their obvious differentiation in rotational spectra analysis. The structure where PhAc acts as the hydrogen bond donor and CH3OH as the acceptor has higher energy at both MP2 and DFT levels of theory. Detailed molecular properties such as rotational constants, dipole moment components and binding energies, corrected for basis set superposition error (BSSE) are presented in Table 1. Calculated atomic coordinates of each structure presented in Fig. 1 are provided in the Tables S6–S11 (ESI†).
Notably, the spectrum reveals the presence of CH3OH dimer and parent and six 13C isotopologues of the PhAc monomer, all of these have been subjects of prior investigation.20,22,59 In particular, the CH3OH dimer exhibits line splitting attributable to internal motions of the methyl group. Each of the two possible structures of PhAc⋯CH3OH (as shown in Fig. 1) displays the feature of a nearly prolate asymmetric top. Computational calculations, as outlined earlier, were employed to anticipate the rotational constants (Ae, Be, and Ce) as well as five distortion constants (DJ, DJK, DK, d1, d2) for the equilibrium configuration of both isomers. Subsequently, based on these computational predictions, the observed spectral transitions were fitted, and the experimental rotational constants agree with the structure optimised at B3LYP-D3/aug-cc-pVDZ level of theory, in which the CH3OH donates the hydrogen bond to the acetylenic π-cloud, and there is a secondary interaction between the ortho hydrogen with the oxygen atom of the CH3OH ((M–P)Ac). We also optimized the structure at B3LYP-D3(BJ)/aug-cc-pVDZ level of theory, which has been found to give accurate structures for chiral tag complexes.60 However, for the PhAc⋯CH3OH complex, the agreement with experimental rotational constants did not improve with B3LYP-D3(BJ)/aug-cc-pVDZ level of theory. The results from this calculation are given in ESI,† Table S17. This fitted structure of the PhAc⋯CH3OH complex is consistent with the structure reported by Karir et al.2 The lower trace of Fig. 2 presents a small section of the broadband spectrum, spanning from 4190 to 4206 MHz and from 4338 to 4348 MHz, showing selected transitions.
For the PhAc⋯CH3OH parent complex, a total of 59 transitions were successfully fitted to Watson's S-reduced asymmetric rotor Hamiltonian61 within the Ir representation using Western's PGOPHER62 program. Notably, most of the assigned transitions in the PhAc⋯CH3OH complex were part of a doublet splitting pattern, indicative of a methyl rotor-type large amplitude motion. The complete list of transitions for both states is given in the ESI,† in Table S1. The assignment includes strong a- and c-type transitions, alongside weaker b-type transitions. The fitted parameters determined by fitting to the frequencies of A-state transitions are shown in Table 2. Notably, in this fit, d2 distortion constant is not determinable. The experimentally determined rotational constants and distortion constants closely match those calculated for the (M–P)Ac isomer using DFT methods. Additionally, E-state transitions, arising from the internal rotation of the methyl group, were identified in the spectra, and subsequently assigned using the XIAM14 program to obtain a global fit. The details of the global fit are provided in Section 4.1. Successively, utilizing the fitted rotational constants, the spectral predictions were regenerated within the 8.0 to 14.0 GHz frequency range, and the predicted transitions were searched using a PN-FTMW spectrometer in IISc Bangalore. Both the A- and E-state transitions were measured. A representative portion of the recorded spectra, displaying A and E-state transitions, is presented in Fig. 3. All the transitions were further split into Doppler doublets.
![]() | ||
Fig. 3 The 422 → 312 transition of PhAc⋯CH3OH complex (recorded using PN-FTMW spectrometer) showing A and E-states. The transitions are further split due to the Doppler effect. |
The observed A-state transitions using the PN-FTMW spectrometer were further fitted alongside transitions observed using the CP-FTMW spectrometer utilizing the ASFIT25 program (fitted parameters are given in Table 2). This fit was found to be more accurate than that performed to the CP-FTMW data alone because it was possible to include a greater number of higher frequency transitions. Fits performed using ASFIT and PGOPHER while using only the CP-FTMW data alone were found to be consistent (see Table S18 of the ESI†). As previously highlighted, the results consistently support the conclusion that the observed spectra correspond to the isomeric configuration where CH3OH is bound to the acetylenic moiety of PhAc, and the ortho-hydrogen of PhAc interacts with the oxygen of CH3OH. It is noteworthy that despite rigorous search, spectral data for the complex in which CH3OH donates a hydrogen bond to the π-cloud of the phenyl ring, as well as the isomer where CH3OH donates a hydrogen bond to the acetylenic moiety, while the methyl group interacts with the π-cloud of the phenyl ring, remained elusive.
