Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Sigma-hole-supported interactions in complexes of group 5 oxyhalides (MOX3) with insights for their extended solids

Donovan Hoilette Jr. , Gabriel F. Stewart and Kelling J. Donald *
Department of Chemistry, Gottwald Center for the Sciences, University of Richmond, Richmond, Virginia 23173, USA. E-mail: kdonald@richmond.edu; Tel: +1-804-484-1628

Received 22nd September 2025 , Accepted 10th November 2025

First published on 12th November 2025


Abstract

The existence and nature of sigma hole type interactions in inorganic extended solids remain largely unexamined, even though the influence of such interactions for heteroorganic compounds is well known. The series of group 5 MOX3 oxyhalides are intriguing in this regard since several of the known crystal structures are molecular solids with highly polarized M centers. We examine computationally the bonding in the isolated MOX3 molecules and their smallest clusters (MOX3)n where M = V, Nb, and Ta, and X = H, F, Cl, Br, and I. That investigation provides us with substantial evidence, based on patterns and quantitative trends exhibited in the bonding and energetics of the oxyhalide clusters, that several of the extended solids are stabilized by sigma-hole-supported M⋯O and more rarely M⋯X bonding interactions between the MOX3 molecules or dimers. In particular, one-dimensional stacks of doubly bridged, X2OMX2MOX2, dimer units in several of the oxyhalide crystal structures are shown to be propagations of sigma-hole-supported bonding interactions with weak charge transfer contributions that are already established at the level of the tetramer. Our results are in harmony with a more recent crystal structure for NbOCl3 over a much older proposal.


Introduction

The relevance of sigma-hole-supported chemical bonding to areas such as crystal engineering is well established.1–4 The term ‘sigma hole interaction’ refers to a stabilizing non-covalent interaction achieved when a lone pair on a base, Y, aligns with a localized region of positive electrostatic potential (a positive sigma hole) on a terminal (X) or central (M) atom opposite a polar R–X or R–M bond, respectively. The emergence of a sigma hole on an atomic center is favored if the R group is strongly electron withdrawing; a sigma hole is induced on iodine by –CF3 in F3C–I, for example, and on carbon (opposite each F–C bond) by F in CF4.5,6

Since the sigma hole on I in F3C–I or on C in CF4 arises and is localized immediately opposite the polar C–I or F–C bond(s), sigma hole interactions are optimized when the lone pair on the base is oriented directly toward the sigma hole. So, the R–X⋯Y or R–M⋯Y bond angle, such as the C–I⋯N angle in the F3C–I⋯NH3 complex, is expected to be linear.7 For bulky and low symmetry donor–acceptor pairs, of course, secondary interactions, such as steric repulsion or competing long range effects, can drive deviations from linearity,8,9 but such instances are exceptions, as it were, that make the rule.

We use the term sigma-hole-supported bonding for interactions that feature positive sigma holes at the acceptor site in donor–acceptor pairs. In that way, one acknowledges that even if electrostatic contributions (including sigma hole effects) dominate an interaction, charge transfer and other influences on chemical bonding can be substantial. Especially since the sigma holes opposite a bond in an electron acceptor tend to coincide with lobes of empty frontier orbitals, charge transfer and electrostatics are partners, not rivals, in stabilizing weak interactions. In SiF4, for instance, the four sigma holes on Si – each located opposite a σSi–F bond – are each coincident with the corresponding image file: d5cp03656a-t1.tif orbital. So a F4Si⋯Y sigma hole interaction is simultaneous with Si ← Y charge transfer such that the bonding in F4Si ← NR3 type complexes, for instance,10–12 is probably best described as a sigma-hole-supported dative bond rather than a sigma hole interaction, especially if NR3 is a strong base.6

Main group systems have received the bulk of the attention in discussions of sigma hole interactions due partly to a significant interest in the properties of organic compounds and complexes and the apparent and potential influences of such interactions in biochemistry, but discussions of sigma hole type interactions involving d-block metals are present in the literature as well.13–16 Indeed, the obsessive naming of weak sigma hole interactions based on the element or group in the periodic table that is involved has not spared the d-block.17–19

One category of transition metal compounds that has become particularly intriguing for the authors is the group 5 oxyhalides. Among those systems, NbOCl3 is one of the best studied. The isolated NbOCl3 molecule may be described as pseudo-tetrahedral about the Nb center such that the Nb[double bond, length as m-dash]O double bond is along the (vertical) C3 principal axis and the basal Cl sites are in a horizontal plane. An early experimental crystal structure for NbOCl320 revealed an assembly of vertical stacks of NbOCl3 dimers. In that solid, the dimer unit is a di-bridged structure in which one Nb–Cl bond from each monomer is involved in a 3-center 4-electron (3-c, 4-e) Nb–Cl–Nb bridge. The other two Cl atoms on each Nb center in the dimer are roughly coplanar with, but point away from, that bridging region, and the two Nb[double bond, length as m-dash]O bonds, which are perpendicular to those basal Cl atoms, are parallel or cis to each other. So, each Nb center in the dimer is locally five-coordinate (square pyramidal) with the exposed position below both Nb[double bond, length as m-dash]O bonds vulnerable to attack.

In the stack of dimers found in the extended solid, the Nb[double bond, length as m-dash]O bonds of one dimer point directly toward those open coordination positions below the Nb[double bond, length as m-dash]O bonds of a second dimer and so forth. That stacking leads, thus, to a pseudo-octahedral coordination environment around each Nb center in the crystal structure with two apical O and four equatorial Cl atoms. Of note, the distances between the Nb atom and those two apical O sites were determined initially20 to be equal – suggesting a symmetrization in the bonding ([double bond, length as m-dash]O⋯Nb[double bond, length as m-dash]O → –O–Nb–O–) along the infinite column of dimers. Four decades later, however, a re-examination of the NbOCl3 crystal structure21 showed that the Nb–O distances are inequivalent, suggesting that the nature of the bonding of Nb to each of the two apical O centers in the pseudo-octahedra is distinct.

In light of the evidence for sigma hole bonding by polarized d-block centers,13,15,16 and the potential role of sigma holes as ordering influences in main group tetrahedral oxyhalides such as POCl3, and POBr322 (and even too in the lower oxidation state systems such as SbOF)23 we became interested in examining the nature of the chemical bonding in group 5 oxyhalides and related systems. Several of the extended solids of group 5 oxyhalides (MOX3; for M = V, Nb, and Ta; X = F,24,25 Cl,20,21,26,27 Br,28 and I28) are known, and Ault et al. have even identified, for instance, ostensibly sigma-hole-supported coordination complexes of VOF3 and VOCl3 with small bases.29,30 We investigate the nature of the bonding in the MOX3 dimers and higher order clusters that lead eventually to the formation of their extended solids. We rationalize the asymmetry identified in ref. 21 and other phenomena in the oxyhalide solids as we probe the role of sigma hole type interactions in fostering if not enforcing structural preferences in extended systems.

