<<

inorganics

Article Decomposition of d- and f-Shell Contributions to Bonding from the Quantum Theory of Atoms in Molecules: Application to Uranium and Uranyl Halides

Jonathan Tanti 1,2 ID , Meghan Lincoln 1 ID and Andy Kerridge 1,* ID 1 Department of Chemistry, Lancaster University, Lancaster LA1 4YB, UK; [email protected] (J.T.); [email protected] (M.L.) 2 School of Chemistry, The University of Manchester, Oxford Road, Manchester M13 9PL, UK * Correspondence: [email protected]; Tel.: +44-1-524-594-770

 Received: 7 August 2018; Accepted: 20 August 2018; Published: 30 August 2018 

Abstract: The electronic structures of a series of uranium hexahalide and uranyl tetrahalide complexes were simulated at the density functional theoretical (DFT) level. The resulting electronic structures were analyzed using a novel application of the Quantum Theory of Atoms in Molecules (QTAIM) by exploiting the high symmetry of the complexes to determine 5f- and 6d-shell contributions to bonding via symmetry arguments. This analysis revealed fluoride ligation to result in strong bonds with a significant covalent character while ligation by and bromide species resulted in more ionic interactions with little differentiation between the . Fluoride ligands were also found to be most capable of perturbing an existing electronic structure. 5f contributions to overlap-driven covalency were found to be larger than 6d contributions for all interactions in all complexes studied while degeneracy-driven covalent contributions showed significantly greater variation. σ-contributions to degeneracy-driven covalency were found to be consistently larger than those of individual π-components while the total π-contribution was, in some cases, larger. Strong correlations were found between overlap-driven contributions, U–O vibrational frequencies, and energetic stability, which indicates that overlap-driven covalency to bond stabilization in these complexes and that uranyl vibrational frequencies can be used to quantitatively probe equatorial bond covalency. For uranium hexahalides, degeneracy-driven covalency was found to anti-correlate with bond stability.

Keywords: uranium; uranyl; halide; covalency; QTAIM; CASSCF; DFT; electron density

1. Introduction The quantification of the covalent contribution to bonding in complexes of the f-elements is an area of great current research being explored via X-ray absorption [1–11], electron paramagnetic [12–15], nuclear magnetic resonance [16,17], emission [18,19], and photoelectron [20] spectroscopies as well as X-ray diffraction [21] and structural studies [22–30]. The latter have often been carried out in combination with theoretical studies and a wealth of purely theoretical data also exists [31–53]. However, covalency is a phenomenon that is open to significant interpretation. It can manifest due to (i) the near-degeneracy of energy levels and/or (ii) the spatial overlap of the electronic wave functions of interacting species and, of these, only the latter leads to the accumulation of electronic charge in the bonding region and, therefore, the potential for energetic stabilization of the bond. Differentiating between these two origins of covalent bond character is, in addition to being of fundamental interest, of significant practical importance. Current European [54] and US [55] strategies for the separation of

Inorganics 2018, 6, 88; doi:10.3390/inorganics6030088 www.mdpi.com/journal/inorganics Inorganics 2018, 6, 88 2 of 19 trivalent from the highly radioactive minor (Np, Am, and Cm) in spent nuclear fuel are based on the exploitation of energetic stabilization of complexes of the latter with the origin of this stability believed to be due to the greater chemical availability of the 5f valence shell. Recently, several groups have adopted Bader’s Quantum Theory of Atoms in Molecules (QTAIM) [56] in order to probe covalency from a computational perspective. The QTAIM has the capacity to differentiate between degeneracy-driven and overlap-driven contributions to bonding interactions [57]. It is an appealing method of analysis since its results are independent of the orbital basis used to express the electronic structure, which cannot be unambiguously defined. This characteristic means that the QTAIM is equally applicable to electronic structures obtained using any quantum chemical methodology and, since it analyzes the physically observable electron density, it can be applied to experimentally-determined data [21,58]. Complexes of the f-elements typically exhibit strong relativistic effects, substantial dynamical electron correlation, and weak crystal fields and these factors combine to produce electronic structures in which the 5f, 6d, and, to a lesser extent, the 7s valence shells exist at similar energies, which complicates the role that these shells play in covalent interactions. Recently, we have shown that the QTAIM can be exploited in systems of high symmetry to differentiate between the roles played by the 5f and 6d shells [59–61]. In this study, we further extend this approach to distinguish between σ-components and π-components of covalent bond character. This approach is applied to a series of 2− homo-halide and hetero-halide complexes of the general form [UO2X2Y2] and UX2Y2Z2 where X, Y, Z ∈ {F, Cl, Br}. These complexes have been selected due to their high symmetry (D2h or higher) when assuming a trans orientation of like ligands and because they include a number of synthetically 2− − realized species: [UCl6] and [UF6] have both been the subject of recent experimental studies of U–X bonding [5,21] while the experimentally determined electronic structure of Cs2UO2Cl4 has been 2− 2− previously analyzed using the QTAIM [58]. Of the remaining complexes, [UO2F4] and [UO2Br4] − − have been observed in the and condensed phases, respectively [62,63]. [UF6] and [UCl6] have − both been experimentally characterized [64–66] although [UBr6] remains unreported.

2. Results

2.1. Identification of Model Chemistry The performance of a number of xc-functionals was assessed in order to determine the most suitable model chemistry for the study. The functionals chosen for this exercise were the GGA functionals PBE [67] and BLYP [68,69] along with the hybrid-GGA functionals PBE0 [70], B3LYP [71,72], and BHLYP [73]. A variation of the chosen basis set was not considered since its large size was assumed to give results sufficiently close to the basis set limit. Additionally, it has recently been shown that the def2-QZVP basis set used for the is sufficient for modeling anions in the absence of diffuse 2− functions [74]. For these test calculations, two representative systems were considered: [UO2Cl4] and UF6, for which accurate experimental data exists. Tables1 and2 summarize selected structural and vibrational parameters of these systems. In particular, only frequencies of vibrational stretching modes are reported since these most clearly relate to bond strength. As expected, hybrid functionals tend to outperform pure GGA functionals. Of the hybrid functionals, BHLYP (incorporating 50% Hartree-Fock exchange) is outperformed by both B3LYP and PBE0 (incorporating 20% and 25% Hartree-Fock exchange, respectively) especially with respect to vibrational frequencies. B3LYP and PBE0 perform comparably well and are in good to excellent agreement with experimental parameters. B3LYP noticeably outperforms PBE0 2− in the simulation of U–O bond lengths and vibrational frequencies in [UO2Cl4] and was chosen as the xc-functional used in the remainder of this study. The accuracy of the B3LYP functional reported in this study is comparable to that previously reported for other simple uranyl coordination complexes [75] and for f-element hexachlorides (M = U, Ce) [61]. Inorganics 2018, 6, 88 3 of 19

2− Table 1. Calculated structural and vibrational parameters of [UO2Cl4] calculated with various exchange-correlation functionals. Values in parenthesis indicate deviations from experimental values. a ref [76], b ref. [77].

Parameter Exp. a,b BLYP PBE PBE0 B3LYP BHLYP 1.817 1.800 1.756 1.776 1.735 r (Å) 1.774 UO (+0.043) (+0.026) (−0.018) (+0.002) (−0.039) 2.765 2.721 2.715 2.749 2.749 r (Å) 2.671 UCl (+0.094) (+0.050) (+0.044) (+0.078) (+0.078) 776 803 896 855 953 −1 834 νUO (cm ) (−58) (−31) (+62) (+21) (+119) 862 887 975 939 1023 922 (−60) (−35) (+53) (+17) (+101) 179 192 199 188 196 206 (−27) (−14) (−7) (−18) (−10) ν (cm−1) UCl 197 238 216 206 213 236 (39) (+2) (−20) (−30) (−23) 213 226 234 224 232 267 (−54) (−41) (−33) (−43) (−25)

Table 2. Calculated structural and vibrational parameters of UF6 calculated with various exchange-correlation functionals. Values in parenthesis indicate deviations from experimental values. a ref [64], b ref. [65].

Parameter Exp. a,b BLYP PBE PBE0 B3LYP BHLYP

rUF (Å) 1.999 2.036 2.018 1.990 2.009 1.984 (+0.037) (+0.019) (−0.009) (+0.010) (−0.055) 511 521 543 532 548 534 (−23) (−13) (+9) (−2) (+14) −1 νUF (cm ) 581 596 633 615 648 626 (−45) (−30) (+7) (−11) (+22) 613 630 682 660 710 667 (−54) (−37) (+15) (+7) (+43)

2.2. Homohalide Complexes of Uranium and Uranyl

2− In this section, the simulation and analysis of complexes of the form [UO2X4] and UX6 (X = F, Cl, Br) is considered. Focusing initially on the uranyl complexes, Table3 summarizes structural parameters obtained using the B3LYP xc-functional. A clear trend of increasing equatorial bond length is found as the halide mass increases, which is expected with the increasing size of the ligand. There is also, however, a commensurate reduction in the axial U–O bond length, which suggests a weakening of the equatorial bonds. Although equatorial bond lengths are found to be greater than the sum of the ionic radii, the difference is least pronounced for X = F, which again increases as the halide mass increases. This suggests some degree of covalent bond stabilization for the lighter halides. 2− 2− As with [UO2Cl4] (Table1), the U–O bond length in [UO 2Br4] is in excellent agreement with the experimentally reported value of 1.766 Å [63]. However, it should be noted that the experimental data are derived from XRD studies of condensed-phase structures. No experimentally determined 2− structural data is available for [UO2F4] , which has only been observed in the gas phase [62]. However, the U–O bond length reported here lies between the computationally-derived gas-phase literature values of 1.808 and 1.851 Å obtained using PBE and CCSD(T), respectively, [62] and compares well to that previously obtained using the B3LYP functional [37,78]. Inorganics 2018, 6, 88 4 of 19

2− Table 3. B3LYP-calculated structural parameters [UO2X4] . R = 6-coordinated Shannon ionic radius.