Rotational constants, distortion constants, and second moment calculated at B3LYP-D3/aug-cc-pVDZ level of theory are taken from Table 1.
Constants | PhAc⋯CH3OH | PhAc⋯CH3OD | PhAc–D⋯CH3OH | PhAc⋯13CH3OH |
---|---|---|---|---|
P
cc is the second moment given by: ![]() |
||||
A 0 | 1955.321(1) | 1962.819(6) | 1891.348(3) | 1935.967(2) |
B 0 | 796.5629(6) | 785.989(2) | 789.224(1) | 785.4982(8) |
C 0 | 626.0072(7) | 619.323(2) | 615.237(1) | 618.6409(8) |
D J | 1.153(3) | 1.06(2) | 1.046(6) | 1.16(1) |
D JK | −6.06(3) | (−5.518) | −4.91(8) | −6.36(6) |
D K | 17.4(1) | 52.5(6) | 18.7(4) | 18.32(2) |
d 1 | −0.142(2) | −0.18(2) | −0.135(8) | −0.172(2) |
D π2J | −322(3) | −274(1) | −315(5) | −327(3) |
D π2K | 844(5) | 760(5) | 787(10) | 858(7) |
V 3 (cm−1) | 235(1) | 255(8) | 236(4) | 237.0(2) |
F 0 (cm−1) | 6.16(5)a | 6.1(2)a | 6.2(1)a | 6.20(6)a |
∠(i, a) (degree) | 68.1(1) | 69.0(6) | 68.3(2) | 68.0(1) |
∠(i, b) (degree) | 22.9(2) | 21.8(6) | 22.8(2) | 22.8(1) |
∠(i, c) (degree) | 96.155(6) | 95.66(2) | 96.777(8) | 95.750(6) |
P cc (μÅ2) | 42.80 | 42.22 | 43.06 | 43.77 |
N A/NE | 72/48 | 31/25 | 29/24 | 44/38 |
σ (kHz) | 10.4 | 12.4 | 8.8 | 9.7 |
It is possible to separately determine F and V3 where a sufficient number of rotational transitions are recorded for multiple torsional states as achieved by De Lucia et al. for the CH3OH monomer.14 However, in general, studies of CH3OH-containing complexes performed by microwave spectroscopy tend to probe rotational transitions of only the ground torsional state and this is also true of the present work. It is not possible to separately determine F and V3 with high precision through fits performed to rotational spectra recorded only for the ground torsional state. An approach often taken by microwave spectroscopists involves fixing F at an assumed value such that a fitted value of V3 can be obtained. Many studies have sought to explore changes in V3 as a function of the substitution position of a CH3 group on a rigid molecular framework (for example, at different substitution positions on an aromatic ring). In such examples, it is well-established that Iα and F0 (where F0 is equal to F for the limiting case where ρa = 0) can respectively be approximated as ∼3.2 μÅ2 and ∼157.9 GHz and this knowledge was exploited by two authors of the present work during recent analyses of internal rotation within isomers of the monohydrate complexes of methylthiazole and methylimidazole.63,64 However, the CH3 group is not bound to a rigid molecular framework within CH3OH so the value of F0 cannot a priori be assumed to be reliable in such cases. In fact, as long ago as the 1980's, papers by Lovas et al. and Suenram et al. analysed large amplitude motions within hydrogen-bonded complexes containing CH3OH while highlighting that their analyses would not reliably distinguish the effects of CH3 internal rotation from vibrational motions of the O–H group.24,25 Authors were careful to report only “effective” V3 barriers which they did not intend to be used in direct comparisons with V3 determined for other molecules and complexes. A recent study of the diphenyl ether⋯CH3OH complex highlighted this continuing constraint.23 The results of two alternative approaches were presented when the spectrum of diphenyl ether⋯CH3OH was analysed: (1) a fit was performed to determine both V3 and F0 even while it was explained that these parameters are unavoidably correlated in the fit that was performed and (2) a fit was performed to determine V3 while assuming a fixed value of F0. To facilitate broad comparisons of data sets which may follow in future works, the same approach will be taken towards the reporting of results for internal rotation parameters herein. The results presented in Table 4 employed the first of the approaches described above while the results obtained by the alternative approach are presented in the ESI† (Table S14).
