DOI:
10.1039/D5SC05500H
(Edge Article)
Chem. Sci., 2025,
16, 21056-21067
Accessing homoleptic neutral and anionic five-coordinate Pr(IV) siloxide complexes
Received
23rd July 2025
, Accepted 3rd October 2025
First published on 3rd October 2025
Abstract
Structurally characterized Pr(IV) complexes are limited to four examples because the ligands and reaction conditions capable of stabilizing Pr(IV) remain elusive. Here we identify reaction conditions allowing the synthesis of Pr(IV) complexes that were originally thought difficult to isolate. The Pr(IV) complexes of the tris(tert-butoxy)siloxide (–OSi(OtBu)3) and triphenylsiloxide (–OSiPh3) ligands, [PrIV(OSi(OtBu)3)4] (2-PrOtBu), [MPrIV(OSiPh3)5] (5M-PrPh) (M = K, Cs), and [KDB18C6][PrIV(OSiPh3)5], (5[KDB18C6-PrPh]) were isolated and fully characterized upon the oxidation of the tetrakis and pentakis(siloxide)praseodymium(III) ate complexes, [KPrIII(OSi(OtBu)3)4] (1-PrOtBu) and [M2PrIII(OSiPh3)5] (4M-PrPh) (M = K, Cs), using the thianthrene radical cation tetrafluoroborate oxidant, thiaBF4. The crucial role of reagents and reaction conditions, like thiaBF4 over the magic blue oxidant and non-coordinating over coordinating solvents, are demonstrated for the isolation of high valent Pr(IV) complexes. The solid state structural and electrochemical properties were studied and further augmented with theoretical calculations. The Pr(IV) oxidation state was further confirmed by electron paramagnetic resonance (EPR) and SQUID magnetometry measurements. Complexes 5M-PrPh and 5[KDB18C6]-PrPh provide the first example of anionic Ln(IV) complexes demonstrating the possibility of accessing charged Pr(IV) complexes as a tool to manipulate the redox potential and therefore access to more stable complexes with the same ligand.
Introduction
The redox chemistry of the f elements has seen remarkable expansion in the last 20 years with new oxidation states having been identified for both lanthanides and actinides.1–14 Molecular compounds containing lanthanides (Ln) in the + II oxidation state were isolated for all Ln ions (except promethium), despite the fact that the predicted reduction potentials suggested that some of these complexes would be too unstable to isolate and their chemistry too difficult to control.1 The discovery of ligands and reaction conditions capable of stabilizing the Ln(II) oxidation state in reactive intermediates or stable complexes has resulted in attractive magnetic properties15–18 and high reactivity towards small molecule activation.19–29 In contrast, molecular complexes of Ln's in the +IV oxidation state were for a long time limited to Ce, which has an accessible Ln(III)/Ln(IV) redox couple that has a broad range of applications.30–40 The first molecular complex of Tb(IV) was isolated and characterized by our group in 2019, using the redox-innocent tris(tert-butoxy)siloxide (–OSi(OtBu)3) ligand,4 which yielded a five-coordinate tetrakis Tb(IV) complex, [TbIV(OSi(OtBu)3)4]. Several other molecular compounds of Tb(IV) have been isolated since, using monodentate and polydentate siloxide ligands41–44 or imidophosphorane ligands.14 The calculated45 Pr(IV)/Pr(III) oxidation potential (+3.4 V vs. NHE) is very close to the potential reported for the Tb(IV)/Tb(III) couple (+3.3 V vs. NHE), suggesting similar thermodynamic accessibility when using similar supporting ligands and oxidizing conditions. However, only one ligand system has been identified so far that allowed the isolation of isostructural molecular complexes of Pr(IV) and Tb(IV) in the solid state.
Notably, in 2020 we reported the solid state molecular structure of a molecular complex of Pr(IV), [PrIV(OSiPh3)4(CH3CN)2] (Scheme 1A), that was prepared using the triphenylsiloxide (–OSiPh3) supporting ligand by oxidation in acetonitrile solution of the trivalent analogue, [KPrIII(OSiPh3)4], using the strong oxidant, tris(4-bromophenyl)amminium salt, [N(C6H4Br)3][SbCl6], also known as “magic blue”.5 A similar procedure allowed the isolation of the isostructural Tb(IV) complex, [TbIV(OSiPh3)4(CH3CN)2].41 The oxidation of the Tb(IV) complex was found to occur at a potential only +0.18 V more than the Pr(IV), as anticipated by calculated redox potentials. We also found that one solvent molecule can be replaced in both Tb(IV) and Pr(IV) complexes by a phosphineoxide ligand (Scheme 1B) leading to higher stability.46 Further, Zheng and coworkers showed that both coordinated solvent molecules can be replaced43,47 by bidentate ligands (Scheme 1C), increasing the solution stability.
 |
| | Scheme 1 Previously isolated and crystallographically characterized Pr(IV) molecular complexes (A–D) and this work. | |
Despite the redox accessibility of the imidophosphorane Pr(IV) analogue (Eox = −0.72 V vs. Fc/Fc+) of the isolated Tb(IV) molecular complex, [TbIV(NP(1,2-bis-tBu-diamidoethane)(NEt2))4],14 the Pr(IV) species generated in situ was reported to be not stable at −35 °C or above and could not be isolated in the solid state.48 The isolation of a thermally stable four-coordinate molecular Pr(IV) complex8 (Scheme 1D) required the use of a different imidophosphorane (NPtBu3) ligand that significantly shifted the oxidation potential towards more negative values (−1.37 V vs. Fc/Fc+).
Therefore, redox accessibility is not the sole parameter that should be considered when pursuing the isolation of Ln(IV) complexes. Indeed, preventing possible decomposition pathways appears to be key for the isolation of Pr(IV) complexes in the solid state, which remain limited to four examples (Scheme 1).5,46,47,49 Considering that only four- and six-coordinate neutral complexes of Pr(IV) have been isolated, we set out to explore alternative synthetic routes to isolate five-coordinate Pr(IV) complexes. Notably, previous attempts to isolate in the solid state five-coordinate Pr(IV) complexes analogues of the [TbIV(OSi(OtBu)3)4] were not successful.5
Here, by choosing the appropriate conditions that prevent rapid decomposition pathways, we synthesised and characterized four examples of homoleptic five-coordinate Pr(IV) complexes that were previously thought to be difficult to isolate.
Result and discussion
Synthesis and structural characterization
Tetrakis-tertbutoxylsiloxide Pr(III/IV) complexes.
We previously reported that all attempts to isolate a tertbutoxylsiloxide complex by oxidation of the Pr(III) complex, [KPrIII(OSi(OtBu)3)4] (1-PrOtBu), with magic blue only resulted in the isolation of a Pr(III) decomposition product, showing the loss of a siloxide ligand and chloride coordination.5 The molecular structure of the decomposition product, [{PrIII(OSi(OtBu)3)3}2(μ-Cl)3(μ-K)3], suggested that chloride binding to the praseodymium is likely to provide a decomposition pathway for the putative Pr(IV) intermediate species. In order to prevent the unwanted halide abstraction, we reasoned that thianthrene tetrafluoroborate (thiaBF4) would be a convenient and innocent oxidizing agent (Eox = 0.86 V vs. Fc/Fc+), which is also stronger than magic blue (Eox = 0.67 V vs. Fc/Fc+). Notably, thiaBF4 would avoid the coordinating halides present in AgI or magic blue and would allow the oxidation reactions in non-coordinating solvents. The non-coordinating nature of the BF4− counter anion and the formation of highly soluble thianthrene side products would also facilitate the isolation of a molecular complex of Pr(IV). When carrying out this work, Zheng and coworkers reported the successful use of thiaBF4 as an oxidizing agent for the synthesis of molecular and supramolecular Tb(IV) complexes.44
First, we found that the addition of thiaBF4 (1.5 equiv.) to 1.0 equiv. of the Pr(III) complex, [KPr(OSi(OtBu)3)4], 1-PrOtBu, in acetonitrile/toluene (4
:
1) at −40 °C, resulted in an immediate color change and formation of an orange-brown suspension (Scheme 2 (top)). Isolation of the orange-brown solid and extraction into n-hexane yielded a red solid identified as [PrIV(OSi(OtBu)3)4] (2-PrOtBu) in 48% yield. Single crystals suitable for X-ray diffraction studies were obtained from a saturated solution of 2-PrOtBu in n-hexane at −40 °C after 24 h.
