Abstract

The stability of polynuclear anions composed of ScF3 building blocks was studied by using ab initio and density functional theory electronic structure methods and flexible basis sets. Thorough exploration of ground state potential energy surfaces of (Sc2F7)−, (Sc3F10)−, and (Sc4F13)− anions which may be viewed as comprising ScF3 fragments and the additional fluorine atom led to determining the isomeric structures thereof. It was found that the most stable isomers which are predicted to dominate at room temperature correspond to the compact structures enabling the formation of a large number of Sc–F–Sc bridging linkages rather than to the chain-like structures. The vertical electron detachment energies of the (ScnF3n+1)− anions were found to be very large (spanning the 10.85–12.29 eV range) and increasing with the increasing number of scandium atoms (n) and thus the ScF3 building blocks involved in the structure. Thermodynamic stability of (ScnF3n+1)− anions (i.e., their susceptibility to fragmentation) was also verified and discussed.
Short abstract
The anions matching the (ScnF3n+1)− formula are thermodynamically stable systems, not susceptible to fragmentations yielding either ScF3 or F− products. The vertical electron detachment energies of the (ScnF3n+1)− (n = 1−4) anions span the 10.01−12.29 eV range and increase when n develops from 1 to 4.
Introduction
There is considerable interest in studying superhalogen anions not only because they represent molecular systems exhibiting very large excess electron binding energies but also they play an important role in ionic liquids by acting as their negatively charged building blocks.1,2 In addition, superhalogens are being employed in the preparation of many new materials, including organic metals and organic superconductors, hydrogen storage materials, solar cells, and Li-ion batteries.3−7 The term “superhalogen” was first used in 1981 by Gutsev and Boldyrev who confirmed the stabilities and large excess electron binding energies of several systems matching the (MXk+1)− formula (where M is a central metal atom of maximal valence k, whereas X stands for halogen atom).8 As recognized during many other studies that followed, the (MXk+1)− formula is much more general than it originally appeared. Namely, it was demonstrated that various molecular fragments (such as acidic functional groups,9 halogenoids,10 electrophilic groups,11 and even superhalogens themselves12−15) may act as suitable ligands X, whereas certain nonmetal atoms (e.g., phosphorus,16 silicon,17−20 hydrogen21,22) can play the central atom M role (see the recent review article23).
The extension of the superhalogen formula led to defining the polynuclear (MnXn·k+1)− superhalogen anions (e.g., (Ca2(CN)5)−, (Na4Cl5)−), whose various applications are constantly being discovered.24−38 However, theoretical studies aimed to determine the properties of such species are difficult due to the large number of isomeric structures that must be considered and characterized. Albeit several (MnXn·k+1)− negatively charged systems have already been described in the literature,23 many more polynuclear superhalogen anions are still waiting to be discovered. In particular, reports describing the (MnXn·k+1)− anions containing transition and rare-earth-metal atoms are very scarce.
In this contribution, we present the results of our theoretical investigation concerning polynuclear superhalogen anions containing scandium (historically classified as a rare-earth element) decorated with fluorine ligands. Since scandium chemistry is almost completely dominated by the compounds involving Sc3+ trivalent cation (with scandium oxide (Sc2O3) and scandium fluoride (ScF3) as prominent examples),39−41 it is commonly assumed that the scandium most predominant oxidation state is +3. Indeed, as demonstrated by Pradhan et al., (ScX4)− tetrahedral anions exhibit the largest vertical electron detachment energies (VDEs) among the (ScXn)− (X = F, Cl, Br; n = 1–5) systems considered.42 Therefore, we decided to investigate the (ScnF3n+1)− anions (for n = 2–4) which can be viewed as comprising n ScF3 moieties and one additional fluorine atom. To the best of our knowledge, the results concerning the (Sc2F7)−, (Sc3F10)−, and (Sc4F13)− anions have not been reported in the literature thus far. We believe that providing the knowledge of the most stable isomeric structures of (ScnF3n+1)− anions (n = 2–4) and their excess electron binding energies will assist the researchers in designing new scandium-containing alloys and other scandium-doped materials.
Methods
Preselection of the Isomeric Structures
The search for the low-energy isomeric structures of (ScnF3n+1)− anions (n = 2–4) and their corresponding neutral ScnF3n+1 parents was carried out (for each n) using the Coalescence Kick (CK) method.43,44 In the CK procedure, a large number of random structures are initially generated. Since these initial random structures often consist of nonbonded molecular fragments, the fragmented parts are simultaneously brought together (by pushing them to the center of mass) to achieve a so-called “coalescence” (connectivity). Once each obtained structure is checked for connectivity, the preliminary geometry optimization process begins. In our case, the CK procedure assumed the initial optimizations with the B3LYP method45,46 together with the Los Alamos National Laboratory (LANL) effective core potentials (ECPs) with the appropriate valence basis set of double-ζ quality (denoted LANL2DZ).47−49
For each (ScnF3n+1)− and ScnF3n+1 (n = 2–4) system considered, ca. 2000–3000 initial structures were generated, coalesced with the CK, and then transformed to represent the nearest local minima by performing the geometry optimization at the B3LYP/LANL2DZ level for each trial structure. In such a way, the CK technique allowed us to preselect the lowest energy isomers in all (ScnF3n+1)− and ScnF3n+1 (n = 2–4) cases, which were further investigated at a more reliable level of theory.
Refinement of the Isomeric Structures
The (ScnF3n+1)− and ScnF3n+1 (n = 2–4) isomeric structures preselected by using the CK procedure were refined by applying the density functional theory with the ωB97XD long-range-corrected functional including empirical dispersion50 together with the aug-cc-pVDZ basis set51,52 for fluorine atoms and Stuttgart RSC 1997 effective core potential (also known as Stuttgart/Dresden effective core potential, SDD) for scandium atoms.53−55 In each case, the harmonic vibrational frequencies characterizing the stationary point structure were evaluated at the same level of theory to ensure that all obtained structures correspond to true minima on the potential energy surface.
