Halogen bond and internal dynamics in the σ–complex of pyridine-chlorotrifluoromethane: A rotational study

https://doi.org/10.1016/j.jms.2020.111323Get rights and content

Highlights

  • Csingle bondCl···N halogen bond.

  • Multiquadrupole problems.

  • Internal dynamics.

Abstract

The rotational spectrum of the 1:1 complex pyridine–chlorotrifluoromethane shows that the two moieties are held together through a Csingle bondCl···N halogen bond (RCl···N = 2.909(5) Å). The quadrupolar effects of the 35Cl (or 37Cl) and 14N nuclei combined with the high density of the rotational transitions give rise to a very congested spectrum. The multidimensional large amplitude internal dynamics is reflected in a two times larger rotational constant A, on an unexpected positive values of the inertial defect of the ground state, and on a negative value of the DJ centrifugal distortion constant. This comes from a combination of a free internal rotation and of a very low energy in plane oscillation of CF3Cl with respect to pyridine. However, a coupled Hamiltonian including Coriolis interactions with an hypothetical (unobserved) v = 1 state solves the initial result of an unphysical negative value for the centrifugal distortion constant.

Introduction

The importance of halogen bond (HaB) has been increasingly noticed in the last years and found applications in the field of crystal engineering [1], nanomaterials [2], supramolecular chemistry [3], liquid crystal [4] and many others. For this reason, IUPAC formalized a definition of this non-bonding interaction: “a halogen bond occurs when there is evidence of a net attractive interaction between an electrophilic region associated with a halogen atom in a molecular entity and a nucleophilic region in another, or the same, molecular entity” [5].

Most of the investigations dedicated to the HaBs are based on solid-state X-ray diffraction [6]. However, more precise information on this kind of interaction, free from solvent effects or solid state linkages, comes from studies of isolated complexes with the subunits held together by such interaction in the gas phase. Accurate details of the HaB nature are provided by rotational spectroscopy of molecular complexes. An overview of properties of the HaB interaction has been given by Legon through Fourier transform microwave (FTMW) spectroscopy studies of a series of B···XY complexes, where B is the electron donor and XY is a dihalogen molecule [7]. It has been proved that in these cases the HaB is almost linear, with B···X-Y angles close to 180°.

Chlorotrifluoromethane (CF3Cl) has been found to be a good prototype molecule to investigate HaBs. Because the fluorine atom is more electron-withdrawing than the chlorine atom, a positive electrostatic region (“σ-hole”) [8], [9] is formed on the outer surface of the chlorine atom in CF3Cl (opposite to the Csingle bondCl bond) and ready to interact with negative sites. CF3Cl forms indeed, as “halogen donor”, a Csingle bondCl···O HaB with H2O [10], formaldehyde [11] and dimethyl ether [12], a Csingle bondCl···N HaB with NH3 [13], and a Csingle bondCl···F HaB with CH3F [14]. Strong dynamic effects considerably alter the rotational constants, with respect to the values of expected for a rigid rotor. The spectrum of CF3Cl-H2O, which is classically an asymmetric top, has been interpreted as that of a symmetric top [10]. The free internal rotation of the CF3 group causes unexpectedly large values of the rotational constants A in the complexes of CF3Cl with formaldehyde [11], dimethyl ether [12] and CH3F [14].

Pyridine (PYR) is the best-known heterocyclic aromatic molecule, used extensively in coordination [15], [16] and surface chemistry [17]. For this reason, the non-covalent interactions between PYR with other molecules have been extensively investigated using high resolution spectroscopy. Depending on the chemical nature of the partner species, both π or σ type complexes could be formed in relation to its negative sites: either the π system of the aromatic ring, or the non-bonding orbital (n) at the nitrogen atom. Combined with laser–vaporization, ZEKE spectroscopy pointed out that Li and Ca complexes prefer σ bonding, whereas the complex with Sc has a π-type linkage [18].

