Next Article in Journal
Three for the Price of One: Concomitant I⋯N, I⋯O, and I⋯π Halogen Bonds in the Same Crystal Structure
Next Article in Special Issue
Synthesis, Characterization, and X-ray Crystallography, of the First Cyclohexadienyl Trifluoromethyl Metal Complex (η5-C6H7)Fe(CO)2CF3
Previous Article in Journal
Process and Anti-Mildew Properties of Tea Polyphenol-Modified Citral-Treated Bamboo
Previous Article in Special Issue
2-Imidazoline Nitroxide Derivatives of Cymantrene
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Reinvestigation of Reactions of HgPh2 with Eu and Yb Metal and the Synthesis of a Europium(II) Bis(tetraphenylborate) †

1
School of Chemistry, Monash University, Clayton, VIC 3800, Australia
2
College of Science & Engineering, James Cook University, Townsville, QLD 4811, Australia
*
Author to whom correspondence should be addressed.
In memory of our colleague and co-author Dr. Michal Wiecko.
Deceased.
Molecules 2022, 27(21), 7547; https://doi.org/10.3390/molecules27217547
Submission received: 21 October 2022 / Revised: 31 October 2022 / Accepted: 1 November 2022 / Published: 3 November 2022

Abstract

:
Europium bis(tetraphenylborate) [Eu(thf)7][BPh4]2⋅thf containing a fully solvated [Eu(thf)7]2+ cation, was synthesized by protolysis of “EuPh2” (from Eu and HgPh2) with Et3NHBPh4, and the structure was determined by single-crystal X-ray diffraction. Efforts to characterize the putative “Ph2Ln” (Ln = Eu, Yb) reagents led to the synthesis of a mixed-valence complex, [(thf)3YbII(μ-Ph)3YbIII(Ph)2(thf)]⋅2thf, resulting from the reaction of Yb metal with HgPh2 at a low temperature. This mixed-valence YbII/YbIII compound was studied by 171Yb-NMR spectroscopy and single-crystal X-ray diffraction, and the oxidation states of the Yb atoms were assigned.

1. Introduction

Rare earth metal compounds containing a carbon-to-metal σ-bond have been of considerable interest for many years [1,2,3,4,5,6,7,8,9,10]. The uniquely high reactivity of this mainly ionic bond not only poses a synthetic challenge from the fundamental research viewpoint but also affords an opportunity for widespread synthetic applications [11,12,13,14,15]. Brönstedt and Lewis basic alkyl or aryl ligands can easily be displaced to afford ionic species [9,16,17], offering access to more reactive systems, in particular for the catalysis of olefin polymerization reactions [18,19,20,21,22,23,24]. In this paper, a weakly coordinating tetraphenylborate BPh4- anion is used to increase the electropositive character of the central metal and to offer available sites for substrate coordination. In addition, complexes with BPh4 offer an interesting reactivity for metalorganic synthesis [25,26,27,28]. In recent years, our group has reported on the synthesis of a number of compounds with BPh4 anions [29]. The ionic perfluoroaryllanthanoid(II) tetraphenylborates [Eu(C6F5)(thf)6]BPh4 and [Yb(C6F5)(thf)5]BPh4 were obtained by treating Yb or Eu metal with HgPh(C6F5) in the presence of Me3NHBPh4 [30]. In the course of the present study, we found that the BPh4 anion has a stabilizing effect, increasing the thermal robustness of the ionic complexes relative to [Eu(C6F5)2(thf)5] and [Yb(C6F5)2(thf)4]. Motivated by these results, we were interested in the possible synthesis of analogous LnII (Ln = Eu, Yb) complexes with an unsubstituted phenyl ligand.