For the selected anionic isomers investigated, namely, three negatively charged (Sc2F7)− systems and six isomeric (Sc3F10)− structures, we determined the stationary point structures at both ωB97XD/SDD/aug-cc-pVDZ and MP2/Sapporo-QZP-diffuse theory levels (where MP2 stands for the second-order Mo̷ller-Plesset perturbation method (MP2),56−58 whereas Sapporo-QZP-diffuse indicates the all-electron quadrupole-zeta Sapporo basis set supplemented with diffuse functions for Sc59,60 and F61,62) to verify whether the former theoretical treatment provides the bond lengths and valence and dihedral angles that can be considered reliable. As it turned out, the structures obtained at the ωB97XD/SDD/aug-cc-pVDZ level differ from those determined at the MP2/Sapporo-QZP-diffuse level only slightly (i.e., the latter treatment leads to bond lengths longer by less than 0.11 Å than those predicted by the former approach, whereas the valence angles obtained at the ωB97XD/SDD/aug-cc-pVDZ theory level differ by less than 3° than those predicted by the MP2/Sapporo-QZP-diffuse treatment. Therefore, we are confident that the ωB97XD/SDD/aug-cc-pVDZ theory level we applied to obtain the isomeric structures of (ScnF3n+1)− and ScnF3n+1 (n = 2–4) systems is sufficient.
Refinement of the Electronic Energies
The electronic energies of the (ScnF3n+1)− and ScnF3n+1 (n = 2–4) isomeric structures obtained at the ωB97XD/SDD/aug-cc-pVDZ theory level were refined by employing the coupled-cluster method with the single and double excitations (CCSD)63−65 using the same SDD/aug-cc-pVDZ basis sets.
In order to verify whether the relative energies of the isomers obtained at the CCSD/SDD/aug-cc-pVDZ theory level can be considered reliable, we performed additional calculations of the relative energies of the (Sc2F7)− isomers using the same CCSD method with the all-electron Sapporo-QZP-diffuse basis set. As it turned out, the energy order of the isomers remains the same regardless the basis sets employed while the differences in relative energies are larger by ca. 0.09–0.13 eV when the CCSD/Sapporo-QZP-diffuse treatment is applied (in comparison to the relative energies obtained at the CCSD/SDD/aug-cc-pVDZ theory level). In addition, we calculated the relative energies of six isomeric (Sc3F10)− structures by employing the CCSD/Sapporo-TZP-diffuse treatment to verify whether the corresponding relative energies obtained at the CCSD/SDD/aug-cc-pVDZ theory level can be considered reliable. As it turned out, the relative energies obtained for the (Sc3F10)− isomers at the CCSD/Sapporo-TZP-diffuse theory level differ from those obtained at the CCSD/SDD/aug-cc-pVDZ theory level by 0.04–0.13 eV. Therefore, we are confident that the refined electronic energies of the (ScnF3n+1)− and ScnF3n+1 isomers (n = 2–4) and thus their relative energies we predicted by employing the CCSD/SDD/aug-cc-pVDZ treatment can be considered reliable.
Vertical Electron Detachment Energies
The VDEs characterizing the isomers of the (ScnF3n+1)− anions (n = 2–4) were calculated by applying the outer valence Green function OVGF method (B approximation)66−74 together with the Stuttgart RSC 1997 effective core potential for scandium and the aug-cc-pVDZ basis set for fluorine. Due to the fact that the OVGF approximation remains valid only for outer valence ionization for which the pole strengths (PS) are greater than 0.80–0.85,75 we verified that the PS values obtained were sufficiently large (i.e., spanning the 0.930–0.938 range) to justify the use of the OVGF method.
The adiabatic electron affinities of the ScnF3n+1 neutral systems (n = 2–4) were obtained by subtracting the CCSD/SDD/aug-cc-pVDZ electronic energies of the (ScnF3n+1)− anions from those of their corresponding neutral parents (all determined for the lowest energy isomeric structure obtained at the ωB97XD/SDD/aug-cc-pVDZ theory level).
In addition, we verified whether the VDEs obtained at the OVGF/SDD/aug-cc-pVDZ theory level can be considered reliable by calculating the vertical electron detachment energy for the tetrahedral (ScF4)− anion using two approaches: (i) the OVGF method and all-electron Sapporo-QZP-diffuse basis set for Sc and F and (ii) the OVGF method and SDD pseudopotentials for Sc and the aug-cc-pVDZ basis set for F. Since we found that the VDE of the Td-symmetry (ScF4)− anion calculated by applying the OVGF/Sapporo-QZP-diffuse treatment (10.040 eV) is larger by only 0.034 eV than that obtained by using the OVGF/SDD/aug-cc-pVDZ approach (10.006 eV), we believe that our vertical electron detachment energies predicted for the (ScnF3n+1)− (n = 2–4) isomers at the OVGF/SDD/aug-cc-pVDZ theory level are reliable yet likely underestimated by ca. 0.5%.
Additional Remarks
The reaction energies and the Gibbs free reaction energies (at T = 298.15 K) for the fragmentation processes of the (ScnF3n+1)− (n = 2–4) anions were calculated using the CCSD/SDD/aug-cc-pVDZ electronic energies and zero-point energy corrections, thermal corrections, and entropy contributions estimated at the ωB97XD/SDD/aug-cc-pVDZ theory level.
Since the neutral ScnF3n+1 (n = 2–4) systems are open-shell molecules, we used methods based on an unrestricted Kohn–Sham or Hartree–Fock starting point. Hence, it was important to make sure that little (if any) artificial spin contamination enters into the final wave functions. We computed the expectation value ⟨S2⟩ for the states studied in this work and found values not exceeding 0.774 for doublet neutral species (at the unrestricted DFT or HF level). Hence, we are confident that spin contamination is not large enough to significantly affect our findings.
All calculations were performed with the GAUSSIAN16 (Rev.C.01) package.76
Results and Discussion
Isomeric Structures of (Sc2F7)− Anions
The search for the geometrically stable structures of the (Sc2F7)− anion led to three isomers labeled 1-3 in Figure 1. The global minimum (1-(Sc2F7)−) corresponds to the C2-symmetry structure with two ScF2 fragments oriented nearly perpendicularly to each other and connected via three F atoms forming a triangle between the scandium atoms. The lengths of the terminal Sc–F bonds (1.883–1.890 Å) are shorter than those of the scandium–fluorine separations in the Sc–F–Sc bridging fragments (1.982–2.245 Å). In the second lowest energy structure (labeled 2-(Sc2F7)− in Figure 1), the Sc atoms are connected via two F atoms (with the Sc–F distances spanning the 1.971–2.303 Å range) which leads to the uneven distribution of the remaining ligands between two central atoms. In consequence, the energy of the Cs-symmetry isomer 2-(Sc2F7)− is slightly (by 0.16 eV) larger than that of the global minimum. Since the remaining isomer 3-(Sc2F7)− whose relative energy (ΔE) is even higher (0.21 eV) contains only one Sc–F–Sc connection (see Figure 1) between two ScF3 moieties, we conclude that the formation of three Sc–F–Sc bridging fragments is energetically favorable in polynuclear superhalogen anions involving scandium central atoms and fluorine ligands.