Computational methods

All of the molecular species and complexes reported in this work have been optimized using the Gaussian 16 (G16) suite of programs.31 The systems identified as local minima on their respective potential energy surfaces have been confirmed to be so by harmonic vibrational frequency analyses carried out at the same level of theory and employing the same (superfine) grid, as deployed for the associated structural optimization. All of the computational data presented in this contribution have been obtained using the ωB97XD method,32 which shows superior performance for the analysis of weak interactions in particular compared to earlier density functional methods.32 The commonly employed correlation-consistent triple-ζ (cc-pVTZ) basis sets were used for all atoms,33,34 except for the group 5 metals and iodine. For those elements, in order to manage computational costs while minimizing loss of accuracy, we employed small-core MDF pseudopotentials35–39 published by the Stuttgart-Cologne group40 in tandem with the associated triple-ζ basis sets.41 Electrostatic potential (ESP) maps plotted on the 0.001 a.u. isodensity surfaces have been generated using the Gaussview 6 software42 within the electrostatic potential range of ±0.05 a.u. on a color spectrum where blue is positive and red is negative. The actual values of the extrema on the potential surfaces were computed using the Multiwfn software package43,44 and the appropriate output files from the Gaussian 16 calculations. In addition to the GaussView 6 software, the Chemcraft graphical user interface45 was used to produce representations of molecules and clusters examined in this work. All molecular orbital pictures included in this work were generated using GaussView 6.42

Results and discussion

The isolated group 5 MOX3 molecules (for X = H, F, Cl, Br, or I) have a pseudo-tetrahedral geometry at the central M atom. The M center has a highly positive +5 formal charge, with a M[double bond, length as m-dash]O double bond and three M–X single bonds. It is expected, therefore, that the M atom will – as is the case for group 14 central atoms in their MF4 compounds, for example6 – host four sigma holes: one opposite each bond to the M center. There is a M[double bond, length as m-dash]O double bond in the systems under consideration here, but the analogy to group 14 molecules is valid since key requirements for four sigma holes, if M is soft enough, are four polarizing σ bonds and no lone pair on the M center.

Given the high electronegativity of O compared to Cl and the heavier halides, the sigma hole opposite the M[double bond, length as m-dash]O bond is expected to be much stronger in general than those opposite the M–X bonds. For X = F, the situation is harder to predict since F is more electronegative than O.

To move beyond those deductions based solely on chemical intuition, pictures of the computed surface electrostatic potentials, Vs, for the 0.001 au isodensity surface of the MOX3 molecules were generated and examined (see Fig. 1). For each molecule, the ESP maps are oriented in two ways to show examples of the different types of sigma holes on M: one orientation (top in Fig. 1) exposes the ‘O[double bond, length as m-dash]M•’ sigma hole – the sigma hole on M opposite the O[double bond, length as m-dash]M bond, on the triangular ‘X–X–X’ face of the MOX3 pseudo-tetrahedron – with the O atom pointing into the plane of the page; the second orientation (bottom in Fig. 1) shows a ‘X–M•’ sigma hole – one X–M bond points into the plane of the page such that the sigma hole on M opposite that X–M bond (at the center of a ‘X–O–X’ face of the pseudo-tetrahedron) is visible. Both views of the molecular surface offer too some faint evidence of the sigma hole on the X atoms (M–X•) as well. Where it exists, that positive halogen atom sigma hole shows up as a faint greenish-blue or light blue region on the halides at the relevant vertices in the ESP maps in Fig. 1, especially for the iodides. The presence of a negative extremum on the ESP surface at the O site is obvious in some of the second (bottom) set of images as a yellow or red region at the top vertex in the ESP maps. F also exhibits a negative extremum, which is betrayed by a faint yellow at the F sites in Fig. 1.


image file: d5cp03656a-f1.tif
Fig. 1 ESP maps of the group 5 MOX3 molecules. Two orientations are shown for each molecule, exposing the O[double bond, length as m-dash]M• sigma hole at the center of the X–X–X face (top), and a X–M• sigma hole on a X–O–X face (bottom) of the molecule.

For a more quantitative picture, we computed the values of the maximum in the surface potentials, Vs,max, at the O[double bond, length as m-dash]M• and X–M• sigma holes in the MOX3 molecules, and those values are listed in Table 1 (top, for Vs,max(O[double bond, length as m-dash]M•), and bottom for Vs,max(X–M•) for each M). We determined the ESP minima as well, and we will bring those data into the discussion shortly.

Table 1 Values of positive extrema, Vs,max, at selected sigma holes in the surface electrostatic potentials on M and X on the 0.001 au isodensity surface of group 5 MOX3 molecules. The value in bold is the most positive of all of the Vs,max values for that compounda
  σ-Hole F Cl Br I H
In each case, the value in bold is the most positive potential in any sigma hole on the isodensity surface of the molecule.a Since there are three X–M• sigma holes, we provide the average of the computed Vs,max(X–M•) values here, and include the standard deviations in the SI: e.g., the average Vs,max(F–V•) value is 37.16 ± 0.08 kcal mol−1.b For VOI3, the potential maximum at the sigma hole on I, Vs,max(V–I•) = 14.87 ± 0.01 kcal mol−1, is more positive than Vs,max(O[double bond, length as m-dash]V•) = 11.4 kcal mol−1, and Vs,max(I–V•) = 7.71 ± 0.02 kcal mol−1.
V O[double bond, length as m-dash]M• 50.4 21.9 16.3 11.4b 57.5
Nb O[double bond, length as m-dash]M• 81.4 44.9 36.2 27.3 72.3
Ta O[double bond, length as m-dash]M• 82.8 47.9 39.4 30.1 77.0
V X–M• 37.2 17.7 12.8 7.7b 25.8
Nb X–M• 64.5 33.0 26.3 18.9 40.5
Ta X–M• 64.2 32.8 26.1 18.6 45.1


The images shown in Fig. 1 and the Vs,max values shown in Table 1 align remarkably well with the deductions outlined earlier based on qualitative chemical intuition. The most positive Vs,max values on the isodensity surface of the MOX3 molecules (the global maxima – in bold in Table 1) arise at the sigma hole on M opposite the O[double bond, length as m-dash]M bond in every case except for VOI3. For that oxyiodide (see Table S1(a) and (b)), the sigma hole on I, Vs,max(V–I•) = 14.9 kcal mol−1, is more positive than any of the other sigma holes induced on that molecular surface: Vs,max(O[double bond, length as m-dash]V•) = 11.4 kcal mol−1, and Vs,max(I–V•) = 7.7 kcal mol−1.

Among all of the oxyhalides, the most positive Vs,max values are obtained for the O[double bond, length as m-dash]M• sigma holes of the oxyfluorides and oxyhydrides (Table 1 and Table S1(a)). Indeed, the Vs,max(O[double bond, length as m-dash]M•) values vary with X as follows: F > H ≫ Cl > Br > I (see Table 1), with the single exception being the vanadium case, where (H > F). For each X, the Vs,max(O[double bond, length as m-dash]M•) values vary as Ta > Nb ≫ V: for the fluorides, for instance, the Vs,max(O[double bond, length as m-dash]V•) is over 30 kcal mol−1 weaker than Vs,max(O[double bond, length as m-dash]Nb•) and Vs,max(O[double bond, length as m-dash]Ta•), which are 81.4 and 82.8 kcal mol−1, respectively – see Table 1. These data are instructive, for they enable us to make certain inferences about the types of bonding interactions that are likely to be most stabilized by sigma hole interactions if the energies and symmetries of frontier orbitals facilitate such interactions. For any given base, for instance, X3OM ← base complexes of Nb and Ta are expected to establish much stronger sigma hole interactions than their V analogue for any X. And for any M, F and H are predicted to return the most stable complexes.

As we mentioned above, VOI3 is exceptional, since the sigma hole on the iodine center (Vs,max(V–I•) = 14.9 kcal mol−1) is larger than both Vs,max(O[double bond, length as m-dash]V•) and Vs,max(I–V•) −11.4 and 7.7 kcal mol−1 respectively – on the VOI3 iso-surface. So, based on sigma hole strength alone, halogen bonding may be favored over alternative modes of interaction by VOI3 as an electron acceptor.