2. Results

Initially, attempts were made to synthesize the desired complexes, [LnPh(thf)n]BPh4 by redox transmetallation/ligand exchange reactions of ytterbium or europium metal, HgPh2, and Me3NHBPh4, a synthesis route similar to that used for the preparation of [Eu(C6F5)(thf)6]BPh4 and [Yb(C6F5)(thf)5]BPh4 [30]. However, along with Hg metal, non-crystalline, poorly soluble products, presumably the bis(tetraphenylborate)s [Ln(thf)n][BPh4]2 (Ln = Eu (1), n = 7; Yb (2); n = 6), were formed (Scheme 1, Equation (1)). The favoured formation of the Yb bis(tetraphenylborate) [Yb(thf)6][BPh4]2 [31] in various ligand exchange attempts with [Yb(C6F5)(thf)5]BPh4 was discussed in [29]. Next, we attempted to synthesize and characterize diphenyl-ytterbium and -europium (Scheme 1, Equation (2)), which could be protolyzed to afford [LnPh(thf)n]BPh4. According to previous reports, synthesis of [LnPh2(thf)n] can be achieved by treating Ln metal with HgPh2 [32,33]. Following this procedure, the compounds “Ph2Yb” and “Ph2Eu” were obtained at room temperature as highly pyrophoric black powders. Layering of a THF solution of both compounds with Et3NHBPh4 resulted in the corresponding bistetraphenylborates as the only isolated product with surprisingly high purity (Scheme 1, Equation (3)). Single crystals of the compounds were collected directly from the reaction mixture, and brown impurities were removed with THF. The ytterbium complex [Yb(thf)6][BPh4]2 2 was reported in [31], and we also obtained this compound from reactions of [Yb(C6F5)(thf)5]BPh4 with protic reagents [29]. The homoleptic complexes 1 and 2 should undergo ligand exchange reactions by displacement of THF. If carried out in non-polar solvents, homoleptic complexes may be targeted.
A single-crystal X-ray diffraction study of [Eu(thf)7][BPh4]2⋅thf (1) (Figure 1, left) showed that it is isostructural to the corresponding SmII complex reported by Evans [31]. [Eu(thf)7][BPh4]2⋅thf (1) crystallized in the monoclinic space group P21/n. Because the coordination chemistry of rare earth metals is primarily governed by ionic interactions and SmII and EuII have virtually the same ionic radius [34], the structures are very similar. The inner coordination sphere of the Eu-centered cation resembles a distorted pentagonal bipyramid. This [Eu(thf)7]2+ core is surrounded by a total of eight BPh4- anions, forming a cube-like shape around the cation (Figure 1 right). The Eu–O distance ranges from 2.5214(17) to 2.6209(16) Å, and the average Eu-O bond length (2.57 Å) in 1 is comparable to that of [Sm(thf)7][BPh4]2⋅thf [31], owing to the similar sizes of Sm2+ and Eu2+ [34].
We were unable to directly characterize the putative “[LnPh2(thf)x]” products from the redox transmetallation reaction (Scheme 1, Equation (2)). Considering the temperature sensitivity of lanthanoid aryls and alkyls and the possibility of ether cleavage reactions [35,36,37,38], the redox transmetallation reaction was repeated at −25 °C. At this temperature, a reaction of HgPh2 with Eu or Yb was only observed at a concentration of at least ~0.5 mol/L. Storage of this slurry mixture at −25 °C for two weeks under daily sonication at r.t. for ~1 min. resulted in the formation of large red crystals of the mixed valent species [Yb2Ph5(thf)4]⋅2thf (3) (Scheme 2), a complex previously isolated by Bochkarev et al. without THF of crystallization, after synthesis by a reaction of ytterbium naphthalenide with HgPh2 [39]. Our new facile synthesis route offers a unique opportunity for future selective syntheses of +2/+3 mixed-valence Yb compounds by protolytic ligand exchange reactions of [Yb2Ph5(thf)4]. No Eu compound similar to 3 or any solely divalent species was isolated from an analogous synthesis.
Compound 3 crystallized in the monoclinic space group C2/c with one whole molecule in the asymmetric unit (Figure 2), whereas the structure reported by Bochkarev was solved and refined in monoclinic space group P21 with two whole molecules in the asymmetric unit (one had two disordered THF ligands) [39]. In the structure, two six-coordinate Yb atoms are bridged by three phenyl ligands, each through a single C atom. Yb1 has two terminal phenyl ligands and one THF donor, whereas Yb2 has only three terminal THF ligands. The simplest explanation for the stoichiometry is that one Yb atom is YbII and the other is YbIII, and these should be distinguishable based on the ionic radius difference of 0.15 Å between six-coordinate Yb2+ and Yb3+ [34]. However, the reality is more complex. Relevant bond lengths are listed in Table 1, together with comparable data reported by Bochkarev for molecule Yb3,4 of [Yb2Ph5(thf)4] [39]. Both Yb1 and Yb2 have terminal THF ligands, where Yb1-O1 is 2.400(6) Å, whereas <Yb-O2,3,4> is 2.45 Å, which is a difference much less than that expected if Yb1 is trivalent and Yb2 is divalent. The corresponding difference for the reported complex is much larger, at 0.14 Å [39], but the difference is not significant according to the 3 e.s.d. criterion. On the other hand, the terminal Yb1-C1,7 bond lengths of 3 (2.446(8), 2.445(7)) correspond closely to those of the six-coordinate anionic YbIII complex [YbPh4(dme)] (2.427(9)-2.475(9) Å) [40]. They are also close to the Yb-C bond lengths (2.398(5)–2.423(5) Å) of six-coordinate [YbIIIPh3(thf)3] [40,41]. In the divalent six-coordinate complex [Yb(C6F5)2(thf)4], Yb-C is considerably longer, at 2.649(3) Å [42]. Accordingly, the Yb-C bond lengths provide evidence that Yb1 is +III; therefore, Yb2 is in the +II oxidation state. The reported data support the view that Yb3 (Table 1) is trivalent, despite the high e.s.d. values [39]. The small difference in the terminal Yb-O bond lengths despite the differing oxidation states proposed above is consistent with the reported data; Yb-O is only marginally shorter in [YbPh3(thf)3] (2.381(3)-2.413(3) Å) [43] than in [Yb(C6F5)2(thf)4] (2.428(2)-2.440(2) Å) [30].
In the case of the bridging phenyl groups, C13, 19, and 25 are bound more closely to Yb1 than Yb2, by 0.176, 0.068, and 0.019 Å, respectively (Table 1, Figure 3), the last difference not being significant; however, overall, these data are consistent with the above oxidation state assignments for Yb1 and Yb2. In addition an ortho carbon (C18, C24, and C30), of the phenyl rings of C13, 19, and 25 approach Yb2 at 2.993(4)-3.130 Å (Table 1), whereas the other ortho carbon (C14, 20, and 26) of the same rings is more distant from Yb1, despite the higher oxidation state of Yb1. The Yb2-ortho-C contacts are in the range for intramolecular YbII-π-C(Ph) interactions [44,45,46,47,48,49], whereas the Yb1-C distances are too long for credible YbIII-C π interactions [44]. There are also possible agostic interactions with the ortho C-H groups on the phenyl rings of C(18,24,30), as the Yb(2)-H distances (Table 1) are just within the sum of the van der Waals radius of hydrogen (1.20 Å) [50] and the metallic radius (pseudo-van der Waals radius) of ytterbium (1.94 Å) [51]. The three bridging phenyl groups lie outside of the plane generated by C(13), C(19), and C(25) by 10, 17, and 9°, respectively, and all twist in the same direction towards Yb(2) in a propellor-like fashion. In summary, π and possible C-H-Yb agostic interactions reinforce the bridging C-Yb2 interaction, leading to a closer approach of the bridging arene rings to divalent Yb2. If the C13 Ph group, which has, by far the greatest difference between the Yb1-C and Yb2-C bond lengths, is allocated to Yb1, then the structure would be an association of [YbPh3(thf)] and [YbPh2(thf)3], an outcome consistent with the proposal of Bochkarev that the complex is formed by an association of YbPh3 and YbPh2 [39]. Applying the 3 e.s.d. criterion to the reported [Yb2Ph5(thf)4] structure, the separations of Yb3 and Yb4 from the same bridging C atom [39] are essentially indistinguishable. In the present structural refinement, the average Yb2-C bond distance is 2.64 Å (2.66 Å for Bochkarev’s structure [39]), whereas the analogous Yb1-C average is 2.51 Å (2.48 Å for that reported in [39]), and the average Yb-O distances are 2.45 and 2.400(6) Å for Yb2 and Yb1, respectively (2.44 and 2.30 Å for those reported in [39]).
The 171Yb NMR spectrum suggests that complex 3 dissociates into two neutral species of differing oxidation states, namely YbPh2 and YbPh3. As paramagnetic YbIII complexes have not been observed in 171Yb NMR spectra, it is unlikely that a mixed-valence Yb2Ph5 species would be observed, owing to bridging of YbII through phenyl groups to YbIII. In the 171Yb NMR spectrum of the compound (Figure 4), a single resonance is observed at δ = 694 ppm. This value is similar to a resonance (686 ppm) in the 171Yb NMR spectrum of the reaction mixture from Yb metal and iodobenzene in THF at -78°C, a system that reacts as “PhYbI(thf)n” [40]. Because the spectrum also has a resonance at 388 ppm attributable to YbI2 [40,41,52], the resonance at 686 ppm can be assigned to [YbPh2(thf)n], the product with YbI2 of the Schlenk equilibrium (Scheme 3). This, in turn, leads to the assignment of 694 ppm of 3 to [YbPh2(thf)n], with coproduct YbPh3 silent in the 171Yb NMR spectrum, owing to paramagnetism. The chemical shift lies between those of four-coordinate [Yb(2,6-Ph2C6H3)2(thf)2] reported by Niemeyer (927 ppm) [52], and that of six-coordinate [Yb(C6F5)2(thf)4] [ (463 ppm) [30,42]. Because coordination number is a key determinant of the chemical shift in the 171Yb NMR spectra within a series of related compounds [53,54] and [YbPh2(thf)n] can be expected to have a coordination number more similar to that of [Yb(C6F5)2(thf)4] than the sterically crowded [Yb(2,6-Ph2C6H3)2(thf)2], the observed value is probably reasonable for n = 3 (5 coordination), given the differing electron-withdrawing properties of C6F5 and Ph, or even n = 4.