 |
| | Scheme 2 Top: synthesis and decomposition of [PrIV(OSi(OtBu)3)4], 2-PrOtBu. Bottom: synthesis of complexes, [M2PrIII(OSiPh3)5], 4M-PrPh, [MPrIV(OSiPh3)5], 5M-PrPh, (M = K, Cs) and 5[KDB18C6]-PrPh. | |
The solid-state molecular structure of complex 2-PrOtBu (Fig. 1) is isostructural to the previously reported Tb(IV) complex, [TbIV(OSi(OtBu)3)4] (2-TbOtBu). Both Ln(IV) ions are 5-coordinated by three κ1-OSi(OtBu)3 and one κ2-OSi(OtBu)3 ligands despite a 0.09 Å difference in their Ln(IV) ionic radii (6-coordinate Shannon radii: Pr(IV), 0.85 Å; Tb(IV), 0.76 Å).50 The same structure was also reported for the analogous Ce(IV) complex.51 The 0.16 Å difference in (Ln–O)avg bond distances between 2-PrOtBu and the previously5 reported 5-coordinate Pr(III) complex 1-PrOtBu (Table 1), is comparable to the 0.15 Å difference in the ionic radii determined for Pr(III) and Pr(IV) 6-coordinate compounds.50 The average Pr–Osiloxide distances in 2-PrOtBu (2.11(4) Å) are comparable to those reported for [PrIV(OSiPh3)4(CH3CN)2] (2.10(1) Å)5 (Table 1).
 |
| | Fig. 1 Molecular structure of [Pr(OSi(OtBu)3)4], 2-PrOtBu, with thermal ellipsoids drawn at the 50% probability level. Hydrogen atoms and methyl groups on the siloxide ligands have been omitted for clarity. | |
Table 1 Selected bond lengths (Å) and angles (deg) of five-coordinate Ln(IV) complex
|
|
2-PrOtBu
|
[TbIV(OSi(OtBu)V3)4]
|
[PrIV(OSiPh3)4(CH3CN)2]
|
5K-PrPh
|
5[KDB18C6]-PrPh
|
5Cs-PrPh
|
| Ln–Osiloxide range |
2.084(4)–2.147(4) |
2.023(3)–2.093(3) |
2.088(4)–2.121(4) |
2.098(6)–2.230(4) |
2.109(4)–2.173(4) |
2.134(8)–2.208(7) |
| Ln–OtBu |
2.563(4) |
2.474(3) |
— |
— |
— |
— |
| Ln–N(MeCN) |
— |
— |
2.599(6)–2.603(6) |
— |
— |
— |
The 1H NMR spectrum of 2-PrOtBu in tol-d8 and CD2Cl2 recorded at −40 °C showed the presence of a single broad resonance for the –OSi(OtBu)3 ligands at δ 1.85 and 1.53 ppm respectively (Fig. S4 and S5). Complex 2-PrOtBu is stable in tol-d8 and CD2Cl2 solution at −40 °C for up to one week but starts to decompose when the mixture is brought to room temperature, resulting in the dimeric PrIII complex, [PrIII(OSi(OtBu)3)3] (3-PrOtBu), and unknown byproducts, identified by the 1H NMR spectroscopy (Fig. S9). Single crystals of 3-PrOtBu were isolated from a toluene solution of 2-PrOtBu left at room temperature for 18 hours (Scheme 2 (top), Fig. 2). The solid-state molecular structure of the Pr(III) decomposition product, 3-PrOtBu, shows the presence of only three siloxide per Pr(III) centre, while the fate of the fourth siloxide ligand remains unknown.
 |
| | Fig. 2 Molecular structure of [Pr(OSi(OtBu)3)3]2, 3-PrOtBu, with thermal ellipsoids drawn at the 50% probability level. Hydrogen atoms and methyl groups on the siloxide ligands have been omitted for clarity. | |
Pr(IV) complexes comprising siloxide ligands were reported to exhibit characteristic absorption bands in the UV-visible region due to parity-allowed 4f–5d transitions. In accordance with these observations, we next examined the electronic absorption spectra of 1-PrOtBu, 2-PrOtBu, and the KLOtBu ligand in toluene. The electronic absorption spectra of 1-PrOtBu and KLOtBu ligand showed a sharp absorption in the UV range at a λmax = 276 nm. The spectrum of 2-PrOtBu displayed a characteristic signal in the UV-visible region with a maximum absorption wavelength (λmax) at 378 nm (Fig. S49 and S50).
Pentakis-triphenylsiloxide Pr(III/IV) complexes.
Considering the successful outcome of the oxidation of the complex 1-PrOtBu, we decided to pursue the synthesis of homoleptic five-coordinate Pr(IV) complexes that were previously unsuccessful using magic blue in THF or MeCN solutions. In particular, we pursued the synthesis of a pentakis-siloxide Pr(IV) complex, with the goal of evaluating differences in redox properties by addition of a fifth anionic ligand in the presence of different cations. It should be noted that all Pr(IV) complexes isolated so far are neutral.
We previously reported that pentakis-Ln(III) complexes46 of the –OSiPh3 ligand could only be prepared in apolar solvents (toluene), while the fifth siloxide ligand rapidly de-coordinates in THF or MeCN preventing the effective use of magic blue as oxidizing agent. Therefore, we set out to use thiaBF4 as an oxidant in toluene to isolate a five coordinate Pr(IV) complex of the –OSiPh3 ligand.
First, the previously reported [PrIII(OSiPh)3)3(THF)3]5 complex was reacted for an hour with 2.0 equiv. of KOSiPh3 in toluene at room temperature. Single crystals of the pentakis Pr(III) complex, [K2PrIII(OSi(Ph)3)5] (4K-PrPh), suitable for X-ray diffraction analysis were isolated in 82% yield from a saturated toluene solution stored at −40 °C. Complex [Cs2PrIII(OSiPh3)5] (4Cs-PrPh), was isolated in 86% yield by analogous reaction conditions as 4K-PrPh, but CsOSiPh3 was used. Additionally, the analogous Ce(III) penta-coordinated analogue, [K2CeIII(OSiPh3)5], (4K-CePh), was also prepared in 78% yield, and single crystals were grown from toluene solution layered with n-hexane at room temperature over 7 days. The quality of the XRD crystal structure allowed to determine the connectivity but does not allow to discuss metric parameters.
The solid-state molecular structure of complexes 4K-PrPh, 4Cs-PrPh and 4K-CePh (Fig. 3a, b and S48) are isostructural to that of the previously reported Nd analogue.46 The solid-state molecular structures of 4K-PrPh, 4Cs-PrPh and 4K-CePh show a five-coordinated Ln(III) metal center bound by five monodentate triphenylsiloxide ligands in a distorted trigonal bipyramidal geometry. The presence of five triphenylsiloxide ligands around the Pr(III) ion in 4K-PrPh does not significantly affect the Pr–Osiloxide distances (2.199(4)–2.315(3) Å) compared to those found in the six-coordinate tetrakis complex, [KPr(OSiPh3)4(THF)3] (2.248(9)–2.304(9) Å). Interestingly the Pr–Osiloxide bond distances (2.279(11)–2.318(9) Å)) for the 4Cs-PrPh are similar to those found for 4K-PrPh (Pr–Osiloxide: 2.199(4)-2.315(3)Å). The –OSiPh3 ligands bind the two potassium (K) cations in close proximity to the Pr(III) metal center through the anionic oxygen and cation–π interactions with the phenyl groups. The two K cations lie at a distance from the Pr(III) center of 3.675(1) Å for K1 and 3.764(1) Å for K2. The Cs cation lies at a significantly longer distance from the PrIII center (Pr(III) –Cs(1) = 4.073(1) and Pr(III) –Cs(2) = 4.092(1) Å) compared to the K ion in 4K-PrPh (3.675(1) Å for K1 and 3.764(1) Å for K2) probably due to steric hindrance of the larger Cs cation. These results suggest that although the lower Lewis acidity of the Cs cation compared to the K ion should result in a stronger PrIII-siloxide interaction, due to the larger size of the Cs, the Pr-siloxide bonding interactions remain similar.