Figure 1.

Isomeric structures of the (Sc2F7)− anion with relative energies (ΔE).
Isomeric Structures of (Sc3F10)− Anions
In the case of the (Sc3F10)− anion, we found six isomeric structures whose relative energies do not exceed 1.7 eV. Although we realize that the formation of isomers whose relative energies are large is highly unlikely at room temperature, we decided to include such structures in our discussion to demonstrate that the chain-like isomers in which the fragments are connected via the single Sc–F–Sc bridging linkage are significantly less stable than the structurally compact isomers containing scandium atoms connected through two or three such linkages.
The most stable isomer of (Sc3F10)− anion (labeled 1-(Sc3F10)− in Figure 2) corresponds to the compact C2v-symmetry structure containing the cage-like fragment formed by the planar 6-member ring (involving alternately aligned Sc and F atoms connected via the Sc–F bonds whose lengths span the 2.015–2.120 Å range) and two F atoms below and above the ring plane. Such an alignment allows each Sc to be involved in two Sc–F–Sc and two Sc-FSc2 motifs, whereas the remaining fluorine ligands are distributed among three Sc atoms and connected to them through shorter (1.868–1.873 Å) bonds. Among the other isomers, the only structure which could be considered competitive (due to its small ΔE value of 0.17 eV) with the lowest energy isomer 1 is the C3v-symmetry 2-(Sc3F10)− isomer which differs from 1-(Sc3F10)− by the position of one fluorine ligand. Namely, instead of being involved in the cage-like fragment as in 1-(Sc3F10)−, one more F atom is connected to a single Sc atom, thus making six F ligands distributed evenly among three central Sc atoms and sticking out of the 6-member ring. In consequence, the ring itself becomes nonplanar and adopts a chair conformation instead (see Figure 2). The third lowest energy isomer, 3-(Sc3F10)−, corresponds to a chain-like C2-symmetry structure with scandium atoms connected via three Sc–F–Sc bridging linkages and thus resembles the global minimum identified for (Sc2F7)− anion (cf. 1-(Sc2F7)− and 3-(Sc3F10)−, depicted in Figures 1 and 2, respectively). Although the relative energy of 3-(Sc3F10)− is larger than that of the global minimum by 0.31 eV, this isomer is substantially more stable than the remaining 4–6 isomers of (Sc3F10)−, likely due to the presence of a large number of Sc–F–Sc bridging fragments which stabilize the structure. As far as these remaining higher energy 4–6 isomers are concerned, none of them contains a pair of Sc atoms connected via three fluorine ligands. In 4-(Sc3F10)− (ΔE = 0.68 eV) and 5-(Sc3F10)− (ΔE = 1.19 eV), only one double Sc–F–Sc linkage can be distinguished, whereas the isomer 6-(Sc3F10)− (ΔE = 1.70 eV) corresponds to the F3Sc–F–ScF2–F-ScF3 structure containing the scandium atoms connected only through single Sc–F–Sc bridges (nota bene, the structures in which the central atoms are linked via one halogen atom were assumed to correspond to the most stable isomers of polynuclear superhalogen anions in early reports concerning these systems77).
Figure 2.

Isomeric structures of the (Sc3F10)− anion with relative energies (ΔE).
Isomeric Structures of (Sc4F13)− Anions
The extensive search of the ground state potential energy surface of the (Sc4F13)− anion led to three isomeric structures whose relative energies are within 1.3 eV and several isomers whose ΔE values span the 1.3–4.1 eV range. As in the preceding section, we decided to briefly mention also these high energy isomers in our discussion because we wanted to include the fully extended chain-like structure while making the comparison of the relative stability. The global minimum (labeled 1-(Sc4F13)− in Figure 3) corresponds to the C2v-symmetry compact structure, resembling a double crown with the F atom in its center and four Sc atoms forming a square. Adopting such a structure seems to maximize the number of Sc–F–Sc connections leaving only four fluorine ligands outside that central substructure. The Sc–F bond lengths span the 2.075–2.178 Å range in a double crown fragment, whereas the lengths of the Sc–F bonds sticking out of that substructure are shorter (1.860 Å).
Figure 3.

Isomeric structures of the (Sc4F13)− anion with their relative energies (ΔE) (see Figure 4 for the remaining isomers of (Sc4F13)−).
The second lowest energy isomer (2-(Sc4F13)−) also corresponds to the C2v-symmetry compact structure yet with a different (i.e., nonplanar) mutual alignment of the Sc atoms. In consequence, the number of Sc–F–Sc bonds in 2-(Sc4F13)− is smaller than that in 1-(Sc4F13)− which results in the higher electronic energy of the former (ΔE = 0.56 eV). Such a large energy gap between the global minimum 1 and the local minimum 2 indicates that the formation of any other isomeric structure of (Sc4F13)− but 1-(Sc4F13)− should be considered highly unlikely at room temperature. Apart from these two low-energy isomers, we found several other geometrically stable structures of (Sc4F13)− having much larger relative energies in the range of 1.30–2.49 eV (see the isomers labeled 3–12 in Figures 3 and 4). Alike in the (Sc2F7)− and (Sc3F10)− cases, the fully extended isomer of the (Sc4F13)− anion (i.e., containing the scandium atoms linked via one fluorine ligand, see 13-(Sc4F13)− in Figure 4) has a very large relative energy (4.05 eV).
Figure 4.

Isomeric structures of the (Sc4F13)− anion with their relative energies (ΔE) (see Figure 3 for the lower energy isomers of that system).