Insights from model heterodimers

To assess the influence of these sigma holes (Fig. 1) on the ability of group 5 oxyhalides and -hydrides to function as acids, and the relationship between the strengths of the sigma holes and the bonding in acid–base complexes of the group 5 compounds, we investigated the structure, bonding and energetics of their trimethylamine (TMA) complexes. To be sure, ammonia is a simpler base for use in model acid–base systems, but less is known about the trimethylamine complexes30,46 and previous efforts to prepare and analyze Cl3OV ← N(CH3)3 met with frustration46 so we considered that investigating the latter systems may be useful for comparison with future experimental observations. Key structural and thermodynamic parameters for the optimized C3v X3OM⋯N(CH3)3 complexes are shown in Table 2 and the structural coordinates are included in the SI.
Table 2 Geometrical, thermodynamic, and bond order data for the trimethylamine (TMA) complexes of group 5 oxyhalide molecules. Here image file: d5cp03656a-t2.tif

image file: d5cp03656a-u1.tif

  M⋯N distances/Å
  F Cl Br I H
V 2.527 2.816 2.953 3.105 2.339
Nb 2.589 2.737 2.788 2.841 2.522
Ta 2.580 2.725 2.770 2.816 2.514
 
  Wiberg bond indices
V 0.15 0.14 0.13 0.12 0.21
Nb 0.16 0.18 0.18 0.18 0.17
Ta 0.16 0.19 0.19 0.19 0.19
 
  Binding free energies, ΔG298.15[thin space (1/6-em)]Kbind./kcal mol−1
V −2.22 3.88 4.33 4.79 −10.25
Nb −9.81 −3.06 −1.60 0.25 −11.83
Ta −11.81 −4.56 −3.24 −1.38 −13.57


Going down group 5, the atomic radii vary as V < Nb ≈ Ta, which tempts the uninitiated to forecast an increase in M⋯N bond distances for X3OM ← N(CH3)3 complexes going from M = V to Ta. But the trend in the Vs,max(O[double bond, length as m-dash]M•) values shown in Fig. 1 and Table 1 must be considered as well. Notwithstanding the trend in atomic radii, if an intensification of M ← N bonding is sufficiently drastic as M gets larger, an associated contraction in the M⋯N distance will follow, which may outstrip the difference in atomic radii, with the possible result that the minimum energy M⋯N distance actually decreases going down group 5 from V to Ta.

The results in Table 2 suggest that precisely such an outcome is achieved in the trimethylamine complexes: the M⋯N bonds are remarkably short for VOH3 and VOF3, but otherwise the interactions strengthen and the M⋯N distance actually decreases as M gets larger.

A general weakening and elongation of the M⋯N bond as X gets larger going from F to I in Table 2 correlates with a waning in the strength of the O[double bond, length as m-dash]M• sigma hole (see Table 1). This is notable, as we consider the prospect for strong charge transfer in the X3OM ← TMA complexes, since we find that the gap between the highest occupied molecular orbital (HOMO) of the base (TMA) and the lowest occupied molecular orbital (LUMO) of the MOX3 acceptor molecule decreases as the halides get larger (see Table S2) – a trend that is expected to promote rather than diminish X3OM ← TMA bonding.

For these and other ‘O[double bond, length as m-dash]MX3 ← base’ complexes, however, one must acknowledge too the increasingly formidable influence of the X atoms in controlling access to the M center by the base as X gets larger. The base in Table 2 approaches the M center via the X–X–X face of the pseudo-tetrahedral O[double bond, length as m-dash]MX3 molecule, and the three X atoms present in effect a physical guard wall to any base aspiring to get beyond that X–X–X face (see Table 2) to the M center of the O[double bond, length as m-dash]MX3 molecule. Going from X = H to F and down group 17 to I, the electron rich X–X–X face becomes increasingly effective in both shielding the M atom – hence masking the O[double bond, length as m-dash]M• sigma hole – as well as repelling the incoming electron rich base. Consequently, the impact that the substantial polarization of the M center has on dative (OX3M ← base) bond formation is weakened as X gets larger, regardless of the suitability of the HOMO–LUMO gap.

That insight into the impact of the X–X–X face in limiting the exposure of the base to the highly positive M center helps us to make sense of some of the details of our results. Why, for example, do the hydrides (see Fig. 1 and Table 1) have the strongest or second strongest sigma holes for any M for the MOX3 molecules? It is definitely not because H is a more electronegative substituent than F, Cl, Br, or even I and thus more effective in enhancing the polarization of M in the MOX3 species. It is not: on the Pauling scale, χP(F) = 3.98, χP(Cl) = 3.16, χP(Br) = 2.96, χP(I) = 2.66, and χP(H) is only 2.20! That phenomenon – the exceptionally strong Vs,max when X = H – is observed because hydrogen is innocent; the H atom is small and, unlike the halogen X atoms, has no lone pair. So the H–H–H face of the O[double bond, length as m-dash]MH3 molecule allows for the most optimal exposure of the O[double bond, length as m-dash]M• sigma hole to (hence, access by) an incoming base compared to the other X–X–X faces of the series of MOX3 molecules.

Indeed, the minimal shielding effect of the H atoms and the slightly smaller size of V relative to Nb and Ta, may account for the curious observation that Vs,max(O[double bond, length as m-dash]V•) is actually more positive for VOH3 than it is for VOF3. Why is that curious? Although the O[double bond, length as m-dash]M• sigma hole is attributed formally to polarization of M by O, sigma holes on central atoms (M in this case) are influenced somewhat by all substituents. So – if shielding of the M center by the X–X–X face is weak enough – the O[double bond, length as m-dash]M• sigma hole on the O[double bond, length as m-dash]MX3 iso-surface is expected to be more positive when X is more electronegative. For M = V, however, the larger size of the F relative to H, and the fact that F has lone pairs, allows the F–F–F face to mask the positive V center enough (Fig. 1) to attenuate the impact of the substantial polarization of V on the net electrostatic potential that arise at the center of the F–F–F face (i.e. at M) on the iso-surface. For Nb and Ta, which are a bit larger than V, the F atoms are somewhat less effective at shielding the M center. Consequently, fuller expressions of the O[double bond, length as m-dash]M• sigma holes (hence more positive Vs,max(O[double bond, length as m-dash]M•) values) register on the F–F–F face of the iso-surface for both NbOF3 and TaOF3 such that Vs,max(O[double bond, length as m-dash]Nb•) and Vs,max(O[double bond, length as m-dash]Ta•) are larger for X = F than they are for X = H (Table 1). For X = Cl, Br, and I, the central M atoms are increasingly more effectively shielded, so the Vs,max(O[double bond, length as m-dash]M•) values for X = H are always larger than the Vs,max(O[double bond, length as m-dash]M•) values for those larger halides (Table 1) and the X3OM ← base interactions (see Table 2) weaken accordingly.

Following that analysis, therefore, and given that as the electron rich halogen atoms get larger going from X = F to I the shielding of M by the X–X–X face becomes more substantial, it is hardly surprising that (i) the elongation of the M⋯N distance in Table 2 is most dramatic (with a change of just over 0.5 Å going from VOF3 to VOI3) for the smallest and most easily shielded M atom – vanadium – and that (ii) the OX3V ← TMA bonding (Table 2) is exergonic for VOH3 and VOF3, but increasingly endergonic (Table 2), if still exothermic (Table S3), for X = Cl, Br, and I.

Secondary influences such as any X⋯H hydrogen bonding between X substituents on Nb and methyl hydrogens in the X3OM ← TMA complexes may enhance the overall stability of the acid–base pair, but there is no evidence that such interactions play any controlling role in the bonding.