With the electron-rich pentakis-siloxide complexes 4M-PrPh (M = K, Cs) in hand, we next investigated their oxidation reactions to yield tetravalent pentakis Pr(IV) complexes. Treatment of a cold toluene suspension of 4M-PrPh (K, Cs) with 1.0 equivalent of thiaBF4 at −40 °C, allowed the isolation of the Pr(IV) complexes, [KPr(OSi(Ph)3)5] (5K-PrPh) and [CsPr(OSi(Ph)3)5] (5Cs-PrPh) in 51% and 54% yields, respectively (Scheme 2 (bottom)).
Dark red single crystals of complexes 5M-PrPh (K, Cs) were isolated from a saturated toluene solution layered with n-hexane at −40 °C after 24 hours.
The solid-state molecular structure of complexes 5M-PrPh (K, Cs) displays a distorted trigonal bipyramidal geometry around the Pr(IV) center (Fig. 3c and d). The Pr(IV) ion in 5K-PrPh is coordinated by five anionic –OSiPh3 ligands with Pr–Osiloxide distances ranging from 2.098(6) to 2.230(4) Å, slightly longer than those found in the neutral six-coordinated tetrakis complex, [PrIV(OSiPh3)4(CH3CN)2] (2.088(4)–2.121(4) Å). The presence of the longer bond distances is most likely the result of steric interactions between the five siloxide ligands. The Pr(IV)–Osiloxide distances in 5Cs-PrPh are slightly shorter (2.134(8)–2.208(7) Å) compared to those in 5K-PrPh, most likely due to the presence of the less electron-withdrawing Cs cation compared to K. The K cation lies in close proximity of the Pr(IV) center (3.805(2) Å) and bridges the siloxide ligands through the anionic oxygen and through cation–π interactions with the siloxide phenyl groups. The Cs cation lies at a much larger distance (Cs–Pr = 4.120(2) Å).
 |
| | Fig. 3 Molecular structure of complexes, (a) [K2Pr(OSiPh)3)5], 4K-PrPh, (b) [Cs2Pr(OSiPh)3)5], 4Cs-PrPh, (c) [KPr(OSiPh)3)5], (5K-PrPh), and (d) [CsPr(OSiPh)3)5], (5Cs-PrPh), with thermal ellipsoids drawn at the 50% probability level. Phenyl groups bound to the alkali ions are drawn as wireframe for clarity. Hydrogen atoms on the siloxide ligands, residual solvent molecules and disordered phenyl groups have been omitted for clarity. | |
In hopes of evaluating inner-versus outer sphere cation effects, we found addition of 1.0 equiv. dibenzo-18-crown-6 (DB18C6) to a dichloromethane (DCM) solution of 5K-PrPh allowed the isolation of orange coloured single crystals of [KDB18C6][PrIV(OSiPh3)5], (5[KDB18C6]-PrPh) in 72% yield upon layering with n-hexane at −40 °C overnight. The solid state molecular structure of 5[KDB18C6]-PrPh shows an ion pair with an anionic five coordinated Pr(IV) complex and a potassium cation coordinated to DB18C6 (Fig. 4). For 5[KDB18C6]-PrPh the Pr–Osiloxide distances ranging from 2.109(4)–2.173(4) Å; are slightly shorter than in 5K-PrPh and closer to those found in 5Cs-PrPh complex, in agreement with stronger bonding of five anionic ligands with Pr(IV) in the absence of inner-sphere cation or in the presence of a weakly coordinating Cs ion.
 |
| | Fig. 4 Molecular structure of [KDB18C6][PrIV(OSiPh3)5], 5[KDB18C6]-PrPh, with thermal ellipsoids drawn at the 50% probability level. Phenyl groups of siloxide ligands in equatorial position are drawn as wireframe for clarity. Hydrogen atoms and the [KDB18C6] counterion have been omitted for clarity. | |
The electronic absorption spectra of 5K-PrPh, 5Cs-PrPh and 5[KDB18C6]-PrPh measured in DCM (0.15 mM) at room temperature showed a maximum absorption wavelengths (λmax) at 371, 364 and 355 nm in the UV-visible region characteristic of Pr(IV) (Fig. S53–S55).
To evaluate the solution stability of 5M-PrPh, we measured the time dependent 1H NMR spectra of 5M-PrPh at −40 °C and room temperature in CD2Cl2 and tol-d8. At −40 °C, we found both 5M-PrPh, complexes are stable for up to 14 days (Fig. S19, and S29). However, at room temperature, the formation of decomposition products begins after two hours (Figure S17, S22, S30 and S31), with full decomposition after 3 days. Further, it was observed that at −40 °C and at room temperature, the 1H NMR spectra of 5K-PrPh and 5Cs-PrPh are similar irrespective of the deuterated solvent suggesting that in both cases the cation remains bound in solution on the NMR time scale (Fig. S35 and S36). The presence of the bound K cation was confirmed by measuring the 1H NMR spectrum of 5K-PrPh in the presence of 1.0 equiv. of DB18C6 in CD2Cl2 at −40 °C and room temperature. The addition of DB18C6 results in the presence of three separate signals which are significantly shifted compared to those found for 5K-PrPh and 5Cs-PrPh.
Electrochemistry
Next, electrochemical studies were performed to evaluate the redox properties of the Pr(IV) complexes. Interestingly, the cyclic voltammogram (CV) of 2-PrOtBu obtained in THF and DCM solvents displayed different redox events (Fig. 5a and b). The CV in THF resulted in an irreversible reduction event at Epc = 0.20 V versus Fc/Fc+. In contrast, for the reported Tb(IV) analogue, [TbIV(OSi(OtBu)3)4],4 both oxidation and reduction events in THF (Epa = 0.83 V and Epc = −0.55 V versus Fc/Fc+) were reported (Fig. 5a), probably as a result of decreased ligand lability compared to the Pr(IV) complex.
 |
| | Fig. 5 Cyclic voltammograms of 2-PrOtBu and [TbIV(OSi(OtBu)3)4] complexes in (a) THF and (b) DCM. (c) Cyclic voltammograms of complexes, 5K-PrPh, 5Cs-PrPh, 5[KDB18C6]-PrPh and [PrIV(OSiPh3)4(CH3CN)2] in DCM. Electrolyte 0.1 M [NBu4][B(C6F5)4]; [analyte] = 0.005 M; υ = 100 mV s−1. | |
However, a contrasting behaviour was observed in non-coordinating DCM, where 2-PrOtBu showed reduction and oxidation events with Epa = 0.93 V and Epc = −0.12 V versus Fc/Fc+. The values measured for 2-PrOtBu are shifted to more positive potentials compared to the values measured in DCM for [TbIV(OSi(OtBu)3)4] (Epa = 0.89 V and Epc = −0.23 V versus Fc/Fc+). The large ΔE between anodic and cathodic peaks in both complexes most likely originates from the different binding modes of the –OSi(OtBu)3 ligands in the absence of coordinating cations in non-coordinating solvents for the Ln(III) (all k1-siloxides) and for the Ln(IV) (3 k1 and 1 k2 siloxide) complexes as already reported for the cerium analogue.34 Notably, upon addition of 2.2.2-cryptand to a solution of 1-PrOtBu, we isolated crystals of the [K(2.2.2cryptand)][PrIII(k1-OSi(OtBu)3)4] complex where all siloxides are binding in k1 mode (See SI and Fig. S42). The oxidation potential is only 0.04 V more positive for 2-PrOtBu and is in line with the calculated difference of 0.1 V between Pr(IV) and Tb(IV) ions.
