Open Access Article
Angus C. G.
Shephard†
a,
Lucie
Pedussaut†
a,
Linda
De Marchi
a,
Luca
Demonti
a,
Thayalan
Rajeshkumar
b,
Nicolas
Casaretto
a,
Laurent
Maron
b,
Grégory
Danoun
a,
Thomas
Simler
*a and
Grégory
Nocton
*a
aLCM, CNRS, École Polytechnique, Institut Polytechnique de Paris, Route de Saclay, 91120 Palaiseau, France. E-mail: thomas.simler@polytechnique.edu; gregory.nocton@polytechnique.edu
bLPCNO, Université de Toulouse-CNRS, INSA, UPS, UMR 5215, Toulouse, France
First published on 14th October 2025
The conversion of dinitrogen into ammonia plays an important role in sustaining life on Earth and serves as a significant building block for our planet's future. The Haber–Bosch process, although a well-established method for converting hydrogen and nitrogen gases into ammonia using metal-based heterogeneous catalysts, requires an extensive industrial infrastructure, limiting its accessibility and flexibility. Molecular systems, whether supported or unsupported, offer the advantage of allowing fine-tuning of the metal properties and the involved elementary steps, which ultimately leads to a better understanding of the transformations. In this context, we present findings on the reactivity of dinitrogen with an organometallic uranium complex featuring the bulky Cpttt ligand (Cpttt = 1,2,4-tris(tert-butyl)cyclopentadienyl). This complex demonstrates the ability to cleave and hydrogenate dinitrogen under mild conditions, at ambient temperature and atmospheric pressure. Most notably, the rich redox chemistry of uranium enables the direct reduction of N2 into a unique formal UIV dimer featuring an end-on coordinated (N2)4− bridging ligand, the cornerstone of the observed reactivity.
000 catalytic cycles.16
In contrast, the molecular direct hydrogenation of dinitrogen, which is similar to the industrial Haber–Bosch process, yields substoichiometric amounts of ammonia after N2 cleavage.23–27 Besides transition metals, actinides are ideal candidates for nitrogen cleavage and functionalization due to their rich redox chemistry and proven ability to fix nitrogen.28–32 Several examples of uranium complexes have illustrated the reduction of dinitrogen and its subsequent cleavage, assisted by nearby cations or amido-phosphine ligands, as demonstrated by the Mazzanti33–36 and Zhu25,37,38 groups. Once formed, the resulting and highly reactive nitrido groups readily engage in reactions with different substrates to form various nitrogen compounds, including ammonium upon protonolysis.28 Yet, the formal hydrogenation of uranium nitrogen or nitrido species,39,40 especially when obtained from molecular dinitrogen,25,33,41 has been rarely witnessed. It should be noted that the formation of amine derivatives from dinitrogen does not always require uranium nitrido intermediates. For example, Arnold and co-workers recently showed that, using a bridged meta-xylyl-tetraphenolate UIV dimer, reduction of the complex under N2 gave a hydrazido (N2H2)2− species with concomitant deprotonation of the ligand scaffold. Cleavage of the remaining N–N bond and formation of ammonium occurred upon protonolysis of the complex.42
Easy access to UIII halide and amido precursors43 has significantly contributed to the development of uranium reductive chemistry44 with various ligand environments, allowing for the tuning of reductive properties and enabling the activation of diverse small molecules and substrates.45–47 Hydrocarbon ligands, e.g. cyclopentadienyl (Cp) and cyclooctatetraenyl (Cot), and related motifs are major assets in this field. In particular, UIII complexes in tris-Cp environments have led to the formation of rare examples of terminal N2 adducts,48 while reversible N2 coordination has been noted in the case of UIII pentalene complexes.49 The reactivity of UIII metallocene complexes typically relies on the rich redox chemistry of uranium, which can access up to the formal +VI oxidation state. In addition, as coordination saturation increases, steric crowding may also play a role in the reactivity. In highly sterically congested molecules, one extra electron can be provided to the metal system upon departure of one Cp ligand as a radical. Such sterically induced reduction (SIR) has allowed multi-electron reductions of various substrates.50–52