Complex rUO (Å) rUX (Å) R = RU + RX (Å) rUX – R (Å) 2− [UO2F4] 1.824 2.229 0.73 + 1.33 = 2.06 0.169 2− [UO2Cl4] 1.776 2.749 0.73 + 1.81 = 2.54 0.209 2− [UO2Br4] 1.769 2.922 0.73 + 1.96 = 2.69 0.232 2+ UO2 1.695 - - -

The U–Cl bond length is overestimated by 0.078 Å but is identical to that reported previously using the same functional [37] and the U–Br bond length is overestimated by 0.108 Å. The U–F bond length of 2.229 Å is very similar to that reported from PBE (2.233 Å), B3LYP (2.231 Å), and CCSD(T) (2.232 Å) simulations [37,62]. The excellent agreement with the CCSD(T) data, in particular, indicates that the origin of the discrepancies between experimental and calculated values for the chloride and bromide complexes is due to crystal packing effects. The calculated structural data for UX6 (Table4) presents a less clear picture. While the U–X bond lengths again increase as the halides mass increases, a comparison with the sum of ionic radii gives no trend. In contrast to the uranyl halides, bond lengths in these species are shorter than the sum of the ionic radii due to the lack of competition with strongly coordinating terminal oxo ligands. However, this shortening is identical in the fluoride and bromide complexes and less than that found in UCl6. The good agreement with the experimental gas-phase U–F bond length (Table2) is apparently not replicated in UCl6 where the U–Cl bond is overestimated by 0.050 Å in comparison to the value of 2.42 Å reported experimentally [66]. However, this value is again derived from an XRD study of 2− a condensed-phase structure and so the discrepancy is comparable to those found for [UO2Cl4] 2− and [UO2Br4] . The calculated U–Cl bond length is in excellent agreement with that calculated by Batista et al. (2.472 Å) using the same xc-functional [79]. There are no experimentally reported simple bromides of uranium [80], but the value reported in this study is in good agreement with and slightly shorter than those reported from GGA-based simulations (2.650–2.687 Å) [81], commensurate with the data presented in Tables1 and2. The good agreement with available literature data, therefore, leads us to tentatively suggest that an interplay of electronic and steric effects serve to mask any simple trend in UX6 bond lengths when compared to ionic radii.

Table 4. B3LYP-calculated structural parameters UX6. R = 6-coordinated Shannon ionic radius.

Complex rUX (Å) R = RU + RX (Å) rUX – R (Å)

UF6 2.009 0.73 + 1.33 = 2.06 −0.051 UCl6 2.470 0.73 + 1.81 = 2.54 −0.070 UBr6 2.639 0.73 + 1.96 = 2.69 −0.051

Bonding analysis of these complexes was performed using the Quantum Theory of Atoms In Molecules (QTAIM) [56]. We have recently discussed the merits of this methodology in analyzing the electronic structure of f-element complexes [57]. In this study, advantage is taken of the high symmetry of these complexes in order to perform a novel decomposition of a bonding contribution to 5f and 6d uranium valence shells. Two QTAIM metrics are commonly used to characterize bonding in molecular systems. The first of these is ρBCP(A, B), which is the magnitude of the electronic density at the bond critical point (BCP) between atoms A and B where the BCP is the saddle point in the electronic density distribution ρ(r) being a local maximum on the interatomic surface and a local minimum perpendicular to the surface. A commonly employed rule of thumb is that covalent interactions are characterized by ρBCP > 0.2 a.u. and closed shell interactions by ρBCP < 0.1 a.u. with intermediate values indicating an increasing degree of a covalent charter. ρBCP can, therefore, be interpreted as a measure of covalent interaction manifested by orbital overlap. Inorganics 2018, 6, 88 5 of 19

The second QTAIM metric used to characterize bonding is the delocalization index, δ(A, B). TheInorganics delocalization 2018, 6, x FOR index PEER isREVIEW an integrated property and, for chemical systems described by a single5 of 18 electronic configuration, can be defined by the equation below:

(A, B) =2 (A) (B) δ(A, B) = 2 ∑ Sij(A)Sij(B) (1)(1) i,,j

where (X) is the overlap between molecular orbitals (MOs) () and () integrated over the where Sij(X) is the overlap between molecular orbitals (MOs) ϕi(r) and ϕj(r) integrated over the atomic basin associated with atom X (as defined by QTAIM). (A, B) is a measure of the number of atomic basin associated with atom X (as defined by QTAIM). δ(A, B) is a measure of the number of electrons shared between two atoms and, in the absence of bond polarisation, correlates strongly with electrons shared between two atoms and, in the absence of bond polarisation, correlates strongly with a formal bond order. More generally, (A, B) can be interpreted as indicating the degree of covalent a formal bond order. More generally, δ(A, B) can be interpreted as indicating the degree of covalent interaction manifested by energetic degeneracy since it can be large in the absence of significant interaction manifested by energetic degeneracy since it can be large in the absence of significant overlap between species. While, typically, large values of (A, B) are accompanied by large overlap between species. While, typically, large values of ρBCP(A, B) are accompanied by large values values of (A, B), the converse is not universally true. of δ(A, B), the converse is not universally true. The complexes considered in this study all possess D2h (or higher) symmetry. Since the systems The complexes considered in this study all possess D2h (or higher) symmetry. Since the systems are closed shell in nature, the electron density possesses the same symmetry as the complex it are closed shell in nature, the electron density possesses the same symmetry as the complex it describes describes and, therefore, the atomic basin associated with the central uranium atom also possesses and, therefore, the atomic basin associated with the central uranium atom also possesses this symmetry this symmetry (see Figure 1). The delocalization index between the uranium center and any other (see Figure1). The delocalization index between the uranium center and any other atom in the complex atom in the complex can then be decomposed by the equation below. can then be decomposed by the equation below. Γ (U, B) =2(U)(B) (2) δ(U, B) = 2 S (U)S (B) (2) ∑ ∑, ij ij Γ i,j where Γ indexes the irreducible representations (irreps) of the point group and the summation over whereMOs isΓ nowindexes limited the irreducibleto those orbitals representations spanning the (irreps) irrep of Γ the. This point definition group andis valid the summationsince the overlap over MOs is now limited to those orbitals spanning the irrep Γ. This definition is valid since the overlap integral (U) is only non-zero when the MOs () and () span the same irreducible integralrepresentation.Sij(U) is onlyWe non-zerohave used when a similar, the MOs thoughϕi(r) and lessϕj (rdeveloped,) span the same approach irreducible in considering representation. the Wecontributions have used ato similar, bonding though in a series less developed, of approach [60]. in considering the contributions to bonding in a series of actinocenes [60].

Figure 1. QTAIM derived an atomic basin of U in UCl6, which exhibits the full symmetry of the Figure 1. QTAIM derived an atomic basin of U in UCl6, which exhibits the full symmetry of the molecule. Figures reproduced from Reference [61]. molecule. Figures reproduced from Reference [61].

Table 5. Irreducible representations spanned by components of the 5f and 6d valence shells with Table 5. Irreducible representations spanned by components of the 5f and 6d valence shells with respect to the principal molecular axis. respect to the principal molecular axis. Irrep ( Γ) Ag B1g B2g B3g Au B1u B2u B3u ComponentIrrep ( Γ)A 6dgσ, 6dδ B1g6dδ B2g 6dπ B3g 6dAπ u 5fδ B1u 5fσ, 5fδ B2u 5fπ, 5fφ B5f3uπ, 5fφ

Component 6dσ, 6dδ 6dδ 6dπ 6dπ 5fδ 5fσ, 5fδ 5fπ, 5fϕ 5fπ, 5fϕ In D2h symmetry, the components of the 5f and 6d valence shells that can engage in σ-bonding, π-bonding, δ-bonding, and φ-bonding interactions with ligands lying on the principal axis are summarized in Table 5. In the systems under consideration here, δ-bonding and φ-bonding interactions are not possible (or are, at least, energetically highly unfavorable) and σ-bonding and π- bonding contributions from the 5f and 6d shell can be decomposed from the total delocalization

index. Similarly, can be decomposed into contributions from the 5f and 6d shells even though further decomposition is not possible since only the σ-bonding contribution is non-zero at the BCP

Inorganics 2018, 6, 88 6 of 19

In D2h symmetry, the components of the 5f and 6d valence shells that can engage in σ-bonding, π-bonding, δ-bonding, and ϕ-bonding interactions with ligands lying on the principal axis are summarized in Table5. In the systems under consideration here, δ-bonding and ϕ-bonding interactions are not possible (or are, at least, energetically highly unfavorable) and σ-bonding and π-bonding contributions from the 5f and 6d shell can be decomposed from the total delocalization index. InorganicsSimilarly, 2018ρBCP, 6, xcan FOR be PEER decomposed REVIEW into contributions from the 5f and 6d shells even though further6 of 18 decomposition is not possible since only the σ-bonding contribution is non-zero at the BCP (all other (allcontributions other contributions to ρ(r) contain to ( a)nodal contain plane a nodal passing plane through passing the through BCP). Lastly, the BCP). it should Lastly, be itnoted should that, be notedeven though that, even this decompositionthough this decomposition is only valid withis on respectly valid to with bonding respect contributions to bonding aligned contributions with the principalaligned with axis the of theprincipal molecule, axis inof allthe complexes molecule, consideredin all complexes here, theconsidered principal here, axis the can principal be chosen axis to canalign be with chosen any to U–O align or U–Xwith bondany U–O while or maintaining U–X bond whileD2h symmetry maintaining and, D therefore,2h symmetry the and, decomposition therefore, thecan decomposition be applied to all can bonding be applied interactions. to all bonding interactions. 2− Figure 22 summarizessummarizes the relevant relevant QTAIM QTAIM metrics metrics for for the the U–O U–O bond bond in in [UO [UO2X42]X2− 4complexes] complexes (see (seeSupplementary Supplementary Materials Materials for numerical for numerical data). data). The magnitude The magnitude of of deviatesρBCP deviates most moststrongly strongly from thatfrom of that free of uranyl free uranyl in the in thecase case of the of thefluoride fluoride complex, complex, which which is isindicative indicative of of a a weakening weakening of of the covalent character of the U–O interaction. As As the the halide halide mass mass increases, increases, so so does does ρBCP, which implies that the heavier halides perturb the U–O bond to a lesser extent. This is commensurate with the data presented in Table3 3 where where covalent covalent stabilizationstabilization ofof thethe equatorialequatorial bondsbonds appearsappears mostmost pronouncedpronounced for the fluoride fluoride complex. complex. The The values values of of ρBCP reported reported inin thisthis study compare well to those previously reported for the fluoridefluoride and chloride complexescomplexes asas wellwell asas forfor thethe freefree uranyluranyl ionion [[37].37].