Excess Electron Binding Energies of (ScnF3n+1)− Anions (n = 1–4)
Since the (ScnF3n+1)− anions (n = 2–4) can be viewed as obtained by subsequent attachments of the ScF3 fragment to the (ScF4)− anion (matching the (ScnF3n+1)− formula for n = 1), we include the (ScF4)− system in our discussion of the electronic stability. In general, it should be emphasized that the vertical electron detachment energies of all (ScnF3n+1)− anions (n = 1–4) are very large, as even the VDE of (ScF4)− exceeds 10 eV. Next, the results gathered in Table 1 indicate that the VDEs of the (ScnF3n+1)− anions increase when n develops from 1 to 4. Namely, taking into account all isomeric structures, we calculated the VDEs to span the 10.44–11.48 eV range (for n = 2), the 11.31–12.06 eV range (for n = 3), and the 11.44–12.38 eV range (for n = 4). Certainly, the increase in the VDE for (ScnF3n+1)− anions as n develops from 1 to 4 can be attributed to the growing number of fluorine ligands. Since we predicted only the lowest energy isomer of each anion to exist at room temperature, it seems reasonable to compare their VDE values as representative for polynuclear (ScnF3n+1)− superhalogen anions of different size. Such a comparison reveals that attaching the first ScF3 subunit to (ScF4)− increases the VDE by 0.84 eV, whereas the subsequent ScF3 fragment attachments further increase the VDE by 0.47 eV and then by 0.97 eV, leading to the 1-(Sc2F7)−, 1-(Sc3F10)−, and 1-(Sc4F13)− anions having their VDEs of 10.85, 11.32, and 12.29 eV, respectively. Such an increase of the excess electron binding energy with the increasing number of ScF3 fragments involved in the structure seems very large and we suspect that even larger vertical electron detachment energies could be achieved for the (ScnF3n+1)− superhalogen anions with n > 4.
Table 1. Vertical Electron Detachment Energies (in eV) and Relative Energies (ΔE in eV) of Isomeric Structures of (ScnF3n+1)− Anions (n = 1–4)a.
| system | ΔE | VDE |
|---|---|---|
| (ScF4)− | 0.00 | 10.01 |
| 1-(Sc2F7)− | 0.00 | 10.85 |
| 2-(Sc2F7)− | 0.16 | 10.44 |
| 3-(Sc2F7)− | 0.21 | 11.48 |
| 1-(Sc3F10)− | 0.00 | 11.32 |
| 2-(Sc3F10)− | 0.17 | 11.80 |
| 3-(Sc3F10)− | 0.31 | 11.53 |
| 4-(Sc3F10)− | 0.68 | 11.31 |
| 5-(Sc3F10)− | 1.19 | 11.45 |
| 6-(Sc3F10)− | 1.70 | 12.06 |
| 1-(Sc4F13)− | 0.00 | 12.29 |
| 2-(Sc4F13)− | 0.56 | 11.80 |
| 3-(Sc4F13)− | 1.30 | 11.44 |
| 4-(Sc4F13)− | 1.43 | 11.73 |
| 5-(Sc4F13)− | 1.64 | 11.56 |
| 6-(Sc4F13)− | 1.66 | 11.57 |
| 7-(Sc4F13)− | 1.68 | 11.93 |
| 8-(Sc4F13)− | 1.71 | 11.92 |
| 9-(Sc4F13)− | 1.85 | 11.51 |
| 10-(Sc4F13)− | 2.03 | 12.28 |
| 11-(Sc4F13)− | 2.38 | 12.02 |
| 12-(Sc4F13)− | 2.49 | 12.07 |
| 13-(Sc4F13)− | 4.05 | 12.38 |
The VDEs were calculated at the OVGF/SDD/aug-cc-pVDZ theory level, and ΔEs were determined at the CCSD/SDD/aug-cc-pVDZ theory level (both for the geometries optimized with the ωB97XD method and SDD/aug-cc-pVDZ basis sets).
Despite the large vertical excess electron binding energies of the (ScnF3n+1)− anions, the adiabatic electron affinities (AEAs) of their corresponding neutral ScnF3n+1 parents are substantially smaller. Namely, we predicted an AEA of 7.46, 7.57, 8.18, and 9.37 eV for ScF4, Sc2F7, Sc3F10, and Sc4F13, respectively. Clearly, the structure reorganization caused by the absence of an excess electron leads to the neutral isomers whose geometrical parameters are significantly different than those of their corresponding daughter anions.
The highest doubly occupied molecular orbitals (HOMO) of the most stable 1-(Sc2F7)−, 1-(Sc3F10)−, and 1-(Sc4F13)− isomeric structure reveal both (i) the effective delocalization of an excess negative charge among electronegative fluorine ligands and (ii) the absence of destabilizing antibonding ligand-central atom interactions (see Figure 5 where the HOMO for the (ScF4)− anion is also shown for comparison).
Figure 5.

Highest energy canonical molecular orbitals and their corresponding eigenvalues for the most stable isomers of (ScnF3n+1)− anions (n = 1–4).
Stability of (ScnF3n+1)− Anions (n = 2–4) against Fragmentation
In order to verify whether the (ScnF3n+1)− anions are thermodynamically stable, we evaluated the Gibbs free reaction energies at T = 298.15 K (ΔG298r) for the most probable fragmentation processes these species might be susceptible to. Namely, we considered two fragmentation channels for each anion: (i) the detachment of the ScF3 system (i.e., (ScnF3n+1)− → (Scn–1F3(n–1)+1)− + ScF3) and (ii) the detachment of the F– anion (i.e., (ScnF3n+1)− → ScnF3n + F–). The results collected in Table 2 indicate that all (ScnF3n+1)− anions considered are thermodynamically stable as the ΔG298r values corresponding to their fragmentation processes are positive and span the 1.95–5.03 eV range. In particular, the Gibbs free reaction energies predicted for the detachment of the ScF3 system from a given (ScnF3n+1)− anion are ca. two times smaller than those predicted for the F– loss, which is likely caused by large electronic stabilities of the (Scn–1F3(n–1)+1)− superhalogen anions produced in the former processes. In addition, we verified that the (ScF4)− anion is also thermodynamically stable as the Gibbs free reaction energy calculated for the (ScF4)− → ScF3 + F– process is equal to 4.26 eV.