However, a balance exists between the bulkiness that either triggers SIR reactivity or kinetically stabilizes low-valent uranium species. The first molecular UII complex was isolated in 2013 as the ion–pair complex [K(2.2.2-cryptand)][(Me3SiC5H4)3U],53 and the use of the bulky pentaisopropylcyclopentadienyl CpiPr5 ligand allowed access to the first example of a neutral UII organometallic sandwich complex.54 Further reduction of the latter led to the formation of a formal UI complex as a charge-separated ion pair.55 The tris-substituted Cpttt ligand (Cpttt = 1,2,4-tris(tert-butyl)cyclopentadienyl), which presents a different steric profile compared to the CpiPr5 ligand, proved effective in stabilizing non-classical divalent rare-earth complexes.56–59 Yet, such a ligand environment still provides an accessible coordination site,60,61 allowing for the coordination and reduction of small molecules such as CO and N2.62–64 In this context, our group recently showed that the isolable [LuCpttt2] complex readily binds N2 to form an end-on (N
N)2− complex. The latter can be hydrogenated under smooth conditions into a Lu–NH2 complex through a unique type of reactivity.64
Inspired by these findings, herein we present the room-temperature binding and cleavage of dinitrogen using an organometallic uranium complex featuring the Cpttt ligand. The redox chemistry of uranium contrasts with the single-electron transfer reactivity of divalent lanthanides and enables the isolation of the first example of an end-on uranium dimer with a formally four-electron reduced dinitrogen complex. Redox assistance by electron transfer from the Cpttt ligand ultimately leads to the six-electron N2 cleavage through an original type of reactivity that does not require the addition of further external reducing agents.34,38,65 As a result, bridged nitrido complexes are formed and can be hydrogenated at room temperature into uranium imido complexes. This work emphasizes the importance of the fine-tuning of the Cp ligand's steric properties for the generation of low-valent uranium species able to bind and achieve multiple (4 to 6) electron reductions of N2.
When a lower excess of KC8 was used (4 equiv.), the same set of 1H NMR signals was transiently observed while the conversion was not complete after 60 min, as evidenced by the presence of remaining amounts of 1. As the starting material was further consumed, the 1H NMR spectrum evolved, over 6 h at room temperature, into a new set of signals. The concentration of both reactions' dark pentane and toluene solutions resulted in dark green and dark brown crystals, respectively, suitable for X-ray diffraction (XRD) studies. The analyses of the structures revealed formation, in the first case, of the dimeric dinitrogen complex [(Cpttt2U)2(μ-N2)] (2) and, in the second case, of a trimeric uranium cluster [Cpttt2U(μ-I)(μ3-N)(μ-N)(UCpttt)2] (3) featuring two nitrido groups (Scheme 1). The 1H NMR spectrum of crystals of 3 (see below for its structural characterization) corresponds to the second set of signals appearing with time when a moderate excess of KC8 (4 equiv.) was used.
When the reduction of 1 was performed under higher pressure of N2 (5 bars) using a large excess of KC8 (10 equiv.), the formation of 2 is transiently observed while a new set of signals appears along with significant amounts of an organic by-product identified as (Cpttt)2 (Scheme 1). The formation of (Cpttt)2 aligns with the departure of one of the Cpttt ligands as its radical form, which leads to the coupled bis(tris-tert-butyl-cyclopentadiene) compound. The loss of one Cpttt ligand as a radical is associated with the reduction of the metal ion, which has already been witnessed in f-block organometallic chemistry in the attempted synthesis of [(Cpttt)2EuIII(X)] (X = Cl, F),67,68 or upon oxidation of [(Cpttt)2UIV
Y] (Y =
O, N(p-tolyl)) centers.69,70 It is worth noting that, in both cases, the electron transfer from the ligand to the metal occurred in the presence of easily reducible metal centers (EuIII, UV/VI).
In parallel, when a toluene solution of 2 is left at room temperature, the same set of 1H NMR signals appear, but more slowly (Fig. S7). The identity of the formed product has been obtained upon re-crystallization from diethyl ether, and XRD studies revealed the formation of the dimeric [Cpttt2U(μ-N)2{U(Cpttt)(OEt2)}] uranium complex 4 (Scheme 1). In this complex, the dinitrogen unit has been cleaved into two nitrido groups, accompanied by the departure of one Cpttt ligand from one uranium center. The coordination sphere of the latter is completed by one diethyl ether molecule. Although 4 was isolated as XRD-suitable crystals, separation from other co-crystallized materials was tedious and did not produce analytically pure material, precluding full characterization of this complex. However, the conversion from 2 to 4 was followed by 1H NMR spectroscopy.
