Is pentavalent Pr(V) feasible in solid CsPrF6?

Lian-Wei Ye a, Zi-He Zhang a, Yang He a, Shi-Ru Wei a, Jun-Bo Lu b, Han-Shi Hu *a and Jun Li *ab
aDepartment of Chemistry and Engineering Research Center of Advanced Rare-Earth Materials of Ministry of Education, Tsinghua University, Beijing 100084, China. E-mail: hshu@mail.tsinghua.edu.cn; junli@tsinghua.edu.cn
bFundamental Science Center of Rare Earths, Ganjiang Innovation Academy, Chinese Academy of Sciences, Ganzhou 341000, China

Received 17th July 2024 , Accepted 19th August 2024

First published on 24th August 2024


Abstract

The oxidation state (OS) holds significant importance in the field of chemistry and serves as a crucial parameter for tracking electrons. Lanthanide (Ln) elements predominately exhibit a +III oxidation state, with a few elements such as Ce, Pr, Nd, Tb, and Dy able to achieve a +IV oxidation state. Over the past century, numerous attempts to synthesize Pr(V) have been made without success until recent reports on Pr(V) oxides and nitride-oxide in the gas phase expanded our understanding of Ln elements. However, the formation of Pr(V) in the condensed phase remains an open question. In this work, based on advanced quantum chemical investigations, we predict that formation of the solid-state CsPrVF6 from Pr(III) and Pr(IV) complexes is exothermic, indicating that CsPrVF6 is stable. The crystal structure comprises [PrF6] octahedral clusters occupying the interstitial spaces of Cs cations. Electronic structure analysis reveals that the CsPrF6 crystal has a closed-shell structure and that Pr reaches its highest oxidation state of +V. The results indicate that the existence of Pr(V) in solid-state Ln fluorides is not impossible, which enriches our understanding of high-valence Ln compounds.


Introduction

The concept of oxidation state (OS) stands out as one of the fundamental principles in chemistry.1–4 Exploring the upper limits of an element's OS has long captured the interest of chemists.5–12 Investigating compounds with high OS deepens our comprehension of chemical elements and paves the way for developing novel oxidants,13,14 catalysts,15,16 and other chemical entities.17,18 In recent decades, the study of Ln chemistry has attracted increased attention, driven by advancements in experimental technologies that facilitate the synthesis of high OS complexes.6,19,20 Ln compounds are predominantly characterized by the prevalence of the +III OS, primarily due to the inert nature of 4f orbitals resulting from contracted radial distribution and the relatively low energy.21–23 Only a handful of elements exhibit a high-valent +IV, such as Ce, Pr, Nd, Tb, and Dy.24–30

Regarding Ln elements with a higher +V OS, Langmuir predicted as early as 1919 that it could appear in the element Pr, which has the fifth lowest ionization energy.31 However, the inertness of the 4f orbitals in Ln elements makes achieving a +V OS in compounds a challenging task. The report in 1938 suggested the possibility of Pr(V) formation at 15 bar pressure.32 However, subsequent research in 1950 conclusively demonstrated that Pr(IV) is the highest OS formed under a higher pressure of 50 bar of oxygen. This was observed when mixtures of Pr2O3–Nd2O3 and Pr2O3–CeO2 were heated at 300 °C.33,34 In 2015, efforts using the highly electronegative element fluorine aimed to generate PrF5 molecules through the laser ablation method. However, conclusive evidence of their formation has not been obtained.25 Up until 2016, both experimental and theoretical studies have unveiled the existence of Pr(V) in the gas phase, exemplified by oxide species PrO4 and PrO2+, and nitride-oxides NPrO.19,20 Theoretical research continues to explore potential gas phase Pr(V) molecules,35,36 while the next objective is to propose an alternative strategy for expanding the Ln OS beyond the gas phase (Scheme 1).


image file: d4dt02063d-s1.tif
Scheme 1 The highest OS of the 4f-block Ln elements. () Highest experimentally known species, (•) the Pr(V) molecule only identified in gas phase.

In pursuit of solid-state Pr(V), we confined octahedral [PrF6] clusters within the Cs+ lattice to achieve CsPrF6 (Scheme 2). A global-minimum search algorithm based on first-principles plane-wave density functional theory (DFT) was employed to elucidate the crystal structure. A comprehensive analysis was conducted, covering stability, electronic structure, chemical bonding, and notably, oxidation state. This discovery provides concrete evidence for the first instance of Pr(V) within a solid state.


image file: d4dt02063d-s2.tif
Scheme 2 A schematic structure of the [PrF6] octahedral cluster.