To gain insight into the V3 barrier within the context of the PhAc⋯CH3OH complex, we conduct a relaxed potential energy scan for the rotation of the CH3 group within the CH3OH monomer and (M–P)Ac structure. These potential energy scans are carried out at the B3LYP-D3/aug-cc-pVDZ level of theory. The resulting potential energy curves for these relaxed scans are shown in Fig. 4, in which the barrier to CH3 rotation is found to be 385.5 cm−1 (solid line) for methanol monomer and 312.8 cm−1 (dashed line) for PhAc⋯CH3OH complex. The experimentally observed (373 cm−1) and computed barrier height (385.5 cm−1) for CH3 rotation in the methanol monomer are in good agreement with each other. The computed barrier height of the PhAc⋯CH3OH complex, when compared to the computed barrier in the methanol monomer, suggests that the methyl (CH3) group within the (M–P)Ac complex is not significantly engaged in intermolecular interactions with PhAc. Analysis of the experimental spectra is facilitated by the XIAM program, which can be used to perform a “global” fit of both A- and E-state transitions to determine parameters in a rotational Hamiltonian, which includes terms to describe the effects of centrifugal distortion and internal rotation. The XIAM program utilizes a Hamiltonian that is formulated within the principal axis system. The program uses a “local” approach, treating each torsional state individually without taking into account any interactions between different torsional states. The analysis yields the values shown in Table 4 for the F0, V3, ∠(i, a), ∠(i, b), ∠(i, c), Dπ2J and Dπ2K parameters (defined within the opening paragraph of this section) for each of four isotopologues of the PhAc⋯CH3OH complex.
![]() | ||
Fig. 4 The potential energy curve for the internal rotation of the CH3 group in the CH3OH monomer (solid line) and PhAc⋯CH3OH complex (dashed line). |
Table 5 presents a comparison of results reported for F and V3 for the series of CH3OH-containing complexes studied to date where analyses of internal rotation have been reported. As noted earlier, F0 and V3 cannot be independently determined from rotational spectra reported for only a single torsional state. On the other hand, the value of the reduced barrier, s = 4V3/9F (a dimensionless quantity), is also sensitive to the splitting between A- and E-state transitions and provides an interesting perspective on the results shown. The complexes featured in Table 5 are presented in order of decreasing s. Many previous works assumed a fixed value of F0 to allow the determination of an “effective” V3. It is immediately apparent from Table 5 that the results for V3 cannot easily be rationalised with any trend in molecular properties or geometry across this set of complexes. For example, the V3 reported for Ar⋯CH3OH and HCl⋯CH3OH are significantly higher than that for the CH3OH monomer for reasons that do not obviously connect with any aspect of the properties of these complexes.14,25,26 In any case, as noted by the authors of the original works, the V3 of these species almost certainly differ because accurate values of F0 were not available when the fits to determine V3 for each of Ar⋯CH3OH and HCl⋯CH3OH were performed, and unreliable assumptions were therefore made. An alternative and more insightful perspective on the underpinning molecular physics is provided by the trend in s as a function of molecular geometry and bonding across the series of complexes featured in Table 5.