:
0.17 ratio, respectively. This ratio did not change significantly over the temperature range of −80 °C to +40 °C. This behaviour precludes performing a van't Hoff analysis to assess the thermodynamic data of a possible equilibrium between the two species. Although end-on N2 coordination has already been witnessed in uranium chemistry,48,71–742-end-on is the first example of an end-on dinitrogen adduct between two uranium metal centers.
The molecular structure of 2 in the solid state reveals two substructures with different coordination modes for the reduced dinitrogen ligand (Fig. 1 and Table S4), specifically the end-on (μ-η1:η1-N2) and side-on (μ-η2:η2-N2) motifs. In the XRD solid-state structure collected at 150 K, the respective ratio was refined to 0.74
:
0.26, which is in fair agreement with the ratio of 0.84
:
0.16 in solution at −40 °C obtained by 1H NMR spectroscopy. This ratio did not evolve either when the data were collected over the 150–230 K temperature range, while higher measurement temperatures resulted in degradation of the crystal.
For the minor component, 2-side-on, the U2–N average distance is 2.20(2) Å, the average U2-Cp(Ctr) distance 2.77 Å, and the N–N distance 1.37(6) Å. The U2–N distances observed in 2-side-on are within the range of those in the uranium side-on dinitrogen complexes reported in the literature, from 2.146 Å33 to 2.465 Å,75 with U⋯U separations ranging from 3.372 to 4.797 Å. The overall bond distances in 2-side-on are in agreement with a 4-electron reduction of dinitrogen and the formation of a side-on coordinated (N2)4− bridging ligand, which has already been witnessed in uranium chemistry (see Table S6).33,34,36,76
For the major component, 2-end-on, the U1–N1 distance is 1.958(12) Å, the average U1-Cp(Ctr) distance is 2.51 Å, and the N–N distance is 1.43(2) Å. In addition to these altered metrics, the change in the coordination mode of the bridging dinitrogen ligand implies significant differences in the U⋯U separations (5.342(6) Å vs. 4.185(16) Å) as well as Cp(Ctr)-U-Cp(Ctr) angles (139° vs. 117°) for 2-end-on and 2-side-on, respectively. Only a few end-on coordinated heterodinuclear U-(μ-η1:η1-N2)-M (M = Fe, Re, Mo, Li) complexes have been reported in the literature,71–74 with U–N and N–N bond distances lying in the ranges of 2.221–2.606 Å and 1.139–1.232 Å, respectively. In these complexes, the bridging dinitrogen ligands have undergone a 2-electron reduction into (N2)2− moieties. In comparison, the U–N bond distance in 2-end-on (1.958(12) Å) is significantly shorter, while the N–N separation of 1.43(2) Å is much larger. Relatively short U–N bond distances are typically found in U–N imido complexes, in which the distances vary with the oxidation state of the uranium center.77–84 The different metrics in 2-end-on compared to the abovementioned end-on coordinated uranium dinitrogen complexes account for a different reduction state of the bridging ligand, more precisely a 4-electron reduction of N2 to an end-on coordinated (N2)4− ligand. Thus, the molecular structure of 2 is consistent with both end-on and side-on (N2)4− ligands bridging two UIV metal centers.
It may be noted that UIV and LuIII have similar crystal radii, which makes a comparison with lanthanide complexes relevant.85 Analysis of the metrics between end-on vs. side-on N2 coordination in lanthanide complexes is shown in the SI (Fig. S59 and Table S7). It reveals that, within identical ligand environments, the Ln–N bond distance is ca. 0.21 Å shorter in the end-on complexes while the Ln⋯Ln separation is ca. 1.1 Å longer.62,86 A similar trend is observed in our case with the U–N bond distance in 2-end-on being 0.24 Å shorter than in 2-side-on while the U⋯U separation is 1.2 Å longer.