Results and discussion

Structure and stability of CsPrF6

To find the stable structure of CsPrF6, the evolutionary algorithm USPEX (Universal Structure Predictor: Evolutionary Xtallography) was used, which has been successfully utilized to search for the most thermodynamically stable structures of systems with a fixed composition. The trial structures generated by USPEX were fully relaxed and their final energies evaluated based on first-principles density functional theory (DFT). The density functional computations were carried out using the Vienna ab initio simulation package (VASP). A total of 246 trial structures were calculated, and Fig. 1 shows the most stable predicted CsPrF6 structure.
image file: d4dt02063d-f1.tif
Fig. 1 Prediction of the crystal structure of CsPrF6. Left: the progress of an evolutionary simulation (black line shows the lowest energy as a function of generation, light blue to dark purple circles present different generations). Right: the most stable R[3 with combining macron] CsPrF6 structure (colour scheme: blue: Cs; yellow: Pr; green: F).

The CsPrF6 structure shown in Fig. 1 belongs to the rhombohedral R[3 with combining macron] phase, which is a KAsF6-type structure and is also observed in KPF6 and CsNbF6 crystals. Detailed lattice parameters are provided in Table S1, and the atomic coordinates are given in the ESI, section VI. The distance between Pr and F atoms is 2.079 Å, representing a moderate bond length compared to the Pr–F distances ranging from 1.994 to 2.109 Å in PrFn(n = 1–5) molecules.25 This observation indicates the formation of [PrF6] clusters within the interstitial spaces of Cs cations. The Madelung energy in this arrangement is calculated to be −145.3 eV for the CsPrF6 primitive cell, stabilizing the [PrF6] cluster and enabling the existence of the pentavalent praseodymium anion cluster [PrF6] in the CsPrF6 crystal. This provides a further explanation for the presence of [PrF6]. The optimal CsPrF6 structure from global-minimum searching has been further confirmed to be stable in dynamics using phonon spectrum calculations along the high-symmetry lines in the Brillouin zone, as shown in Fig. S2. There are no obvious imaginary vibrational frequencies in the phonon spectrum, suggesting the dynamic stability of the structure.

To demonstrate the thermodynamic stability of the CsPrF6 crystal, a possible synthetic pathway is depicted in Table 1, eqn (1) and (2). The Gibbs free energy of formation (ΔG300 K) at a temperature of 300 K was calculated using the formula ΔG300 K = G(CsPrF6) − G(PrF3) − G(F2) or ΔG300 K = G(CsPrF6) + G(CsF) − G(Cs2PrF6) − 1/2 G(F2), respectively. Here, G represents the energy with entropy and zero-point energy correction for different species. Importantly, all calculated ΔG300 K values are negative, signifying that the formation of the CsPrF6 solid is energetically favourable under these conditions. Notably, the reactants F2, CsF37 and PrF3[thin space (1/6-em)]38 mentioned in eqn (1) are naturally occurring, with PrF3 being a common raw material for synthesizing Pr fluorides in the laboratory,39 which outlines plausible methods for synthesizing CsPrF6. Comprehensive thermodynamic free energy data for the reactants in the equations can be found in Table S2.

Table 1 Plausible methods for synthesizing CsPrF6 and its reaction free energy (kcal mol−1)
Reaction ΔG300 K  
PrF3(s) + F2(g) + CsF(s) → CsPrF6(s) −74.1 (1)
Cs2PrF6(s) + ½F2(g) → CsPrF6(s) + CsF(s) −133.2 (2)


To further examine the thermal stability of the predicted CsPrF6, a series of AIMD simulations was carried out at various temperatures of 300 K, 600 K, 900 K and 1200 K for a duration of 10 ps, respectively. Since the distances between Pr and other atoms reflect the structural integrity, the radial distribution function (RDF) of the neighbouring Pr–F and Pr–Cs distances is a good parameter for quantitatively evaluating the structural deformation of CsPrF6. The RDFs of AIMD simulations at different temperatures are shown in Fig. 2, with a sharp Pr–F peak around 2.1 Å and clear Pr–Cs peak around 4.0 and 4.6 Å at 300 K after 10 ps simulations, indicating stability at room temperature. As expected, at a relatively high temperature of 600 K, the peaks show a little more dispersion, and much more disorder appears at 900 K. Once the temperature reaches 1200 K, the Pr–Cs peak becomes indistinguishable and the Pr–F peak disperses even further, indicating that the structure is not stable at this high temperature.


image file: d4dt02063d-f2.tif
Fig. 2 The radial distribution function (RDF) of Pr and other atoms such as F and Cs respectively at different temperatures for spin-polarized calculations.