Molecule | V 3 (cm−1) | F (cm−1) | s | Ref. |
---|---|---|---|---|
a The original works fixed F0 rather than F, so the value for F0 is quoted in this table. For the heavier complexes (phenyl vinyl ether⋯CH3OH, diphenyl ether⋯CH3OH), it can reliably be assumed that the value of F0 will be very similar to (within a few percent of) that of F. This will not necessarily be true for complexes of CH3OH with lighter molecules such as HCl, Ar or formamide. b Allowed to float during the fitting procedure. c Mon I and Mon II are the two CH3OH monomers within the (CH3OH)2 dimer, where Mon I is the hydrogen bond donor whereas Mon II is hydrogen bond acceptor. | ||||
CH
3
OH ![]() |
||||
Phenyl vinyl ether⋯CH3OH | 261.8 | 5.3a | 21.95 | 34 |
Diphenyl ether⋯CH3OH | 480 | 10.16ab | 21.00 | 23 |
Diphenyl ether⋯CH3OH | 250.74 | 5.3a | 21.03 | 23 |
Formamide⋯CH3OH | 231.01 | 5.26a | 19.52 | 24 |
Aniline⋯CH3OH | 215 | 5.28 | 18.10 | 29 |
H2CO⋯CH3OH | 240.5 | 6.17 | 17.32 | 31 |
PhAc⋯CH3OH | 200.0 | 5.24 | 16.96 | Present work |
PhAc⋯CH3OH | 235 | 6.16b | 16.96 | Present work |
CH3OH dimer (Mon I)c | 201c | 5.4b | 16.54 | 20 |
Trimethylamine⋯CH3OH | 174 | 5.30 | 14.59 | 28 |
CO⋯CH3OH | 183 | 5.76 | 14.12 | 30 |
Other molecules and complexes | ||||
Phenol⋯CH3OH | 170 | 5.27 | 14.34 | 65 |
CO2⋯CH3OH | 174.8 | 5.45 | 14.25 | 31 |
SO2⋯CH3OH | 128.7 | 5.3 | 10.79 | 27 |
CH3OH dimer (Mon II)c | 118.7c | 5.16b | 10.22 | 20 |
HCl⋯CH3OH | 74.0 | 5.3a | 6.21 | 26 |
CH3OH monomer | 373 | 27.63b | 6.00 | 14 |
Ar⋯CH3OH | 68.5 | 5.3a | 5.74 | 66 |
CH3OH-containing complexes where the O of CH3OH acts as hydrogen bond acceptor (HCl⋯CH3OH or monomer II of (CH3OH)2) or where the complex is bound by another form of weak interaction (e.g. SO2⋯CH3OH, Ar⋯CH3OH) have values of s which are very similar to that determined (s = 6.0) for the isolated CH3OH monomer.14,21,22,25,27 Those complexes where the O–H of CH3OH acts as a hydrogen bond donor have significantly higher s that range from 14.12 for CO⋯CH3OH to 21.95 for phenyl vinyl ether⋯CH3OH.30,34 There also appears to be a correlation between the relative strength of the hydrogen bond(s) formed within each complex and the value of s. Where a complex contains only a single hydrogen bond, and especially where this bond can be expected to be weak, the s for the complex appears to lie at the lower end of the range. Complexes with stronger hydrogen bonds generally associate with higher values of s and those at the high end of the range are held together by multiple interactions. For example, formamide⋯CH3OH contains multiple intermolecular hydrogen bonds whereas the diphenyl ether and CH3OH monomers interact via dispersion interactions in addition to the hydrogen bond present within the diphenyl ether⋯CH3OH complex. An important confirmation of the described relationship is provided by the s values attributed to each of the CH3OH monomers present within the CH3OH dimer. This complex is primarily the result of a hydrogen bond formed between the O−H of one monomer and the oxygen atom of the second. That monomer which acts as hydrogen bond donor within the CH3OH dimer has a value for s that is consistent with its role. The other, which acts as the hydrogen bond acceptor, has an s value similar to that seen for the isolated CH3OH monomer.
The findings of the IR/UV double resonance experiments conducted on the PhAc⋯H2O and PhAc⋯CH3OH complexes have been previously reported, with the determination of their structures relying on the analysis of the fluorescence dip infrared (FDIR) spectra in the O–H and C–H stretching regions.4 The vibrational frequencies corresponding to the O–H stretch (νO–H) of the H2O and CH3OH complexes of PhAc, both calculated and experimentally obtained (as cited in ref. 2), are given in Table S15 of the ESI.† In the case of the PhAc⋯H2O complex, the experimentally recorded νO–H value for the PhAc⋯H2O complex was identified at 3629 cm−1 for the hydrogen bonded O–H and 3724 cm−1 for the free OH. The predicted values for the hydrogen bonded νO–H in acetylenic and phenyl bound structures were 3587 and 3621 cm−1, respectively. Clearly, the observed value of 3629 cm−1 is larger than what is predicted for both conformers, though closer to the phenyl-bound structure. Still, the acetylenic bound structure was favoured for the following reasons. The resonant ion-dip infrared spectra of the benzene⋯H2O π hydrogen bonded complex in the O–H stretch region exhibit over six transitions, attributed to the large amplitudes tumbling motion of the H2O molecule.67 As previously mentioned, the FDIR spectrum of the PhAc⋯H2O complex displays two distinctive transitions corresponding to the hydrogen bonded and the free O–H vibrations. Consequently, it was deduced that, unlike benzene⋯H2O, the H2O molecule forms a strong bond with the PhAc in the PhAc⋯H2O complex. Given that the acetylenic bound structure involves O–H⋯π and C–H⋯O hydrogen bonds leading to a cyclic structure, it suggests the plausibility of the acetylenic bound configuration. More importantly, the FDIR spectra in the C–H stretch region were utilized to gain more comprehensive insights. In the PhAc monomer, peaks at 3325 cm−1 and 3343 cm−1 were detected, indicating the occurrence of Fermi resonance. This Fermi resonance entails the vibration of C–H stretching and a combination band of one quantum of CC stretch and two quanta of out-of-plane acetylenic C–H bend. Disturbances impacting the C
C bond stretching vibration, such as intermolecular bonding, lead to the disappearance of Fermi resonance bands and the appearance of a single band corresponding to the free C–H oscillator. The FDIR spectrum of the PhAc⋯H2O complex reveals a single band at 3331 cm−1, which further confirms the interaction of the O–H group with the π electron density of the acetylenic group.