The Raman spectrum was not informative, with only a broad feature and no difference with 2 prepared from 15N2, which contrasts with the well-defined Raman signal detected for the diazenido (N2)2− ligand in the [(Cpttt2Ln)2(μ-η1:η1-N2)] complexes.62,64 The different reduction state of the dinitrogen ligand as N24− in 2 could rationalize this observation, with a significant shift in the energy of the bond stretch and difference in oscillator strength.33,42 The elongated distance of the N–N moiety agrees with this statement,87 which accordingly lies in the range of hydrazido complexes.42,88
The temperature-dependent solid-state magnetic data of 2 were recorded and feature a χT value of 1.53 cm3 K mol−1, e.g. 0.765 cm3 K mol−1 (2.47 μB) for each U center (Fig. 2, S56 and S57). The χT value decreases monotonically with temperature to reach 0.09 χT cm3 K mol−1 at low temperature. This behavior indicates the presence of a non-magnetic ground state with low-lying magnetic crystal field states. Indeed, the absence of a maximum Curie temperature and the nearly linear evolution of χT until 70 K is typical for van Vleck paramagnetism arising from a second-order Zeeman effect, which is independent of the temperature (χT is linear with T). Van Vleck paramagnetism (TIP)89–91 is typically reported for UIV complexes and thus agrees with a [(Cpttt2UIV)2(μ-η1:η1-N24−)] formula for 2.92–96
![]() | ||
| Fig. 2 Solid-state temperature-dependent magnetic data for 2 (red unfilled dots) and 3 (blue filled dots). An aberrant point has been removed at 290 K for 2 (see Fig. S56). | ||
DFT calculations (B3PW91 functional) were carried out to determine the electronic structure of 2 and the energy difference between the two N2 coordination modes. Computationally, both the side-on and end-on coordinations proved stable, in line with the experiment, with the end-on coordination appearing to be more favorable than the side-on by 9.8 kcal mol−1. Three different spin states were considered in the end-on case (septet, quintet, and open-shell singlet). The open-shell singlet appears to be the ground state (Table S8), as evidenced by the unpaired spin density plot (Fig. 3), and an (N2)4− bridging ligand. Still, the quintet state is only lying slightly higher in energy (2.0 kcal mol−1 in enthalpy and 0.8 kcal mol−1 in Gibbs Free energy).
For both the 2-end-on and the 2-side-on complexes, the optimized structures are in fairly good agreement with the experimental ones. For 2-end-on, the structure is symmetrical, and the two U–N distances are 2.03 Å long for the open-shell singlet and 1.99 Å in the quintet (1.96 Å experimentally) with an N–N bond distance of 1.31 Å and 1.33 Å, respectively (1.43 Å experimentally). For the 2-side-on complex, the optimized structure is unsymmetrical with the shortest U–N bond distance of 2.21 Å (2.20 Å experimentally) with an N–N bond distance of 1.36 Å (1.37 Å experimentally). Although complexes bearing an (N2)4− ligand are known in uranium chemistry,33,34,36,76 they have always exhibited side-on coordination since it allows a better overlap between the two N–N π* and the 5f orbital on the uranium center.
The situation is somewhat different here, and the bonding situation in 2-end-on was analyzed using Molecular Orbitals (MO), Nature Bonding Orbitals (NBO), and Quantum Theory of Atoms in Molecules (QTAIM) methods (Tables S10–S19 and Fig. S66). The HOMO and HOMO-1 are doubly occupied, where the alpha and beta parts are occupied, and are displaying two U–N π interactions, meaning an overlap between the 5f-6d hybrid orbital of U and the two π* of N2. This is also observed at the NBO level, where two formal U–N triple bonds are found. These bonds are polarized toward N(80%) and involve overlap between hybrid 5f–6d orbitals on U and either an sp hybrid on N (for the σ bond) or pure 2p orbitals (for the two π). A single N–N bond remains, being highly covalent (50–50 at the NBO) and of σ-type (overlap between two sp orbitals). The nature of the bonds in the system is finally confirmed by a QTAIM plot and an analysis of the critical points of the bond (BCP). Two U–N BCP are found with a large density in line with an iono-covalent bond, and the ellipticity value indicates the presence of a π character. On the other hand, the N–N BCP with a large density, a negative Laplacian, and almost zero ellipticity is indicative of a σ-type single bond. Therefore, all these analyses align with an N24− ligand. The preference for the end-on coordination over the side-on one for the four-electron reduction of N2 in 2 is due to the steric hindrance induced by the Cpttt ligand. Indeed, with a side-on coordination, the U⋯U distance is short so that the repulsion between the tert-butyl substituents is significant, while in an end-on coordination, the U⋯U distance remains long, decreasing the steric repulsion between the tert-butyl ligands. This repulsion is responsible for the unsymmetrical coordination in the 2-side-on complex (Tables S20–S25). The formation of both the 2-end-on and 2-side-on complexes was computed to be thermodynamically favorable by, respectively, 17.8 and 8.0 kcal.mol−1 in enthalpy in line with the experimental observation.