Chemical bonding and oxidation state analysis

The stability of the CsPrF6 crystal relies on its electronic structure. A detailed examination of the electronic structure was conducted, using the band structure, density of states (DOS), spin density, crystal orbital Hamiltonian population (COHP), and Bader charge analysis. Furthermore, the electronic structural properties of CsPrF6 were compared with those of Pr-containing complexes featuring different OS, specifically CsPrIIF3,40 Cs3PIIIrF6,41 and Cs2PrIVF6.41 In the subsequent analyses, we conducted a comprehensive comparison of the properties of Pr across formal OS +II to +V compounds, delving into electron occupancy, orbital interactions, and chemical bonding in various Pr complexes

The band structure and related properties of CsPrF6 are presented in Fig. 3. A sizable band gap appears in the CsPrF6 crystal between the filled and empty states, of approximately 3.5 eV at the HSE06 hybrid functional level, indicating that it is a potential semiconductor material. The band structures of different Pr-containing complexes were calculated along the high-symmetry directions in the Brillouin zone (BZ) as illustrated in Fig. S11–13.


image file: d4dt02063d-f3.tif
Fig. 3 The HSE06 band and partial density of states (pDOS) of the CsPrVF6 crystal. The Fermi level is assigned at 0 eV.

The interactions between different orbitals can be illustrated through the DOS analysis shown in Fig. 3. The inner 4f orbitals are a distinctive characteristic of Ln elements, deemed “core-like” due to their minimal participation in chemical bonding, as evident from the sharp distribution in DOS. Until Pr reaches its highest OS, unpaired electrons occupy the 4f orbitals below the Fermi level (Fig. S14–16). However, in the CsPrF6 compound, the majority of Pr 4f orbitals extend beyond the Fermi level (approximately 4 eV), indicating significant hybridization with the F 2p orbitals (Fig. 3). This implies that the 4f orbitals of Pr are unfilled, contributing to Pr attaining its highest +V OS. The results are attributed to the higher energy of Pr 4f orbitals compared to F 2p in CsPrF6, facilitating electron transfer from Pr to F. The Pr 5d orbitals exhibit a greater dispersion than 4f orbitals, allowing for enhanced interaction with the F element. In the CsPrF6 compound, the bonding and antibonding orbitals of 5d show significant separation. The 5d antibonding orbitals of Pr are located approximately 11 eV above the Fermi level, while the bonding orbitals are distributed in the energy range of −4 to −1 eV. This dispersed band indicates a non-negligible interaction with F 2p (Fig. 3). Similar trends are observed in other Pr compounds (Fig. S14–16), indicating a prevalent utilization of Pr 5d orbitals for bonding. Further quantitative analysis of the COHP to emphasize the importance of the 5d orbitals in the bonding mechanism of CsPrF6 can be found in the next section.

Due to the inherently inner and unpaired nature of 4f electrons, the analysis of spin density serves as another powerful tool to illustrate their localization. The spin densities of different Pr-containing compounds are depicted in Fig. S17–20. Unpaired electrons are not observed in CsPrVF6, whereas they are prominently displayed in CsPrIIF3, Cs3PIIIrF6, and Cs2PrIVF6. This suggests that the inner 4f electrons of Pr in CsPrF6 have been removed, indicating a +V OS.

To elucidate the bonding mechanism in the CsPrF6 crystal, COHP analysis was applied to evaluate the population of wavefunctions on the atomic orbitals’ overlap of selected atom pairs. The integrated COHP (ICOHP) for the bonded Pr–F in CsPrF6 is −5.94 eV per pair up to the Fermi level, indicating some covalent interaction between the Pr and F atoms, as shown in Fig. 4. Specifically, the covalent interaction between Pr 5d and F 2p is stronger than that of Pr 4f, where the 4f orbitals of Pr play an insignificant role in chemical bonding due to their inner nature. The COHP analysis for other compounds with different OS of Pr is provided in Fig. S21–23. For example, the ICOHP value for the Pr–F bond is −2.58 eV for CsPrF3, −2.09 eV for Cs3PrF6, and −5.04 eV for Cs2PrF6. The CsPrF6 crystal exhibits a moderate ICOHP value among them.


image file: d4dt02063d-f4.tif
Fig. 4 The crystal orbital Hamilton population (COHP) analysis of Pr–F interactions in solid-state CsPrF6. Zero line (dotted) represents the Fermi level. The ICOHP values (in eV per bond) are listed here to show the corresponding interactions (black) and main orbital-pair contributions to these (coloured). There are six Pr–F bonds in a CsPrF6 cell.