2− Figure 2. Decomposed QTAIM metrics of the U-O bond in [UO2X4] complexes (X = F, Cl, Br) Figure 2. Decomposed QTAIM metrics of the U-O bond in [UO2X4]2− complexes (X = F, Cl, Br) obtained obtained from B3LYP-derived densities. Values from free uranyl are also provided for comparison. from B3LYP-derived densities. Values from free uranyl are also provided for comparison. f-shell f-shell contributions are given in yellow and d-shell contributions are given in blue. contributions are given in yellow and d-shell contributions are given in blue.

The 5f contribution to ρ accounts for ~60% of the total value irrespective of the species under The 5f contribution to BCP accounts for ~60% of the total value irrespective of the species under consideration, which indicates that the 5f-shell is more involved in covalent interactions with the oxo ligand than the 6d-shell. The percentage variations in the 5f and 6d contributions to ρ are extremely ligand than the 6d-shell. The percentage variations in the 5f and 6d contributionsBCP to are extremelysimilar with similar values with in the values fluoride in the complex fluoride being complex 71% andbeing 73% 71% of theand free73% uranyl of the values,free uranyl respectively. values, respectively.The corresponding δ(U, O) data demonstrate broadly the same trend, but provide further insight. The totalThe corresponding number of electrons (U,O) shared data isdemonstrate substantially broadly lower the in thesame halide trend, complexes but provide than further in free insight.uranyl inThe agreement total number with of previous electrons work shared [37], is but substantially again increases lower as in the the halide halide mass complexes increases than and in freethe perturbationuranyl in agreement of the U–O with bond previous is reduced. work δ[37](U,, Obut) is again 78% ofincreases the free as uranyl the halide value mass in the increases fluoride andcomplex the perturbation but the 5f contribution of the U–O accounts bond foris reduced. 65% of the (U,O) total, which is 78% isgreater of the thanfree theuranyl 63% value contribution in the fluoridein free uranyl complex and but indicates the 5f contribution a stability of accounts this contribution for 65% of to the the total, U–O which bond is with greater respect than to the that 63% of contributionfree uranyl. Thein free ratio uranyl of 5f/6d and contributionsindicates a stability is reasonably of this contribution constant across to the the U–O halides. bond Forwith example, respect tothe that 5f component of free uranyl.is 64% The of ratio the totalof 5f/6d value contribu in the bromidetions is reasonably complex. Itconstant should beacross noted the that halides. there For are example,two components the 5f component of the 5f shell is 64% (and of likewisethe total forvalue the in 6d the shell) bromide that can complex. engage It in shouldπ-interactions be noted withthat there are two components of the 5f shell (and likewise for the 6d shell) that can engage in π- interactions with the oxo ligand and the contributions reported in this paper are the composite of these two distinct interactions. Therefore, the degree of electron sharing is greater for the σ- component of each bond than for each of the π-components. While there is an overall reduction in (U, O) upon completion, this reduction is not distributed evenly among the four bonding components. Taking the fluoride complex as an example, the 5fσ contribution is largely unchanged and is reduced to 95% of its free value while the 5fπ contribution reduces to 72% of its free value. Corresponding 6d contributions reduce to 80% and 69%, respectively. This variation can be understood in terms of the availability of these shells to equatorial bonding

Inorganics 2018, 6, 88 7 of 19 the oxo ligand and the contributions reported in this paper are the composite of these two distinct interactions. Therefore, the degree of electron sharing is greater for the σ-component of each bond than for each of the π-components. While there is an overall reduction in δ(U, O) upon completion, this reduction is not distributed evenly among the four bonding components. Taking the fluoride complex as an example, the 5fσ contribution is largely unchanged and is reduced to 95% of its free value while the 5fπ contribution reduces to 72% of its free value. Corresponding 6d contributions reduce to 80% and 69%, respectively. This variation can be understood in terms of the availability of these shells to equatorial bonding interactions: the 5fσ and 6dσ components are orientated along the U-O bond and are, therefore, relatively unavailable for equatorial interactions while the 5fπ and 6dπ components are available for equatorial σ-bonding and π-bonding interactions, respectively, and are, therefore, most significantly affected by the presence of equatorial ligands. Qualitatively similar behavior is found for the other halide complexes. More broadly, the quantitative difference in the reduction of ρBCP (which is only sensitive to σ-interactions) and the σ-contributions to δ(U, O) demonstrate how these metrics provide independent but related data regarding bonding. This is most obvious for the 5f-shell: the 5f contribution to ρBCP reduces to 71% of its free value upon equatorial coordination by fluoride while the corresponding 5fσ contribution to δ(U, O) maintains 95% of its free value, i.e., overlap-induced covalency is reduced while degeneracy-induced covalency is maintained. 2− Figure3 summarizes QTAIM metrics for equatorial U-X bonds in [UO 2X4] and UX6 complexes (see Supplementary Materials for numerical data). This data was obtained by taking a U–X bond to define the principal molecular axis and reference to σ-contributions and π-contributions to bonding is given in this study with respect to the U–X bond. This approach is taken throughout this contribution. Considering first ρBCP(U, X), overall magnitudes are larger in the UX6 complexes due to the lack of competition with the strongly coordinating terminal oxo species and the consequently shorter U–X bonds. The ρBCP(U, X) values are, however, much lower than those of the U–O bond, which indicates, as expected, significantly reduced covalent contributions to bonding. Nonetheless, ρBCP(U, X) has a maximum value (0.155 a.u.) in UF6, consistent with substantial covalent bond character. There is a clear reduction in ρBCP(U, X) in both sets of complexes as the halide mass increases, which shows a reduction in overlap-induced equatorial covalency. Interestingly, the reduction is significantly more pronounced for the 5f shell. The 5f contribution to ρBCP is 34% and 38% larger than the 2− 2− 6d contribution in [UO2F4] and UF6, respectively, but is just 4% and 7% larger in [UO2Br4] and UBr6, respectively. There is a larger contribution to bonding from the 5f shell in fluoride complexes while 5f and 6d contributions are comparable for the heavier halides. − The value of ρBCP in UF6 can be indirectly compared to a recent X-ray diffraction study of [UF6] in which topological analysis of the experimentally determined electron density gave ρBCP ~0.135 a.u. for the U–F bond [21]. This is approximately 13% lower than the value reported here, but it should be − borne in mind that uranium is in the +5 in [UF6] as opposed to the +6 oxidation state in UF6. We have previously investigated the dependence of QTAIM parameters on uranium oxidation state in uranium hexachloride [61], finding a reduction in ρBCP of 17% when comparing the +6 and +5 oxidation states. These recent experimental findings, therefore, appear to be in accord with the present study, which suggests a significant decrease in the overlap-driven covalent bond character upon the reduction of the uranium center in both hexachloride and hexafluoride species. Inorganics 2018, 6, 88 8 of 19 Inorganics 2018, 6, x FOR PEER REVIEW 8 of 18

2− Figure 3. Decomposed QTAIM metrics of the U-X bond in [UO2X42−] and UX6 complexes (X = F, Figure 3. Decomposed QTAIM metrics of the U-X bond in [UO2X4] and UX6 complexes (X = F, Cl, Cl, Br) obtained from B3LYP-derived densities. f-shell contributions are given in yellow and d-shell Br) obtained from B3LYP-derived densities. f-shell contributions are given in yellow and d-shell contributions are in blue. contributions are in blue.