3. Materials and Methods

3.1. General Considerations

All manipulations were performed with the rigorous exclusion of oxygen and moisture in flame-dried glassware with greaseless stopcocks, either on a dual-manifold Schlenk line or in an argon-filled Saffron glovebox. All solvents were predried over Na wire, stored over Na/K benzophenone ketyl, and distilled prior to use. “Ph2Ln” (Ln = Eu, Yb) and “PhYbI(thf)n” were prepared according to the procedure described in [32,33,40,41]. NMR spectra were obtained using a Bruker DPX 300 MHz spectrometer. 171Yb-NMR spectra are referenced to 0.15 M [Yb(C5Me5)2(thf)] in THF/10 % C6D6 [53].

3.2. Synthesis of [Eu(thf)7][BPh4]2⋅thf (1)

“Ph2Eu” (60 mg) was dissolved in THF (3 mL) in a glass tube. The solution was slowly topped with a solution of Et3NHBPh4 (100 mg, 0.24 mmol) in 7 mL of THF. The solutions were allowed to mix over a period of 3 days. The remaining dark solution was decanted, evaporated and the resulting off-white crystalline solid was washed with 2 mL of THF, affording 71 mg (55 % based on Et3NHBPh4) of compound 1. Single crystals suitable for X-ray diffraction were found in this solid. IR (Nujol, cm1): 1577 (m), 1542 (m), 1307 (w), 1261 (s), 1156 (s), 1069 (w), 1030 (w), 856 (w), 800 (w), 747 (s), 711 (s). Anal. Calcd. for C76H96B2EuO7 (1295.66 g/mol, loss of THF of crystallization): C 70.48, H 7.47; found: C 70.85, H 7.68.

3.3. Synthesis of [Yb(thf)6][BPh4]2 (2)

“Ph2Yb” (60 mg) was dissolved in THF (3 mL) in a glass tube. The solution was slowly topped with a solution of Et3NHBPh4 (100 mg, 0.24 mmol) in 7 mL of THF. The solutions were allowed to mix over a period of 3 days. The remaining dark solution was decanted, evaporated and the resulting yellow crystalline solid was washed with 2 mL of THF, affording compound 2. Single crystals suitable for X-ray diffraction were found in this solid, and 2 was identified by unit cell [31].