Usually, the formal OS of a central atom in a coordination sphere is defined as the charge of the central atom when every ligand of the coordination sphere is removed in its most stable form. The bonding electron pairs between the metal center and the ligand are therefore exclusively assigned to the more electronegative fragment, resulting in a negative OS.6 Therefore, the OS of fluorine and cesium are considered as −I and +I, respectively. According to these rules, the oxidation number of praseodymium in the CsPrF6 solid is +V. To further support this, we calculated the charges using Bader's quantum theory of atoms in molecules (QTAIM) for the CsPrF6 solid.42–44 In QTAIM analysis, the charges of an atom are calculated by integrating the charge density in the surrounding basin, partitioned by the stationary points of the charge density.

In Fig. 5, various Pr-containing complexes are presented. The Bader charges for the central Pr atom are 1.97, 2.20, 2.37, and 2.66 for CsPrIIF3, Cs3PrIIIF6, Cs2PrIVF6, and CsPrVF6, respectively. The CsPrF6 solid exhibits the highest Bader charge, which is consistent with the formal OS of Pr in the CsPrF6 solid which is +V as mentioned above. It is common for the Bader charge to increase as the OS increases in similar compounds. This trend is evident in Bader charges of 1.35, 1.86, 2.28, and 2.66 for the standard compounds, PrIIF2, PrIIIF3, PrIVF4, and CsPrVF6, respectively. Additionally, the bond lengths between Pr and F are reported to be 2.411, 2.275, 2.141, and 2.097 Å for the respective complexes. It is observed that as the OS increases, the bond length decreases, and the charge on the praseodymium atom becomes more positive. These findings further confirm that CsPrF6 exists with the Pr(V) formal OS.


image file: d4dt02063d-f5.tif
Fig. 5 The calculated Pr Bader charge and Pr–F bond lengths in CsPrIIF3, Cs3PrIIIF6, Cs2PrIVF6, CsPrVF6 crystals and PrFn (n = 2–4) compounds.

Conclusions

In summary, an unexpected solid-state CsPrF6 with Pr(V) was designed and predicted using a global-minimum search method. The crystal structure features a hexahedral [PrVF6] anion cluster embedded within the Cs+ lattice. Cs+ contributes to the Madelung energy, enhancing the stability of the Pr(V) cluster. Gibbs free energy calculations confirm the thermodynamic stability of the crystal, suggesting its potential for experimental synthesis. Phonon calculations provide evidence of its dynamic stability, while AIMD simulations conducted at different temperatures indicate that the crystal remains stable up to 900 K. As anticipated, Bader charge analysis indicates that Pr in CsPrF6 exhibits a +V OS, confirmed by comparing it with different OS of Pr-containing crystals. This finding is further supported by its closed shell electronic structure, where nearly all of the 4f bands are situated above the Fermi level, and no spin density is located at the Pr atom. The prediction of the CsPrF6 crystal suggests that the highest known OS of Ln(V) is achievable in the solid state and proposes a stabilizing strategy for obtaining high-valent fluorides.

Computational details

The structure of complexes CsF (mp-1784),37 PrF3 (ICSD-77741),38 CsPrF3 (OQMD-1706464),40 Cs3PrF6 (mp-1206084),41 and Cs2PrF6 (OQMD-1282041)41 were obtained from different databases. The lowest-enthalpy CsPrF6 structure was searched using the evolutionary algorithm USPEX 9.4.4.45–48 The calculations were carried out for the best structure that no longer changed for 8 generations and each generation of structures was obtained by applying heredity (50%), softmutation (20%), and lattice mutation (10%) operators, while some were produced randomly from the space group (20%).