The XRD structure of 3 spans three uranium centers, two having one Cpttt ligand η5-coordinated (U1 and U2) and the last one (U3) coordinated to two Cpttt rings (Fig. 4 and Table S4). The latter uranium center coordinates to one iodine atom, bridging with the U1 center, with the U1–I1 and U3–I1 bond distances of 3.1065(3) and 3.2293(3) Å, respectively. One μ3-nitrido ligand (N2) is centered between the three uranium centers. One μ-nitrido ligand (N1) bridges the two uranium centers, which only bear one Cpttt ligand (U1 and U2). The U3-Cp(Ctr) average distance is 2.63 Å, and the Cpttt-U3-Cpttt angle is 134°. The U1-Cp(Ctr) and U2-Cp(Ctr) distances are comparatively shorter, 2.51 Å and 2.55 Å, respectively, which may partly account for a lower coordination number compared to U3. The μ3-nitrido (N2) distances to the uranium centers are 2.221(3), 2.170(3), and 2.284(3) Å for U1–U3, respectively, comparable to those found in other μ3-nitrido uranium complexes, and are consistent with U–N single bonds.37,97 In contrast, the μ-nitrido (N1) distances are 2.027(3) and 2.033(3) Å for U1 and U2, respectively, in agreement with U–N multiple bond character.34 The differences in the U–I, U-Cp(Ctr), and U–N distances are globally in agreement with a different oxidation state for U3 compared to U1 and U2. From the observed bond distances and to balance the 11 negative charges, a tentative assignment of the oxidation states for the different uranium centers in 3 leads to formal UIV centers for U1 and U2, while U3 corresponds to a formal UIII center. However, in the literature, similar distances have been reported for both UIII and UIV complexes, and this sole metric should be taken cautiously.66,69,98
The temperature-dependent solid-state magnetic data of 3 present a χT value of 2.67 cm3 K mol−1 (Fig. 2). With decreasing temperature, the χT value decreases to 2.42 cm3 K mol−1 at 75 K and then rapidly to 0.53 cm3 K mol−1 at 2 K. The low-temperature value agrees with the presence of one UIII center, while the two other uranium centers are UIV with nearly zero χT as observed for 2, which overall agrees with a formula of [Cpttt2UIII(μ-I)(μ3-N)(μ-N)(UIVCpttt)2] for 3.
The molecular structure of 4 in the solid state reveals a dinuclear uranium complex bearing two bridging nitrido groups (Fig. 5 and Table S4). The first uranium ion, U1, is coordinated by two Cpttt ligands, and the second one, U2, by one Cpttt ligand–disordered over two positions–and one diethyl ether molecule. Complex 4, [Cpttt2U(μ-N)2{U(Cpttt)(OEt2)}], is a rare example of a molecular uranium bridged bis-nitrido complex, which is formed upon direct cleavage of dinitrogen.35–37,65,97,100 The short and almost equivalent U–N bond distances (2.031(7)–2.094(7) Å) are consistent with multiple bond character and agree with μ-nitrido groups,28 resulting in a planar and slightly distorted U1–N1–U2–N2 diamond core. The U1-Cp(Ctr) distances are 2.59 Å and 2.61 Å, which are intermediate between the U-Cp(Ctr) distances in 2-end-on and 2-side-on. The U2-Cp(Ctr) separations for both disordered components have similar distances (2.52 and 2.61 Å). From the charge balance in 4, with two N3− nitrido groups and three Cpttt ligands, an average mixed valent +4.5 oxidation state is noted. However, the metrics alone do not allow for the assignment of the extent of the delocalization of the spin density over the two uranium centers.