Variation in δ(U, X) is markedly different in the two sets of complexes. For UX species, an overall Variation in (U,X) is markedly different in the two sets of complexes. For6 UX6 species, an increase is found as the halide mass increases, which opposes the trend in ρ and shows the danger overall increase is found as the halide mass increases, which opposes the BCPtrend in and shows ofthe attempting danger of toattempting use δ as an to indicator use as of an bond indicator order inof thebond presence order in of the bond presence polarization. of bond 5f polarization. components ( ) of5f δcomponentsU, X are consistently of (U,X) are higher consistently than 6d contributionshigher than 6d with contributions the 5fσ contribution with the dominating5fσ contribution the fluoridedominating complex the fluoride and all complex components and increasingall components as the increasing halide mass as the increases. halide Thismass behaviorincreases. again This opposes that exhibited by ρ and demonstrates that overlap-induced and degeneracy-induced behavior again opposes thatBCP exhibited by and demonstrates that overlap-induced and covalencydegeneracy-induced can (i) manifest covalency independently can (i) manifest and (ii) independently both be identified and by (ii) the both methodology be identified employed by the 2− here.methodology For [UO employed2X4] species, here. For the [UO trend2X4]2− in species, 5fσ, 5f theπ, andtrend 6d inπ 5fcomponentsσ, 5fπ, and 6dπ mirrors components that ofmirrorsρBCP while the variation in 6dσ strongly opposes this trend and results in the overall magnitude of that of while the variation in 6dσ strongly opposes this trend and results in the overall ( ) δmagnitudeU, X remaining of (U,X) largely remaining unchanged largely with unchanged the 6dσ withcomponent the 6dσ overtakingcomponent 5fovertakingσ as the dominant5fσ as the contributiondominant contribution when the when halide the mass halide increases. mass increases. The qualitative The qualitative difference difference in δ( inU, (U,X) X) between between the twothe setstwo ofsets complexes of complexes presumably presumably reflects reflects the strong the influencestrong influence of the axial of the oxo axial species oxo on species (formally) on uranium-based(formally) uranium-based energy levels. energy levels. 2.3. Heterohalide Complexes 2.3. Heterohalide Complexes 2− In this section, complexes of the form [UO2X2Y2] and UX2Y2Z2 (where X, Y, Z ∈ {F, Cl, Br}) are In this section, complexes of the form [UO2X2Y2]2− and UX2Y2Z2 (where X, Y, Z ∈ {F, Cl, Br}) are considered. Again, these complexes possess D2h symmetry or higher, which implies that like ligands considered. Again, these complexes possess D2h symmetry or higher, which implies that like ligands are oriented trans to each other and allows for the equivalent analysis to that reported in Section 2.2 are oriented trans to each other and allows for the equivalent analysis to that reported in Section 2.2 to be performed. In the case of the UX2Y2Z2 complexes, only the “axial” U–X bond is considered to be performed. In the case of the UX2Y2Z2 complexes, only the “axial” U–X bond is considered since the permutation of the ligands means that each of the bonding interactions in the 10 possible complexes is thus included.

Inorganics 2018, 6, 88 9 of 19

sinceInorganics the 2018 permutation, 6, x FOR PEER of REVIEW the ligands means that each of the bonding interactions in the 10 possible9 of 18 complexes is thus included.

Figure4 4 summarizes summarizes calculated calculated ρBCPvalues values of of the the axial axial bond bond (numerical (numerical values values are are presented presented in thein the Supplementary Supplementary Materials). Materials). Focusing Focusing initially initia on thelly uranylon the halides, uranyl the halides, characteristics the characteristics identified in Sectionidentified 2.2 in are Section maintained, 2.2 are i.e., maintained, that equatorial i.e., that ligand equatorial sets including ligand thesets lighter, including more the electronegative lighter, more halideselectronegative perturb thehalides axial U–Operturb bond the more axial with U–O the bond 5f and more 6d contributions with the 5f beingand 6d approximately contributions equally being affected.approximately The effect equally of alteringaffected. the The equatorial effect of altering ligand the set equatorial in the uranium ligand halides set in the is lessuranium pronounced, halides presumablyis less pronounced, due to the presumably more ionic due nature to ofthe the more interactions. ionic natureρBCP ofattains the interactions. a maximum value 0.160attains a.u. a formaximum the U–F value bond 0.160 with a.u. an equatorialfor the U–F ligand bond setwith comprised an equatorial of Cl ligand and/or set Br. comprised This is, however, of Cl and/or just 3%Br. largerThis is, than however, the value just found 3% larger in UF 6than. Broadly the value speaking, found there in UF is a6. modestBroadly perturbation speaking, there to ρBCP is awhen modest the equatorialperturbation ligand to set includes when the F while equatorial calculated ligand metrics set includes are almost F while completely calculated insensitive metrics to are replacing almost Clcompletely by Br. insensitive to replacing Cl by Br.

2−2− Figure 4.4. ρBCP values ofof axialaxial U–OU–O andand U–XU–X (X (X = = F, F, Cl, Cl, Br) Br) bonds bonds in in [UO [UO2X2XY22Y] 2] andand UXUX2Y2Y22ZZ22 complexes derived from B3LYP-generated densities.densities. AllAll valuesvalues areare inin a.u.a.u.

Delocalization indices exhibit more variation (See(See FigureFigure5 :5: numericalnumerical valuesvalues presentedpresented inin (U,O) Supplementary Materials). δ(U, O) exhibits exhibits thethe samesame variationvariation as reportedreported in Section 2.2 with larger values found when the halide mass increases. The 5fσ contribution is, in all cases, close to the value values found when the halide mass increases. The 5fσ contribution is, in all cases, close to the value for freefor free uranyl uranyl and and insensitive insensitive to the to equatorialthe equatorial ligand ligand set. Otherset. Other contributions contributions vary vary in accordance in accordance with with the total value, with the 6dπ component exhibiting the greatest sensitivity. As previously the total value, with the 6dπ component exhibiting the greatest sensitivity. As previously discussed, thisdiscussed, component this component of the 6d shellof the is 6d also shell able is toalso engage able to in engage equatorial in equatorialπ-bonding π-bonding interactions interactions and the variationand the variation seen here seen is indicative here is indicative of the formation of the formation of such bonding of such interactions bonding interactions even though even this though should bethis interpreted should be ininterpreted the context in ofthe degeneracy-driven context of degeneracy-driven covalency. covalency. All UX2Y2Z2 complexes exhibit qualitatively similar trends. Total values of (U,X) are largely All UX2Y2Z2 complexes exhibit qualitatively similar trends. Total values of δ(U, X) are largely insensitive to the equatorial environment, which increases slightly as the equatorialequatorial ligands become heavier and less electronegative. A trend is also exhibited by the 5fπ and 6dσ contributions. The 5fσ heavier and less electronegative. A trend is also exhibited by the 5fπ and 6dσ contributions. contribution decreases as the halide mass increases, but, in contrast, the 6dπ component increases by a greater magnitude. Since both the 5fσ and 6dπ components are available for equatorial π-bonding interactions, the variation in (U,X) can, therefore, be attributed to increased equatorial π- interactions with the 6d shell, which is partially balanced by a reduction in such interactions with the

Inorganics 2018, 6, 88 10 of 19

The 5fσ contribution decreases as the halide mass increases, but, in contrast, the 6dπ component increases by a greater magnitude. Since both the 5fσ and 6dπ components are available for equatorial

πInorganics-bonding 2018 interactions,, 6, x FOR PEER the REVIEW variation in δ(U, X) can, therefore, be attributed to increased equatorial10 of 18 π-interactions with the 6d shell, which is partially balanced by a reduction in such interactions with the5f shell. 5f shell. This This is indicative is indicative of ofcloser closer energetic energetic matching matching of of fluorine fluorine 2p 2p and and uranium uranium 5f 5f levels and ofof brominebromine 4p4p andand uraniumuranium 6d6d levels.levels.

2− Figure 5. Delocalization indices of axial U–O and U–X (X = F, Cl, Br) bonds in [UO2X2Y2] and Figure 5. Delocalization indices of axial U–O and U–X (X = F, Cl, Br) bonds in [UO2X2Y2]2− and UX2Y2Z2 UX Y Z complexes derived from B3LYP-generated densities. All values are in a.u. complexes2 2 2 derived from B3LYP-generated densities. All values are in a.u.

2.4.2.4. Covalency andand BondBond StabilizationStabilization Lastly,Lastly, the relationship relationship between between equatorial equatorial bond bond covalency covalency and and bond bond stability stability is considered. is considered. For Foruranyl, uranyl, the thebinding binding energy energy of the of theequatorial equatorial ligand ligand set is set defined is defined by the by theequation equation below. below.   =UO2+ +2(X−)+2(Y −)− (UOXY 2−) (3) EB = E UO2 + 2E X + 2E Y − E([UO2X2Y2] ) (3)

so that is positive for a stabilizing interaction.

Inorganics 2018, 6, 88 11 of 19

Inorganics 2018, 6, x FOR PEER REVIEW 11 of 18 so that EB is positive for a stabilizing interaction.

Figure 6. CorrelationCorrelation between between QTAIM QTAIM parameters of the U–O bond, U–O stretching vibrational 2− 2− frequencies, and and the the binding binding energy energy of of the the equatorial equatorial ligand ligand set set in in [UO [UO2X22XY2]Y2 ]complexes.complexes. (a) (correlatesa) correlates νUO andand ρBCP, (,(b)b )correlates correlates νUO and δ(U,O)(U, O),,( (cc)) correlatescorrelatesE Band andρBCP ,(d, )(d correlates) correlatesEB and andδ(U, (U,O) O). .