3.4. Synthesis of [Ph5Yb2(thf)4]⋅2thf (3)

Ytterbium metal (260 mg, 1.5 mmol) and HgPh2 (355 mg, 1.0 mmol) were suspended in 2 mL of THF. Upon sonication at r.t. for 2 h, the mixture developed a red color and was thereafter cooled to −25 °C. Daily sonication for 1 min at r.t. and storage at −25 °C resulted in the formation of mercury metal and a dark red crystalline solid. The remaining solution was decanted, and the residue was treated with 10 mL of THF and decanted from the mercury metal. Concentration of the solution to 5 mL resulted in the formation of large red crystals of 3∙(thf)2 suitable for X-ray diffraction. Isolated yield: 209 mg (41 %). 171Yb-NMR (thf, 52.6 MHz, −30 °C): δ = 694 ppm.

3.5. X-ray Crystallography

Complexes 1 and 3 were measured on a Bruker APEX-II CCD diffractometer equipped with graphite-monochromated Mo-Kα radiation (λ = 0.71073 Å) at 123 K and mounted on a fiber loop in crystallography oil. Absorption corrections were completed using the Apex II program suite with SADABS [55]. Structural solutions were obtained by SHELXT [56] using full-matrix least-squares methods against F2 with SHELX2018 [56] in conjunction with the Olex2 [57] graphical user interface. All hydrogen atoms were placed in calculated positions according to the riding model.
Crystal data for 1 and 3:
1: C80H104B2EuO8, M = 1367.21, yellow block, 0.5 × 0.4 × 0.2 mm3, monoclinic, space group P21/n (No. 14), a = 18.2175(6) Å, b = 19.1237(6) Å, c = 20.2661(6) Å, β = 94.6320(10)°, V = 7037.4(4) Å3, Z = 4, ρc = 1.290 g/cm3, μ = 0.947 mm1, F000 = 2884, Bruker X8 Apex II CCD, MoKα radiation, λ = 0.71073 Å, T = 123(1)K, 2θmax = 52.0°, 202,209 reflections collected, 13,450 unique (Rint = 0.0622), Final GooF = 1.035, R1 = 0.0323 [I > 2σ(I)], wR2 = 0.0681.
3: C54H73O6Yb2, M = 1164.20, dark red block, 0.50 × 0.30 × 0.30 mm3, monoclinic, space group C2/c (No. 15), a = 40.713(8) Å, b = 11.271(2) Å, c = 22.724(5) Å, β = 104.70(3)°, V = 10,086(4) Å3, Z = 8, ρc = 1.533 g/cm3, μ = 3.733 mm1, F000 = 4680, Bruker X8 Apex II CCD, MoKα radiation, λ = 0.71073 Å, T = 123(1) K, 2θmax = 55.8°, 39,023 reflections collected, 11,781 unique (Rint = 0.0324), Final GooF = 1.006, R1 = 0.0282 [I > 2σ(I)], wR2 = 0.0590.

4. Conclusions

In conclusion, we prepared the ionic Eu(II) complex [Eu(thf)7][BPh4]2⋅thf. In an attempt to characterize the previously described “Ph2Ln” compounds, which can be used to prepare [Ln(thf)n][BPh4]2 (Ln = Eu, Yb), an improved synthesis of the YbII/YbIII compound [Yb2Ph5(thf)4] was developed. The Eu complex 1 (and the Yb complex 2) should be a source of homoleptic complexes via displacement of THF by neutral ligands in non-polar solvents, whereas 3 may yield other mixed-oxidation-state Yb complexes via protolysis reactions, whereby the phenyl groups of 3 are replaced by ligands (L) (e.g., L = OAr, OR, pyrazolate, amidinate, etc.) by reaction with the corresponding proligand (HL).

Author Contributions

The complexes were synthesized and characterized by M.W. and Z.G; the original draft of the manuscript was written by M.W. and subsequently reviewed and edited by Z.G., G.B.D. and P.C.J.; supervision, project administration and funding acquisition, G.B.D. and P.C.J. All authors have read and agreed to the published version of the manuscript.

Funding

Australian Research Council funding (DP190100798).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Crystal data can be obtained free of charge from the Cambridge Crystallographic Data Centre (CCDC 2211517 for 1 and 2211518 for 3).