Table 2. Reaction Energies (ΔEr in eV) and Gibbs Free Reaction Energies (ΔG298r in eV) Predicted for the Fragmentation Processes of the (ScnF3n+1)− Anions (n = 2–4) at T = 298.15 K.
| fragmentation path | ΔEr | ΔG298r |
|---|---|---|
| (Sc2F7)− → (ScF4)− + ScF3 | 2.34 | 1.95 |
| (Sc2F7)− → Sc2F4 + F– | 4.57 | 4.29 |
| (Sc3F10)− → (Sc2F7)− + ScF3 | 3.03 | 2.25 |
| (Sc3F10)− → Sc3F9 + F– | 5.60 | 5.03 |
| (Sc4F13)− → (Sc3F10)− + ScF3 | 3.41 | 2.59 |
| (Sc4F13)− → Sc4F12 + F– | 4.70 | 4.26 |
Hence, we are confident that (Sc2F7)−, (Sc3F10)−, and (Sc4F13)− anions are stable against fragmentation processes at room temperature.
Conclusions
On the basis of the ab initio and density functional theory electronic structure calculations carried out using the ωB97XD, CCSD, and OVGF methods with the SDD effective core potentials and the aug-cc-pVDZ basis sets performed for the (ScnF3n+1)− (n = 2–4) anions and their corresponding neutral parents, we arrive at the following conclusions:
(Sc2F7)−, (Sc3F10)−, and (Sc4F13)− anions are thermodynamically stable systems, not susceptible to fragmentations yielding either ScF3 or F– products;
The adiabatic electron affinities characterizing neutral Sc2F7, Sc3F10, and Sc4F13 molecules are relatively large (7.57–9.37 eV);
The vertical electron detachment energies of the (ScnF3n+1)− (n = 1–4) anions are very large (10.01–12.29 eV) and increase when n develops from 1 to 4 (the largest VDE of 12.29 eV corresponds to the most stable isomer of (Sc4F13)− anion);
The analysis of the isomeric structures of the (ScnF3n+1)− (n = 2–4) anions reveals that adopting compact rather than chain-like structures is favored as the former enable the formation of larger number of Sc–F–Sc bridging linkages.
Acknowledgments
The authors dedicate this article to the memory of Prof. Alexander I. Boldyrev. This research was supported by the Polish Ministry of Science and Higher Education grant No. DS 531-T110-D844-23 (to P.S.). The calculations have been carried out using resources provided by Wroclaw Centre for Networking and Supercomputing (http://wcss.pl) grant No. 435 and 455.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.inorgchem.3c02937.
Cartesian coordinates of all systems (PDF)
The authors declare no competing financial interest.
Supplementary Material
References
- Watanabe M.; Thomas M. L.; Zhang S.; Ueno K.; Yasuda T.; Dokko K. Application of Ionic Liquids to Energy Storage and Conversion Materials and Devices. Chem. Rev. 2017, 117, 7190–7239. 10.1021/acs.chemrev.6b00504. [DOI] [PubMed] [Google Scholar]
- Izgorodina E. I.; Seeger Z. L.; Scarborough D. L. A.; Tan S. Y. S. Quantum Chemical Methods for the Prediction of Energetic, Physical, and Spectroscopic Properties of Ionic Liquids. Chem. Rev. 2017, 117, 6696–6754. 10.1021/acs.chemrev.6b00528. [DOI] [PubMed] [Google Scholar]
- Wudl F. From organic metals to superconductors: managing conduction electrons in organic solids. Acc. Chem. Res. 1984, 17, 227–232. 10.1021/ar00102a005. [DOI] [Google Scholar]
- Giri S.; Behera S.; Jena P. Superhalogens as building blocks of halogen-free electrolytes in lithium-ion batteries. Angew. Chem., Int. Ed. Engl. 2014, 53, 13916–13919. 10.1002/anie.201408648. [DOI] [PubMed] [Google Scholar]
- Jena P. Superhalogens: A Bridge between Complex Metal Hydrides and Li Ion Batteries. J. Phys. Chem. Lett. 2015, 6, 1119–1125. 10.1021/acs.jpclett.5b00006. [DOI] [PubMed] [Google Scholar]
- Fang H.; Jena P. Super-ion inspired colorful hybrid perovskite solar cells. J. Mater. Chem. A 2016, 4, 4728–4737. 10.1039/C5TA09646D. [DOI] [Google Scholar]
- Fang H.; Jena P. Molecular Origin of Properties of Organic–Inorganic Hybrid Perovskites: The Big Picture from Small Clusters. J. Phys. Chem. Lett. 2016, 7, 1596–1603. 10.1021/acs.jpclett.6b00435. [DOI] [PubMed] [Google Scholar]
- Gutsev G. L.; Boldyrev A. I. DVM-Xα calculations on the ionization potentials of MXk+1– complex anions and the electron affinities of MXk+1 “superhalogens. Chem. Phys. 1981, 56, 277–283. 10.1016/0301-0104(81)80150-4. [DOI] [Google Scholar]
- Anusiewicz I. Superhalogen Anions Utilizing Acidic Functional Groups As Ligands. J. Phys. Chem. A 2009, 113, 11429–11434. 10.1021/jp907246w. [DOI] [PubMed] [Google Scholar]
- Smuczyńska S.; Skurski P. Halogenoids as Ligands in Superhalogen Anions. Inorg. Chem. 2009, 48, 10231–10238. 10.1021/ic901253r. [DOI] [PubMed] [Google Scholar]
- Anusiewicz I. Electrophilic Substituents as Ligands in Superhalogen Anions. J. Phys. Chem. A 2009, 113, 6511–6516. 10.1021/jp901910q. [DOI] [PubMed] [Google Scholar]
- Paduani C.; Wu M. M.; Willis M.; Jena P. Theoretical Study of the Stability and Electronic Structure of Al(BH4)n=1→4 and Al(BF4)n=1→4 and Their Hyperhalogen Behavior. J. Phys. Chem. A 2011, 115, 10237–10243. 10.1021/jp206330d. [DOI] [PubMed] [Google Scholar]