Figure 66 reportsreports correlationscorrelations betweenbetween QTAIMQTAIM parametersparameters ofof thethe UOUO bond,bond, U–OU–O stretchingstretching vibrational frequencies ( ), and the binding energy ( ) of the equatorial ligand set (see vibrational frequencies (νUO), and the binding energy (EB) of the equatorial ligand set (seeSupplementary Supplementary Materials Materials for fornumerical numerical data). data). Ex Extremelytremely strong strong correlations correlations are are found found between vibrational frequencies and both and (U,O) (Figure 6a and 6b, respectively), which indicates vibrational frequencies and both ρBCP and δ(U, O) (Figures6a and6b, respectively), which indicates that thethe covalencycovalency of of the the U–O U–O bond bond is stronglyis strongly stabilizing. stabilizing. We haveWe have previously previously reported reported similar similar strong strongcorrelations correlations with a morewith varieda more ligand varied set ligand [75]. Perhapsset [75]. more Perhaps interestingly, more interestingly, strong anti-correlations strong anti- correlationsare also found are betweenalso found these between topological these parameterstopological parameters and binding and energies, binding which energies, is defined which inis definedEquation in (3). Equation These anti-correlations (3). These anti-correlations (Figure6c,d) illustrate(Figure 6c,d) the perturbation illustrate the of perturbation the U-O bond of covalency the U-O bondby the covalency presence by of thethe equatorialpresence of ligands, the equatorial which suggestsligands, which that U-O suggests bond vibrationsthat U-O bond can bevibrations used as cana quantitative be used as probea quantitative of equatorial probe ligand of equatorial stability. ligand stability. It might be expected that binding energies woul wouldd also correlate with QTAIM parameters of the axial bonds.bonds. FigureFigure7 summarizes7 summarizes these these relationships relationships (see (see Supplementary Supplementary Materials Materials for numerical for numerical data). data).Note thatNote since that both since homoleptic both homoleptic and heteroleptic and heteroleptic complexes complexes are considered, are considered, averaged topological averaged valuestopological are reported. values are We havereported. taken We this approachhave take previouslyn this approach and found previously strong correlations and found between strong correlationsaveraged values between and bindingaveraged energies values and [75]. binding energies [75]. Figure 7a illustrates an extremely strong correlation between ̅ and , which provides Figure7a illustrates an extremely strong correlation between ρBCP and EB, which provides further furtherjustification justification for considering for considering these averaged these averag values.ed Conversely, values. Conversely, no correlation no correlation is found between is foundδ ̅ betweenand E (Figure and7 b). This (Figure can be7b). rationalized This can be if rationalized one considers if one the aspectsconsiders of covalencythe aspects that of covalencyρ and B ̅ BCP that and probe [57]: identifies overlap-driven covalency, i.e., charge accumulation in δ probe [57]: ρBCP identifies overlap-driven covalency, i.e., charge accumulation in the bonding region, the bonding region, which might be expected to have a stabilizing effect on the interaction. , on the other hand, can be considered a measure of degeneracy-driven covalency for which no energetic stabilization of the bond is necessarily manifested. The absence of correlation presented in Figure 7b is, therefore, indicative of the fact that, in these complexes, degeneracy-driven covalency has no

Inorganics 2018, 6, 88 12 of 19 which might be expected to have a stabilizing effect on the interaction. δ, on the other hand, can be Inorganicsconsidered 2018, a 6 measure, x FOR PEER of REVIEW degeneracy-driven covalency for which no energetic stabilization of the12 bondof 18 is necessarily manifested. The absence of correlation presented in Figure7b is, therefore, indicative of ̅ clearly-definedthe fact that, in energetic these complexes, effect. It degeneracy-driven should also be noted covalency that variation has no clearly-defined in is rather energetic modest effect.with maximumIt should alsodeviations be noted less that than variation 2% from in theδ is mean rather value. modest In contrast, with maximum maximum deviations deviations less than are 2% ~40% from the mean. from the mean value. In contrast, maximum ρBCP deviations are ~40% from the mean.

FigureFigure 7. 7. CorrelationCorrelation between between QTAIM QTAIM parameters parameters of the of the equatorial equatorial bonds, bonds, the thebinding binding energy energy of the of 2− 2− equatorialthe equatorial ligand ligand set, set,and andthe theU–O U–O stre stretchingtching vibrational vibrational frequencies frequencies in [UO in [UO2X22YX2]2Y 2complexes.] complexes. (a) ̅ correlates(a) correlates EandB and ̅ρBCP, (b,() bcorrelates) correlates E Bandand δ,,( (cc)) correlates correlates νUO andand ρBCP̅,(, (dd)) correlatescorrelatesν UOand andδ . ̅.

Since νUO correlated strongly with EB and EB itself correlated strongly with ρBCP for the Since correlated strongly with and itself correlated strongly with ̅ for the equatorial bonds, the direct relationship between νUO and equatorial bond parameters was considered equatorial(see Figures bonds,7c and 7thed). Asdirect expected, relationship a very strongbetween (albeit slightly and weaker)equatorial anticorrelation bond parameters was found was consideredwith ρ while(see Figure correlation 7c and with 7d). δAswas expected, not identified. a very strong The anti-correlation (albeit slightly weaker) with ρ anticorrelationlends further BCP ̅ BCP was found with ̅ while correlation with was not identified. The anti-correlation with ̅ support to our previous assertion that νUO could be used to probe equatorial bond covalency [75]. lends further support to our previous assertion that could be used to probe equatorial bond The success of correlating the averaged equatorial ρBCP values with binding energies in uranyl covalencyhalides suggested [75]. that a similar approach might be applicable to the uranium hexahalides considered in thisThe study success where of correlating the average the would averaged be with equatorial respect to all coordinating values with species.binding Theenergies binding in uranyl energy halidesis defined suggested by the equationthat a similar below. approach might be applicable to the uranium hexahalides considered in this study where the average would be with respect to all coordinating species. The binding energy is defined by the equation below. 6+ − − − EB = E U + 2E X + 2E Y + 2E Z − E(UX2Y2Z2). (4) =(U ) +2(X ) +2(Y ) +2(Z ) − (UXYZ). (4) Figure8 presents the results of correlating EB with ρBCP and δ (See̅ Supplementary Materials for Figure 8 presents the results of correlating with ̅ and (See Supplementary Materials numerical data). As seen for the uranyl halides, ρBCP exhibits an excellent correlation with EB and for numerical data). As seen for the uranyl halides, ̅ exhibits an excellent correlation with again demonstrates that ρBCP quantitatively measures stabilizing covalent interactions in these species. and again demonstrates that ̅ quantitatively measures stabilizing covalent interactions in these ̅ species. Remarkably, shows a very strong anti-correlation with , i.e., as degeneracy-driven covalency increases, bond stability reduces.

Inorganics 2018, 6, 88 13 of 19

Remarkably, δ shows a very strong anti-correlation with EB, i.e., as degeneracy-driven covalency Inorganicsincreases, 2018 bond, 6, x FOR stability PEER reduces.REVIEW 13 of 18

Figure 8. Correlation between the averaged QTAIM parameters of U–X, U–Y & U–Z bonds and the Figure 8. Correlation between the averaged QTAIM parameters of U–X, U–Y & U–Z bonds and the total binding energy in UX2Y2Z2 complexes. (a) correlates EB and ρBCP,(b) correlates EB and δ. ̅ total binding energy in UX2Y2Z2 complexes. (a) correlates and ̅, (b) correlates and , 3. Computational Details 3. Computational Details AllAll calculations calculations were were performed performed using using version version 6.6 6.6of the of TURBOMOLE the TURBOMOLE quantum quantum chemistry chemistry code xc [82,83]code [82 using,83] using scalar-relativistic scalar-relativistic density density functional functional theory theory (DFT). (DFT). Several Several exchange-correlation exchange-correlation ( (xc-)-) functionalsfunctionals were were considered considered in in order order to to identify identify which which was was most most suitable suitable for for these these simulations simulations and, and, asas reported reported above, above, the the hybrid-GGA hybrid-GGA B3LYP B3LYP function functionalal [71,72] [71,72] was was found found to to give give the the best best agreement agreement withwith experimental experimental data data in in test test systems. systems. There There ar aree numerous numerous examples examples in in the the literature literature demonstrating demonstrating thethe suitability suitability of of B3LYP B3LYP in in the the study or U(VI) coordination complexes [34,37,78,75]. [34,37,75,78]. All All simulations simulations ζ werewere performed performed using using the the Ahlrichs Ahlrichs def2-QZVP def2-QZVP basis basis sets sets [84] [84] of of polarized quadruple- quadruple-ζ qualityquality for for lightlight atoms atoms (O, (O, F, F, Cl, Cl, Br). Br). The The small-core small-core pseudopotential pseudopotential of ofDolg Dolg and and co-workers[85], co-workers [ 85along], along with with the correspondingthe corresponding (14s13p10d8f6g)/[10s9p5d4f3g] (14s13p10d8f6g)/[10s9p5d4f3g] basis set basis [86], set was [86 ],used was to used model to modelthe uranium the uranium center bycenter incorporating by incorporating scalar scalarrelativistic relativistic effects. effects. This This pseudopotential pseudopotential is isconstructed constructed from from calculations calculations employingemploying the the Wood-Boring Wood-Boring equation equation (which (which may may be be derived derived directly directly from from the the Dirac Dirac equation) equation) and and includesincludes both both a a mass-velocity mass-velocity and and a a Darwin Darwin term. term. Th Thee effects effects of of the the spin-orbit spin-orbit coupling coupling have have not not been been 0 0 consideredconsidered in in these these closed closed shell, shell, formally formally 5f 5f06d6d0 species.species. De De Jong Jong et etal. al.have have shown shown that that the the effects effects of spin-orbitof spin-orbit coupling coupling are are rather rather moderate moderate in inthe the high high energy energy valence valence region region of of uranyl uranyl but but becomes becomes moremore pronounced pronounced when when considering considering the the pseudo pseudo core core 6p-shell 6p-shell [87]. [87]. However, However, we we have have recently recently shown shown thatthat the the impact impact of of the the 6p-shell 6p-shell on on QTAIM QTAIM metrics metrics is is modest modest [88]. [88]. We We would would also also expect expect the the degree degree of of SO-couplingSO-coupling to to be be comparable comparable in in analogous analogous systems systems and, and, therefore, therefore, have have little little impact impact on on the the trends trends reportedreported here. here. GeometryGeometry optimizations werewere performedperformed in in the the gas-phase gas-phase using using default default convergence convergence criteria criteria and andlocal local energetic energetic minima minima were identifiedwere identified using analytical using analytical frequency frequency analysis. Inanalysis. order toIn decompose order to decompose bonding into contributions from different electronic shells, all optimizations were constrained to the D2h point group. Topological and integrated properties of the electron density were evaluated using version 3.3.9 of the Multiwfn code [89].