Acknowledgments

G.B.D. and P.C.J. gratefully acknowledge ARC for funding (DP190100798). M.W. thanks Deutsche Forschungs-gemeinschaft (DFG) for a research fellowship.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ortu, F. Rare Earth Starting Materials and Methodologies for Synthetic Chemistry. Chem. Rev. 2022, 122, 6040–6116. [Google Scholar] [CrossRef] [PubMed]
  2. Guo, Z.; Huo, R.; Tan, Y.Q.; Blair, V.; Deacon, G.B.; Junk, P.C. Syntheses of reactive rare earth complexes by redox transmetallation/protolysis reactions–A simple and convenient method. Coord. Chem. Rev. 2020, 415, 213232. [Google Scholar] [CrossRef]
  3. Zimmermann, M.; Anwander, R. Homoleptic rare-earth metal complexes containing Ln− C σ-bonds. Chem. Rev. 2010, 110, 6194–6259. [Google Scholar] [CrossRef] [PubMed]
  4. Johnson, K.R.; Hayes, P.G. Synthesis and reactivity of dialkyl lutetium complexes supported by a novel bis (phosphinimine) carbazole pincer ligand. Organometallics 2009, 28, 6352–6361. [Google Scholar] [CrossRef]
  5. Petrov, A.R.; Rufanov, K.A.; Harms, K.; Sundermeyer, J. Re-investigation of ortho-metalated N, N-dialkylbenzylamine complexes of rare-earth metals. First structurally characterized arylates of neodymium and gadolinium Li [LnAr4]. J. Organomet. Chem. 2009, 694, 1212–1218. [Google Scholar] [CrossRef]
  6. Yan, K.; Upton, B.M.; Ellern, A.; Sadow, A.D. Lewis acid-mediated β-hydride abstraction reactions of divalent M(C(SiHMe2)3)2THF2 (M= Ca, Yb). J. Am. Chem. Soc. 2009, 131, 15110–15111. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Ge, S.; Meetsma, A.; Hessen, B. Highly efficient hydrosilylation of alkenes by organoyttrium catalysts with sterically demanding amidinate and guanidinate ligands. Organometallics 2008, 27, 3131–3135. [Google Scholar] [CrossRef] [Green Version]
  8. Lyubov, D.M.; Fukin, G.K.; Trifonov, A.A. N,N′-Diisopropyl-N′′-bis(trimethylsilyl)guanidinate Ligand as a Supporting Coordination Environment in Yttrium Chemistry. Synthesis, Structure, and Properties of Complexes [(Me3Si)2NC(Ni-Pr)2]YCl2(THF)2,[(Me3Si)2NC(Ni-Pr)2]Y(CH2SiMe3)2(THF)2, and [(Me3Si)2NC(Ni-Pr)2]Y[(μ-H)(μ-Et)2BEt]2(THF). Inorg. Chem. 2007, 46, 11450–11456. [Google Scholar]
  9. Bambirra, S.; Meetsma, A.; Hessen, B. Lanthanum tribenzyl complexes as convenient starting materials for organolanthanum chemistry. Organometallics 2006, 25, 3454–3462. [Google Scholar] [CrossRef] [Green Version]
  10. Buschmann, D.A.; Schumacher, L.; Anwander, R. Rare-earth-metal half-sandwich complexes incorporating methyl, methylidene, and hydrido ligands. Chem. Commun. 2022, 58, 9132–9135. [Google Scholar] [CrossRef]
  11. Molander, G.A.; Romero, J.A.C. Lanthanocene catalysts in selective organic synthesis. Chem. Rev. 2002, 102, 2161–2186. [Google Scholar] [PubMed]
  12. Gromada, J.; Carpentier, J.F.; Mortreux, A. Group 3 metal catalysts for ethylene and α-olefin polymerization. Coord. Chem. Rev. 2004, 248, 397–410. [Google Scholar] [CrossRef]
  13. Hou, Z.; Wakatsuki, Y. Recent developments in organolanthanide polymerization catalysts. Coord. Chem. Rev. 2002, 231, 1–22. [Google Scholar] [CrossRef]
  14. Trambitas, A.G.; Panda, T.K.; Jenter, J.; Roesky, P.W.; Daniliuc, C.; Hrib, C.G.; Jones, P.G.; Tamm, M. Rare-earth metal alkyl, amido, and cyclopentadienyl complexes supported by imidazolin-2-iminato ligands: Synthesis, structural characterization, and catalytic application. Inorg. Chem. 2010, 49, 2435–2446. [Google Scholar] [CrossRef]
  15. Wooles, A.J.; Mills, D.P.; Lewis, W.; Blake, A.J.; Liddle, S.T. Lanthanide tri-benzyl complexes: Structural variations and useful precursors to phosphorus-stabilised lanthanide carbenes. Dalton Trans. 2010, 39, 500–510. [Google Scholar] [CrossRef]
  16. Ge, S.; Meetsma, A.; Hessen, B. Neutral and Cationic Rare Earth Metal Alkyl and Benzyl Compounds with the 1, 4, 6-Trimethyl-6-pyrrolidin-1-yl-1,4-diazepane Ligand and Their Performance in the Catalytic Hydroamination/Cyclization of Aminoalkenes. Organometallics 2008, 27, 5339–5346. [Google Scholar] [CrossRef] [Green Version]
  17. Zeimentz, P.M.; Okuda, J. Cationic aryl complexes of the rare-earth metals. Organometallics 2007, 26, 6388–6396. [Google Scholar] [CrossRef]
  18. Arndt, S.; Okuda, J. Cationic Alkyl Complexes of the Rare-Earth Metals: Synthesis, Structure, and Reactivity. Adv. Synth. Catal. 2005, 347, 339–354. [Google Scholar]