The redox properties of 5M-PrPh (K, Cs) were also studied by cyclic voltammetry in DCM (Fig. 5c). The cyclic voltammogram of 5K-PrPh displays two irreversible reduction events at Epc1 = 0.13 V and Epc2 = −0.19 vs. Fc/Fc+ in DCM, most likely due to chemical equilibria resulting from the strong binding of the K cation. However, removal of K with DB18C6 resulted in the appearance of an oxidation (0.86 V) and reduction event (−1.33 V, ΔE = 2.19), resulting in a significantly greater peak separation for 5[KDB18C6]-PrPh. The large peak separation observed for 5[KDB18C6]-PrPh is most likely due to changes in the coordination sphere concurrent with changes in the Pr(IV) oxidation state and ion-pair formation effects occurring in the absence of coordinating cations. Furthermore, a quasi-reversible redox event was recorded for the Cs complex in DCM (Epc = 0.68 V and Epa = 0.89 vs. Fc/Fc+ in DCM). The oxidation potential measured for 5Cs-PrPh is 0.04 V less positive than that measured for 2-PrOtBu. A similar shift had also been reported for [TbIV(OSi(OtBu)3)4] (ref. 4) compared to [PrIV(OSiPh3)4(CH3CN)2].5 The higher reversibility and the shift of the reduction event to a more positive potential for complex 5Cs-PrPh compared to complex 5[KDB18C6]-PrPh is most likely due to the presence of bound Cs that facilitates the reduction process.
Surprisingly, the redox potentials measured for 5Cs-PrPh are similar to those reported for the six-coordinate Pr(IV) complex, [PrIV(OSiPh3)4(CH3CN)2] (Epa = 0.85 V and reduction potential Epc = 0.62 V vs. Fc/Fc+), suggesting that the replacement of two acetonitrile ligands by an anionic –OSiPh3 ligand does not significantly affect the redox properties (Table 2). The similarity of oxidation potential may be correlated to the presence of an overall weaker Pr(IV)–Osiloxide interaction in pentakis complexes due to the steric encumbrance created by the presence of the five bulky ligands. Notably, the Pr–Osiloxide distances are significantly elongated in 5Cs-PrPh compared to those found in [PrIV(OSiPh3)4(CH3CN)2] (Table 1) and 5[KDB18C6]-PrPh; DFT computational studies corroborated this interpretation (vide infra).
Table 2 Electrochemical data in V vs. Fc0/Fc+ for Ln(IV) complexes in DCM and/or THF(a) and previously reported data measured in THF(a*)4,5 CH3CN (b) with scan rate 100 mV s−1
| Complexes |
2-PrOtBu
|
[TbIV(OSi(OtBu)3)4]
|
[PrIV(OSiPh3)4(CH3CN)2]
|
5K-PrPh
|
5[KDB18C6]-PrPh
|
5Cs-PrPh
|
|
E
pc
|
0.20a; −0.12 |
−0.55a*; −0.23 |
−0.38a*; 0.48b; 0.62 |
0.13; −0.19 |
−1.33 |
0.68 |
|
E
pa
|
—; 0.93 |
0.83a*; 0.89 |
0.67a*; 0.69b; 0.85 |
— |
0.86 |
0.89 |
| ΔE |
1.05 |
1.38a*;1.12 |
1.05a; 0.21b; 0.23 |
— |
2.19 |
0.21 |
Magnetic measurements
The χMT versus T data (χM = molar magnetic susceptibility) were measured for 2-PrOtBu and compared with the previously reported51 4f1 isoelectronic cerium analogue [KCe(OSi(OtBu)3)4] in the temperature range 300 to 2 K (Fig. 6a). The χMT = 0.689 and 0.623 (at 300 K) and 0.079, 0.331 emu K mol−1 (at 2 K) were measured for 2-PrOtBu and [KCe(OSi(OtBu)3)4], respectively. The χMT values are in agreement with predicted values for a 4f1 complex (χMT = 0.8 emu K mol−1) and much lower than the 1.7 emu K mol−1 found for the 4f2 Pr(III) in 2-PrOtBu (predicted 1.6 emu K mol−1) (Fig. 6b). The χMT versus T data of pentakis 5K-PrPh, 4K-PrPh and 4K-CePh were also measured in the temperature range 300 to 2 K. and compared with previously reported [PrIV(OSiPh3)4(CH3CN)2] complex. The χMT = 0.516, 0.624 (at 300 K), and 0.126 0.297 emu K mol−1 (at 2 K) measured for 5K-PrPh and 4K-CePh respectively, are in agreement with the predicted 4f1 and are comparable to the values reported for [PrIV(OSiPh3)4 (CH3CN)2]5 (χMT = 0.622 emu K mol−1 at 300 K) (Fig. 6a). A value of χMT = 1.57 emu K mol−1 was measured for 4K-PrPh at 300 K in agreement with the presence of a 4f2 ion (Fig. 6b). Similar values of χMT were also measured for the recently reported Pr(IV) imidophosphorane ([Pr(NPtBu3)4]) (χMT = 0.64 emu K mol−1)8 although this value was found to be smaller than the analogous Ce(III) imidophosphorane ([Cs(THFx)Ce(NPtBu3)4]) (χMT = 0.89 emu K mol−1). These measurements further support the Pr(IV) oxidation state for 2-PrOtBu and 5K-PrPh. The comparative study of five coordinate Pr(IV) siloxide complexes with six and four coordinated analogues suggests that the geometry and coordination number has only marginal effect on the magnetic properties of Pr(IV) complexes with –OSiPh3 ligands. These results are in line with the presence of an overall similar magnitude of the crystal field for the Ce(III) and all Pr(IV) complexes.
 |
| | Fig. 6 (a) Plot of χMT versus temperature data for solid samples of 1-PrOtBu (red) 4K-PrPh (black), 2-PrOtBu (green), 5K-PrPh (olive green) and [PrIV(OSiPh3)4(CH3CN)2] (pink) collected under an applied magnetic field of 1 T. (b) Plot of χMT versus temperature data for solid samples of Pr(IV) and Ce(III) complexes collected under an applied magnetic field of 1 T. | |
To further confirm the Pr(IV) oxidation state of complex 2-PrOtBu, X-band electron paramagnetic resonance (EPR) spectra were measured at 6 K in the solid and solution-state (n-hexane) (10 mM). (Fig. S64). Further, EPR spectra of 5M-PrPh and 5[KDB18C6]-PrPh complexes were also recorded in the solid state at 6 K (Fig. S65–S67). All spectra of 2-PrOtBu, 5M-PrPh and 5[KDB18C6]-PrPh showed intense hyperfine structure as expected for the Pr4+ (4f1, Seff = ½, I = 5/2) ion52–54 that are consistent with the spectra reported by La Pierre and coworkers for complexes containing Pr(IV).8,48,55 We also found that the [PrIV(OSiPh3)4(CH3CN)2] complex, which we previously reported EPR silent, also showed a characteristic Pr(IV) spectrum when the measurements were carried out at relatively high concentration (20 mM) in acetonitrile at 6 K (Fig. S68).