- Willis M.; Götz M.; Kandalam A. K.; Ganteför G.; Jena P. Hyperhalogens: Discovery of a New Class of Highly Electronegative Species. Angew. Chem., Int. Ed. 2010, 49, 8966–8970. 10.1002/anie.201002212. [DOI] [PubMed] [Google Scholar]
- Paduani C.; Jena P. A Recipe for Designing Molecules with Ever-Increasing Electron Affinities. J. Phys. Chem. A 2012, 116, 1469–1474. 10.1021/jp211736e. [DOI] [PubMed] [Google Scholar]
- Paduani C.; Jena P. Structure, stability and superhalogen properties of sodium and magnesium borohydrides. Chem. Phys. Lett. 2013, 556, 173–177. 10.1016/j.cplett.2012.11.038. [DOI] [Google Scholar]
- Gutsev G. L.; Boldyrev A. I.. The theoretical investigation of the electron affinity of chemical compounds, Advances in Chemical Physics, Vol. LXI, Ed. Prigogine I.; Rice S. A.; John Wiley & Sons, Inc., 1985, 169–221. [Google Scholar]
- Gutsev G. L.; Boldyrev A. I. DVM α calculations on the electronic structure of complex chlorine anions. Chem. Phys. Lett. 1981, 84, 352–355. 10.1016/0009-2614(81)80362-4. [DOI] [Google Scholar]
- Gutsev G. L. A theoretical investigation on the structure of the hypervalent carbon and silicon pentahalogenides as well as their singly charged anions. Chem. Phys. 1992, 166, 57–68. 10.1016/0301-0104(92)87005-T. [DOI] [Google Scholar]
- King R. A.; Mastryukov V. S.; Schaefer H. F. III The electron affinities of the silicon fluorides SiFn (n = 1–5). J. Chem. Phys. 1996, 105, 6880–6886. 10.1063/1.471846. [DOI] [Google Scholar]
- Marchaj M.; Freza S.; Skurski P. Why Are SiX5– and GeX5– (X = F, Cl) Stable but Not CF5– and CCl5–?. J. Phys. Chem. A 2012, 116, 1966–1973. 10.1021/jp300251t. [DOI] [PubMed] [Google Scholar]
- Smuczyńska S.; Skurski P. Is hydrogen capable of playing a central atom role in superhalogen anions?. Chem. Phys. Lett. 2007, 443, 190–193. 10.1016/j.cplett.2007.06.075. [DOI] [Google Scholar]
- Freza S.; Skurski P. Enormously large (approaching 14 eV!) electron binding energies of [HnFn+1]− (n = 1–5, 7, 9, 12) anions. Chem. Phys. Lett. 2010, 487, 19–23. 10.1016/j.cplett.2010.01.022. [DOI] [Google Scholar]
- Skurski P.Superhalogens – Enormously Strong Electron Acceptors, in Superatoms: Principles, Synthesis and Applications, ed. Jena P.; Sun Q.; John Wiley & Sons Ltd.2022. (53–84). [Google Scholar]
- Alexandrova A. N.; Boldyrev A. I.; Fu Y.-J.; Yang X.; Wang X.-B.; Wang L.-S. Structure of the NaxClx+1– (x = 1–4) clusters via ab initio genetic algorithm and photoelectron spectroscopy. J. Chem. Phys. 2004, 121, 5709–5719. 10.1063/1.1783276. [DOI] [PubMed] [Google Scholar]
- Sobczyk M.; Sawicka A.; Skurski P. Theoretical Search for Anions Possessing Large Electron Binding Energies. Eur. J. Inorg. Chem. 2003, 2003, 3790–3797. 10.1002/ejic.200300180. [DOI] [Google Scholar]
- Sikorska C.; Skurski P. The saturation of the excess electron binding energy in AlnF3n+1– (n = 1–5) anions. Chem. Phys. Lett. 2012, 536, 34–38. 10.1016/j.cplett.2012.03.089. [DOI] [Google Scholar]
- Anusiewicz I. Mg2Cl5– and Mg3Cl7– Superhalogen Anions. Aust. J. Chem. 2008, 61, 712–717. 10.1071/CH08212. [DOI] [Google Scholar]
- Ko Y. J.; Wang H.; Pradhan K.; Koirala P.; Kandalam A. K.; Bowen K. H.; Jena P. Superhalogen properties of CumCln clusters: Theory and experiment. J. Chem. Phys. 2011, 135, 244312 10.1063/1.3671457. [DOI] [PubMed] [Google Scholar]
- Wu M. M.; Wang H.; Ko Y. J.; Wang Q.; Sun Q.; Kiran B.; Kandalam A. K.; Bowen K. H.; Jena P. Manganese-Based Magnetic Superhalogens. Angew. Chem., Int. Ed. 2011, 50, 2568–2572. 10.1002/anie.201007205. [DOI] [PubMed] [Google Scholar]
- Yin B.; Li J.; Bai H.; Wen Z.; Jiang Z.; Huang Y. The magnetic coupling in manganese-based dinuclear superhalogens and their analogues. A theoretical characterization from a combined DFT and BS study. Phys. Chem. Chem. Phys. 2012, 14, 1121–1130. 10.1039/C1CP22928A. [DOI] [PubMed] [Google Scholar]
- Li Y.; Zhang S.; Wang Q.; Jena P. Structure and properties of Mn4Cl9: An antiferromagnetic binary hyperhalogen. J. Chem. Phys. 2013, 138, 054309 10.1063/1.4776217. [DOI] [PubMed] [Google Scholar]
- Yu Y.; Li C.; Yin B.; Li J. L.; Huang Y. H.; Wen Z. Y.; Jiang Z. Y. Are trinuclear superhalogens promising candidates for building blocks of novel magnetic materials? A theoretical prospect from combined broken-symmetry density functional theory and ab initio study. J. Chem. Phys. 2013, 139, 054305 10.1063/1.4817189. [DOI] [PubMed] [Google Scholar]
- Ding L. P.; Kuang X. Y.; Shao P.; Zhong M. M.; Zhao Y. R. Formation and properties of iron-based magnetic superhalogens: A theoretical study. J. Chem. Phys. 2013, 139, 104304. 10.1063/1.4819912. [DOI] [PubMed] [Google Scholar]