As for the PhAc⋯CH3OH complex, the experimental νO–H for the hydrogen bonded O–H was identified at 3615 cm−1. The theoretically predicted νO–H for the acetylenic (3633 cm−1) and phenyl bound (3637 cm−1) structures of the PhAc⋯CH3OH complex were very similar, with a marginal difference of only 4 cm−1. Consequently, these slight infrared band shifts observed in the weak O–H⋯π complexes do not offer sufficient evidence for definitive structural determinations. In the C–H stretch region of the FDIR spectrum, two bands were observed at 3323 cm−1 and 3334 cm−1, indicating the presence of Fermi resonance. Hence, based on the Fermi resonance bands, it was deduced that the PhAc⋯H2O complex involves O–H⋯π interactions through the acetylenic-π, while the PhAc⋯CH3OH complex exhibits O–H⋯π interactions through the phenyl π-system. The PhAc⋯H2O complex has already been previously investigated by microwave spectroscopy and has confirmed the unambiguous structure of the complex to the acetylenic bound conformer.10
Karir et al. explored the docking preference of CH3OH in the PhAc and its mono-methylated derivative (3-methylphenylacetylene) using FTIR spectroscopy of supersonic jet expansions.2 The FTIR jet spectra of the 3-methylphenylacetylene⋯CH3OH complex displayed two additional peaks at 3620 cm−1 and 3639 cm−1 in the O–H stretch region (see Fig. S7 of ref. 1, ESI†) compared to the spectrum of bare methanol (3686 cm−1), therefore these new bands were attributed to the acetylene and phenyl docking of the CH3OH. It is well-established that deuteration strengthens the hydrogen bond due to reduction of zero point vibrational energy (ZPVE). The observed spectrum of 3-methylphenylacetylene⋯CH3OD complex was very similar (OD stretch region was scaled to the OH stretch region by using a scaling factor of 1.356) with a reduction in intensity noted for the less shifted band (Fig. 4 of ref. 1). Thus, the band at 3620 cm−1 was assigned to the acetylenic bound whereas the band at 3639 cm−1 was assigned to the phenyl bound structure. The observation of both the phenyl and acetylene bound structure in 3-methylphenylacetylene⋯CH3OH complex is possible if the barriers between the docking sites are too high to be overcome. The DFT calculations (refer to Table 1 of ref. 1) also suggested that the acetylenic bound structure is the more favourable case in 3-methylphenylacetylene⋯CH3OH complex therefore the binding preference in this complex is shifted towards the acetylenic coordination by deuteration and also shifted in the same direction in the PhAc⋯CH3OH complex. In the FTIR jet spectra of the PhAc⋯CH3OH complex, a vibrational frequency associated with hydrogen bonded O–H stretching was detected only at 3622 cm−1. This value notably deviates from that previously reported by Singh et al. (3615 cm−1 for CH3OH complex). This disparity in the experimentally obtained vibrational frequencies within the same complex are attributed to inaccuracies in calibration during the laser experiment. Additionally, there is a notable discrepancy between the result of 3622 cm−1 and the measured vO–H of 3613 cm−1 for the PhAc⋯H2O complex, for which no explanation was provided. The DFT-predicted and experimentally observed vibrational frequencies (taken from ref. 1) are given in Table S15 of the ESI,† for the hydrogen bonded O–H in the acetylenic and phenyl bound conformations for PhAc⋯H2O, PhAc⋯CH3OH and 3-methylphenylacetylene⋯CH3OH. As described above, the assignments for PhAc⋯CH3OH complexes were based on consistent deviations between DFT predictions and experimental results across various systems, further supported by isotope effects in the methylated derivative of PhAc.