The geometry of these structures was further optimized based on first-principles density functional theory (DFT) electronic-structure calculations involving the projector-augmented wave (PAW) method49 as implemented in the Vienna ab initio simulation package (VASP)50 version 5.4.4. The Perdew–Burke–Ernzerhof (PBE) functional51 was utilized for both structural search and structural relaxation. The energy cutoff for the plane-wave basis was set to 450 eV. The Brillouin zone was sampled using uniform Γ-centered meshes with a resolution of 2π × 0.04 Å−1. Electronic degrees of freedom were considered converged when the total energy change between two steps was smaller than 10−7 eV. All atoms were fully relaxed until the Hellmann–Feynman forces fell below 0.01 eV Å−1. The Madelung energy of the optimized CsPrF6 primitive structure was calculated using VESTA software.52 To explore the thermal stability, ab initio molecular (AIMD) simulations were carried out using VASP with the canonical (NVT) ensemble and a Nosé–Hoover thermostat.53,54 The simulations lasted for 10 ps at temperatures of 300, 600, 900, and 1200 K, with a time step of 1 fs. To assess the dynamic stability, the finite displacement method was used for phonon dispersion calculations within the Phonopy program.55 For the chemical bonding analysis, we utilized the COHP method56,57 as calculated using the LOBSTER package.58,59 During the orbital projection, the basis sets pbeVasp201559 were used with additional functions fitted to atomic VASP GGA-PBE wave functions. For charge population analysis, Bader's quantum theory of atoms in molecules (QTAIM) was used for the solids.42–44

In order to account for the correlation effects involving 4f electrons and obtain a more accurate electronic structure, the Heyd–Scuseria–Ernzerhof (HSE06) hybrid functional was applied to obtain more accurate band structures and density of states (DOS). The Brillouin zone was sampled using uniform Γ-centered meshes with a resolution of 2π × 0.04 Å−1. For the band structure calculations, 78, 65, 50, and 56 k-points were used for the CsPrF3, Cs3PrF6, Cs2PrF6, and CsPrF6 crystal structures, respectively. The self-consistent field convergence criterion was set to 10−5 eV.60

Author contributions

J. L. designed the project. L. W. Y. performed the computational calculations. L. W. Y., Z. H. Z., Y. H., S. R. W., J. B. L., H. S. H. and J. L. wrote and edited the manuscript. All authors helped to analyse the theoretical results.

Data availability

The data supporting this article have been included in the ESI. All data from DFT calculations involved in this work can be found in the article and ESI.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work was supported by the National Key Research and Development Project (No. 2022YFA1503900, 2022YFA1503000), the National Natural Science Foundation of China (NSFC No. 22033005, 22222605 and 22076095), and the Guangdong Provincial Key Laboratory of Catalysis (No. 2020B121201002). The calculations were performed using supercomputers at the High Performance Computing Center of Tsinghua University. The Tsinghua Xuetang Talents Program is acknowledged for providing computational resources.