Inorganics 2018, 6, 88 14 of 19 bonding into contributions from different electronic shells, all optimizations were constrained to the D2h point group. Topological and integrated properties of the electron density were evaluated using version 3.3.9 of the Multiwfn code [89].

4. Summary and Conclusions The electronic structures of a series of uranium hexahalide and uranyl tetrahalide complexes were simulated at the density functional theoretical (DFT) level. A comparison with previous experimental structural and vibrational data indicated that the B3LYP exchange-correlation functional was best able to accurately simulate these complexes. B3LYP-derived electronic structures were subsequently analyzed using a novel application of the Quantum Theory of Atoms in Molecules (QTAIM). This approach exploited the high (D2h or greater) symmetry of the complexes to determine 5f-shell and 6d-shell contributions to bonding through symmetry arguments. The analysis also allowed differentiation between overlap-driven and degeneracy-driven contributions to covalency and, for the latter, to further differentiate between σ-type and π-type interactions. This analysis revealed that, of the halogen ligands considered (X = F, Cl, Br), fluoride ligation resulted in the strongest bonds with the most significant covalent character and with values of the density at the U-F bond critical point, ρBCP, as high as 0.16 a.u. As a reference, ρBCP ≥ 0.20 a.u. is often taken to be an indicator of a predominantly covalent interaction. Ligation by chloride and bromide species resulted in more ionic interactions with little differentiation between the ligands. Fluoride ligands were also found to be most 2− capable of perturbing the electronic structure of other bonds in the complex. Taking [UO2F4] as an example, QTAIM metrics of the U–O bond were reduced to ~75% of the values for free uranyl. The QTAIM analysis allows for the distinction to be made between covalent bond contributions whose origin lies in either orbital overlap or energy degeneracy. Analysis of the complexes considered here showed that 5f contributions to overlap-driven covalency were larger than 6d contributions for all interactions in all complexes studied while degeneracy-driven covalent contributions showed significantly greater variation. σ-contributions to degeneracy-driven covalency were found to be consistently larger than those of each individual π-component even though the total π-contribution was, in some cases, larger. We have previously reported on strong correlations between uranyl stretching vibrational frequencies and equatorial bond covalency [75]. In this study, strong correlations were found between overlap-driven covalent bond contributions, U–O vibrational frequencies, and energetic stability, which indicates that overlap-driven covalency leads to bond stabilization in these complexes and that uranyl vibrational frequencies can be used to quantitatively probe equatorial bond covalency. For uranium hexahalides, overlap-driven covalency was shown to strongly correlate with bond stability while degeneracy-driven covalency was found to strongly anti-correlate. This anti-correlation illustrates the care that must be taken to differentiate between overlap-driven and degeneracy-driven covalent contributions to bond character when employing such analysis to rationalize experimental data. This work demonstrates that the QTAIM can be used to provide a highly detailed characterization of bonding interactions, which distinguishes between the different origins of covalent contributions and provides quantitative data with which to rationalize observable phenomena.

Supplementary Materials: The following are available online at http://www.mdpi.com/2304-6740/6/3/88/s1. Tables S1–S10: numerical data using which graphs and figures have been generated. Author Contributions: Conceptualization, A.K. Methodology, A.K. Formal Analysis, J.T., M.L., and A.K. Writing-Original Draft Preparation, J.T., M.L., and A.K. Writing-Review & Editing, J.T., M.L., and A.K. Visualization, A.K. Supervision, A.K. Project Administration, A.K. Funding: This research received no external funding. Conflicts of Interest: The authors declare no conflict of interest. Inorganics 2018, 6, 88 15 of 19

References

1. Jensen, M.P.; Bond, A.H. Comparison of covalency in the complexes of trivalent and cations. J. Am. Chem. Soc. 2002, 124, 9870–9877. [CrossRef][PubMed] 2. Gregson, M.; Lu, E.; Tuna, F.; McInnes, E.J.L.; Hennig, C.; Scheinost, A.C.; McMaster, J.; Lewis, W.; Blake, A.J.; Kerridge, A.; et al. Emergence of Comparable Covalency in Isostructural (IV)- and Uranium(IV)- Multiple Bonds. Chem. Sci. 2016, 7, 3286–3297. [CrossRef][PubMed] 3. Gregson, M.; Lu, E.; Mills, D.; Tuna, F.; McInnes, E.; Hennig, C.; Scheinost, A.; McMaster, J.; Lewis, W.; Blake, A.; et al. The inverse-trans-influence in tetravalent lanthanide and actinide bis(carbene) complexes. Nat. Commun. 2017, 8, 14137. [CrossRef][PubMed] 4. Kozimor, S.A.; Yang, P.; Batista, E.R.; Boland, K.S.; Burns, C.J.; Clark, D.L.; Conradson, S.D.; Martin, R.L.; Wilkerson, M.P.; Wolfsberg, L.E. Trends in covalency for d- and f-element metallocene dichlorides identified using K-edge X-ray absorption spectroscopy and time-dependent density functional theory. J. Am. Chem. Soc. 2009, 131, 12125–12136. [CrossRef][PubMed] 5. Minasian, S.G.; Keith, J.M.; Batista, E.R.; Boland, K.S.; Clark, D.L.; Conradson, S.D.; Kozimor, S.A.; Martin, R.L.; Schwarz, D.E.; Shuh, D.K.; et al. Determining relative f and d orbital contributions to M–Cl 2− − covalency in MCl6 (M = Ti, Zr, Hf, U) and UOCl5 using Cl K-edge X-ray absorption spectroscopy and time-dependent density functional theory. J. Am. Chem. Soc. 2012, 134, 5586–5597. [CrossRef][PubMed] 6. Spencer, L.P.; Yang, P.; Minasian, S.G.; Jilek, R.E.; Batista, E.R.; Boland, K.S.; Boncella, J.M.; Conradson, S.D.; 2+ Clark, D.L.; Hayton, T.W.; et al. Tetrahalide complexes of the [U(NR)2] : Synthesis, theory, and chlorine K-edge X-ray absorption spectroscopy. J. Am. Chem. Soc. 2013, 135, 2279–2290. [CrossRef][PubMed] 7. Minasian, S.G.; Keith, J.M.; Batista, E.R.; Boland, K.S.; Clark, D.L.; Kozimor, S.A.; Martin, R.L.; Shuh, D.K.; Tyliszczak, T. New evidence for 5f covalency in actinocenes determined from carbon K-edge XAS and electronic structure theory. Chem. Sci. 2014, 5, 351–359. [CrossRef] 8. Altman, A.B.; Pacold, J.I.; Wang, J.; Lukens, W.W.; Minasian, S.G. Evidence for 5d-σ and 5d-π covalency in lanthanide sesquioxides from K-edge X-ray absorption spectroscopy. Dalt. Trans. 2016, 45, 9948–9961. [CrossRef][PubMed] 9. Löble, M.W.; Keith, J.M.; Altman, A.B.; Stieber, S.C.E.; Batista, E.R.; Boland, K.S.; Conradson, S.D.; Clark, D.L.; Lezama Pacheco, J.; Kozimor, S.A.; et al. Covalency in Lanthanides. An X-ray Absorption Spectroscopy and x− Density Functional Theory Study of LnCl6 (x= 3, 2). J. Am. Chem. Soc. 2015, 137, 2506–2523. [CrossRef] [PubMed] 10. Minasian, S.G.; Batista, E.R.; Booth, C.H.; Clark, D.L.; Keith, J.M.; Kozimor, S.A.; Lukens, W.W.; Martin, R.L.;

Shuh, D.K.; Stieber, S.C.E.; et al. Quantitative Evidence for Lanthanide-Oxygen Orbital Mixing in CeO2, PrO2, and TbO2. J. Am. Chem. Soc. 2017, 139, 18052–18064. [CrossRef][PubMed] 11. Dumas, T.; Guillaumont, D.; Fillaux, C.; Scheinost, A.; Moisy, P.; Petit, S.; Shuh, D.K.; Tyliszczak, T.; Auwer, C. Den. The nature of chemical bonding in actinide and lanthanide ferrocyanides determined by X-ray absorption spectroscopy and density functional theory. Phys. Chem. Chem. Phys. 2016, 18, 2887–2895. [CrossRef][PubMed] 12. Formanuik, A.; Ariciu, A.M.; Ortu, F.; Beekmeyer, R.; Kerridge, A.; Tuna, F.; McInnes, E.J.L.; Mills, D.P. Actinide covalency measured by pulsed electron paramagnetic resonance spectroscopy. Nat. Chem. 2017, 9, 578–583. [CrossRef][PubMed] 13. Seaman, L.A.; Wu, G.; Edelstein, N.; Lukens, W.W.; Magnani, N.; Hayton, T.W. Probing the 5f orbital contribution to the bonding in a U(V) ketimide complex. J. Am. Chem. Soc. 2012, 134, 4931–4940. [CrossRef] [PubMed] 14. Lukens, W.W.; Edelstein, N.M.; Magnani, N.; Hayton, T.W.; Fortier, S.; Seaman, L.A. Quantifying the σ and π interactions between U(V) f orbitals and halide, alkyl, alkoxide, amide and ketimide ligands. J. Am. Chem. Soc. 2013, 135, 10742–10754. [CrossRef][PubMed] 15. Lukens, W.W.; Speldrich, M.; Yang, P.; Duignan, T.J.; Autschbach, J.; Kögerler, P. The roles of 4f- and 5f-orbitals in bonding: A magnetochemical, crystal field, density functional theory, and multi-reference wavefunctional study. Dalt. Trans. 2016, 11508–11521. [CrossRef][PubMed] 16. Smiles, D.E.; Wu, G.; Hrobárik, P.; Hayton, T.W. Use of 77Se and 125Te NMR Spectroscopy to Probe Covalency − of the Actinide- Bonding in [Th(En){N(SiMe3)2}3] (E = Se, Te; n = 1, 2) and Their Oxo-Uranium(VI) Congeners. J. Am. Chem. Soc. 2016, 138, 814–825. [CrossRef][PubMed] Inorganics 2018, 6, 88 16 of 19

