The experimental Xe-O bond length in XeO4 determined by electron diffraction is 1.736 Å and the O-O distance is 2.832 Å.13 The geometries of NgO4 (Ng = Xe,Rn) were optimised at the CCSD(T)/DKH-TZVPP level and the bond distances are displayed on Fig. 1. The calculated values are in good agreement with the experimental ones for XeO4 (1.756 Å and 2.867 Å, respectively) differing at the most by 0.035Å.
RnO4 exhibits as expected Td point group symmetry, and is a confirmed minimum on the potential energy surface i.e. possessing all real vibrational modes. The calculated bond lengths of this compound are, as expected, slightly larger than those of the xenon congener.
The bond strengths in RnO4 however are palpably inferior to those of XeO4, as is apparent from comparison of the calculated vibrational frequencies (Table 1 and Fig. S1). The Ng-O stretches with t2 symmetry in particular are both IR active and appear at a lower frequency for RnO4 than for XeO4.
Table 1
– Calculated vibrational modes of NgO4 at the CCSD(T)/DKH-TZVPP level compared to the experimental spectra of XeO4 measured by Selig et al.12
Mode
|
RnO4 (cm− 1)
|
XeO4 (cm− 1)
|
XeO4 Exp.12 (cm− 1)
|
e
|
205
|
264
|
|
t2
|
225
|
306
|
306
|
a1
|
706
|
777
|
|
t2
|
799
|
873
|
877
|
To understand why the Ng-O bonds in RnO4 are weaker than in XeO4, a more detailed electronic structure analysis of the CCSD orbitals was carried via Natural Bond Orbital analysis. The Natural Population Analysis (NPA) in particular affords some insight into the distribution of the electrons within the levels of the formally NgVIII tetroxides (Table 2) .
Table 2
– NBO analysis of the CCSD/DKH-TZVPP natural orbitals: electron occupation and composition of orbitals.
Natural Population
|
XeO4
|
RnO4
|
NAO Ng 5s/6s
|
1.416
|
1.534
|
NAO Ng 5p/6p
|
2.889
|
2.691
|
|
NLMO composition
|
|
σ(Ng-O)
|
47.1% Xe + 52.9% O
|
47.0% Rn + 53.0% O
|
π1(Ng-O)
|
6.2% Xe + 93.8% O
|
5.4% Rn + 94.6% O
|
π2(Ng-O)
|
6.5% Xe + 93.5% O
|
5.8% Rn + 94.2% O
|
It is found that in the XeO4 case the electrons assigned to Xe are 2.889e− in the 5p orbital set while the 5s orbital holds 1.416e−. By comparison the Rn atom in RnO4 holds more electrons in the 6s orbitals (1.534 e−) and fewer electrons in the 6p orbitals (2.691e−). The Natural Localised Molecular Orbitals (NLMO) indicate a strong covalent Ng-O σ bond in both tetroxides and a very faint O→Ng π donor-acceptor interaction.
Thus far, the difference in the formation of Ng-O bonds appears to be down to the different radial structure of the valence orbitals in Rn with respect to Xe.
To probe into the origin of the weaker Rn-O bonds, a computational experiment was performed: the formation enthalpies were calculated with and without the DKH2 relativistic Hamiltonian. The experimental14 standard enthalpy of formation of XeO4 is + 642.2 kJ mol− 1 which is extremely well matched by the CCSD(T)/CBS//CCSD(T)/DKH-TZVPP approach that yields + 642.6 kJ mol− 1 (Fig. 2). Radon tetroxide is less stable by 88.6 kJ mol− 1 (21.2 kcal mol− 1).
Since there is no reason to assume that the kinetics of the decomposition process of RnO4 is substantially different to XeO4, the former can safely be predicted to exist on account of these results with a degree of confidence.
The non-relativistically optimised structures have very different thermodynamic features from the relativistic models. Without relativity XeO4 would be more explosive by almost two-fold (521.3 kJ mol− 1) and RnO4 would be less endothermic by nearly 200 kJ mol− 1. This provides evidence that relativistic effects have a drastic influence on the valence orbitals in both Xe and Rn compounds and consequently play a major role in their high valent chemistry.
But this effect can best be exemplified by the mean radial expectation values < r > that provide a measure of the changes that relativity imposes on the valence shells at the Hartree-Fock level (Table S1).
The 5s orbital in Xe already suffers a considerable radial contraction (0.111 a.u.) with respect to its non-relativistic counterpart but Rn by far exhibits the biggest changes with considerable contraction on both the 6s and 6p orbitals, -0.368 and − 0.191 a.u. respectively. This major contraction of the 6s orbital in Rn accounts for the augmented NPA of RnO4 with relation to the XeO4 NPA.