References

  1. L. J. Malone and T. Dolter, Basic concepts of chemistry, John Wiley & Sons, 2008 Search PubMed .
  2. F. A. Cotton, G. Wilkinson, C. A. Murillo and M. Bochmann, Advanced inorganic chemistry, John Wiley and Sons, Inc., 1999 Search PubMed .
  3. N. N. Greenwood and A. Earnshaw, Chemistry of the Elements, Elsevier, 2012 Search PubMed .
  4. P. Karen, P. McArdle and J. Takats, Comprehensive definition of oxidation state (IUPAC Recommendations 2016), Pure Appl. Chem., 2016, 88, 831–839 CrossRef CAS .
  5. S. N. Achary and A. K. Tyagi, in Handbook on Synthesis Strategies for Advanced Materials: Volume-III: Materials Specific Synthesis Strategies, ed. A. K. Tyagi and R. S. Ningthoujam, SpringerSingapore, Singapore, 2021, pp. 137–158 Search PubMed .
  6. G. Wang, M. Zhou, J. T. Goettel, G. J. Schrobilgen, J. Su, J. Li, T. Schlöder and S. Riedel, Identification of an iridium-containing compound with a formal oxidation state of IX, Nature, 2014, 514, 475–477 CrossRef CAS PubMed .
  7. P. Pyykkö and W.-H. Xu, The formal oxidation states of iridium now run from −III to +IX, Angew. Chem., Int. Ed., 2015, 54, 1080–1081 CrossRef PubMed .
  8. Z.-L. Xue, [IrO4]+ at the IX (+9) oxidation state observed, report states, Sci. China: Chem., 2015, 58, 4–5 CrossRef CAS .
  9. P. Pyykkö and W.-H. Xu, On the extreme oxidation states of iridium, Chem. – Eur. J., 2015, 21, 9468–9473 CrossRef PubMed .
  10. S.-X. Hu, W.-L. Li, J.-B. Lu, J. L. Bao, H. S. Yu, D. G. Truhlar, J. K. Gibson, J. Marçalo, M. Zhou, S. Riedel, W. H. E. Schwarz and J. Li, On the upper limits of oxidation states in chemistry, Angew. Chem., Int. Ed., 2018, 57, 3242–3245 CrossRef CAS .
  11. P. Pyykkö, N. Runeberg, M. Straka and K. G. Dyall, Could uranium(XII)hexoxide, UO6 (Oh) exist?, Chem. Phys. Lett., 2000, 328, 415–419 CrossRef .
  12. H. Xiao, H.-S. Hu, W. H. E. Schwarz and J. Li, Theoretical investigations of geometry, electronic structure and stability of UO6: Octahedral uranium hexoxide and its isomers, J. Phys. Chem. A, 2010, 114, 8837–8844 CrossRef CAS PubMed .
  13. J. H. Canterford and T. A. O'Donnell, Reactivity of transition metal fluorides. IV. Oxidation-reduction reactions of vanadium pentafluoride, Inorg. Chem., 1967, 6, 541–544 CrossRef CAS .
  14. B. C. Bales, P. Brown, A. Dehestani and J. M. Mayer, Alkane oxidation by osmium tetroxide, J. Am. Chem. Soc., 2005, 127, 2832–2833 CrossRef CAS PubMed .
  15. N. Bartlett, The oxidizing properties of the third transition series hexafluorides and related compounds, Angew. Chem., Int. Ed. Engl., 1968, 7, 433–439 CrossRef CAS .
  16. B. L. Pagenkopf and E. M. Carreira, Transition metal fluoride complexes in asymmetric catalysis, Chem. – Eur. J., 1999, 5, 3437–3442 CrossRef CAS .
  17. W. W. Dukat, J. H. Holloway, E. G. Hope, M. R. Rieland, P. J. Townson and R. L. Powell, High oxidation state binary transition metal fluorides as selective fluorinating agents, J. Chem. Soc., Chem. Commun., 1993, 1993, 1429–1430 RSC .
  18. J. H. Holloway, E. G. Hope, P. J. Townson and R. L. Powell, Selective fluorination of dichloromethane by highest oxidation state transition-metal oxide fluorides, J. Fluor. Chem., 1996, 76, 105–107 CrossRef CAS .
  19. Q. Zhang, S.-X. Hu, H. Qu, J. Su, G. Wang, J.-B. Lu, M. Chen, M. Zhou and J. Li, Pentavalent lanthanide compounds: formation and characterization of praseodymium(V) oxides, Angew. Chem., Int. Ed., 2016, 55, 6896–6900 CrossRef CAS PubMed .
  20. S.-X. Hu, J. Jian, J. Su, X. Wu, J. Li and M. Zhou, Pentavalent lanthanide nitride-oxides: NPrO and NPrO complexes with NPr triple bonds, Chem. Sci., 2017, 8, 4035–4043 RSC .
  21. P. B. Hitchcock, M. F. Lappert, L. Maron and A. V. Protchenko, Lanthanum does form stable molecular compounds in the +2 oxidation state, Angew. Chem., Int. Ed., 2008, 47, 1488–1491 CrossRef CAS PubMed .