Computational studies
To gain further insights, DFT calculations (B3PW91 functional) including dispersion effects have been carried out on complexes 2-PrOtBu and 5M-PrPh (K, Cs). In the latter, the effect of the alkali cation was also investigated by computing the putative [Pr(OSi(Ph)3)5]−, 5-PrPh anion. The optimized geometry of 2-PrOtBu is in very good agreement with the experimental values (see Table S11 in SI) with and without dispersion included. The ground state is a doublet with the unpaired spin located at the Pr center in line with a Pr(IV) complex. The bonds were thus analyzed using Natural Bonding Orbital (NBO) method. The presence of one Pr–Osiloxide single bond and three Pr–Osiloxide double bonds is found at the first order NBO. These bonds are found to be strongly polarized toward O (90%) and involve df (50–50) atomic orbital at the Pr center. These polarizations explain the Pr–Osiloxide Wiberg Bond Index of 0.70–0.77 despite the presence of double bonds. The extra Pr–OtBu interaction is even weaker since only weak donation is found at the second order donor–acceptor level with a WBI of 0.24. For sake of comparison, the Ce(IV) analog of 2-PrOtBu was optimized and compared since with a closed-shell singlet ground state the oxidation state + IV of Ce is obvious. The geometry and bonding situation are very similar to that described above which is further corroborating the presence of a Pr(IV) center. Finally, the decomposition of 2-PrOtBu onto 3-PrOtBu was computed to be marginally favorable (−2.5 kcal.mol−1) assuming that the fourth siloxide ligand forms a dimer. The bonding of this Pr(III) complex is similar to that found for 1-PrOtBu and its Ce(III) analog (see SI), where the Pr–Osiloxide are only single bond (polarized toward O). This can be easily explained by the fact that due to the increased charge at the metal center in 2-PrOtBu the Ln–O distances are shorter for the tetravalent state allowing π-type interactions.
A similar analysis was carried out on complexes 5M-PrPh (M = K, Cs). In both cases, the optimized geometry obtained for the doublet ground state compare very well with the experimental one (see Table S26 and S30 in SI). The unpaired spin located at the Pr centre is again in line with the presence of Pr(IV) complexes. The bonding analysis is somewhat similar to that observed for 2-PrOtBu or [PrIV(OSiPh3)4(CH3CN)2] (see SI) with the presence of one Pr–Osiloxide single bond and four (rather than three) Pr–Osiloxide double bonds, strongly polarized toward O (92–94%). However, the associated WBI appears to differ significantly from one alkali atom to the other and from that found for 2-PrOtBu. Therefore, for sake of comparison, the penta-coordinated alkali free anion was computed and analyzed ([5-PrPh]−). In 5Cs-PrPh and [5-PrPh]−, the Pr–Osiloxide are in the 0.64–0.77 range which is somewhat similar to that found for 2-PrOtBu and its Ce(IV) analog but also to [PrIV(OSiPh3)4(CH3CN)2]. This clearly indicates that the presence of the Cs cation does not perturb the bonding, meaning that Cs does not interact significantly with the oxygen of the siloxide ligands, in line with the electrochemistry observation. In a same way, the similarity between the CVs of [PrIV(OSiPh3)4(CH3CN)2] and of 5Cs-PrPh and [5-PrPh]− is explained by the similar bonding situation in the three complexes largely due to steric effects preventing a strong interaction with the five siloxide ligands in 5Cs-PrPh and [5-PrPh]− as evidenced by the increased Pr–O distances in the latter (around 2.15 Å in average) compared to [PrIV(OSiPh3)4(CH3CN)2] (2.08 Å In average). On the other hand, in the 5K-PrPh, two Pr–Osiloxide WBI are reduced to 0.57–0.60 while the others remain in the 0.74–0.75 range. This is due to the presence of strong K–Osiloxide interactions, that reduce the electron density at the oxygen of the siloxide ligand resulting in a less efficient overlap with the Pr(IV) centre.
Experimental
Synthesis of [PrIV(OSi(OtBu)3)4] (2-PrOtBu)
In an argon filled glovebox, a cold (−40 °C) purple solution of thiaBF4 (0.030 g, 0.097 mmol, 1.5 equiv.) in CH3CN (0.8 mL) was added dropwise, under stirring to a cold (−40 °C) colorless solution of [KPr(OSi(OtBu)3)4] (1-PrOtBu), prepared as previously reported,5 (0.080 g, 0.065 mmol, 1 equiv.) in toluene (0.2 mL). The resulting purple suspension was stirred at −40 °C in freezer for 15 min. After 15 min the color of the suspension changed to red purple. The suspension was filtered and the red solid collected over a porosity 4 filter-frit, at −40 °C and then washed with MeCN (2 × 3 mL) to remove the byproducts. The red solid was then dissolved in cold (−40 °C) n-hexane (0.8 mL), resulting in a dark red solution which was filtered over a 0.22 μm porosity filter frit. The cold red n-hexane solution was concentrated under reduced pressure to yield 49% (0.038 g, 0.032 mmol) a red solid. X-ray quality single red crystals of 2-PrOtBu were obtained overnight from concentrated n-hexane solution of 2-PrOtBu at −40 °C. 1H NMR (400 MHz, Tol-d8, 233 K) δ 1.85 (br, OC(CH3)3), ppm. 1H NMR (400 MHz, CD2Cl2, 233 K) δ 1.53 (br, OC(CH3)3), ppm. Anal. Cal. For C48H108O16Si4Pr: C, 48.26; H, 9.11; N, 0.00. Found: C, 47.64; H, 9.13; N, 0.00.
Synthesis of [K2PrIII(OSi(Ph)3)5] (4K-PrPh)
A solution of KOSiPh3 (56 mg, 0.178 mmol, 2.1 equiv.) in toluene (1 mL) was added at room temperature to a pale green solution of [Pr(OSiPh3)3(THF)3], prepared as previously reported5,56 (100 mg, 0.85 mmol, 1 equiv.) in toluene (1 mL). The resultant pale green solution was stirred for 1 hour at room temperature. The volatiles were removed under vacuum, the residue was triturated with n-hexane (3 × 2 mL) and dried for 30 min affording white solid. Toluene (1.0 mL) was added, and the colorless solution storage at −40 °C affording white crystalline powder of [K2Pr(OSi(Ph)3)5], 4K-PrPh, in 81% yield (110 mg, 0.069 mmol). Anal. Calcd for [K2Pr(OSi(Ph)3)5]. (Toluene) (1688.10 g mol−1): C97H83K2O5PrSi5: C, 69.01; H, 4.96; N, 0.00. Found: C, 68.91; H, 4.94; N, 0.00. The fractional toluene content is residual co-crystallized solvent left from partial drying also observed in 1H NMR spectrum. 1H NMR (CD2Cl2, 400 MHz, 298 K): δ = 6.73 (s, 15H, Ph), 5.66 (s, 30H, Ph), 4.47 (br, s, 30H, Ph) ppm. Single crystals suitable for X-ray diffraction analysis were obtained from a toluene solution of 4K-PrPh layered with n-hexane at rt.
Synthesis of [KPrIV(OSi(Ph)3)5] (5K-PrPh)
To a colorless solution of 4K-PrPh (100 mg, 0.063 mmol, 1 equiv.) in toluene (1 mL), dark purple solid of thiaBF4 (21 mg, 0.069 mmol, 1.1 equiv.) was added at −40 °C and reaction mixture was stirred for 30 min at −40 °C. During the stirring color of reaction mixture was changed to deep red. The red color reaction mixture was filtered off over porosity 4 filter-frit. Deep red filtrate was evaporated to dryness under reduced pressure to get dark red solid. The dark red solid was redissolved in toluene (1 mL) and layered with n-hexane at −40 °C to get dark red crystals of 5K-PrPh in 51% yield. (50 mg, 0.032 mmol). Anal. Calcd for [KPr(OSi(Ph)3)5]. (Toluene)2 (1741.22 g mol−1): C104H91O5Si5PrK: C, 71.73; H, 5.27; N, 0.00. Found: C, 71.50; H, 5.29; N, 0.00. The toluene content is residual co-crystallized solvent left from partial drying and also observed in 1H NMR spectrum. 1H NMR (Tol-d8, 400 MHz, 233 K): δ = 6.95–6.91 (t, 30H, Ph), 6.74 (s, 45H, Ph) ppm. 1H NMR (Tol-d8, 400 MHz, 298 K): δ = 6.95–6.91 (t, 30H, Ph), 6.73 (s, 45H, Ph) ppm. 1H NMR (CD2Cl2, 400 MHz, 233 K): δ = 7.04 (br, s, 15H, Ph), 6.68 (br, s, 60H, Ph) ppm. 1H NMR (CD2Cl2, 400 MHz, 298 K): δ = 7.07 (br, s, 15H, Ph), 6.72 (br, s, 60H, Ph) ppm.