The nature of the orbital interactions contributing to the bonding in the N(CH3)3 complexes is demonstrated by selected molecular orbitals shown in Fig. 2. The M⋯N contacts are all in the range 2.33 and 3.11 Å (Table 2), so they are noticeably longer in many cases, especially for the larger halides, than the sums of the covalent radii of M and N47 – V–N, (1.34 + 0.71) Å = 2.05 Å; Nb–N, (1.47 + 0.71) Å = 2.18 Å; and Ta–N, (1.46 + 0.71) Å = 2.17 Å – but well below the sums of their van der Waals radii – V–N, (2.05 + 1.6) Å = 3.7 Å; Nb–N, (2.15 + 1.6) Å = 3.8 Å; and Ta–N, (2.2 + 1.6) Å = 3.8 Å.48 So electrostatic interactions are likely to dominate these inter-molecular interactions, and increasingly so as the halogen atom gets larger, M⋯N elongates, and the already low bond order slowly deteriorates (see Table 2).


image file: d5cp03656a-f2.tif
Fig. 2 Representations of X3ONb ← N(CH3)3 molecular orbitals that are bonding between the NbOX3 and N(CH3)3 fragments. The position of each orbital relative to the highest occupied molecular orbital (HOMO) for the complex is indicated. All of the MO pictures share the same orientation. That common orientation is depicted on the left.

Homodimers: a rolling landscape

Our discussion of the X3OM ← TMA complexes are of interest since some of those pairs have already been studied experimentally29,30,46 but several others remain to be considered. For this work, however, the homodimers of the metal oxyhalides (MOX3)2 are of particular interest. As we will see presently, dimer units appear in many of the extended solids of group 5 oxyhalides. ‘Vertical’ homodimers analogous structurally to the trimethylamine complexes just described (Fig. 2), in which the two monomers establish a X3M[double bond, length as m-dash]O⋯MOX3 bond mediated by sigma hole and charge transfer interactions, are obtained in this work at the ωB97XD level for only some of the oxyhalide systems. Where it is located, typically for the halides, though never for the fluorides, that homodimer comes with twist. It is somewhat bent as shown in the example on the left (top) in Fig. 3. For X = H and F, the ‘vertical’ starting structure optimizes with imaginary frequencies that lead us, through subsequent searches, to one of the other structures shown in Fig. 3.
image file: d5cp03656a-f3.tif
Fig. 3 Diverse minimum energy isomers of MOX3 dimers obtained from a ‘vertical’ starting structure. The bent structure (i) is stable for all cases where X ≠ H or F. *The other isomeric forms (ii) to (v) were obtained by following imaginary vibrational frequencies from optimizations for X = F and H. Those forms are not examined further in this work. For M = Ta, this bridging structure is quite twisted. Far from C2v, the O–M–M–O dihedral is 13.6°. Located only for TaOH3.

We do not examine the dimer forms in Fig. 3 in detail in this work (see coordinates in the SI), nor do we undertake a more exhaustive or systematic search of the potential energy surface of the homodimers. That is because several of the group 5 MOX3 extended solid structures feature a distinct form of dimer unit and we take our cue on the bonding motifs that we consider herein from there. Those dimers are held together by a pair of covalent bridging bonds (similar to but more symmetric than the distorted MOH3 structure in Fig. 3(iii)), and, importantly, sigma hole type interactions appear to be crucial in stabilizing a one-dimensional tower of those dimers, which is a dominant motif in several of the MOX3 crystal structures.

The extended solids of MOX3

We have been interested in the donor acceptor properties of group 5 oxyhalides because they hold for us potential insights into the nature of the bonding present in the NbOCl3 extended solid, for which there have been some ambiguities in the literature.20,21 We have also been interested more broadly in understanding the potential roles of sigma hole interactions in stabilizing acid–base type interactions in the larger class of group 5 oxyhalides and inorganic extended solids in general.

We have considered the oxyhydride molecules in our discussion so far for completeness, but we have been unable to find crystal structures for any of the group 5 MOH3 compounds. And many of the oxyhalide crystal structures are missing as well: the oxy-bromides and iodides of V and Ta, and tantalum oxychloride are evidently still unavailable.49 Representations of published oxyhalide crystal structures are shown in Fig. 4.49,50 The crystal structures available for NbOF3 and TaOF3 are disordered in the O and F sites and representations of those disordered systems generated using the CrystalMaker modelling program50 are included in the SI. A published structure with a significant degree of uncertainty in the Nb position in NbOBr3,28 leading to an average structure that is a symmetrical version of that shown in Fig. 4 is included in the SI as well.


image file: d5cp03656a-f4.tif
Fig. 4 Images of know crystal structures of group 5 oxyhalides. In most case, two views are presented to brings into focus the layout of one-dimensional stacks as well as the long inter-monomer or -dimer M⋯O contacts in the solids (striped bonds). Disordered structures published for NbOF3, TaOF3, and NbOBr3 are shown in the SI.

The structures of the MOX3 solids are interesting (Fig. 4) since the involvement of sigma-hole-supported bonding is apparent in each. Of the known crystal structures, the VOF3 structure is the only one that exhibits no evidence of M[double bond, length as m-dash]O⋯M (or the symmetrical M–O–M) inter-monomer or inter-dimer bonding about the oxygen centers. That vanadium system (top left in Fig. 4) is composed of a network of VOF3 dimers interlinked by V⋯F bonds (see the striped bonds in Fig. 4(i)). The dimers are similar to the doubly bridged systems described before (F2OVF2VOF2) but with the V[double bond, length as m-dash]O bonds trans relative to a pair of asymmetric ‘V–F–V’ bridges with bond lengths of 1.920 Å, and 1.972 Å. Those dimers are interlinked by long V⋯F sigma hole type bonds (2.294 Å) between a terminal F on one dimer and the vacant apical position on the square-pyramidal V site of an adjacent dimer (Fig. 4(i)). That is, the electron rich terminal F atom on one dimer interacts directly with the positive O[double bond, length as m-dash]V• sigma hole on the second dimer to form a O[double bond, length as m-dash]V(F4)⋯F distorted octahedron.

This observation that VOF3 is the only crystal structure in Fig. 4 devoid of M⋯O type inter-monomer or -dimer bonding motivated us to look again at the ESPs on the iso-surfaces of the MOX3 molecules, focusing in this case on the distribution of ESP minima (Vs,min). And, remarkably (see Table 3), we find that, in line with the exceptional role of F in the VOF3 crystal structure, where M⋯O bonding is absent and M⋯F interactions dominate, VOF3 is the only molecule in the whole series in Table 3 in which the negative extremum at F (Vs,min(M–F▶)) is more negative than Vs,min(M[double bond, length as m-dash]O▶). So, VOF3 is the single instance for which the global minimum – the most negative spot on the ESP surface – is at the X rather than the O site (see Table S1(a) and (b)).