  22. M. R. MacDonald, J. E. Bates, J. W. Ziller, F. Furche and W. J. Evans, Completing the series of +2 ions for the lanthanide elements: Synthesis of molecular complexes of Pr2+, Gd2+, Tb2+, and Lu2+, J. Am. Chem. Soc., 2013, 135, 9857–9868 CrossRef CAS PubMed .
  23. W.-L. Li, C. Ertural, D. Bogdanovski, J. Li and R. Dronskowski, Chemical bonding of crystalline LnB6 (Ln = La–Lu) and its relationship with Ln2B8 gas-phase complexes, Inorg. Chem., 2018, 57, 12999–13008 CrossRef CAS PubMed .
  24. A. Schulz and J. F. Liebman, Paradoxes and paradigms: High oxidation states and neighboring rows in the periodic table—Lanthanides, actinides, exotica and explosives, Struct. Chem., 2008, 19, 633–635 CrossRef CAS .
  25. T. Vent-Schmidt and S. Riedel, Investigation of praseodymium fluorides: A combined matrix-isolation and quantum-chemical study, Inorg. Chem., 2015, 54, 11114–11120 CrossRef CAS PubMed .
  26. A. F. Lucena, C. Lourenço, M. C. Michelini, P. X. Rutkowski, J. M. Carretas, N. Zorz, L. Berthon, A. Dias, M. Conceição Oliveira, J. K. Gibson and J. Marçalo, Synthesis and hydrolysis of gas-phase lanthanide and actinide oxide nitrate complexes: A correspondence to trivalent metal ion redox potentials and ionization energies, Phys. Chem. Chem. Phys., 2015, 17, 9942–9950 RSC .
  27. S. Gabelnick, G. Reedy and M. Chasanov, Infrared spectra and structure of some matrix–isolated lanthanide and actinide oxides, J. Chem. Phys., 1974, 60, 1167–1171 CrossRef CAS .
  28. Z. Mazej, Room temperature syntheses of lanthanoid tetrafluorides (LnF4, Ln = Ce, Pr, Tb), J. Fluor. Chem., 2002, 118, 127–129 CrossRef CAS .
  29. V. I. Spitsyn, Y. M. Kiselev, L. I. Martynenko, V. N. Prusakov and V. B. Sokolov, Synthesis of praseodymium tetrafluoride, Dokl. Akad. Nauk SSSR (USSR), 1974, 219, 621–624 CAS .
  30. T. Vent-Schmidt, Z. Fang, Z. Lee, D. Dixon and S. Riedel, Extending the row of lanthanide tetrafluorides: A combined matrix-isolation and quantum-chemical study, Chem. – Eur. J., 2016, 22, 2406–2416 CrossRef CAS PubMed .
  31. I. Langmuir, The arrangement of electrons in atoms and molecules, J. Am. Chem. Soc., 1919, 41, 868–934 CrossRef CAS .
  32. W. Prandtl and G. Rieder, Über die Wertigkeit des Praseodyms und des Terbiums, Z. Anorg. Allg. Chem., 1938, 238, 225–235 CrossRef CAS .
  33. J. D. McCullough, An X-ray study of the rare-earth oxide systems: CeIV—NdIII, CrIV—PrIII, CeIV—PrIV and PrIV—NdIII, J. Am. Chem. Soc., 1950, 72, 1386–1390 CrossRef CAS .
  34. J. Kleinberg, Unfamiliar oxidation states: Recent progress, J. Chem. Educ., 1952, 29, 324 CrossRef CAS .
  35. W.-J. Zhang, G.-J. Wang, P. Zhang, W. Zou and S.-X. Hu, The decisive role of 4f-covalency in the structural direction and oxidation state of XPrO compounds (X: group 13 to 17 elements), Phys. Chem. Chem. Phys., 2020, 22, 27746–27756 RSC .
  36. P. S. Piotr, G. Szkudlarek and W. Grochala, Limits of stability for compounds of pentavalent praseodymium, arXiv, 2023,  DOI:10.48550/arXiv.2306.13939.
  37. E. Posnjak and R. W. G. Wyckoff, The crystal structures of the alkali halides. II, J. Wash. Acad. Sci., 1922, 12, 248–251 CAS .
  38. O. Greis, R. Ziel, B. Breidenstein, A. Haase and T. Petzel, Structural data of the tysonite-type superstructure modification β-PrF3 from X-ray powder and electron single-crystal diffraction, Powder Diffr., 1995, 10, 44–46 CrossRef CAS .
  39. R. Hoppe and W. Liebe, Komplexe Fluoride des vierwertigen Praseodyms, Z. Anorg. Allg. Chem., 1961, 313, 221–227 CrossRef CAS .
  40. J. E. Saal, S. Kirklin, M. Aykol, B. Meredig and C. Wolverton, Materials design and discovery with high-throughput density functional theory: The Open Quantum Materials Database (OQMD), JOM, 2013, 65, 1501–1509 CrossRef CAS .