The relativistic radial contraction17 of the s and p shells on heavy elements has been known for some time. In particular, the inert pair effect initially formulated by Sidgwick18 in the 1930s and later found19 to be a consequence of relativity has been a mainstay in the chemical behaviour of the heavy elements of groups 13–15. Generally, the electron pair in the 6s2 shell of the heavy elements becomes progressively less chemically accessible to engage in chemical bonds that would result in TlIII, PbIV and BiV compounds. Such compounds exhibit a more unfavourable standard enthalpy of formation than each of the lighter congeners InIII, SnIV and SbV. This feature is best exemplified when comparing the values of the standard reduction potentials E(SnII/SnIV)=-0.088 V vs E(PbII/PbIV) = + 1.69 V for heavy elements of group 14.
The inert pair effect is generally not considered for elements in groups 16–18 since the chemistry of polonium, astatine and radon is so poorly known. These results clearly show that the chemistry of RnVIII is also affected by the inert pair effect.
The paper by Slepkov et al15 provided a detailed account into the possible decomposition pathway of XeO4 predicting the existence of a di-haptic isomer Xe(η2-O2)2 as intermediate. It is therefore natural to address the next leading question which is how the energetics of this process compares in the case of RnO4.
While the authors discuss the frontier orbitals of Xe(η2-O2)2 they do not provide an in-depth discussion into the oxidation state of the di-haptic oxides. Furthermore, the calculated HOMO-LUMO gap is notoriously small (0.38 eV) making it a species with high chemical potential (reactive) and possible multi-configurational character. Our examination of the CCSD or even MP2 natural orbitals does indeed confirm that there are two competing configurations in the Xe(η2-O2)2 molecule. For a more orthodox and accurate treatment of the electronic structure in this molecule, a CASPT2(22,15) geometry optimisation was conducted for the ground states of both Xe(η2-O2)2 and Rn(η2-O2)2. Both the quintet and triplet states were explored but lead to dissociation of either both or one of the O2•• ligands, respectively, so only the singlet state potential energy surface presents a bound minimum.
The outcome of the calculation is that the structure can be described as a superoxide with spin coupling mediated by the Ng fragment (cf. Section 3 of the SI). This indicates that there is significant O2−…O2− through bond interaction despite the fragments being over 4 Å apart. Formally this amounts to oxidation state II, RnII(η2-O2)2 and XeII(η2-O2)2, in both species.
One possibility not covered in the Slepkov paper15 is the existence of a stepwise decomposition intermediate NgO2(η2-O2). Thus, the structure of these dioxides was optimised and their electronic structure analysed herein. Both structures are minima and their ground singlet states exhibit a single configurational description. Triplet states were explored with this geometry but this leads to a local minimum with a dissociated oxygen atom.
Either species is quite unique in their geometry as they show a distorted tetrahedral shape with no symmetry consistent with an AX3E type VSEPR stereochemistry;20 in addition, one bond of the dihaptic O2 fragment is weaker than the other (cf. Figure 3 and Section 4 of the SI). The dihaptic η2-O2 ligand may best be described as a peroxide ligand (O22−). As such the formal oxidation state of the Ng atom in NgO2(η2-O2) may appropriately be considered + VI.
The decomposition processes of the tetroxides will therefore involve a successive descent of oxidation number in the order VIII→VI→II→0.
In a general overview of the xenon and radon species studied so far (Fig. 4), it may be seen that the energetics of the Ng(η2-O2)2 and NgO2(η2-O2) species is surprisingly similar between the two elements. The biggest contrast remains still the disparate enthalpies of formation of the two noble gas tetroxides. The joining together of two oxygen atoms to form XeO2(η2-O2) is endergonic (+ 69.3 kJ mol− 1) which may be a hint as to why XeO4 is still isolable given a low temperature or solvent.
The value reached by Slepkov et al for the enthalpy of formation of Xe(η2-O2)2 was + 633 kJ mol− 1; this value is fairly close to the one reached in this work by a proper multi-reference method, + 600.3 kJ mol− 1. This is indicative that approximate density functionals of the GGA type exhibit some degree of tolerance for multi-reference character if the single determinant represents around 60% of the wavefunction.
In finalising it is important to stress that when performing an in silico prediction of the existence of any species the term ‘stability’ is not very helpful,21 particularly if the systems in question are endothermic with respect to decomposition. But it may be concluded that RnO4 is a more difficult system to synthesise and isolate. XeO4 is typically obtained from the acid dehydration of a metal perxenonate (XeO64−).16 Presumably, a perradonate RnO64− with high enough lattice energy might be isolated and likewise transformed using very low temperature matrices. The e symmetry vibrational modes, corresponding to a scissor like motion of the oxygen atoms in RnO4 display a lower energy (205 cm− 1) than in XeO4 (265 cm− 1) which is indication that thermal decomposition may occur more easily for RnO4.