Synthesis of [KDB18C6][PrIV(OSiPh3)5] (5[KDB18C6]-PrPh)
To a dark red solution of 5K-PrPh (50 mg, 0.032 mmol, 1.0 equiv.) in DCM (0.5 mL), white solid of dibenzo-18-crown-6 (DB18C6) (11.5 mg, 0.032 mmol, 1.0 equiv.) was added and reaction mixture was stirred for 5 min at rt. During the stirring color of reaction mixture was changed to orange from dark red. The orange-colored solution was layered with n-hexane and kept at −40 °C to get orange crystals of 5[KDB18C6]-PrPh in 72% yield. (44 mg, 0.022 mmol). Anal. Calcd for 5[KDB18C6]-PrPh. (CH2Cl2) (2002.33 g mol−1): C111H101O11Si5PrCl2: C, 66.58; H, 5.08; N, 0.00. Found: C, 66.15; H, 5.08; N, 0.01. The DCM content is residual co-crystallized solvent left from partial drying.1H NMR (CD2Cl2, 400 MHz, 233 K): δ = 7.32 (s, 15H, Ph), 6.94 (t, 30H, Ph), 6.90 (m, 8H, DB18C6), 6.57 (d, 30H, Ph), 4.14 (s, 8H, DB18C6), 3.91 (s, 8H, DB18C6) ppm. 1H NMR (CD2Cl2, 400 MHz, 298 K): δ = 7.26 (s, 15H, Ph), 7.04–7.01 (m, 8H, DB18C6), 6.94 (t, 30H, Ph), 6.63 (s, 30H, Ph), 4.21 (t, 8H, DB18C6), 3.98 (t, 8H, DB18C6) ppm.
Conclusions
In summary, we have identified reaction conditions (solvent, oxidizing agent, and cation effect) that have allowed the synthesis and full characterization of four new complexes of Pr(IV), while presenting an unprecedented penta-coordination environment. The use of the strong oxidizing agent, thiaBF4, prevented previously observed decomposition pathways of the complex, [PrIV(OSi(OtBu)3)4] (2-PrOtBu), and allowed its isolation in the solid state, which was found to be isostructural with the previously reported Tb(IV) analogue. Despite the very similar oxidation potential measured for 2-PrOtBu, and for its previously reported Tb(IV) analogue, the Pr(IV) complex is significantly more labile and more prone to undergo fast decomposition pathways rendering the choice of the oxidizing agents crucial for its isolation. The isolated complex is relative stable in cold toluene and DCM but decomposes rapidly at room temperature to afford the Pr(III) dimeric complex, [Pr(OSi(OtBu)3)3]2 (3-PrOtBu). The oxidation of the Pr(III) pentakis-triphenylsiloxide complexes, 4M-PrPh (K, Cs) with thiaBF4 in toluene allowed the isolation of the first example of anionic penta-coordinated Pr(IV) complexes. The complexes were both isolated with inner-sphere bound K and Cs cations as complexes, 5M-PrPh (K, Cs), and as an ion pair, 5[KDB18C6]-PrPh, for complex 5K-PrPh. The solution stability of these pentakis complexes is slightly increased compared to the previously reported tetrakis complex, [PrIV(OSiPh3)4(CH3CN)2] although we found that the presence of an additional anionic ligand does not result in a complex that is easier to oxidize. Computational and structural analysis suggests that steric hindrance prevents the formation of strong π-bonding interactions between the –OSiPh3 ligands and the Pr(IV) center resulting in an overall similar bonding mode in both tetrakis-and pentakis-complexes. All Pr(IV) complexes show similar magnetic properties independent of the geometry and number of –OSiPh3 ligands, which are also similar to the Ce(III) analogues in agreement with the computed electronic structures. The possibility of accessing charged Pr(IV) complexes provide a tool to manipulate the redox potential and therefore access to more stable complexes with the same ligand, but care needs to be taken in preventing unfavorable steric interactions. Future studies will be directed to explore this route to access a broad range of Pr(IV) complexes.
Author contributions
P. P. designed and carried out all the experiments and analysed the data; M. K. performed preliminary experiments. M. M. designed and supervised the project and analysed the data; T. R. and L. M. carried out the computational study; R. S. measured and analyzed the X-ray data, I. Z. measured and analysed the magnetic data; A. S. measured and analysed the EPR data; P. P, L. M. and M. M. wrote the manuscript with contributions of all authors, and all authors have given approval for the final version of the manuscript.
Conflicts of interest
There are no conflicts to declare.
Data availability
The data that support the findings of this study are openly available in the Zenodo repository, https://doi.org/10.5281/zenodo.17280527.
CCDC 2470050–2470056 and 2471366 contain the supplementary crystallographic data for this paper.57a–h
Supplementary information: synthetic details, analytical data including depictions of all spectra and coordinate data of all computationally optimised species, are documented in the supplementary information (SI). See DOI: https://doi.org/10.1039/d5sc05500h.
Acknowledgements
We acknowledge support from the Swiss National Science Foundation grant number 231638 and the Ecole Polytechnique Fédérale de Lausanne (EPFL). We also thank Dr Anne-Sophie Chauvin for her assistance with the synthesis of thiaBF4. LM is a senior member of the Institut Universitaire de France. CalMip is acknowledged for a generous grant of computing time.
References
- J. C. Wedal and W. J. Evans, A Rare-Earth Metal Retrospective to Stimulate All Fields, J. Am. Chem. Soc., 2021, 143, 18354–18367 CrossRef CAS PubMed.
-
D. H. Woen, and W. J. Evans, U. Expanding the +2 Oxidation State of the Rare-Earth Metals, and Thorium in Molecular Complexes, in Handbook on the Physics and Chemistry of Rare Earths, Elsevier B.V., 2016, pp. 1–57, DOI:10.1016/bs.hpcre.2016.08.002.
- L. Barluzzi, S. R. Giblin, A. Mansikkamäki and R. A. Layfield, Identification of Oxidation State+1 in a Molecular Uranium Complex, J. Am. Chem. Soc., 2022, 144, 18229–18233 CrossRef CAS PubMed.
- C. T. Palumbo, I. Zivkovic, R. Scopelliti and M. Mazzanti, Molecular Complex of Tb in the +4 Oxidation State, J. Am. Chem. Soc., 2019, 141, 9827–9831 CrossRef CAS PubMed.
- A. R. Willauer, C. T. Palumbo, F. Fadaei-Tirani, I. Zivkovic, I. Douair, L. Maron and M. Mazzanti, Accessing the plus IV Oxidation State in Molecular Complexes of Praseodymium, J. Am. Chem. Soc., 2020, 142, 5538–5542 CrossRef CAS PubMed.
- A. C. Boggiano, C. M. Studvick, S. R. Chowdhury, J. E. Niklas, H. Tateyama, H. W. Wu, J. E. Leisen, F. Kleemiss, B. Vlaisavljevich, I. A. Popov and H. S. La Pierre, Praseodymium in the formal +5 oxidation state, Nat. Chem., 2025, 17, 1005–1010 CrossRef CAS PubMed.
- M. R. MacDonald, M. E. Fieser, J. E. Bates, J. W. Ziller, F. Furche and W. J. Evans, Identification of the +2 Oxidation State for Uranium in a Crystalline Molecular Complex, K(2.2.2-Cryptand) (C5H4SiMe3)(3)U, J. Am. Chem. Soc., 2013, 135, 13310–13313 CrossRef CAS PubMed.
- A. C. Boggiano, S. R. Chowdhury, M. D. Roy, M. G. Bernbeck, S. M. Greer, B. Vlaisavljevich and H. La Pierre, A Four-Coordinate Pr4+ Imidophosphorane Complex, Angew. Chem., Int. Ed. Engl., 2024, e202409789 CAS.
- M. J. Beltran-Leiva, W. N. G. Moore, T. F. Jenkins, W. J. Evans, T. E. Albrecht and C. Celis-Barros, A comprehensive approach for elucidating the interplay between 4fn+1 and 4fn5d1 configurations in Ln2+ complexes, Chem. Sci., 2025, 16, 2024–2033 RSC.