Table 3 Minima of the surface electrostatic potentials in kcal mol−1 at O (top) and X (bottom), Vs,min, on the 0.001 au isodensity surface of group 5 MOX3 molecules. The value is in bold if it is the most negative (least positive) Vs,min value anywhere on the surface of that compoundab
    F Cl Br I H
a The ‘▶’ symbol is used in this work to indicate an ESP minimum. b For the more polarizable X atoms, the local ESP minimum on the atom (M–X▶) arises not opposite the M–X bond but about the electron-rich equator or belt of the X atom. Where different minima appeared at points on that belt, the most negative value was selected. F has no sigma hole (no Vs,max) in the MOF3 compounds; in fact, the negative extremum, Vs,min(M–F▶), is at the pole on F opposite the M–F bond. Each MOX3 molecule has three essentially identical Vs,min(M–X▶) values; the average is given in this table and the standard deviation to two decimal places is given in Table S1b.
V M[double bond, length as m-dash]O▶ −7.9 −9.0 −10.7 −13.1 −19.0
Nb M[double bond, length as m-dash]O▶ −18.8 −16.3 −17.2 −18.3 −23.6
Ta M[double bond, length as m-dash]O▶ −26.1 −23.1 −23.4 −23.9 −27.8
V M–X▶b −9.0 −1.3 −1.8 −1.1
Nb M–X▶b −10.1 −1.2 −1.1 −1.4 −5.6
Ta M–X▶b −9.1 −1.2 −1.4 +1.7 −5.6


In the only known VOCl3 crystal structure, dimerization, which was already asymmetrical in the fluoride, is abandoned. The solid is an assembly of antiparallel chains of VOCl3 monomers weakly linked to each other by repeating V[double bond, length as m-dash]O⋯V bonds (though published structures disagree a bit on the long intermonomer O⋯V distance: 3.402 Å26 and 3.45527). In VOF3, the most substantial ESP extrema (the maximum at O[double bond, length as m-dash]V• and the minimum at V–F▶) drive oligomerization; and that is the case for VOCl3 as well, but in the latter case, the global maximum is again O[double bond, length as m-dash]V• while the global minimum is V[double bond, length as m-dash]O▶ instead of V–Cl▶ (see Fig. 4(ii), and Table 3).

All of the other crystal structures in Fig. 4 (NbOCl3, NbOBr3, and NbOI3) feature stacks of symmetric dimers with two Nb–X–Nb bridges of equal Nb–X bond lengths. The dimers are arranged in vertical stacks with each dimer linked to the next dimer above it by two Nb[double bond, length as m-dash]O⋯Nb contacts. Those long Nb⋯O bonds – 2.203 Å, 2.210 Å, and 2.233 Å for X = Cl, Br, and I, respectively – are much shorter than the intermonomer V⋯O contacts in VOCl3 (vide supra) a feature that might be anticipated given the escalations in the magnitudes of both Vs,max(O[double bond, length as m-dash]M•) and Vs,min(M[double bond, length as m-dash]O▶) going from M = V to Ta for each X (see Tables 1 and 3). So, the M[double bond, length as m-dash]O⋯M interactions between monomer or dimer units in the solids (Fig. 4) are expected to strengthen based on the sigma hole trends alone going down group 5.

Even so, the structural evidence suggests that the double bond character of the M[double bond, length as m-dash]O bonds is largely preserved in the extended systems (Fig. 4). The short metal–oxygen (Nb[double bond, length as m-dash]O) distances in the NbOX3 dimer units in the crystal structures, for example, are only 1.758 Å, 1.743 Å, and 1.763 Å for NbOCl3, NbOBr3, and NbOI3, respectively, which are all more than 0.4 Å shorter than the long Nb⋯O contacts in the crystal structures (vide supra; see the striped bonds in Fig. 4), and are within 0.07 Å of the experimental gas phase electron diffraction (ra) Nb[double bond, length as m-dash]O double bond distances in the isolated molecules: 1.693(4) Å, 1.693(7) Å, and 1.717(3) Å.51 Covalent bonds to central atoms often elongate upon oligomerization or solids formation, especially if the coordination (crowding) around atomic centers increase (such as at M, going from four-coordinate in the MOX3 molecule to six-coordinate in the solids), so the rather slight increases in the M[double bond, length as m-dash]O bonds of less than 0.07 Å going from the molecule to the solid signals no grand change in the basic nature of that double bond. What of the nature of the long M⋯O interaction? As we show in the next section, focusing on the oxyhalides since we have no data for MOH3 solids, the alternating M[double bond, length as m-dash]O⋯M bonding pattern is not exclusively a solid-state phenomenon at all.

MOX3 bridged dimers

The halogen-bridged dimers identified in many of the extended solids are local minima on the potential surfaces of all of the oxyhalide pairs (see Table 4).52 In line with a reviewer request, the basic nature of the chemical bonding in such di-bridged structures, and the influence of bridge formation on the M–X distances, is discussed briefly in the SI. The coordinates of the optimized dimers are provided as .xyz files in the SI. The free energy changes going from the monomers to the dimers are included in Table 4, and the corresponding enthalpies and zero-point corrected energies are available in Table S4. The strengths of the extrema in the electrostatic potentials at M and O on the dimer iso-surface are included in Table 4 as well.
Table 4 V s,max(O[double bond, length as m-dash]M•) and Vs,min(M[double bond, length as m-dash]O▶) data, all in kcal mol−1 units, and binding free energies for the dimers of group 5 oxyhalides relative to isolated monomers

image file: d5cp03656a-u2.tif

  F Cl Br I
a The sigma hole on I has the most positive Vs,max on the iso-surface of the VOI3 monomer. In the dimer, however, the sigma hole on the terminal iodides (Vs,max(V–I•) = 13.5 kcal mol−1) is weaker than Vs,max(O[double bond, length as m-dash]V•) = 27.5 kcal mol−1. b For the VOF3 dimer, the most negative sites on the iso-surface is that on the pole of the terminal F atoms (opposite the V–F bond) where Vs,min(M–X▶) = −9.5 kcal mol−1. In all other cases, the most negative Vs,min value is Vs,min(V[double bond, length as m-dash]O▶).
  O[double bond, length as m-dash]M•
V 70.6 38.9 33.0 27.5 a
Nb 98.7 67.4 58.2 48.3
Ta 102.2 71.4 62.0 51.9
 
  M[double bond, length as m-dash]O▶
V −2.0b −4.7 −7.3 −10.8
Nb −15.1 −13.5 −15.0 −16.9
Ta −21.5 −20.0 −21.1 −22.5
 
  Binding free energies/kcal mol−1
V 13.3 28.0 27.7 23.0
Nb −7.6 10.0 12.5 12.9
Ta −8.6 9.0 11.6 12.8


The bonding is endergonic in these side-on di-bridged dimers, except for Nb and Ta oxyfluorides, but the enthalpies are much more negative (see Table S4). And that structural motif (Table 4) is a minimum on the potential energy surface for all of the oxyhalide molecules, and persist, as we have shown above, in several of the extended solids at ambient conditions. The Vs,max and Vs,min data summarized in Table 4 (see also Table S5 for the global ESP extrema for each dimer) show that the stacking of the dimer pairs, an acid–base type interaction mediated by two parallel O⋯M bonds that characterizes the NbOCl3, NbOBr3, and NbOI3 crystal structures in Fig. 4, is strongly favored by the distribution of the electron density in the dimer units.

In those systems, the ‘acidic’ Vs,max(O[double bond, length as m-dash]M•) and ‘basic’ Vs,min(M[double bond, length as m-dash]O▶) sites involved in the O⋯M inter-dimer interactions in the extended solid are the most positive and most negative extrema, respectively, on almost all of the MOX3 surfaces (Table S5). VOF3 is the only exception. In that case, Vs,max(O[double bond, length as m-dash]V•) is still the most positive extremum, but the minimum on F (Vs,min(V–F▶) = −9.5 kcal mol−1) is much more negative than Vs,min(V[double bond, length as m-dash]O▶) = −2.0 kcal mol−1. As we pointed out above, a similar observation of a relatively strong negative extremum at F in the free VOF3 molecule helps us to make sense of the preference for the exceptional and relatively complicated M⋯X mediated structure in Fig. 4(i) which has no apparent inter-dimer M⋯O bonding.