- Wileńska D.; Skurski P.; Anusiewicz I. High-symmetry compact structures are preferred equilibrium configurations of LinFn+1– (n = 2–5) superhalogen anions. J. Fluorine Chem. 2014, 168, 99–104. 10.1016/j.jfluchem.2014.09.015. [DOI] [Google Scholar]
- Li J. F.; Sun Y. Y.; Bai H.; Li M. M.; Li J. L.; Yin B. Are superhalogens without halogen ligand capable of transcending traditional halogen-based superhalogens? Ab initio case study of binuclear anions based on pseudohalogen ligand. AIP Adv. 2015, 5, 067143 10.1063/1.4922958. [DOI] [Google Scholar]
- Ding L.-P.; Shao P.; Lu C.; Zhang F.-H.; Liu Y.; Mu Q. Prediction of the Iron-Based Polynuclear Magnetic Superhalogens with Pseudohalogen CN as Ligands. Inorg. Chem. 2017, 56, 7928–7935. 10.1021/acs.inorgchem.7b00646. [DOI] [PubMed] [Google Scholar]
- Shi Y.; Bian S.; Ma Y.; Wang Y.; Ren J.; Kong X. Structures and Superhalogen Properties of Pt2Cln (2 ≤ n ≤ 10) Clusters. J. Phys. Chem. A 2019, 123, 187–193. 10.1021/acs.jpca.8b08349. [DOI] [PubMed] [Google Scholar]
- Srivastava A. K. A Simple Strategy to Design Polycyclic Superhalogens. J. Phys. Chem. A 2023, 127, 4867–4872. 10.1021/acs.jpca.3c02254. [DOI] [PubMed] [Google Scholar]
- Melnikov P.; Nalin M.; Messaddeq Y. Scandium fluorides. J. Alloys Compd. 1997, 262–263, 296–298. 10.1016/S0925-8388(97)00407-6. [DOI] [Google Scholar]
- Wang C.; Chen X.; Wei R.; Gong Y. Raman Spectroscopic and Theoretical Study of Scandium Fluoride and Oxyfluoride Anions in Molten FLiNaK. J. Phys. Chem. B 2020, 124, 6671–6678. 10.1021/acs.jpcb.0c06021. [DOI] [PubMed] [Google Scholar]
- De araujo F. M. F.; Duarte-ruiz D.; Saßnick H.-D.; Gentzmann M. C.; Huthwelker T.; Cocchi C. Electronic Structure and Core Spectroscopy of Scandium Fluoride Polymorphs. Inorg. Chem. 2023, 62, 4238–4247. 10.1021/acs.inorgchem.2c04357. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pradhan K.; Gutsev G. L.; Jena P. Negative ions of transition metal-halogen clusters. J. Chem. Phys. 2010, 133, 144301 10.1063/1.3489117. [DOI] [PubMed] [Google Scholar]
- Sergeeva A. P.; Averkiev B. B.; Zhai H. J.; Boldyrev A. I.; Wang L. S. All-boron analogues of aromatic hydrocarbons: B17– and B18–. J. Chem. Phys. 2011, 134, 224304. 10.1063/1.3599452. [DOI] [PubMed] [Google Scholar]
- The source code of the CK software can be obtained at https://github.com/averkiev75/Coalescence-Kick.
- 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 Condens. Matter. 1988, 37, 785–789. 10.1103/PhysRevB.37.785. [DOI] [PubMed] [Google Scholar]
- Becke A. D. Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A 1988, 38, 3098–3100. 10.1103/PhysRevA.38.3098. [DOI] [PubMed] [Google Scholar]
- Hay P. J.; Wadt W. R. Ab Initio Effective Core Potentials for Molecular Calculations. Potentials for the Transition Metal Atoms Sc to Hg. J. Chem. Phys. 1985, 82, 270–283. 10.1063/1.448799. [DOI] [Google Scholar]
- Wadt W. R.; Hay P. J. Ab Initio Effective Core Potentials for Molecular Calculations. Potentials for Main Group Elements Na to Bi. J. Chem. Phys. 1985, 82, 284–298. 10.1063/1.448800. [DOI] [Google Scholar]
- Hay P. J.; Wadt W. R. Ab Initio Effective Core Potentials for Molecular Calculations. Potentials for K to Au Including the Outermost Core Orbitals. J. Chem. Phys. 1985, 82, 299–310. 10.1063/1.448975. [DOI] [Google Scholar]
- Chai J.-D.; Head-gordon M. Long-range corrected hybrid density functionals with damped atom-atom dispersion corrections. Phys. Chem. Chem. Phys. 2008, 10, 6615–6620. 10.1039/b810189b. [DOI] [PubMed] [Google Scholar]
- Dunning T. H. Jr. Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen. J. Chem. Phys. 1989, 90, 1007–1023. 10.1063/1.456153. [DOI] [Google Scholar]
- Kendall R. A.; Dunning T. H.; Harrison R. J. Electron affinities of the first-row atoms revisited. Systematic basis sets and wave functions. J. Chem. Phys. 1992, 96, 6796–6806. 10.1063/1.462569. [DOI] [Google Scholar]
- Dolg M.; Wedig U.; Stoll H.; Preuss H. Energy-adjusted ab initio pseudopotentials for the first row transition elements. J. Chem. Phys. 1987, 86, 866–872. 10.1063/1.452288. [DOI] [Google Scholar]
- Dolg M.; Stoll H.; Savin A.; Preuss H. Energy-adjusted pseudopotentials for the rare earth elements. Theor. Chim. Acta 1989, 75, 173–194. 10.1007/BF00528565. [DOI] [Google Scholar]
- Andrae D.; Häussermann U.; Dolg M.; Stoll H.; Preuss H. Energy-adjusted ab initio pseudopotentials for the 2nd and 3rd row transition-elements. Theor. Chim. Acta 1990, 77, 123–141. 10.1007/BF01114537. [DOI] [Google Scholar]
- Mo̷ller C.; Plesset M. S. Note on an Approximation Treatment for Many-Electron Systems. Phys. Rev. 1934, 46, 618–622. 10.1103/PhysRev.46.618. [DOI] [Google Scholar]
- Head-gordon M.; Pople J. A.; Frisch M. J. MP2 energy evaluation by direct methods. Chem. Phys. Lett. 1988, 153, 503–506. 10.1016/0009-2614(88)85250-3. [DOI] [Google Scholar]