- T. N. Poe, H. Ramanantoanina, J. M. Sperling, H. B. Wineinger, B. M. Rotermund, J. Brannon, Z. L. Bai, B. Scheibe, N. Beck, B. N. Long, S. Justiniano, T. E. Albrecht-Schönzart and C. Celis-Barros, Isolation of a californium(II) crown-ether complex, Nat. Chem., 2023, 15, 722–728 CrossRef CAS PubMed.
- W. L. Li, T. T. Chen, W. J. Chen, J. Li and L. S. Wang, Monovalent lanthanide(I) in borozene complexes, Nat. Commun., 2021, 12, 6467 CrossRef CAS PubMed.
- R. R. Langeslay, M. E. Fieser, J. W. Ziller, F. Furche and W. J. Evans, Synthesis, structure, and reactivity of crystalline molecular complexes of the {[C5H3(SiMe3)2]3Th}1- anion containing thorium in the formal +2 oxidation state, Chem. Sci., 2015, 6, 517–521 RSC.
- J. Su, C. J. Windorff, E. R. Batista, W. J. Evans, A. J. Gaunt, M. T. Janicke, S. A. Kozimor, B. L. Scott, D. H. Woen and P. Yang, Identification of the Formal +2 Oxidation State of Neptunium: Synthesis and Structural Characterization of {Np-II C5H3(SiMe3)(2) (3)}(1-), J. Am. Chem. Soc., 2018, 140, 7425–7428 CrossRef CAS PubMed.
- N. T. Rice, I. A. Popov, D. R. Russo, J. Bacsa, E. R. Batista, P. Yang, J. Telser and H. S. La Pierre, Design, Isolation, and Spectroscopic Analysis of a Tetravalent Terbium Complex, J. Am. Chem. Soc., 2019, 141, 13222–13233 CrossRef CAS PubMed.
- M. Liu, Y. C. Chen, H. Wang, T. Shang, M. L. Tong, R. A. Layfield, A. Mansikkamäki and F. S. Guo, A Linear Dysprosium(II) Metallocene with a High Effective Energy Barrier and Magnetic Hysteresis up to 70 Kelvin, J. Am. Chem. Soc., 2025, 147, 18307–18316 CrossRef CAS PubMed.
- J. Moutet, J. Schleinitz, L. La Droitte, M. Tricoire, F. Pointillart, F. Gendron, T. Simler, C. Clavaguéra, B. Le Guennic, O. Cador and G. Nocton, Bis-Cyclooctatetraenyl Thulium(II): Highly Reducing Lanthanide Sandwich Single-Molecule Magnets, Angew Chem. Int. Ed. Engl., 2021, 60, 6042–6046 CrossRef CAS PubMed.
- C. A. Gould, K. R. McClain, J. M. Yu, T. J. Groshens, F. Furche, B. G. Harvey and J. R. Long, Synthesis and Magnetism of Neutral, Linear Metallocene Complexes of Terbium(II) and Dysprosium(II), J. Am. Chem. Soc., 2019, 141, 12967–12973 CrossRef CAS PubMed.
- K. R. Meihaus, M. E. Fieser, J. F. Corbey, W. J. Evans and J. R. Long, Record High Single-Ion Magnetic Moments Through 4fn5d1 Electron Configurations in the Divalent Lanthanide Complexes (C5H4SiMe3)3Ln, J. Am. Chem. Soc., 2015, 137, 9855–9860 CrossRef CAS PubMed.
- R. A. K. Shivaraam, T. Rajeshkumar, R. Scopelliti, I. Zuivkovic, L. Maron and M. Mazzanti, Dinitrogen Reduction and Functionalization by a Siloxide Supported Thulium-Potassium Complex for the Formation of Ammonia or Hydrazine Derivatives, Angew Chem. Int. Ed. Engl., 2025, 64, e202414051 CrossRef CAS PubMed.
- E. Papangelis, L. Demonti, I. del Rosal, A. Shephard, L. Maron, G. Nocton and T. Simler, Room-Temperature H2 Splitting and N2-Hydrogenation Induced by a Neutral LuII Complex, J. Am. Chem. Soc., 2025, 147, 9752–9763 CrossRef CAS PubMed.
- A. Mondal, C. G. T. Price, J. K. Tang and R. A. Layfield, Targeted Synthesis of End-On Dinitrogen-Bridged Lanthanide Metallocenes and Their Reactivity as Divalent Synthons, J. Am. Chem. Soc., 2023, 145, 20121–20131 CrossRef CAS PubMed.
- W. J. Evans, G. Zucchi and J. W. Ziller, Dinitrogen reduction by Tm(II), Dy(II), and Nd(II) with simple amide and aryloxide ligands, J. Am. Chem. Soc., 2003, 125, 10–11 CrossRef CAS PubMed.
- W. J. Evans, N. T. Allen and J. W. Ziller, Expanding divalent organolanthanide chemistry: The first organothulium(II) complex and the in situ organodysprosium(II) reduction of dinitrogen, Angew Chem. Int. Ed. Engl., 2002, 41, 359 CrossRef CAS PubMed.
- W. J. Evans, N. T. Allen and J. W. Ziller, Facile dinitrogen reduction via organometallic Tm(II) chemistry, J. Am. Chem. Soc., 2001, 123, 7927–7928 CrossRef CAS PubMed.
- F. Jaroschik, A. Momin, F. Nief, X. F. Le Goff, G. B. Deacon and P. C. Junk, Dinitrogen Reduction and C-H Activation by the Divalent Organoneodymium Complex (C5H2tBu3)2Nd(μ-I)K( 18 crown-6), Angew Chem. Int. Ed. Engl., 2009, 48, 1117–1121 CrossRef CAS PubMed.
- C. M. Kotyk, M. E. Fieser, C. T. Palumbo, J. W. Ziller, L. E. Darago, J. R. Long, F. Furche and W. J. Evans, Isolation of+2 rare earth metal ions with three anionic carbocyclic rings: bimetallic bis(cyclopentadienyl) reduced arene complexes of La(2+)and Ce2+ are four electron reductants, Chem. Sci., 2015, 6, 7267–7273 RSC.
- T. Simler, K. N. McCabe, L. Maron and G. Nocton, CO reductive oligomerization by a divalent thulium complex and CO2-induced functionalization, Chem. Sci., 2022, 13, 7449–7461 RSC.
- A. R. Willauer, D. Toniolo, F. Fadaei-Tirani, Y. Yang, M. Laurent and M. Mazzanti, Carbon dioxide reduction by dinuclear Yb(II) and Sm(II) complexes supported by siloxide ligands, Dalton Trans., 2019, 48, 6100–6110 RSC.
- G. B. Deacon, P. C. Junk, J. Wang and D. Werner, Reactivity of Bulky Formamidinatosamarium(II or III) Complexes with C=O and C=S Bonds, Inorg. Chem., 2014, 53, 12553–12563 CrossRef CAS PubMed.
- A. J. Gremillion, J. Ross, X. J. Yu, P. Ishtaweera, R. Anwander, J. Autschbach, G. A. Baker, S. P. Kelley and J. R. Walensky, Facile Oxidation of Ce(III) to Ce(IV) Using Cu(I) Salts, Inorg. Chem., 2024, 63, 9602–9609 CrossRef CAS PubMed.
- Y. S. Qiao, H. L. Yin, L. M. Moreau, R. L. Feng, R. F. Higgins, B. C. Manor, P. J. Carroll, C. H. Booth, J. Autschbach and E. J. Schelter, Cerium(iv) complexes with guanidinate ligands: intense colors and anomalous electronic structures, Chem. Sci., 2021, 12, 3558–3567 RSC.
- L. A. Solola, A. V. Zabula, W. L. Dorfner, B. C. Manor, P. J. Carroll and E. J. Schelter, Cerium(IV) Imido Complexes: Structural, Computational, and Reactivity Studies, J. Am. Chem. Soc., 2017, 139, 2435–2442 CrossRef CAS PubMed.