The second moment, Pcc, calculated from the optimized structures at the DFT and MP2 levels of theories are presented in Table 1. It is apparent that the Pcc values for the (M–P)Ac structure optimized at MP2 and DFT exhibit substantial discrepancies. This discrepancy can be attributed to the presence of a secondary interaction in the DFT-optimized (M–P)Ac structure, which significantly impacts the overall configuration, despite both structures involving CH3OH donating a hydrogen bond to the acetylenic moiety. The experimentally-determined values of Pcc for the parent and other isotopologues are provided in Tables 2–4, obtained during A-state only fits and global fitting containing both A- and E-state transitions. The experimental second moments closely align with the DFT-calculated (M–P)Ac structure. The replacement of a bonded hydrogen atom of CH3OH causes minimal changes to the Pcc value, whereas the substitution of all four hydrogen atoms of the CH3OH moiety results in a notable increase.
By using the Kraitchman approach,68 the singly substituted isotopologues enable a direct structural examination of the complex. These equations could be used to find the position of the substituted atom from the centre of mass. For a non-planar asymmetric top molecule, the coordinate (rs) of the substituted atom is given as follows:
In Fig. 6, we present the labelling of atoms within the molecular complex, along with an approximate orientation of the inertial axis framework. The rotational constants obtained using XIAM fitting approach, which contains both A- and E-species transitions, were utilised for the derivation of rs coordinates for the acetylenic H atom (H14) and the carbon atom of CH3OH (C15).
![]() | ||
Fig. 5 Labelling of the atoms used in the structural analysis for the PhAc⋯CH3OH complex. The approximate location of the principal axes is shown. |
![]() | ||
Fig. 6 Atoms in Molecules (AIM) topology study for the PhAc⋯CH3OH and PhAc⋯H2O complexes. The green dots and red dots refer to the bond critical point and ring critical points respectively. |
These coordinates, along with associated Costain errors,71 are provided in Table 6. It is noteworthy that the Kraitchman method does not yield the signs of atomic coordinates, which have been inferred from the outcomes of DFT calculations. The rs coordinate of the bonding hydrogen atom (H20) of the CH3OH sub-unit was found to deviate from the corresponding coordinate in the calculated structural model. This discrepancy can be attributed to the interplay between the methyl internal rotation and the large amplitude motion of the OH group in the CH3OH.
Atom number | Method | Coordinates | ||
---|---|---|---|---|
a | b | c | ||
H14 | Theory | 2.4776 | 2.9328 | −0.2894 |
Experiment | 2.3578(7) | 2.9401(6) | −0.492(3) | |
C15 | Theory | 2.8611 | −1.2257 | 0.9626 |
Experiment | 2.8282(5) | −1.296(1) | 0.986(2) |
Following a rigorous examination of the spectra of all five isotopologues, the structural configuration where CH3OH forms a hydrogen bond with the phenyl π system could not be observed. Since the spectral assignment in the case of results obtained by Singh et al. was exclusively based on the presence and absence of Fermi resonance, one possibility is that the PhAc⋯CH3OH complex has a structure (CH3OH is interacting with acetylenic-π system) that is observed by both FTIR spectroscopic study and the present study by microwave spectroscopy, while the Fermi resonance is unaffected. If this is the case, then all three experimental results give a consistent structure. The current study using microwave spectroscopy provides an unambiguous structure for PhAc⋯CH3OH complex, wherein MeOH is forming hydrogen bond with the acetylenic π-system followed by a secondary weak C–H⋯O interaction. The present study further supports the conclusions drawn through FTIR spectroscopic results.2
The molecular graph for the PhAc⋯CH3OH complex closely resembles that of the PhAc⋯H2O complex except for the one additional bond critical point (BCP) and ring critical point (RCP) found in the PhAc⋯CH3OH complex. The first BCP occurs between the hydroxyl hydrogen atom of CH3OH/H2O and the midpoint of the π-bond in PhAc, while the second BCP is situated between the oxygen atom of CH3OH/H2O and the ortho-hydrogen atom of PhAc. The third BCP in case of PhAc⋯CH3OH complex is present between one of the hydrogen atoms of the methyl group to the ortho-carbon atom of the PhAc. The electron densities (ρ) and Laplacian (∇2ρ) values for all BCPs are provided in Table 7. The molecular graphs showing BCP's and RCP's are given in Fig. 6. These ρ and ∇2ρ values for both complexes fall within the range defined by Koch and Popelier for C–H⋯O hydrogen bonds.72 A comparison of the electron densities at the BCP for O–H⋯π and C–H⋯O interactions in PhAc⋯CH3OH and PhAc⋯H2O complexes reveals that both interactions contribute equally to the stability of the observed geometries. Consistent with the findings in the PhAc⋯H2O complex, the C–H⋯O interaction in the PhAc⋯CH3OH complex is as strong as the O–H⋯π interaction.