  19. Zeimentz, P.M.; Arndt, S.; Elvidge, B.R.; Okuda, J. Cationic organometallic complexes of scandium, yttrium, and the lanthanoids. Chem. Rev. 2006, 106, 2404–2433. [Google Scholar] [CrossRef]
  20. Robert, D.; Spaniol, T.P.; Okuda, J. Neutral and Monocationic Half-Sandwich Methyl Rare-Earth Metal Complexes: Synthesis, Structure, and 1,3-Butadiene Polymerization Catalysis. Eur. J. Inorg. Chem. 2008, 2008, 2801–2809. [Google Scholar]
  21. Nishiura, M.; Mashiko, T.; Hou, Z. Synthesis and styrene polymerisation catalysis of η5-and η1-pyrrolyl-ligated cationic rare earth metal aminobenzyl complexes. Chem. Commun. 2008, 2019–2021. [Google Scholar] [CrossRef] [PubMed]
  22. Döring, C.; Kretschmer, W.P.; Bauer, T.; Kempe, R. Scandium Aminopyridinates: Synthesis, Structure and Isoprene Polymerization. Eur. J. Inorg. Chem. 2009, 2009, 4255–4264. [Google Scholar] [CrossRef]
  23. Zimmermann, M.; Toörnroos, K.W.; Waymouth, R.M.; Anwander, R. Structure-Reactivity Relationships of Amido-Pyridine-Supported Rare-Earth-Metal Alkyl Complexes. Organometallics 2008, 27, 4310–4317. [Google Scholar] [CrossRef]
  24. Li, X.; Nishiura, M.; Mori, K.; Mashiko, T.; Hou, Z. Cationic scandium aminobenzyl complexes. Synthesis, structure and unprecedented catalysis of copolymerization of 1-hexene and dicyclopentadiene. Chem. Commun. 2007, 4137–4139. [Google Scholar] [CrossRef] [PubMed]
  25. Evans, W.J.; Seibel, C.A.; Ziller, J.W. Unsolvated Lanthanide Metallocene Cations [(C5Me5)2Ln][BPh4]: Multiple Syntheses, Structural Characterization, and Reactivity Including the Formation of (C5Me5)3Nd1. J. Am. Chem. Soc. 1998, 120, 6745–6752. [Google Scholar] [CrossRef]
  26. Evans, W.J.; Champagne, T.M.; Ziller, J.W. Synthesis and reactivity of mono(pentamethylcyclopentadienyl) tetraphenylborate lanthanide complexes of ytterbium and samarium: Tris (ring) precursors to (C5Me5) Ln moieties. Organometallics 2007, 26, 1204–1211. [Google Scholar] [CrossRef]
  27. Evans, W.J.; Walensky, J.R.; Champagne, T.M.; Ziller, J.W.; DiPasquale, A.G.; Rheingold, A.L. Displacement, reduction, and ligand redistribution reactivity of the cationic mono-C5Me5 Ln2+ complexes (C5Me5) Ln (BPh4)(Ln= Sm, Yb). J. Organomet. Chem. 2009, 694, 1238–1243. [Google Scholar] [CrossRef]
  28. Wiecko, M.; Roesky, P.W. A Cationic Bis (phosphiniminomethanide) Europium (II) Complex. Organometallics 2009, 28, 1266–1269. [Google Scholar] [CrossRef]
  29. Deacon, G.B.; Evans, D.J.; Forsyth, C.M.; Junk, P.C. Lanthanoid (II) tetraphenylborate complexes: From discrete ions to pseudo metallocenes. Coord. Chem. Rev. 2007, 251, 1699–1706. [Google Scholar] [CrossRef]
  30. Deacon, G.B.; Forsyth, C.M. Synthesis and structures of the first cationic perfluoroaryllanthanoid (II) complexes. Chem. Eur. J. 2004, 10, 1798–1804. [Google Scholar] [CrossRef]
  31. Evans, W.J.; Johnston, M.A.; Greci, M.A.; Gummersheimer, T.S.; Ziller, J.W. Divalent lanthanide complexes free of coordinating anions: Facile synthesis of fully solvated dicationic [LnLx]2+ compounds. Polyhedron 2003, 22, 119–126. [Google Scholar]
  32. Starostina, T.A.; Shifrina, R.R.; Rybakova, L.F.; Petrov, E.S. Diphenyl ytterbium. Zh. Obshch. Khim. 1987, 57, 2402. [Google Scholar]
  33. Rybakova, L.F.; Tsygankova, S.V.; Rojtershtejn, D.M.; Petrov, E.S. Metallation of 1,3-diphenyl-2-benzylpropene by diphenylytterbium. Russ. J. Gen. Chem. 2004, 74, 141. [Google Scholar]
  34. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta crystallographica section A: Crystal physics, diffraction, theoretical and general crystallography. Acta. Crystallogr. Sect. A 1976, 32, 751–767. [Google Scholar] [CrossRef]
  35. Cole, M.L.; Deacon, G.B.; Forsyth, C.M.; Junk, P.C.; Konstas, K.; Wang, J. Steric modulation of coordination number and reactivity in the synthesis of lanthanoid (III) formamidinates. Chem. Eur. J. 2007, 13, 8092–8110. [Google Scholar] [PubMed]
  36. Evans, W.J.; Allen, N.T.; Ziller, J.W. Facile dinitrogen reduction via organometallic Tm (II) chemistry. J. Am. Chem. Soc. 2001, 123, 7927–7928. [Google Scholar] [CrossRef]
  37. Guo, Z.; Blair, V.; Deacon, G.B.; Junk, P.C. Can Bismuth Replace Mercury in Redox Transmetallation/Protolysis Syntheses from Free Lanthanoid Metals? Chem. Eur. J. 2018, 24, 17464–17474. [Google Scholar] [CrossRef]
  38. Guo, Z.; Blair, V.; Deacon, G.B.; Junk, P.C. Widely contrasting outcomes from the use of tris (pentafluorophenyl) bismuth or pentafluorophenylsilver as oxidants in the reactions of lanthanoid metals with N, N′-diarylformamidines. Dalton Trans. 2020, 49, 13588–13600. [Google Scholar] [CrossRef]