- Frisch M. J.; Head-gordon M.; Pople J. A. A direct MP2 gradient method. Chem. Phys. Lett. 1990, 166, 275–280. 10.1016/0009-2614(90)80029-D. [DOI] [Google Scholar]
- Koga T.; Tatewaki H.; Matsuyama H.; Satoh Y. Contracted Gaussian-type basis functions revisited. III. Atoms K through Kr. Theor. Chem. Acc. 1999, 102, 105–111. 10.1007/s002140050479. [DOI] [Google Scholar]
- Noro T.; Sekiya M.; Koga T.; Matsuyama H. Valence and correlated basis sets for the first-row transition atoms from Sc to Zn. Theor. Chem. Acc. 2000, 104, 146–152. 10.1007/s002140000126. [DOI] [Google Scholar]
- Tatewaki H.; Koga T. Contracted Gaussian-type basis functions revisited. J. Chem. Phys. 1996, 104, 8493–8499. 10.1063/1.471599. [DOI] [Google Scholar]
- Noro T.; Sekiya M.; Koga T. Contracted polarization functions for the atoms helium through neon. Theor. Chem. Acc. 1997, 98, 25–32. 10.1007/s002140050274. [DOI] [Google Scholar]
- Bartlett R. J.; Purvis G. D. III Many-body perturbation theory, coupled-pair many-electron theory, and the importance of quadruple excitations for the correlation problem. Int. J. Quantum Chem. 1978, 14, 561–581. 10.1002/qua.560140504. [DOI] [Google Scholar]
- Purvis G. D. III; Bartlett R. J. A full coupled-cluster singles and doubles model: The inclusion of disconnected triples. J. Chem. Phys. 1982, 76, 1910–1918. 10.1063/1.443164. [DOI] [Google Scholar]
- Scuseria G. E.; Janssen C. L.; Schaefer H. F. III An efficient reformulation of the closed-shell coupled cluster single and double excitation (CCSD) equations. J. Chem. Phys. 1988, 89, 7382–7387. 10.1063/1.455269. [DOI] [Google Scholar]
- Zakrzewski V. G.; Ortiz J. V.; Nichols J. A.; Heryadi D.; Yeager D. L.; Golab J. T. Comparison of perturbative and multiconfigurational electron propagator methods. Int. J. Quantum Chem. 1996, 60, 29–36. . [DOI] [Google Scholar]
- Simons J. Direct Calculation of First- and Second-Order Density Matrices. The Higher RPA Method. J. Chem. Phys. 1971, 55, 1218–1230. 10.1063/1.1676208. [DOI] [Google Scholar]
- Ortiz J. V. Electron binding energies of anionic alkali metal atoms from partial fourth order electron propagator theory calculations. J. Chem. Phys. 1988, 89, 6348–6352. 10.1063/1.455401. [DOI] [Google Scholar]
- Rowe D. J. Equations-of-Motion Method and the Extended Shell Model. Rev. Mod. Phys. 1968, 40, 153–166. 10.1103/RevModPhys.40.153. [DOI] [Google Scholar]
- Cederbaum L. S. One-body Green’s function for atoms and molecules: theory and application. J. Phys. B: At. Mol. Phys. 1975, 8, 290–303. 10.1088/0022-3700/8/2/018. [DOI] [Google Scholar]
- Simons J. Energy-Shift Theory of Low-Lying Excited Electronic States of Molecules. J. Chem. Phys. 1972, 57, 3787–3792. 10.1063/1.1678845. [DOI] [Google Scholar]
- Simons J.; Smith W. D. Theory of electron affinities of small molecules. J. Chem. Phys. 1973, 58, 4899–4907. 10.1063/1.1679074. [DOI] [Google Scholar]
- Zakrzewski V. G.; Ortiz J. V. Semidirect algorithms for third-order electron propagator calculations. Int. J. Quantum Chem. 1995, 53, 583–590. 10.1002/qua.560530602. [DOI] [Google Scholar]
- Zakrzewski V. G.; Ortiz J. V. Semidirect algorithms in electron propagator calculations. Int. J. Quantum Chem. 1994, 28, 23–27. 10.1002/qua.560520806. [DOI] [Google Scholar]
- Zakrzewski V. G.; Dolgounitcheva O.; Ortiz J. V. Ionization energies of anthracene, phenanthrene, and naphthacene. J. Chem. Phys. 1996, 105, 8748–8753. 10.1063/1.472654. [DOI] [Google Scholar]
- Frisch M. J.; Trucks G. W.; Schlegel H. B.; Scuseria G. E.; Robb M. A.; Cheeseman J. R.; Scalmani G.; Barone V.; Petersson G. A.; Nakatsuji H.; Li X.; Caricato M.; Marenich A. V.; Bloino J.; Janesko B. G.; Gomperts R.; Mennucci B.; Hratchian H. P.; Ortiz J. V.; Izmaylov A. F.; Sonnenberg J. L.; Williams-young D.; Ding F.; Lipparini F.; Egidi F.; Goings J.; Peng B.; Petrone A.; Henderson T.; Ranasinghe D.; Zakrzewski V. G.; Gao J.; Rega N.; Zheng G.; Liang W.; Hada M.; Ehara M.; Toyota K.; Fukuda R.; Hasegawa J.; Ishida M.; Nakajima T.; Honda Y.; Kitao O.; Nakai H.; Vreven T.; Throssell K.; Montgomery J. A. Jr.; Peralta J. E.; Ogliaro F.; Bearpark M. J.; Heyd J. J.; Brothers E. N.; Kudin K. N.; Staroverov V. N.; Keith T. A.; Kobayashi R.; Normand J.; Raghavachari K.; Rendell A. P.; Burant J. C.; Iyengar S. S.; Tomasi J.; Cossi M.; Millam J. M.; Klene M.; Adamo C.; Cammi R.; Ochterski J. W.; Martin R. L.; Morokuma K.; Farkas O.; Foresman J. B.; Fox D. J.. Gaussian 16, Revision C.01; Gaussian, Inc.: Wallingford CT, 2016. [Google Scholar]
- Gutsev G. L.; Boldyrev A. I. The Electronic Structure of Superhalogens and Superalkalies. Russ. Chem. Rev. 1987, 56, 519–531. 10.1070/RC1987v056n06ABEH003287. [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