![]() | ||
| Fig. 5 Molecular structure of one of the disordered positions of 4 in the solid state. Hydrogen atoms are omitted for clarity. Thermal ellipsoids are shown at the 50% probability level. | ||
Two spin states were computed for 4 (a doublet and a quartet), and both are very close in energy (0.6 kcal mol−1 in enthalpy favoring the quartet). However, in both cases, the system is a mixed-valence complex, UIV–UV, with either ferromagnetic coupling (quartet) or antiferromagnetic coupling (doublet) between the two uranium centers (Table S26–S32 and Fig. S68). The slight energy difference between the two situations aligns with a weak coupling.
:
9
:
9
:
9 ratio for the first one (A), and two principal features at δ −6.5, −20.4 in a 1
:
2 ratio for the second one (B). Although A appears stable over time, we could not isolate an analytically pure material. Despite numerous crystallization attempts, crystals suitable for X-ray diffraction studies could only be obtained once.
The molecular structure of A (Fig. 6) unambiguously reveals the formation of the dinuclear bis-imido UIV complex [(Cpttt2UIV)2(μ-NH)2]. In particular, the hydrogen atoms on all imido units could be successfully located in the electron density map. Two independent molecules of A featuring very similar metrical data are found in the asymmetric unit (Fig. S63–S65). The U–N bond distances in A, from 2.188(6) to 2.214(6) Å, are longer than those in 4, in agreement with U–N single bonds, which supports hydrogenation of 2 and formation of a bridged bis-imido uranium complex. Accordingly, the U⋯U separation of ca. 3.62 Å in A is considerably longer than that in 4 (3.27 Å) and the U–N–U angles are ca. 5° larger in A. Interestingly, the U⋯C separations involving the Cpttt ligands in A span an unusually large range of distances, from 2.695(6) to 3.229(7) Å, which is more consistent with η3- rather than η5-coordination for the cyclopentadienyl ligands. Accordingly, relatively long U–Cp(Ctr) distances (2.671–2.705 Å) can be noted.
Protonolysis of 2 by the addition of excess HCl(Et2O) on 2 led to an immediate color change and the formation of HCpttt but no NH4Cl was detected in the 1H NMR spectrum (Fig. S11). However, the same protonolysis experiment performed either after hydrogenation of 2 into A or after letting a toluene solution of 2 degrade at room temperature led to the formation of NH4Cl (Fig. S13). In the corresponding 1H NMR spectra, the characteristic deshielded 1
:
1
:
1 triplet of NH4Cl with 1J(14N–H) = 48 Hz was observed (14N, I = 1, 99.6% natural abundance) and easily quantified by quantitative NMR spectroscopy (see the SI for details). Accordingly, when the isotopically labelled 2-15N was prepared from 15N2, its hydrogenation under a pressure of 5 bars, to form A-15N, and subsequent protonolysis led to the formation of 15NH4Cl featuring a deshielded doublet with 1J(15N–H) = 71 Hz (Fig. S16).
Similar hydrogenation and protonolysis experiments were performed on 3. The hydrogenation under 1.2 bar H2 proceeds smoothly over 16 h to yield a clean compound crystallized from toluene as the hydrogenated uranium trimer 5 (Scheme 3). The XRD structure presents similar coordination environments for the three uranium centers as in the parent complex 3, with similar U–Cp(ctr) distances (2.51, 2.53, and 2.63 Å, for U1–U3, respectively, in 5, compared to 2.51, 2.55, and 2.63 Å in 3) (Fig. 5 and Table S4). The principal difference is that the U1–N1–U2–N2 core is slightly bent in agreement with a pyramidalization of the protonated nitrogen (N1) (Fig. 4c and d.).