17. Adam, C.; Kaden, P.; Beele, B.B.; Müllich, U.; Trumm, S.; Geist, A.; Panak, P.J.; Denecke, M.A. Evidence for covalence in a N-donor complex of (III). Dalton Trans. 2013, 42, 14068–14074. [CrossRef][PubMed] 18. Polinski, M.J.; Garner, E.B.; Maurice, R.; Planas, N.; Stritzinger, J.T.; Parker, T.G.; Cross, J.N.; Green, T.D.; Alekseev, E.V.; Van Cleve, S.M.; et al. Unusual structure, bonding and properties in a californium borate. Nat. Chem. 2014, 6, 387–392. [CrossRef][PubMed] 19. Cary, S.K.; Vasiliu, M.; Baumbach, R.E.; Stritzinger, J.T.; Green, T.D.; Diefenbach, K.; Cross, J.N.; Knappenberger, K.L.; Liu, G.; , M.A.; et al. Emergence of californium as the second transitional element in the actinide series. Nat. Commun. 2015, 6, 6827. [CrossRef][PubMed] 20. Su, J.; Dau, P.D.; Liu, H.T.; Huang, D.L.; Wei, F.; Schwarz, W.H.E.; Li, J.; Wang, L.S. Photoelectron spectroscopy q− and theoretical studies of gaseous uranium hexachlorides in different oxidation states: UCl6 (q = 0–2). J. Chem. Phys. 2015, 142, 134308. [CrossRef][PubMed] 21. Gianopoulos, C.G.; Zhurov, V.V.; Minasian, S.G.; Batista, E.R.; Jelsch, C.; Pinkerton, A.A. Bonding in Uranium(V) Hexafluoride Based on the Experimental Electron Density Distribution Measured at 20 K. Inorg. Chem. 2017, 56, 1775–1778. [CrossRef][PubMed] 22. Arnold, P.L.; Turner, Z.R.; Kaltsoyannis, N.; Pelekanaki, P.; Bellabarba, R.M.; Tooze, R.P. Covalency in CeIV and UIV halide and N-heterocyclic carbene bonds. Chem. Eur. J. 2010, 16, 9623–9629. [CrossRef][PubMed] 23. King, D.M.; Tuna, F.; McInnes, E.J.L.; McMaster, J.; Lewis, W.; Blake, A.J.; Liddle, S.T. Isolation and characterization of a uranium(VI)–nitride triple bond. Nat. Chem. 2013, 5, 482–488. [CrossRef][PubMed] 24. Brown, J.L.; Fortier, S.; Wu, G.; Kaltsoyannis, N.; Hayton, T.W. Synthesis and Spectroscopic and Computational Characterization of the Chalcogenido-Substituted Analogues of the Uranyl Ion, [OUE]2+ (E = S, Se). J. Am. Chem. Soc. 2013, 135, 5352–5355. [CrossRef][PubMed] 25. Arliguie, T.; Belkhiri, L.; Bouaoud, S.; Thuery, P.; Villiers, C.; Boucekkine, A.; Ephritikhine, M. Lanthanide(III)

and Actinide(III) Complexes [M(BH4)2(THF)5][BPh4] and [M(BH4)2(18-crown-6)][BPh4] (M = Nd, Ce, U): Synthesis, , and Density Functional Theory Investigation of the Covalent Contribution to Metal-Borohydride Bonding. Inorg. Chem. 2009, 48, 221–230. [CrossRef][PubMed] 26. Jones, M.B.; Gaunt, A.J.; Gordon, J.C.; Kaltsoyannis, N.; Neu, M.P.; Scott, B.L. Uncovering f-element bonding differences and electronic structure in a series of 1:3 and 1:4 complexes with a diselenophosphinate ligand. Chem. Sci. 2013, 4, 1189–1203. [CrossRef] 27. Gregson, M.; Wooles, A.J.; Cooper, O.J.; Liddle, S.T. Covalent Uranium Carbene Chemistry. Comments Inorg. Chem. 2015, 35, 262–294. [CrossRef] 28. Walensky, J.R.; Martin, R.L.; Ziller, J.W.; Evans, W.J. Importance of energy level matching for bonding in 3+ 3+ i i Th -Am actinide metallocene amidinates, (C5Me5)2[ PrNC(Me)N Pr]An. Inorg. Chem. 2010, 49, 10007–10012. [CrossRef][PubMed] 29. Behrle, A.C.; Barnes, C.L.; Kaltsoyannis, N.; Walensky, J.R. Systematic investigation of (IV)- and uranium(IV)-ligand bonding in dithiophosphonate, thioselenophosphinate, and diselenophosphonate complexes. Inorg. Chem. 2013, 52, 10623–10631. [CrossRef][PubMed] 30. Behrle, A.C.; Kerridge, A.; Walensky, J.R. Dithio- and Diselenophosphinate Thorium(IV) and Uranium(IV) Complexes: Molecular and Electronic Structures, Spectroscopy, and Transmetalation Reactivity. Inorg. Chem. 2015, 54, 11625–11636. [CrossRef][PubMed] 31. Kaltsoyannis, N. Does covalency increase or decrease across the actinide series? Implications for minor actinide partitioning. Inorg. Chem. 2013, 52, 3407–3413. [CrossRef][PubMed]

32. Tassell, M.J.; Kaltsoyannis, N. Covalency in AnCp4 (An = Th–Cm): A comparison of molecular orbital, natural population and atoms-in-molecules analyses. Dalt. Trans. 2010, 39, 6719–6725. [CrossRef][PubMed] 33. Kirker, I.; Kaltsoyannis, N. Does covalency really increase across the 5f series? A comparison of molecular 5 orbital, natural population, spin and electron density analyses of AnCp3 (An = Th−Cm; Cp = η -C5H5). Dalton Trans. 2011, 40, 124–131. [CrossRef][PubMed] 34. Huang, Q.R.; Kingham, J.R.; Kaltsoyannis, N. The strength of actinide–element bonds from the quantum theory of atoms-in-molecules. Dalt. Trans. 2015, 44, 2554–2566. [CrossRef][PubMed] 35. Trumm, M.; Schimmelpfennig, B.; Geist, A. Structure and separation quality of various N- and O-donor ligands from quantum-chemical calculations. Nukleonika 2015, 60, 847–851. [CrossRef] 36. Trumm, M.; Schimmelpfennig, B. Towards the origin of effective An(III)/Ln(III) separation by tridentate N-donor ligands: A theoretical study on atomic charges and polarisabilities for Cm(III)/Gd(III) separation. Mol. Phys. 2016, 114, 876–883. [CrossRef] Inorganics 2018, 6, 88 17 of 19

37. Vallet, V.; Wahlgren, U.; Grenthe, I. Probing the nature of chemical bonding in uranyl(VI) complexes with quantum chemical methods. J. Phys. Chem. A 2012, 116, 12373–12380. [CrossRef][PubMed] 38. Szabo, Z.; Toraishi, T.; Vallet, V.; Grenthe, I. Solution coordination chemistry of actinides: Thermodynamics, structure and reaction mechanisms. Coord. Chem. Rev. 2006, 250, 784–815. [CrossRef] 39. Di Pietro, P.; Kerridge, A. Assessing covalency in equatorial U–N bonds: Density based measures of bonding in BTP and isoamethyrin complexes of uranyl. Phys. Chem. Chem. Phys. 2016, 18, 16830–16839. [CrossRef] [PubMed] 40. Di Pietro, P.; Kerridge, A. Ligand size dependence of U–N and U–O bond character in a series of uranyl hexaphyrin complexes: Quantum chemical simulation and density based analysis. Phys. Chem. Chem. Phys. 2017, 19, 7546–7559. [CrossRef][PubMed] 41. Fryer-Kanssen, I.; Austin, J.; Kerridge, A. Topological Study of Bonding in Aquo and Bis(triazinyl)pyridine Complexes of Trivalent Lanthanides and Actinides: Does Covalency Imply Stability? Inorg. Chem. 2016, 55, 10034–10042. [CrossRef][PubMed] 42. Woodall, S.D.; Swinburne, A.N.; Kerridge, A.; Di Pietro, P.; Adam, C.; Kaden, P.; Natrajan, L.S. Neptunyl(VI) centred visible LMCT emission directly observable in the presence of uranyl(VI). Chem. Commun. 2015, 51, 5402–5405. [CrossRef][PubMed] 43. Guillaumont, D. Quantum Chemistry Study of Actinide(III) and Lanthanide(III) Complexes with Tridentate Ligands. J. Phys. Chem. A 2004, 108, 6893–6900. [CrossRef] 44. Petit, L.; Joubert, L.; Maldivi, P.; Adamo, C. A comprehensive theoretical view of the bonding in actinide molecular complexes. J. Am. Chem. Soc. 2006, 128, 2190–2191. [CrossRef][PubMed] 45. Petit, L.; Adamo, C.; Maldivi, P. Toward a Clear-Cut Vision on the Origin of Insights from Theory. Society 2006, 45, 8517–8522. 46. Lan, J.; Shi, W.; Yuan, L.; Zhao, Y.; Li, J.; Chai, Z. Trivalent Actinide and Lanthanide Separations by Tetradentate Nitrogen Ligands: A Quantum Chemistry Study. Inorg. Chem. 2011, 50, 9230–9237. [CrossRef][PubMed] 47. Roy, L.E.; Bridges, N.J.; Martin, L.R. Theoretical insights into covalency driven f element separations. Dalton Trans. 2013, 42, 2636–2642. [CrossRef][PubMed] 48. Dolg, M.; Cao, X.; Ciupka, J. Misleading evidence for covalent bonding from EuIIIX and AmIIIX density functional theory bond lengths. J. Electron Spectros. Relat. Phenomena 2014, 194, 8–13. [CrossRef]