PYR is the only aromatic molecule for which, thanks to its permanent dipole moment, the rotational spectra of all van der Waals complexes with rare gases (PYR-RG, except radon) have been reported, providing experimental evidences of the magnitude and directionality of the dispersive interactions [19], [20], [21], [22], [23], [24], [25]. In all the complexes, a π-type structure is observed, even for complexes with two RG atoms, which exhibit a “double π” arrangement with one atom above and one atom below the ring [24], [26], [27]. Conversely, most studies of molecular adducts of PYR with any other partners revealed σ-type arrangements [28], [29], [30], [31], [32], [33]. When reviewing the complexes of PYR with the family of CHnF4-n, one can note an interesting behaviour. In the case of PYR-CF4 (n = 0), the two subunits are held together by a CF3···N HaB, and CF4 undergoes a free rotation with respect to PYR [34]. In PYR-CHF3 [35], PYR-CH3F [36] and CH2F2 [37], two kinds of weak hydrogen bonds (WHB), Csingle bondH···N and Csingle bondH···F, have been found to connect the two constituent moieties. The barriers to internal rotation of the CHF3 and of the CH3F groups have been determined from the A-E tunnelling splittings of the rotational transitions. PYR-CH4 constitutes an exception: CH4 prefers to link with PYR through the π electronic system rather than the n-orbital, with CH4 acting as a pseudo rare gas [38].

Herein, we decided to investigate the rotational spectrum of PYR-CF3Cl, which represents the largest molecular complex held by a HaB, to date, by using pulsed jet Fourier transform microwave (PJ-FTMW) spectroscopy. There are two interesting aspects: (1) which negative site of PYR, n-orbital or π-system, is favoured for the positive region of CF3Cl? (2) How the internal dynamics will affect the value of the rotational constant of the complex with a much heavier molecule PYR than formaldehyde, CH3F and dimethyl ether? The results are reported below.

Section snippets

Experimental

The molecular clusters were generated in a supersonic expansion, under conditions optimized for the formation of the adduct PYR-CF3Cl. Details of the Fourier transform microwave spectrometer [39] (COBRA-type [40]), which covers the range 6.5–18 GHz, have been described previously [41].

A gas mixture of ca. 1% of CF3Cl in helium at a stagnation pressure of ∼0.5 MPa was passed over a sample of pyridine (cooled to 0 °C) and expanded through a solenoid valve (General Valve, Series 9, nozzle diameter

Theoretical calculations

Two possible isomers, stabilized by Csingle bondCl···n (σ-type) or Csingle bondCl···π (π-type) HaB are expected, by chemical intuition, to be the most stable forms of the title complex, as shown in the upper part of Table 1. Ab initio calculations at MP2/6–311++G(d,p) level of theory by using the Gaussian09 Suite of Programs [42] confirmed this hypothesis. Table 1 also reports their relative energies and the spectroscopic parameters useful for the investigation of the microwave spectra. The intermolecular binding

Rotational spectrum

According to the small values of calculated rotational constants B and C, the a-type R-bands of rotational transitions were expected to appear in narrow frequency regions separated by B + C. The presence of two quadrupolar nuclei with nuclear spins I(35Cl or 37Cl) = 3/2 and I(14N) = 1 was expected to complicate the rotational spectrum (as shown in Fig. 1 for the 131,12 ← 121,11 transition).

Following extensive spectral searches for the two isomers, we could identify, based on their 35Cl/14N

Conformational information

Comparing the experimental spectroscopic parameters in Table 2 with the theoretical values of the two conformations in Table 1, the rotational (except A) and quadrupole coupling constants match only those of σ-type isomer, which is stabilized by a Csingle bondCl···N HaB. We could not observe any lines belonging to π-type isomer, despite the relatively small complexation energy difference. The small negative value of vibrational frequency for the latter isomer means that it could be not real minimum at the

Molecular structure

In the observed isomer of PYR-CF3Cl, the two subunits are held together through a Csingle bondCl···N HaB (principal inertial axes in Fig. 4). The full ab initio geometry of this isomer is available in the Supplementary Material.