  39. Bochkarev, M.N.; Khramenkov, V.V.; Rad’kov, Y.F.; Zakharov, L.N.; Struchkov, Y.T. Synthesis and characterization of pentaphenyldiytterbium Ph2Yb(THF)(μ-Ph)3Yb(THF)3. J. Organomet. Chem. 1992, 429, 27–39. [Google Scholar]
  40. Wiecko, M.; Deacon, G.B.; Junk, P.C. Organolanthanoid-halide synthons−a new general route to monofunctionalized lanthanoid (ii) compounds? Chem. Commun. 2010, 46, 5076–5078. [Google Scholar]
  41. Ali, S.H.; Deacon, G.B.; Junk, P.C.; Hamidi, S.; Wiecko, M.; Wang, J. Lanthanoid Pseudo−Grignard Reagents: A Major Untapped Resource. Chem. Eur. J. 2018, 24, 230–242. [Google Scholar] [CrossRef] [PubMed]
  42. Deacon, G.B.; Forsyth, C.M. A half-sandwich perfluoro-organoytterbium(II) complex from a simple redox transmetallation/ligand exchange synthesis. Organometallics 2003, 22, 1349–1352. [Google Scholar] [CrossRef]
  43. Bochkarev, L.N.; Zheleznova, T.A.; Safronova, A.V.; Drozdov, M.S.; Zhil′tsov, S.F.; Zakharov, L.N.; Fukin, G.K.; Khorshev, S.Y. Synthesis and crystal structure of triphenyl[tris(tetrahydrofuran)]ytterbium, Ph3Yb(THF)3. Russ. Chem. Bull. 1998, 47, 163–166. [Google Scholar] [CrossRef]
  44. Deacon, G.B.; Shen, Q. Complexes of lanthanoids with neutral π donor ligands. J. Organomet. Chem. 1996, 511, 1–17. [Google Scholar] [CrossRef]
  45. Cloke, F.G.N. Zero oxidation state compounds of scandium, yttrium, and the lanthanides. Chem. Soc. Rev. 1993, 22, 17–24. [Google Scholar] [CrossRef]
  46. Liang, H.; Shen, Q.; Jin, S.; Lin, Y. Synthesis and X-ray crystal structure of [Eu(η6-C6Me6)(AlCl4)2]4; the first cyclotetrameric lanthanide(II) complex with a neutral π-ligand. J. Chem. Soc. Chem. Commun. 1992, 480–481. [Google Scholar] [CrossRef]
  47. Deacon, G.B.; Forsyth, C.M.; Junk, P.C.; Skelton, B.W.; White, A.H. The striking influence of intramolecular lanthaoid π-arene interactions on the structural architecture of the homoleptic aryloxolanthanoid(II) complexes, [Eu2(Odpp)(μ-Odpp)3] and [Yb2(Odpp)2(μ-Odpp)2], and the Yb(II)–Yb(III) trimetallic [Yb2(μ-Odpp)3][Yb(Odpp)4]- (Odpp = 2,6-diphenylphenolato). Chem. Eur. J. 1999, 5, 1452–1459. [Google Scholar]
  48. Deacon, G.B.; Forsyth, C.M. Linkage isomerism and C-H activation in an ytterbium(II) tetraphenylborate. Chem. Commun. 2002, 2522–2523. [Google Scholar]
  49. Deacon, G.B.; Junk, P.C.; Moxey, G.J.; Ruhlandt-Senge, K.; Prix, C.S.; Zuniga, M.F. Charge-separated and molecular heterobimetallic rare earth-rare earth and alkaline-earth-rare earth aryloxo complexes featuring intramolecular metal -π-arene interactions. Chem. Eur. J. 2009, 15, 5503–5519. [Google Scholar] [CrossRef]
  50. Bondi, A. van der Waals volumes and radii. J. Phys. Chem. 1964, 68, 441–451. [Google Scholar] [CrossRef]
  51. Wells, A.F. Structural Inorganic Chemistry, 5th ed; Oxford University Press: Oxford, UK, 1984; p. 1288. [Google Scholar]
  52. Heckmann, G.; Niemeyer, M. Synthesis and first structural characterization of lanthanide (II) aryls: Observation of a Schlenk equilibrium in europium(II) and ytterbium(II) chemistry. J. Am. Chem. Soc. 2000, 122, 4227–4228. [Google Scholar] [CrossRef]
  53. Avent, A.G.; Edelman, M.A.; Lappert, M.F.; Lawless, G.A. The first high resolution direct NMR observation of an f-block element. J. Am. Chem. Soc. 1989, 111, 3423–3425. [Google Scholar] [CrossRef]
  54. Deacon, G.B.; Fallon, G.D.; Forsyth, C.M.; Schumann, H.; Weimann, R. Organoamido- and Aryloxo-lanthanoids. 15. Syntheses of low coordination number divalent lanthanoid organoamide complexes and the X-ray crystal structures of Bis[(N-2,6-di-iso-propylphenyl)(N-trimethylsilyl)amido]bis(tetrahydrofuran)-samarium(II) and -ytterbium(II). Chem. Ber. 1997, 130, 409–415. [Google Scholar]
  55. Sheldrick, G.M. SADABS, Software for Empirical Absorption Corrections; Standard Software Reference; University of Göttingen: Göttingen, Germany, 1996. [Google Scholar]
  56. Sheldrick, G.M. SHELXT–Integrated space-group and crystal-structure determination. Acta Cryst. 2015, C71, 3–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.; Puschmann, H. OLEX2: A complete structure solution, refinement and analysis program. J. Appl. Cryst. 2009, 42, 339–341. [Google Scholar] [CrossRef]
Scheme 1. Synthesis of complexes 12.
Scheme 1. Synthesis of complexes 12.
Molecules 27 07547 sch001