49. Wang, C.; Cheng, C.; Su, J.; Huai, P. Bonding nature of the actinide tetrafluorides AnF4 (An = Th−Cm). Mol. Phys. 2015, 113, 3450–3458. [CrossRef] 50. Kaneko, M.; Miyashita, S.; Nakashima, S. Bonding Study on the Chemical Separation of Am(III) from Eu(III) by S-, N-, and O-Donor Ligands by Means of All-Electron ZORA-DFT Calculation. Inorg. Chem. 2015, 54, 7103–7109. [CrossRef][PubMed] 51. Bryantsev, V.S.; Hay, B.P. Theoretical prediction of Am(III)/Eu(III) selectivity to aid the design of actinide-lanthanide separation agents. Dalt. Trans. 2015, 7935–7942. [CrossRef][PubMed] 52. Duignan, T.; Autschbach, J. Impact of the Kohn-Sham Delocalization Error on the 4f Shell Localization and Population in Lanthanide Complexes. J. Chem. Theory Comput. 2016, 12, 3109–3121. [CrossRef][PubMed] 53. Hu, S.X.; Li, W.L.; Dong, L.; Gibson, J.K.; Li, J. Crown ether complexes of actinyls: A computational assessment 2+ of AnO2 (15-crown-5) (An = U, Np, Pu, Am, Cm). Dalt. Trans. 2017, 46, 12354–12363. [CrossRef][PubMed] 54. Madic, C.; Hudson, M.J.; Liljenzin, J.O.; Glatz, J.P.; Nannicini, R.; Facchini, A.; Kolarik, Z.; Odoj, R. Recent achievements in the development of partitioning processes of minor actinides from nuclear wastes obtained in the frame of the NEWPART European Programme (1996–1999). Prog. Nucl. Energy 2002, 40, 523–526. [CrossRef] 55. Nilsson, M.; Nash, K.L. Review article: A review of the development and operational characteristics of the TALSPEAK process. Solvent Extr. Ion Exch. 2007, 25, 665–701. [CrossRef] 56. Bader, R.F.W. Atoms in Molecules: A Quantum Theory; Oxford University Press: Oxford, UK, 1990. 57. Kerridge, A. Quantification of f-element covalency through analysis of the electron density: Insights from simulation. Chem. Commun. 2017, 53, 6685–6695. [CrossRef][PubMed] 58. Zhurov, V.V.; Zhurova, E.A.; Stash, A.I.; Pinkerton, A.A. Characterization of bonding in cesium uranyl chloride: Topological analysis of the experimental charge density. J. Phys. Chem. A 2011, 115, 13016–13023. [CrossRef][PubMed] 59. Kerridge, A. Oxidation state and covalency in f-element metallocenes (M = Ce, Th, Pu): A combined CASSCF and topological study. Dalton Trans. 2013, 42, 16428–16436. [CrossRef][PubMed] Inorganics 2018, 6, 88 18 of 19

60. Kerridge, A. f-orbital covalency in the actinocenes (An = Th–Cm): Multiconfigurational studies and topological analysis. RSC Adv. 2014, 4, 12078–12086. [CrossRef] 61. Beekmeyer, R.; Kerridge, A. Assessing Covalency in Cerium and Uranium Hexachlorides: A Correlated Wavefunction and Density Functional Theory Study. Inorganics 2015, 3, 482–499. [CrossRef] 62. Dau, P.D.; Su, J.; Liu, H.T.; Liu, J.B.; Huang, D.L.; Li, J.; Wang, L.S. Observation and investigation of the 2− uranyl tetrafluoride dianion (UO2F4 ) and its solvation complexes with water and acetonitrile. Chem. Sci. 2012, 3, 1137–1146. [CrossRef] 63. Van den Bossche, G.; Spirlet, M.R.; Rebizant, J. Structure of Diammonium Tetrabromodioxouranate(VI) Dihydrate. Acta Crystallogr. C 1987, 43, 383–384. [CrossRef] 64. Kimura, M.; Schomaker, V.; Smith, D.W.; Weinstock, B. Electron-Diffraction Investigation of the Hexafluorides of , , , Uranium, , and . J. Chem. Phys. 1968, 48, 4001–4012. [CrossRef] 65. Mcdowell, R.S.; Asprey, L.B.; Paine, R.T. Vibrational spectrum and force field of uranium hexafluoride. J. Chem. Phys. 1974, 61, 3571–3580. [CrossRef] 66. Zachariasen, W.H. Crystal chemical studies of the 5f-series of elements. V. The crystal structure of uranium hexachloride. Acta Crystallogr. 1948, 1, 285–287. [CrossRef] 67. Perdew, J.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [CrossRef][PubMed] 68. Becke, A.D. Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A 1988, 38, 3098–3100. [CrossRef] 69. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. [CrossRef] 70. Adamo, C.; Barone, V. Toward reliable density functional methods without adjustable parameters: The PBE0 model. J. Chem. Phys. 1999, 110, 6158–6170. [CrossRef] 71. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648. [CrossRef] 72. Stephens, P.; Devlin, F.; Chabalowski, C.; Frisch, M. Ab initio calculation of vibrational absorption and circular dichroism spectra using density functional force fields. J. Phys. Chem. 1994, 98, 11623–11627. [CrossRef] 73. Becke, A.D. A new mixing of Hartree–Fock and local density-functional theories. J. Chem. Phys. 1993, 98, 1372–1377. [CrossRef] 74. Bauzá, A.; Quiñonero, D.; Deyaì, P.M.; Frontera, A. Is the use of diffuse functions essential for the properly description of noncovalent interactions involving anions? J. Phys. Chem. A 2013, 117, 2651–2655. [CrossRef] [PubMed] 75. Di Pietro, P.; Kerridge, A. U-Oyl Stretching Vibrations as a Quantitative Measure of the Equatorial Bond Covalency in Uranyl Complexes: A Quantum-Chemical Investigation. Inorg. Chem. 2016, 55, 573–583. [CrossRef] [PubMed] 76. Watkin, D.J.; Denning, R.G.; Prout, K. Structure of dicaesium tetrachlorodioxouranium(VI). Acta Crystallogr. Sect. C Cryst. Struct. Commun. 1991, 47, 2517–2519. [CrossRef] 77. Bullock, J.I. Raman and infrared spectroscopic studies of the uranyl ion: The symmetric stretching frequency, force constants, and bond lengths. J. Chem. Soc. A 1969, 781–784. [CrossRef] 78. Schreckenbach, G.; Hay, P.J.; Martin, R.L. Density functional calculations on actinide compounds: Survey of 2− recent progress and application to [UO2X4] (X = F, Cl, OH) and AnF6 (An = U, Np, Pu). J. Comput. Chem. 1999, 20, 70–90. [CrossRef] 79. Batista, E.R.; Martin, R.L.; Hay, P.J. Density functional investigations of the properties and thermochemistry of UFn and UCln (n = 1,...,6). J. Chem. Phys. 2004, 121, 11104–11111. [CrossRef][PubMed] 80. Grenthe, I.; Drozd˙ zy´nski,J.;˙ Fujino, T.; Buck, E.C.; Albrecht-Schmitt, T.E.; Wolf, S.F. Uranium. In The Chemistry of the Actinide and Transactinide Elements; Morss, L.R., Edelstein, N.M., Fuger, J., Katz, J.J., Eds.; Springer: Dordrecht, The Netherlands, 2006; pp. 253–698.

81. Yun-guang, Z.; Yu-de, L. Relativistic density functional investigation of UX6 (X = F, Cl, Br and I). Chin. Phys. B 2010, 19, 033302. [CrossRef] 82. TURBOMOLE, version 6.6, A development of University of Karlsruhe and Forschungszentrum Karlsruhe GmbH. 2014. Available online: http://www.turbomole.com (accessed on 29 August 2018). Inorganics 2018, 6, 88 19 of 19

83. Ahlrichs, R.; Bär, M.; Häser, M.; Horn, H.; Kölmel, C. Electronic structure calculations on workstation computers: The program system turbomole. Chem. Phys. Lett. 1989, 162, 165–169. [CrossRef] 84. Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297–3305. [CrossRef] [PubMed] 85. Küchle, W.; Dolg, M.; Stoll, H.; Preuss, H. Energy-adjusted pseudopotentials for the actinides. Parameter sets and test calculations for thorium and thorium monoxide. J. Chem. Phys. 1994, 100, 7535–7542. [CrossRef] 86. Cao, X.; Dolg, M. Segmented contraction scheme for small-core actinide pseudopotential basis sets. J. Mol. Struct. THEOCHEM 2004, 673, 203–209. [CrossRef] 87. De Jong, W.A.; Visscher, L.; Nieuwpoort, W.C. On the bonding and the electric field gradient of the uranyl ion. J. Mol. Struc. 1999, 458, 41–52. [CrossRef] 88. Fryer-Kanssen, I.; Kerridge, A. Elucidation of the inverse trans influence in uranyl, its imido and carbene analogues via quantum chemical simulation. Chem. Commun. 2018, 54, 9761–9764. [CrossRef][PubMed] 89. Lu, T.; Chen, F. Multiwfn: A multifunctional wavefunction analyzer. J. Comput. Chem. 2012, 33, 580–592. [CrossRef][PubMed]

© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (http://creativecommons.org/licenses/by/4.0/).