MOX3 tetramers

An assessment of the initial step along the path to the infinite stack of dimers seen in many of the crystal structures is instructive. We find that the basic mode of bonding observed in those solids (Fig. 4) is already present in the tetramer (see Fig. 5). The O sites in the lower bridged dimer in Fig. 5 align with the exposed base of the M atoms of the upper dimer, where the antibonding σ* orbital of the O[double bond, length as m-dash]M σ bond and O[double bond, length as m-dash]M• sigma hole of that upper dimer coincide. A model of that type of tetramer structure is shown in Table 5 along with the binding free energies for the pairing process (ΔGbind = G(tetramer) − 2 × G(dimers)), with the corresponding enthalpies and zero-point corrected energies provided in Table S4. The extent of any charge transfer between the two dimers has been quantified as well in the form Wiberg bond indices for the inter-dimer M⋯O contacts (Table 5).
image file: d5cp03656a-f5.tif
Fig. 5 MOX3 tetramers built up from bridged dimers; showing NbOF3 (left) and NbOCl3 (right). The tetramer is tilted for VOCl3, NbOF3, and TaOF3. This basic motif is not a minimum for VOF3.
Table 5 Selected bonding, geometrical, and free energy data for the group 5 oxyhalide tetramers (stacked di-bridge dimer pairs). The distances between the metal atom (uM) in the ‘upper’ dimer and the oxygen site (lO) in the ‘lower’ dimer are listed. The binding free energies are computed relative to isolated dimer pairs (ΔG298.15[thin space (1/6-em)]Kbind. = G298.15[thin space (1/6-em)]Ktetramer − 2G298.15[thin space (1/6-em)]Kdimer)

image file: d5cp03656a-u3.tif

  Fa Cl Br I
a The Nb and Ta fluorides and VOCl3 have Cs symmetry. In those cases, one dimer is tilted relative to the other – a tetrameric version of the bent vertical dimers (Fig. 3(i)). The two inter-dimer M⋯O bonds are retained as well as the mirror plane passing through the four bridging X atoms. b This tetrameric structure is not a minimum for VOF3. c There was a slight difference between the two V⋯O bonds with an average of 2.9826 ± 0.0015 Å.
  uM⋯lO bond distance/Å
V b 2.983ac 2.825 2.810
Nb 2.566 2.573 2.559 2.563
Ta 2.493 2.481 2.463 2.463
 
  uM⋯lO Wiberg bond indices
V b 0.06a 0.10 0.13
Nb 0.13 0.19 0.20 0.21
Ta 0.16 0.22 0.24 0.24
 
  Binding free energies/kcal mol−1
V b +5.4a +3.6 0.0
Nb −5.2 −4.4 −5.3 −8.5
Ta −16.1 −13.7 −16.8 −22.6


These results provide the first hint that the patterns found in the extended solids have their roots in some of the smallest clusters of the molecular systems. Indeed, the perception that the oxyfluorides exhibit a distinct behavior compared to the other oxyhalides is also evident in the tetramer as a non-linear stacking of the dimer pairs for M = Nb and Ta (Fig. 5). The tetramer is also tilted for VOCl3, but for VOF3 the departure is even more elaborate. Neither the linear nor the tilted tetramer is a minimum. Instead, a lower symmetry tetramer (see the SI) was found in which one dimer is displaced horizontally relative to the other. So, we begin to see already that, as oligomerization progresses toward the extended solid, VOF3 will go its own way.

Insights from higher order oligomers for NbOCl3

The published NbOCl3 crystal structures shown in Fig. 4 (cases (iii) and (iv)) disagree on the (a)symmetry of the Nb–O–Nb inter-dimer contacts. That ambiguity was likely due to high uncertainties in the atomic positions in the early symmetric proposal,20 which were resolved several decades later.21 That the bonding in the isolated tetramer of NbOCl3 (Fig. 5) is clearly remembered in the extended solid (Fig. 4), is strong evidence that the observed distortion or asymmetry in the columns of dimers in the crystal structures is not due to any phenomenon specific to the solid state. Rather, it is a culmination of a pattern initiated at the tetramer based ostensibly on sigma-hole-supported bonding interactions with weak though not insignificant charge transfer components (see bond indices in Table 5). To assess that inference, we obtained and examined optimized geometries of the hexamer, octamer, and decamer of NbOCl3 in the stacked dimer motif (Table 6).
Table 6 Bonding and thermodynamic data for higher oligomers of the NbOCl3 moleculea
a The corresponding values for the tetramer are included in Table 5 but are echoed here for direct comparison: Nb⋯O distance: 2.573 Å; bond order: 0.19; ΔG298.15[thin space (1/6-em)]Kolig. = −4.4 kcal mol−1.
Hexamer Octamer Decamer
image file: d5cp03656a-u4.tif image file: d5cp03656a-u5.tif image file: d5cp03656a-u6.tif
     
Shortest intermonomer contacts/Å
2.479 2.417 2.384
 
Wiberg bond indices    
0.23 0.26 0.27
image file: d5cp03656a-t3.tif/kcal mol−1    
−11.6 −21.2 −33.4


The coordinates for all of those oligomers are included in the SI. Costly density functional calculations that we performed on the NbOCl3 oligomers up to the decamer prove illuminating. They afford us some useful insights into the evolution of structural and energetic properties of the clusters en route from the doubly bridged dimer units to the extended solids. A clear trend towards shorter average interatomic separations, a slow but steady increase in bond orders, and an escalation of the binding free energies relative to the free dimers are illustrated in Fig. 6.


image file: d5cp03656a-f6.tif
Fig. 6 Illustrations of the dependence of certain bonding and thermodynamic properties of NbOCl3 clusters on oligomer size. (i) Total oligomerization free energies at 298.15 K, ΔG298.15[thin space (1/6-em)]K, for oligomerization relative to the dimer units (ΔG = Gn-mer − (n/2) × Gdimer), and per dimer (ΔG/(n/2)). (ii) Wiberg bond orders, and (iii) Inter-dimer M⋯O distances.

There is a limited number of data points (from the tetramer to the decamer) for the bond order and bond distance values, but a logarithmic relationship provides the best fit in those cases. A quadratic function affords a surprisingly impressive fit to the free energy as a function of n, and a linear relationship with a high coefficient of determination is found for a theoretical average increase in the stabilization per dimer as the oligomer grows. We find that for each successive dimer added – up to the decamer at least – more stability is conferred on the entire system (Fig. 6(i)). And that stabilization comes with a marginal increase in charge transfer, estimated here by computed bond orders, and a net contraction in the mediating Nb⋯O contacts.

The structural motif in NbOCl3 and in the isomorphic NbOBr3, and NbOI3 systems shown in Fig. 4 may be described, therefore, as the assembly of oligomeric towers of dimers, where each tower is an expansion on a pattern of bonding already inaugurated at the tetramer and propagated by exergonic sigma-hole-supported interactions. Those interactions become more stabilizing and ever so slightly more dative as stacking progresses. The double bond character of the M[double bond, length as m-dash]O bonds in the dimer units, however, persists. So covalent inter-dimer M–O–M bridges are never achieved,53 but the sub-covalent M⋯O distance definitely contracts as oligomerization progresses (Fig. 6(iii)) through to the extended solid, culminating, for NbOCl3, at 2.203 Å.