Figure 1. Left: Structure of the cation of 1 in the solid state drawn with ellipsoids at 50% probability and hydrogen atoms and anions omitted. Right: Space-filling representation of the environment around the [Eu(thf)7]2+ cation with one BPh4 and one THF molecule removed for clarity. Average Eu-O distance: <Eu-O> = 2.57 Å. Selected angles [°]: O1-Eu-O2 106.18(6), O1-Eu-O3 74.84(6), O1-Eu-O4 78.39(6), O1-Eu-O5 131.01(6), O1-Eu-O6 75.14(5), O1-Eu-O7 147.31(6), O2-Eu-O3 85.76(5), O2-Eu-O4 170.19(6), O2-Eu-O5 103.63(5), O2-Eu-O6 78.08(5), O2-Eu-O7 80.49(6), O3-Eu-O4 87.14(6), O3-Eu-O5 145.63(5), O3-Eu-O6 140.13(5), O3-Eu-O7 73.80(6), O4-Eu-O5 78.72(6), O4-Eu-O6 111.66(6), O4-Eu-O7 91.04(6), O5-Eu-O6 74.12(5), O5-Eu-O7 75.29(5), O6-Eu-O7 136.91(6).
Figure 1. Left: Structure of the cation of 1 in the solid state drawn with ellipsoids at 50% probability and hydrogen atoms and anions omitted. Right: Space-filling representation of the environment around the [Eu(thf)7]2+ cation with one BPh4 and one THF molecule removed for clarity. Average Eu-O distance: <Eu-O> = 2.57 Å. Selected angles [°]: O1-Eu-O2 106.18(6), O1-Eu-O3 74.84(6), O1-Eu-O4 78.39(6), O1-Eu-O5 131.01(6), O1-Eu-O6 75.14(5), O1-Eu-O7 147.31(6), O2-Eu-O3 85.76(5), O2-Eu-O4 170.19(6), O2-Eu-O5 103.63(5), O2-Eu-O6 78.08(5), O2-Eu-O7 80.49(6), O3-Eu-O4 87.14(6), O3-Eu-O5 145.63(5), O3-Eu-O6 140.13(5), O3-Eu-O7 73.80(6), O4-Eu-O5 78.72(6), O4-Eu-O6 111.66(6), O4-Eu-O7 91.04(6), O5-Eu-O6 74.12(5), O5-Eu-O7 75.29(5), O6-Eu-O7 136.91(6).
Molecules 27 07547 g001
Scheme 2. Synthesis of complex 3.
Scheme 2. Synthesis of complex 3.
Molecules 27 07547 sch002
Figure 2. X-ray crystal structure of compound 3. Thermal ellipsoids are shown at 40%. Hydrogen atoms have been removed for clarity.
Figure 2. X-ray crystal structure of compound 3. Thermal ellipsoids are shown at 40%. Hydrogen atoms have been removed for clarity.
Molecules 27 07547 g002
Figure 3. Comparison of the numbering of complex 3 and [Yb2Ph5(thf)4] [39].
Figure 3. Comparison of the numbering of complex 3 and [Yb2Ph5(thf)4] [39].
Molecules 27 07547 g003
Figure 4. 171Yb-NMR spectra of “PhYbI(thf)n” (top, −78 °C) and complex 3 (bottom, −30 °C).
Figure 4. 171Yb-NMR spectra of “PhYbI(thf)n” (top, −78 °C) and complex 3 (bottom, −30 °C).
Molecules 27 07547 g004
Scheme 3. Schlenk equilibrium in the PhYbI system (coordinated solvent not shown) [41].
Scheme 3. Schlenk equilibrium in the PhYbI system (coordinated solvent not shown) [41].
Molecules 27 07547 sch003
Table 1. Selected bond lengths of complex 3 and the reported structure of [Yb2Ph5(thf)4].
Table 1. Selected bond lengths of complex 3 and the reported structure of [Yb2Ph5(thf)4].
This WorkRef [39] *
BondLength (Å)BondLength (Å)
Yb1-C12.446 (8)Yb3-C652.39 (4)
Yb1-C72.445 (7)Yb3-C712.46 (3)
Yb1-O12.400 (6)Yb3-O52.30 (2)
Yb1-C132.489 (9)Yb3-C592.48 (3)
Yb1-C192.573 (8)Yb3-C532.54 (3)
Yb1-C252.583 (8)Yb3-C472.51 (4)
Yb2-C132.665 (8)Yb4-C592.68 (3)
Yb2-C192.641 (10)Yb4-C532.55 (3)
Yb2-C252.602 (8)Yb4-C472.75 (4)
Yb1-O42.452 (6)Yb4-O62.44 (2)
Yb2-O22.434 (6)Yb4-O72.38 (3)
Yb2-O32.462 (6)Yb4-O82.50 (3)
Yb2C183.034 (4)Yb4C483.08 (4)
Yb2C242.995 (4)Yb4C642.93 (3)
Yb2C303.130 (4)Yb4C543.25 (4)
Yb1-C14 a3.350 (4)Yb3-C52 a3.45 (4)
Yb1-C20 a3.244 (4)Yb3-C60 a3.16 (4)
Yb1-C26 a3.412 (4)Yb3-C58 a3.39 (5)
Yb2H242.785
Yb2H182.921
Yb2H303.060
* There are two similar molecules of [Yb2Ph5(thf)4] in the reported structure. As some THF ligands were disordered in the molecules labelled Yb1 and Yb2, we only made a comparisonwith the non-disordered molecule Yb3 and Yb4. a Nonbonding.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wiecko, M.; Guo, Z.; Deacon, G.B.; Junk, P.C. Reinvestigation of Reactions of HgPh2 with Eu and Yb Metal and the Synthesis of a Europium(II) Bis(tetraphenylborate). Molecules 2022, 27, 7547. https://doi.org/10.3390/molecules27217547

AMA Style

Wiecko M, Guo Z, Deacon GB, Junk PC. Reinvestigation of Reactions of HgPh2 with Eu and Yb Metal and the Synthesis of a Europium(II) Bis(tetraphenylborate). Molecules. 2022; 27(21):7547. https://doi.org/10.3390/molecules27217547

Chicago/Turabian Style

Wiecko, Michal, Zhifang Guo, Glen B. Deacon, and Peter C. Junk. 2022. "Reinvestigation of Reactions of HgPh2 with Eu and Yb Metal and the Synthesis of a Europium(II) Bis(tetraphenylborate)" Molecules 27, no. 21: 7547. https://doi.org/10.3390/molecules27217547

Article Metrics

Back to TopTop