Interactions | PhAc⋯CH3OH | Interactions | PhAc⋯H2O | ||
---|---|---|---|---|---|
ρ (a.u) | ∇2ρ (a.u) | ρ (a.u) | ∇2ρ (a.u) | ||
O19–H20⋯π | 0.013 | +0.033 | O15–H16⋯π | 0.013 | +0.036 |
C6–H11⋯O19 | 0.011 | +0.031 | C3–H11⋯O15 | 0.011 | +0.032 |
C15–H17⋯C6 | 0.005 | +0.020 | — | — | — |
Interaction | E (2) (kJ mol−1) | |
---|---|---|
PhAc⋯CH3OH | PhAc⋯H2O | |
πC![]() |
10.8 | 13.3 |
nO → σC–H* | 7.1 | 10.0 |
Method | Structures | E Electrostatics | E Induction | E Dispersion | E Exchange | E Interaction |
---|---|---|---|---|---|---|
SAPT‘2 + 3’ | OH⋯πPh | −17.6 | −6.8 | −29.7 | 37.8 | −16.3 |
MP2/aug-cc-pVDZ | OH⋯πAc | −20.2 | −7.0 | −29.1 | 39.4 | −16.9 |
SAPT‘DFT’ | OH⋯πPh | −12.7 | −4.1 | −20.4 | 21.7 | −15.5 |
B3LYP-D3/aug-cc-pVDZ | OH⋯πAc | −25.6 | −7.5 | −19.8 | 35.8 | −17.1 |
It should be noted that Singh et al.'s conclusion that PhAc⋯CH3OH contains a hydrogen bond donated by the O–H group of CH3OH to the phenyl π-cloud has been very well supported by IR/UV double resonance spectroscopic experiments and ab initio computations. However, the evidence of the CH3OH donating hydrogen bond to the phenyl π-system was concluded based on the presence or absence of the Fermi resonance in the FDIR spectra of the C–H stretch region. The Fermi resonance bands which involve the acetylenic π-cloud, was not affected in the PhAc⋯CH3OH complex, whereas in the PhAc⋯H2O complex, the Fermi resonance disappeared and gave only one peak corresponding to the C–H oscillator. In case, the interaction in PhAc⋯CH3OH complex having similar structure as PhAc⋯H2O complex does not affect the Fermi resonance, it is likely that there is no discrepancy in the experimental results about the global minimum structure.
There is clearly a need for further investigation which might rationalise all the observations made during this and previous works. One possible explanation can be immediately stated: it is possible that the isomer of PhAc⋯CH3OH complex observed by Singh et al. authentically has a different structure from that observed during the present work. However, given the extensive similarities between the experimental conditions of this work and those used by Singh et al., which each involved co-expansion of the PhAc and CH3OH precursors within a buffer gas of helium, this seems very unlikely.
Footnote |
† Electronic supplementary information (ESI) available: Tables of microwave transitions measured for parent and five isotopologues of the title complex (Tables S1–S6), equilibrium geometries of the complex optimized at various levels (Tables S7–S11), normal mode vibrational wave numbers at two different levels (Tables S12–S13), experimental rotational constants from the global fit (Table S14), experimental and computed C–H/O–H stretching wavenumbers from ref. 1 and 2 (Table S15) and results from SAPT energy decomposition analysis (Table S16) are available online as ESI. See DOI: https://doi.org/10.1039/d4cp01916d |
This journal is © the Owner Societies 2024 |