Summary and conclusion

Molecular clusters and solids of inorganic compounds remain a rich area for exploration in efforts to enhance our understanding of the nature, role, and potential applications of weak interactions. The group 5 MOX3 oxyhalides utilize such interactions to stabilize molecular aggregation and extended solid formation. We have shown here that the bonding motif exhibited by many of the group 5 oxyhalide solids are already evident at the level of the tetramer, which is a stack of doubly bridged dimers held together by electrostatic (sigma hole) and weak charge transfer interactions. The charge transfer contribution, based on bond order and structural data, increases gradually as oligomerization progresses, one additional dimer after another, though the alternating non-covalent and double bond [⋯M[double bond, length as m-dash]O]n motif never symmetrize to a chain [–M–O–]n of identical M–O bonds. Sigma hole interactions play a significant role in sustaining the stacked dimer structures. We surmise, however, that weak orbital interactions may help to account for subtle structural features such as the tilting or bending in certain dimer and tetramer isomers. Vanadium oxyfluoride is in a class of its own as an extended solid compared to the other known group 5 oxyhalides. And its exceptional features – in particular an exclusive role for V⋯F inter-dimer interactions – are traced back to the distribution of electron density in the VOF3 monomer, for which the global minimum in the electrostatic potential resides on the F center (rather than on the O site, as in the other group 5 MOX3 molecules).

Author contributions

D. H. Jr.: methodology, computational studies, data curation, analysis, writing preliminary drafts of sections of manuscript, and reviewing: G. F. S.: methodology, computational studies, data curation, analysis, and reviewing K. J. D.: conceptualization, supervision, methodology, computational studies, data curation, analysis, writing, reviewing, editing.

Conflicts of interest

There are no conflicts to declare.

Data availability

The data supporting this article have been included as part of the SI. Supplementary information: includes several visualizable .xyz files for systems considered in this work, additional thermodynamic and other data mentioned in the text, images of crystal structures, and the full long reference of the Gaussian 16 software used in completing work leading to results presented in this manuscript. See DOI: https://doi.org/10.1039/d5cp03656a.

Acknowledgements

Our work was supported by the National Science Foundation [NSF-RUI Award (CHE-2055119) and NSF-MRI grants (CHE-0958696 (University of Richmond) and CHE-1662030 (the MERCURY consortium))] and by the Henry Dreyfus Teacher-Scholar Awards Program [TH-16-015]. K. J. D. acknowledges as well the support of the University of Richmond (UR) and the Clarence E. Denoon Jr. Endowment in the Natural Sciences at UR.