  41. A. Jain, S. P. Ong, G. Hautier, W. Chen, W. D. Richards, S. Dacek, S. Cholia, D. Gunter, D. Skinner, G. Ceder and K. A. Persson, Commentary: The Materials Project: A materials genome approach to accelerating materials innovation, APL Mater., 2013, 1, 011002 CrossRef .
  42. G. Henkelman, A. Arnaldsson and H. Jónsson, A fast and robust algorithm for Bader decomposition of charge density, Comput. Mater. Sci., 2006, 36, 354–360 CrossRef .
  43. E. Sanville, S. D. Kenny, R. Smith and G. Henkelman, Improved grid-based algorithm for Bader charge allocation, J. Comput. Chem., 2007, 28, 899–908 CrossRef CAS PubMed .
  44. W. Tang, E. Sanville and G. Henkelman, A grid-based Bader analysis algorithm without lattice bias, J. Phys.: Condens. Matter, 2009, 21, 084204 CrossRef CAS PubMed .
  45. A. R. Oganov and C. W. Glass, Crystal structure prediction using ab initio evolutionary techniques: Principles and applications, J. Chem. Phys., 2006, 124, 244704 CrossRef PubMed .
  46. A. O. Lyakhov, A. R. Oganov, H. T. Stokes and Q. Zhu, New developments in evolutionary structure prediction algorithm USPEX, Comput. Phys. Commun., 2013, 184, 1172–1182 CrossRef CAS .
  47. A. R. Oganov, A. O. Lyakhov and M. Valle, How evolutionary crystal structure prediction works—And why, Acc. Chem. Res., 2011, 44, 227–237 CrossRef CAS PubMed .
  48. A. R. Oganov, Y. Ma, A. O. Lyakhov, M. Valle and C. Gatti, Evolutionary crystal structure prediction as a method for the discovery of minerals and materials, Rev. Mineral. Geochem., 2010, 71, 271–298 CrossRef CAS .
  49. P. E. Blöchl, Projector augmented-wave method, Phys. Rev. B: Condens. Matter Mater. Phys., 1994, 50, 17953–17979 CrossRef PubMed .
  50. G. Kresse and J. Furthmüller, Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54, 11169–11186 CrossRef CAS PubMed .
  51. J. P. Perdew, K. Burke and M. Ernzerhof, Generalized gradient approximation made simple, Phys. Rev. Lett., 1996, 77, 3865–3868 CrossRef CAS PubMed .
  52. K. Momma and F. Izumi, VESTA 3 for three-dimensional visualization of crystal, volumetric and morphology data, J. Appl. Crystallogr., 2011, 44, 1272–1276 CrossRef CAS .
  53. S. Nosé, A unified formulation of the constant temperature molecular dynamics methods, J. Chem. Phys., 1984, 81, 511–519 CrossRef .
  54. W. G. Hoover, Canonical dynamics: Equilibrium phase-space distributions, Phys. Rev. A, 1985, 31, 1695–1697 CrossRef PubMed .
  55. A. Togo and I. Tanaka, First principles phonon calculations in materials science, Scr. Mater., 2015, 108, 1–5 CrossRef CAS .
  56. R. Dronskowski and P. E. Bloechl, Crystal orbital Hamilton populations (COHP): Energy-resolved visualization of chemical bonding in solids based on density-functional calculations, J. Chem. Phys., 1993, 97, 8617–8624 CrossRef CAS .
  57. V. L. Deringer, A. L. Tchougréeff and R. Dronskowski, Crystal orbital Hamilton population (COHP) analysis as projected from plane-wave basis sets, J. Phys. Chem. A, 2011, 115, 5461–5466 CrossRef CAS PubMed .
  58. S. Maintz, V. L. Deringer, A. L. Tchougréeff and R. Dronskowski, Analytic projection from plane-wave and PAW wavefunctions and application to chemical-bonding analysis in solids, J. Comput. Chem., 2013, 34, 2557–2567 CrossRef CAS PubMed .
  59. S. Maintz, V. L. Deringer, A. L. Tchougréeff and R. Dronskowski, LOBSTER: A tool to extract chemical bonding from plane-wave based DFT, J. Comput. Chem., 2016, 37, 1030–1035 CrossRef CAS PubMed .
  60. J. Heyd, G. E. Scuseria and M. Ernzerhof, Hybrid functionals based on a screened Coulomb potential, J. Chem. Phys., 2003, 118, 8207–8215 CrossRef CAS .

Footnote

Electronic supplementary information (ESI) available: theoretical and computational details. See DOI: https://doi.org/10.1039/d4dt02063d

This journal is © The Royal Society of Chemistry 2024