![]() | ||
| Scheme 3 Hydrogenation and hydrolysis reaction from 3. The +3.5 oxidation state results from charge balance and is given as an average. The extent of delocalization was not determined. | ||
Additionally, several U–N distances are significantly elongated in 5 compared to those in 3, particularly the U1–N1 and U2–N1 bond distances of 2.173(4) and 2.158(4) Å, respectively (compared to 2.027(3) Å and 2.033(3) Å in 3), consistent with a bridging imido (μ-NH).39 In contrast, the U–N bond distances involving the μ3-nitrido ligand (N2) are less altered: while the U3–N2 distance in 5 (2.315(3) Å) is slightly elongated compared to that in 3 (2.284(3) Å), the U1–N2 and U2–N2 distances (2.186(4) and 2.141(4) Å, respectively) are slightly shortened (2.221(3) and 2.170(3) Å in 3, respectively). The only minor changes in bond distances around N2 tend to indicate that the N2 atom is not protonated upon hydrogenation. However, a clear-cut conclusion cannot be formed from mere analysis of the XRD structures.
The 1H NMR spectrum of 5 shows 15 identifiable and paramagnetically shifted signals (Fig. S37–S40), two of them being broad with tentative integration, ten of them accounting for the twelve tert-butyl groups (two overlapping signals), and five signals for the aromatic protons of the Cp ring (instead of 8). Considering the significant isotropic shifts of some protons, several may likely experience too fast relaxation to be easily observed. The synthesis of 5 from D2 allowed assignment of the protonated nitrogen at δ 156.4 ppm (Fig. S46). Interestingly, in the solid-state, 3 also readily reacts with H2 at 60 °C to lead to 5 (Fig. S30). Compound 5 appears stable in the solid state at room temperature for several days. Yet, if 5 is left to stand in solution for one week at room temperature, the 1H NMR spectrum evolves with partial conversion back to 3 and formation of HCpttt (Fig. S44). The formation of 3 indicates that either the H2 addition is reversible or that the acidity of the imido proton is sufficient to protonate the Cpttt ligand. No H2 was detected in the solution by 1H NMR studies or by analyzing the volatiles by gas chromatography. Upon heating a toluene solution of 5 at 60 °C, a 90% conversion was observed over 16 h, leading to the formation of 3 and HCpttt (Fig. S45). This observation suggests that the degradation of 5 is triggered by the protonation of Cpttt ligands by the NH group,23 leading to partial conversion back into 3.
The controlled protonolysis of 3 was performed to evaluate the basicity of the N3− ligands compared to that of the Cpttt ligands. The addition of 0.5 equiv. of pyridinium triflate or water led to partial consumption of 3 with the formation of several new species and only small amounts of HCpttt (Fig. S27 and S28). The stoichiometric addition of 1 equiv. of acidic protons led to the disappearance of 3 with pyridinium triflate but not with water. In both cases, adding 2 equiv. of H+ led to complete consumption of 3 with formation of new species, different for pyridinium triflate and water, along with minor amounts of HCpttt. The formation of the new species, rather than HCpttt, contrary to what was observed for 2, indicates preferential protonation of the basic N3− ligands, which can explain the partial thermal reversibility of the hydrogenation of 3. Complex 5 may engage in acid–base reactions with itself, leading to partial regeneration of 3 and formation of by-products which degrade with loss of HCpttt. Protonolysis of 3 and 5 with excess HCl(Et2O) led to the generation of NH4Cl in 69 and 44% yields, respectively, with the formation of the characteristic 1
:
1
:
1 triplet (1JNH = 51 Hz) in the corresponding 1H NMR spectra (Fig. S29 and S47). All attempts to detect ammonia from the hydrogenation reactions of 2–5 were unsuccessful. As observed with the protonation of Cpttt ligands by the imido NH2− group, the acidity of ammonia may lead to similar protonation events whenever formed.23
Examples of similar linear M
N–N
M bis-imido complexes have been previously observed in early transition metal (group IV (Ti), group V (Nb, Ta), and group VI (W)) chemistry.30 In particular, a mixed cyclopentadienyl/amidate TaIV
N–N
TaIV complex has been obtained upon reduction of the corresponding trichloride precursor under an N2 atmosphere and found to convert, above 0 °C, to the corresponding bis(μ-nitrido) TaV dinuclear complex.101 A similar pathway might be accessible in the case of 2, leading to the generation of the putative [(Cpttt2UV)2(μ-N)2] complex featuring two bridging nitrido units (Scheme 4).