References

  1. C. B. Aakeroy and K. R. Seddon, The Hydrogen Bond and Crystal Engineering, Chem. Soc. Rev., 1993, 22, 397 RSC.
  2. P. Metrangolo, G. Resnati, T. Pilati and S. Biella, Halogen Bonding in Crystal Engineering, Struct. Bonding, 2008, 126, 105–136 CrossRef CAS.
  3. A. Mukherjee, S. Tothadi and G. R. Desiraju, Halogen Bonds in Crystal Engineering: Like Hydrogen Bonds yet Different, Acc. Chem. Res., 2014, 47, 2514–2524 CrossRef CAS.
  4. A. Frontera and A. Bauzá, On the Importance of σ-Hole Interactions in Crystal Structures, Crystals, 2021, 11, 1205 CrossRef CAS.
  5. K. J. Donald, B. K. Wittmaack and C. Crigger, Tuning σ-Holes: Charge Redistribution in the Heavy (Group 14) Analogues of Simple and Mixed Halomethanes Can Impose Strong Propensities for Halogen Bonding, J. Phys. Chem. A, 2010, 114, 7213–7222 CrossRef CAS PubMed.
  6. K. J. Donald and M. Tawfik, The Weak Helps the Strong: Sigma-Holes and the Stability of MF4•Base Complexes, J. Phys. Chem. A, 2013, 117, 14176–14183 CrossRef CAS.
  7. S. L. Stephens, N. R. Walker and A. C. Legon, Internal Rotation and Halogen Bonds in CF3I⋯NH3 and CF3I⋯N(CH3)3 Probed by Broadband Rotational Spectroscopy, Phys. Chem. Chem. Phys., 2011, 13, 20736–20744 RSC.
  8. A. J. Parker, J. Stewart, K. J. Donald and C. A. Parish, Halogen Bonding in DNA Base Pairs, J. Am. Chem. Soc., 2012, 134, 5165–5172 CrossRef CAS PubMed.
  9. O. S. Bushuyev, D. Tan, C. J. Barrett and T. Friscic, Fluorinated Azobenzenes with Highly Strained Geometries for Halogen Bond-Driven Self-Assembly in the Solid State, CrystEngComm, 2015, 17, 73–80 RSC.
  10. A. A. Ennan and B. M. Kats, Silicon Tetrafluoride Adducts, Russ. Chem. Rev., 1974, 43, 539–550 CrossRef.
  11. B. S. Ault, Matrix-Isolation Studies of Lewis Acid/Base Interactions: Infrared Spectra of the 1:1 Adduct SiF4•NH3, Inorg. Chem., 1981, 20, 2817–2822 CrossRef CAS.
  12. T. J. Lorenz and B. S. Ault, Matrix-Isolation Studies of Lewis Acid/Base Interactions. 2. 1/1 Adduct of SiF4 with Methyl-Substituted Amines, Inorg. Chem., 1982, 21, 1758–1761 CrossRef CAS.
  13. C. Chieh, Synthesis and Structure of Dichlorobis(Thiosemicarbazide)Mercury(II), Can. J. Chem., 1977, 55, 1583–1587 CrossRef CAS.
  14. Ref. 13 is mentioned in ref. 15 as an earlier example of what is putatively a sigma hole interaction to a Hg center. In that system, the Hg center is pseudo-tetracoordinate with two short Hg–S bonds (<2.42 Å) and two short Hg–Cl bonds (<2.83 Å) but features as well two somewhat longer Cl⋯Hg contacts (3.250(3) Å), stabilized evidently by sigma hole type interactions.
  15. J. Joy and E. D. Jemmis, Designing M-Bond (X-M⋯Y, M = Transition Metal): σ-Hole and Radial Density Distribution, J. Chem. Sci., 2019, 131, 117 CrossRef CAS.
  16. S. Scheiner, Participation of Transition Metal Atoms in Noncovalent Bonds, Phys. Chem. Chem. Phys., 2024, 26, 27382–27394 RSC.
  17. J. H. Stenlid, A. J. Johansson and T. Brinck, σ-Holes and σ-Lumps Direct the Lewis Basic and Acidic Interactions of Noble Metal Nanoparticles: Introducing Regium Bonds, Phys. Chem. Chem. Phys., 2018, 20, 2676–2692 RSC.
  18. A. Bauzá, I. Alkorta, J. Elguero, T. J. Mooibroek and A. Frontera, Spodium Bonds: Noncovalent Interactions Involving Group 12 Elements, Angew. Chem., Int. Ed., 2020, 59, 17482–17487 CrossRef.
  19. I. Alkorta, J. Elguero and A. Frontera, Not Only Hydrogen Bonds: Other Noncovalent Interactions, Crystals, 2020, 10, 180 CrossRef CAS.
  20. D. E. Sands, A. Zalkin and R. E. Elson, The Crystal Structure of NbOCl3, Acta Crystallogr., 1959, 12, 21–23 CrossRef CAS.
  21. M. Ströbele and H.-J. Meyer, Neubestimmung Der Kristallstruktur von NbOCl3, Z. Anorg. Allg. Chem., 2002, 628, 488–491 CrossRef.
  22. P. R. Varadwaj, A. Varadwaj, H. M. Marques and K. Yamashita, The Phosphorus Bond, or the Phosphorus-Centered Pnictogen Bond: The Covalently Bound Phosphorus Atom in Molecular Entities and Crystals as a Pnictogen Bond Donor, Molecules, 2022, 27, 1487 CrossRef CAS.
  23. A. Åström and S. Andersson, The Crystal Structure of L-SbOF, J. Solid State Chem., 1973, 6, 191–194 CrossRef.
  24. J. Supeł, U. Abram, A. Hagenbach and K. Seppelt, Technetium Fluoride Trioxide, TcO3F, Preparation and Properties, Inorg. Chem., 2007, 46(14), 5591–5595 CrossRef.
  25. J. Köhler, A. Simon, L. van Wüllen, S. Cordier, T. Roisnel, M. Poulain and M. Somer, Structures and Properties of NbOF3 and TaOF3—with a Remark to the O/F Ordering in the SnF4 Type Structure, Z. Anorg. Allg. Chem., 2002, 628, 2683–2690 CrossRef.
  26. J. Galy, R. Enjalbert, G. Jugie and J. Strähle, VOCl3: Crystallization, Crystal Structure, and Structural Relationships: A Joint X-Ray and 35Cl-NQR Investigation, J. Solid State Chem., 1983, 47, 143–150 CrossRef CAS.
  27. S. I. Troyanov, Crystal Structures of SbCl5, VCl4, and VOCl3, Russ. J. Inorg. Chem., 2005, 50, 1727–1732 Search PubMed.
  28. S. Hartwig and H. Hillebrecht, Crystal Structures of NbOI3 and NbOBr3 – Polar Double Chains in Different Non-Centrosymmetric Structures, Z. Anorg. Allg. Chem., 2008, 634(1), 115–120 CrossRef CAS.
  29. B. S. Ault, Infrared Matrix Isolation Study of the 1:1 Molecular Complex of OVCl3 with (CH3)2O, Spectrochim. Acta, Part A, 2003, 59, 1989–1994 CrossRef PubMed.
  30. M. D. Hoops and B. S. Ault, Matrix Isolation Infrared Spectroscopic Investigation of the Coordination Chemistry and Reactivity of OVF3, J. Mol. Struct., 2002, 616, 91–101 CrossRef CAS.
  31. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, G. A. Petersson and H. Nakatsuji, Gaussian 16, rev. C.02, Gaussian, Inc., Wallingford, CT, 2019 Search PubMed.
  32. J.-D. Chai and M. Head-Gordon, Long-range corrected hybrid density functionals with damped atom-atom dispersion corrections, Phys. Chem. Chem. Phys., 2008, 10, 6615–6620 RSC.
  33. R. A. Kendall, T. H. Dunning Jr. and R. J. Harrison, Electron Affinities of the First-row Atoms Revisited. Systematic Basis Sets and Wave Functions, J. Chem. Phys., 1992, 96, 6796–6806 CrossRef CAS.
  34. D. E. Woon and T. H. Dunning Jr, Gaussian Basis Sets for Use in Correlated Molecular Calculations. III. The Atoms Aluminum through Argon, J. Chem. Phys., 1993, 98, 1358–1371 CrossRef CAS.
  35. M. Dolg, U. Wedig, H. Stoll and H. Preuss, Energy-adjusted Ab Initio Pseudopotentials for the First Row Transition Elements, J. Chem. Phys., 1987, 86, 866–872 CrossRef CAS.
  36. K. A. Peterson, D. Figgen, E. Goll, H. Stoll and M. Dolg, Systematically Convergent Basis Sets with Relativistic Pseudopotentials. II. Small-Core Pseudopotentials and Correlation Consistent Basis Sets for the Post-d Group 16–18 Elements, J. Chem. Phys., 2003, 119, 11113–11123 CrossRef CAS.
  37. K. A. Peterson, B. C. Shepler, D. Figgen and H. Stoll, On the Spectroscopic and Thermochemical Properties of ClO, BrO, IO, and Their Anions, J. Phys. Chem. A, 2006, 110, 13877–13883 CrossRef CAS PubMed.
  38. K. A. Peterson, D. Figgen, M. Dolg and H. Stoll, Energy-Consistent Relativistic Pseudopotentials and Correlation Consistent Basis Sets for the 4d Elements Y–Pd, J. Chem. Phys., 2007, 126, 124101 CrossRef.
  39. D. Figgen, K. A. Peterson, M. Dolg and H. Stoll, Energy-Consistent Pseudopotentials and Correlation Consistent Basis Sets for the 5d Elements Hf–Pt, J. Chem. Phys., 2009, 130, 164108 CrossRef PubMed.
  40. Energy-consistent pseudopotentials of the Stuttgart-Cologne group. https://www.tc.uni-koeln.de/PP/clickpse.en.html (last accessed August 28, 2025).
  41. J. M. Martin and A. Sundermann, Correlation Consistent Valence Basis Sets for Use with the Stuttgart–Dresden–Bonn Relativistic Effective Core Potentials: The Atoms Ga–Kr and In–Xe, J. Chem. Phys., 2001, 114(8), 3408–3420 CrossRef CAS.
  42. R. Dennington, T. A. Keith and J. M. Millam, GaussView, 6.0, Semichem Inc., Shawnee Mission, KS, 2016 Search PubMed.
  43. T. Lu and F. Chen, Multiwfn: a multifunctional wavefunction analyzer, J. Comput. Chem., 2012, 33, 580–592 CrossRef CAS.
  44. T. Lu and F. Chen, Quantitative analysis of molecular surface based on improved Marching Tetrahedra algorithm, J. Mol. Graphics Modell., 2012, 38, 314–323 CrossRef CAS PubMed.
  45. Chemcraft – Graphical Software for Visualization of Quantum Chemistry Computations https://www.chemcraftprog.com.
  46. B. S. Ault, Matrix isolation study of the thermal and photochemical reaction of OVCl3 with NH3: Spectroscopic and theoretical characterization of Cl2V(O)NH2, J. Phys. Chem. A, 2001, 105, 4758–4764 CrossRef CAS.
  47. P. Pyykkö and M. Atsumi, Molecular Single-Bond Covalent Radii for Elements 1–118, Chem. – Eur. J., 2009, 15, 186–197 CrossRef.
  48. S. S. Batsanov, Inorg. Mater., 2001, 37, 871–885 CrossRef CAS (translated from Neorg. Mater., 2001, 37, 1031–1046). We use here for consistency the crystallographic van der Waals radii for all of the elements in question as presented in Table 9 of this reference.
  49. NIST Inorganic Crystal Structure Database, NIST Standard Reference Database Number 3, National Institute of Standards and Technology, Gaithersburg MD, 20899 DOI:10.18434/M32147, (retrieved August 2025).
  50. Crystal structure images were generated using CrystalMaker®: a crystal and molecular structures program for Mac and Windows. CrystalMaker Software Ltd, Oxford, England (https://www.crystalmaker.com).
  51. I. Nowak, E. M. Page, D. A. Rice, A. D. Richardson, R. J. French, K. Hedberg and J. S. Ogden, Chalcogenide-Halides of Niobium(V). 1. Gas-Phase Structures of NbOBr3, NbSBr3, and NbSCl3. 2. Matrix Infrared Spectra and Vibrational Force Fields of NbOBr3, NbSBr3, NbSCl3, and NbOCl3, Inorg. Chem., 2003, 42, 1296–1305 CrossRef CAS PubMed.
  52. A version of the C2v doubly bridged dimer (see Fig. 3(iii)) is obtained for (MOH3)2, but it is very distorted for M = Ta (see also the.xyz coordinates for the MOH3 dimer in the SI).
  53. The computed Nb[double bond, length as m-dash]O double bond distance in the NbOCl3 molecule is 1.668 Å (comparable to 1.693(4) from experiment), and the Nb–O single bond in the computed trigonal bipyramidal NbCl4OH molecule is 1.863 Å, which is still much shorter than the computed inter-dimer distance of 2.573 in the tetramer or even the experimental 2.203 Å in the extended solid.

This journal is © the Owner Societies 2026
Click here to see how this site uses Cookies. View our privacy policy here.