- N. A. Piro, J. R. Robinson, P. J. Walsh and E. J. Schelter, The electrochemical behavior of cerium(III/IV) complexes: Thermodynamics, kinetics and applications in synthesis, Coord. Chem. Rev., 2014, 260, 21–36 CrossRef CAS.
- J. Friedrich, Y. S. Qiao, C. Maichle-Mossmer, E. J. Schelter and R. Anwander, Redox-enhanced hemilability of a tris(tert-butoxy) siloxy ligand at cerium, Dalton Trans., 2018, 47, 10113–10123 RSC.
- R. Anwander, M. Dolg and F. T. Edelmann, The difficult search for organocerium(IV) compounds, Chem. Soc. Rev., 2017, 46, 6697–6709 RSC.
- P. Drose, A. R. Crozier, S. Lashkari, J. Gottfriedsen, S. Blaurock, C. G. Hrib, C. Maichle-Mossmer, C. Schadle, R. Anwander and F. T. Edelmann, Facile Access to Tetravalent Cerium Compounds: One-Electron Oxidation Using Iodine(III) Reagents, J. Am. Chem. Soc., 2010, 132, 14046–14047 CrossRef CAS PubMed.
- H. L. Yin, P. J. Carroll, B. C. Manor, J. M. Anna and E. J. Schelter, Cerium Photosensitizers: Structure-Function Relationships and Applications in Photocatalytic Aryl Coupling Reactions, J. Am. Chem. Soc., 2016, 138, 5984–5993 CrossRef CAS PubMed.
- Q. M. Yang, Y. H. Wang, Y. S. Qiao, M. Gau, P. J. Carroll, P. J. Walsh and E. J. Schelter, Photocatalytic C-H activation and the subtle role of chlorine radical complexation in reactivity, Science, 2021, 372, 847–852 CrossRef CAS PubMed.
- Y. M. So and W. H. Leung, Recent advances in the coordination chemistry of cerium(IV) complexes, Coord. Chem. Rev., 2017, 340, 172–197 CrossRef CAS.
- T. P. Gompa, A. Ramanathan, N. T. Rice and H. S. La Pierre, The chemical and physical properties of tetravalent lanthanides: Pr, Nd, Tb, and Dy, Dalton Trans., 2020, 49, 15945–15987 RSC.
- A. R. Willauer, C. T. Palumbo, R. Scopelliti, I. Zivkovic, I. Douair, L. Maron and M. Mazzanti, Stabilization of the Oxidation State plus IV in Siloxide-Supported Terbium Compounds, Angew Chem. Int. Ed. Engl., 2020, 59, 3549–3553 CrossRef CAS PubMed.
- M. Tricoire, F. C. Hsueh, M. Keener, T. Rajeshkumar, R. Scopelliti, I. Zivkovic, L. Maron and M. Mazzanti, Siloxide tripodal ligands as a scaffold for stabilizing lanthanides in the+4 oxidation state, Chem. Sci., 2024, 15, 6874–6883 RSC.
- T. Xue, Y.-S. Ding, X.-L. Jiang, L. Tao, J. Li and Z. Zheng, Tetravalent Terbium Chelates: Stability Enhancement and Property Tuning, Precis. Chem., 2023, 1, 583–591 CrossRef CAS PubMed.
- T. J. Xue, Q. S. Yang, L. Li, X. Y. Chang, Y. S. Ding and Z. P. Zheng, Supramolecular assemblies of tetravalent terbium complex units: syntheses, structure, and materials properties, Chem. Sci., 2025, 16, 6805–6811 RSC.
-
S. Cotton, Lanthanides and Actinides Chemistry, John Wiley & Sons, Chichester, 2006 Search PubMed.
- A. R. Willauer, I. Douair, A. S. Chauvin, F. Fadaei-Tirani, J. C. Bünzli, L. Maron and M. Mazzanti, Structure, reactivity and luminescence studies of triphenylsiloxide complexes of tetravalent lanthanides, Chem. Sci., 2022, 13, 681–691 RSC.
- T. J. Xue, Y. S. Ding and Z. P. Zheng, A tetravalent praseodymium complex with field-induced slow magnetic relaxation, Dalton Trans., 2024, 53, 5779–5783 RSC.
- N. T. Rice, I. A. Popov, R. K. Carlson, S. M. Greer, A. C. Boggiano, B. W. Stein, J. Bacsa, E. R. Batista, P. Yang and H. S. La Pierre, Spectroscopic and electrochemical characterization of a Pr4+ imidophosphorane complex and the redox chemistry of Nd3+ and Dy3+ complexes, Dalton Trans., 2022, 51, 6696–6706 RSC.
- Z. F. Guo, G. B. Deacon and P. C. Junk, When is a reported air-stable PrIV complex not a PrIV complex? Answer: when it is PrIII, J. Coord. Chem., 2025, 78, 25–30 CrossRef CAS.
- R. D. Shannon, Revised Effective Ionic-Radii and Systematic Studies of Interatomic Distances in Halides and Chalcogenides, Acta Crystallogr. A, 1976, 32, 751–767 CrossRef.
- R. P. Kelly, L. Maron, R. Scopelliti and M. Mazzanti, Reduction of a Cerium(III) Siloxide Complex To Afford a Quadruple-Decker Arene-Bridged Cerium(II) Sandwich, Angew Chem. Int. Ed. Engl., 2017, 56, 15663–15666 CrossRef CAS PubMed.
- Y. Hinatsu and N. Edelstein, Electron paramagnetic resonance spectra of Pr4+ ions doped in BaMO3 (M=Ce, Zr, Sn) and SrCeO3, J. Alloys Compd., 1997, 250, 400–404 CrossRef CAS.
- K. Tezuka and Y. Hinatsu, Electron paramagnetic resonance spectra of Pr4+ ions doped in pyrochlore-type compounds La2Sn2O7 and La2Zr2O7, J. Solid State Chem., 1999, 143, 140–143 CrossRef CAS.
- Y. Hinatsu, Magnetic Susceptibilities and Electron Paramagnetic Resonance Spectra of Lithium-Rich Lanthanide Oxides Li8PrO6 and Li8TbO6, J. Solid State Chem., 1997, 128, 228–232 CrossRef CAS.
- A. Ramanathan, E. D. Walter, M. Mourigal and H. S. La Pierre, Increased Crystal Field Drives Intermediate Coupling and Minimizes Decoherence in Tetravalent Praseodymium Qubits, J. Am. Chem. Soc., 2023, 145, 17603–17612 CrossRef CAS PubMed.
- W. J. Evans, R. E. Golden and J. W. Ziller, A Comparative Synthetic and Structural Study of Triphenylmethoxide and Triphenylsiloxide Complexes of the Early Lanthanides, Including X-Ray Crystal-Structures of La2(OCPh3)6 and Ce2(OSiPh3)6, Inorg. Chem., 1991, 30, 4963–4968 CrossRef CAS.
-
(a)
CCDC 2470050, Experimental Crystal Structure Determination, 2025, DOI:10.5517/ccdc.csd.cc2nx91y;
(b)
CCDC 2470051, Experimental Crystal Structure Determination, 2025, DOI:10.5517/ccdc.csd.cc2nx92z;
(c)
CCDC 2470052, Experimental Crystal Structure Determination, 2025, DOI:10.5517/ccdc.csd.cc2nx930;
(d)
CCDC 2470053, Experimental Crystal Structure Determination, 2025, DOI:10.5517/ccdc.csd.cc2nx941;
(e)
CCDC 2470054, Experimental Crystal Structure Determination, 2025, DOI:10.5517/ccdc.csd.cc2nx952;
(f)
CCDC 2470055, Experimental Crystal Structure Determination, 2025, DOI:10.5517/ccdc.csd.cc2nx963;
(g)
CCDC 2470056, Experimental Crystal Structure Determination, 2025, DOI:10.5517/ccdc.csd.cc2nx974;
(h)
CCDC 2471366, Experimental Crystal Structure Determination, 2025, DOI:10.5517/ccdc.csd.cc2nynhs.
|
| This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.