Indeed, the formation of the nitrido fragment “(Cpttt2UV)(N)” from 2 was computed to be thermodynamically favorable with a weak N–N bond dissociation energy in 2 (17.0 kcal mol−1), resulting in an exergonic transformation (−1.3 kcal mol−1 in favor of the nitrido fragment). The DFT calculations also support thermodynamically driven transformations of the “(Cpttt2UV)(N)” nitrido fragment into complexes A and 4 (Scheme S1). The hydrogenation of 2 under 5 bars H2, leading to the bis(imido) complex A provides further support for the transient generation of the bis(μ-nitrido) [(Cpttt2UV)2(μ-N)2] complex. A similar hydrogenation reactivity on a dinuclear nitrido-bridged UV complex supported by siloxide ligands has been reported by Mazzanti and co-workers.102
Owing to the oxidizing propensities of UV centers, the loss of one Cpttt ligand as a radical from the putative [(Cpttt2UV)2(μ-N)2] dimer, via an intramolecular single-electron transfer from one (Cpttt)− ligand to one uranium center, results in the formation of 4 along with the (Cpttt)2 dimer (Scheme 4). Related single-electron transfer reactions involving the Cpttt ligand have already been witnessed in f-block organometallic chemistry, in association with easily reducible metal centers (EuIII, UV/VI).67–70 This step, which is accompanied by a release of steric pressure, can also be seen as an SIR.51,52 Indeed, as observed in the XRD structure of A, the strong steric crowding induced by the four Cpttt ligands results in an unusual η3-coordination mode for the cyclopentadienyl rings, with long U–Cp(Ctr) separations.
Similarly, the loss of a second Cpttt radical from 4 would result in the putative [(CptttUIV)2(μ-N)2] complex, which upon coordination to one molecule of 1 yields complex 3. This possible pathway is consistent with the formation of 3 when a lower amount of KC8 is used, i.e. when the kinetics of the reduction of 1 enter in competition with those of the degradation of 2 into nitrido species.
Complexes 3 and 4 are rare examples of direct 6-electron cleavage of N2 by well-defined uranium dinitrogen complexes, without further addition of an external reducing agent.32 One system has been reported by the Zhu group, involving the assistance of the phosphine ligand side-arm on a dinuclear UIII complex.38 Besides, Mazzanti and co-workers have shown that reducing dinuclear UIII nitrido or oxo complexes under N2 can lead to cleavage of N2 and formation of bis-nitrido complexes.34,36,65 Here, we show that this challenging reactivity in uranium chemistry is not restricted to coordination complexes supported by N- or O-donor ligands but can be extended to organometallic uranium complexes featuring Cp-type ligands. In addition to their ability to stabilize low-valent species in f-block chemistry,47 substituted Cp ligands can provide possible redox assistance via single-electron transfer reactions, sterically induced51 or not, leading to a rich chemistry. Overall, the 6-electron cleavage of N2, resulting in the formation of 4 from the transient divalent [Cpttt2UII], formally involves 5 electrons from the two uranium centers and one additional electron from an intramolecular ligand-to-metal electron transfer with the loss of one Cpttt radical. The loss of another Cpttt radical, leading to 3, provides one additional electron to the system.
It is noteworthy that the hydrogenation reactivity of 2-5 differs from that of the lanthanide complex [(Cpttt2Lu)2(μ-η1:η1-N2)], for which formation of [Cpttt2Lu(NH2)] occurred via direct hydrogenation of the reduced dinitrogen ligand.64 The rich redox chemistry of uranium, in which high-valent +V or +VI oxidation states are accessible, allows direct splitting of N2 into nitrido groups before hydrogenation. The other possible pathways to N2 hydrogenation, i.e., by activation of H2 and reaction of the resulting hydride complexes with N2,26 or by direct synergistic H2/N2 activation,27 may be possible by tuning the system and are currently being investigated.
The data supporting this article have been included as part of the supplementary information (SI). Supplementary information: full experimental details, 1H NMR spectra, IR spectra, X-ray crystallographic details, magnetic analysis, and DFT calculations. See DOI: https://doi.org/10.1039/d5sc07194a.
Footnote |
| † These authors made equal contributions. |
| This journal is © The Royal Society of Chemistry 2025 |