The ab initio value obtained for the halogen bond length RCl···N was 2.993 Å By adjusting the value of RCl….…N to 2.909(5) Å, we obtained an experimental value which could satisfactorily reproduce the rotational constants B and C of all isotopologues (the largest discrepancy is

Internal dynamics

Back to the observation of a rotational constant A larger than that predicted for a rigid complex, the enormously large discrepancies can be accounted by the effects of free (or almost free) internal rotations of the two moieties, CF3Cl and PYR, with respect to each other around their C3 and C2 symmetry axes, respectively. Similar effects have been observed for the asymmetric top 13C-benzene-CHF3 [50], [51] PYR-CF4 [34] and benzotrifluoride [52], where the –CF3 rotor freely or almost freely

Conclusions

This Fourier transform microwave studies of PYR-CF3Cl and its isotopologues pointed out that the “σ-hole” on the Cl atom of CF3Cl prefers to interact with PYR through the n-orbital rather than the π-system. This revealed result is in contrast to that of PYR-CH4, where the positive site of CH4, Csingle bondH group, linking to the π-system of PYR is favoured [39].

Unlike the complexes of PYR with other C3v symmetric tops CH3F [38] and CHF3 [36], the free internal rotation of the -CF3 group changes

Acknowledgements

We acknowledge Italian MIUR (PRIN project 2010ERFKXL_001) and the University of Bologna (RFO) for financial support. This work was also supported by the National Natural Science Foundation of China (Grant No. 21703021), Fundamental and Frontier Research Fund of Chongqing (Grant No. cstc2018jcyjAX0050), the Spanish MINECO (CTQ2017-89150-R and CTQ2015-68148-C2-2-P), Basque Government (IT1162-19 and PIBA 2018-11), the UPV/EHU (PPG17/10 and GIU18/207), CSIC (PIC2018 and LINKA20249). M.V.L.

References (53)

  • B. Velino et al.

    J. Mol. Spectrosc.

    (2008)
  • S. Melandri et al.

    Chem. Phys. Lett.

    (1996)
  • J.L. Doran et al.

    J. Mol. Struct.

    (2012)
  • M.S. Labarge et al.

    Chem. Phys. Lett.

    (1989)
  • M. Vallejo-López et al.

    Chem. Phys. Lett.

    (2014)
  • W. Caminati et al.

    Chem. Phys. Lett.

    (2004)
  • A. Mukherjee et al.

    Acc. Chem. Res.

    (2014)
  • See for example: W. Wang, D. Wang, Y. Zhang, B. Ji, A. Tian, J. Chem. Phys. 134 (2011)...
  • L.C. Gilday et al.

    Chem. Rev.

    (2015)
  • H.L. Nguyen et al.

    J. Am. Chem. Soc.

    (2004)
  • G.R. Desiraju et al.

    Pure Appl. Chem.

    (2013)
  • See, for example, G. R. Desiraju, D. Braga, J. S. Miller, A. G. Orpen, S. L. Price, CrystEngComm. 4 (2002) 500, and...
  • A.C. Legon

    Angew. Chem. Int. Ed.

    (1999)
  • A.C. Legon

    Phys. Chem. Chem. Phys.

    (2010)
  • P. Politzer et al.

    J. Mol. Model.

    (2007)
  • L. Evangelisti et al.

    Angew. Chem. Int. Ed.

    (2011)
  • Q. Gou et al.

    Chem. Eur. J.

    (2012)
  • L. Evangelisti et al.

    J. Phys. Chem. A

    (2014)
  • G. Feng et al.

    Chem. Eur. J.

    (2012)
  • Q. Gou et al.

    J. Phys. Chem. Lett.

    (2014)
  • R. Kempe

    Eur. J. Inorg. Chem.

    (2003)
  • J. Ashmore et al.

    J. Chem. Soc. Dalton Trans.

    (1995)
  • See, for example, S. Haq, D. A. King, J. Phys. Chem. 100 (1996) 16957, and refs...
  • S.A. Krasnokutski et al.

    J. Chem. Phys.

    (2009)
  • C. Tanjaroon et al.

    J. Chem. Phys.

    (2007)
  • A. Maris et al.

    Chem. Commun.

    (1998)
  • Cited by (9)

    View all citing articles on Scopus
    1

    Present address.

    View full text