Next Article in Journal / Special Issue
Expanding the Chemistry of Actinide Metallocene Bromides. Synthesis, Properties and Molecular Structures of the Tetravalent and Trivalent Uranium Bromide Complexes: (C5Me4R)2UBr2, (C5Me4R)2U(O-2,6-iPr2C6H3)(Br), and [K(THF)][(C5Me4R)2UBr2] (R = Me, Et)
Previous Article in Journal / Special Issue
Synthesis and Characterization of Cerium(IV) Metallocenes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Molecular Pnictogen Activation by Rare Earth and Actinide Complexes

Chemistry Research Laboratory, Department of Chemistry, University of Oxford, Mansfield Road, Oxford OX1 3TA, UK
Submission received: 1 October 2015 / Revised: 4 December 2015 / Accepted: 9 December 2015 / Published: 21 December 2015
(This article belongs to the Special Issue Rare Earth and Actinide Complexes)

Abstract

:
This review covers the activation of molecular pnictogens (group 15 elements) by homogeneous rare earth and actinide complexes. All examples of molecular pnictogen activation (dinitrogen, white phosphorus, yellow arsenic) by both rare earths and actinides, to date (2015), are discussed, focusing on synthetic methodology and the structure and bonding of the resulting complexes.

Graphical Abstract

1. Introduction

Rare earth (scandium, yttrium and the lanthanides) and actinide complexes remain underexplored with respect to the transition metals and main group elements but often demonstrate both unique reactivity and molecular properties. Understanding of the bonding and electronic structure of these complexes has particular significance for separation of metals in nuclear waste streams [1].
Activation of molecular pnictogens (group 15 elements) is an area of growing importance; atmospheric dinitrogen (N2) and white phosphorus (P4) are principal sources of N- and P-containing compounds (e.g., polymers, pharmaceuticals, agrochemicals, explosives, and specialty chemicals) but are both very challenging to selectively activate. Metal-arsenic, -antimony and -bismuth complexes remain rare [2,3,4], while the study of metal-pnictogen complexes, including these heavier pnictogen homologues, is also of fundamental importance with respect to Ln and An-pnictogen bonding and electronic structure.
Fixation of N2, the six electron reduction to two molecules of more reactive ammonia, is necessary for further formation of N-element bonds. In nature, nitrogenase enzymes containing metalloproteins (Fe, Mo or V) fix N2 through proton-coupled electron transfer under ambient conditions [5,6]. In industry, the Haber–Bosch process combines N2 and high purity H2 at high temperatures and pressures over heterogeneous iron- or ruthenium-based catalysts [7,8,9,10]. This highly efficient process produces 100 million tons of ammonia per year but is the largest energy-consuming process in the modern world today; the need for direct activation and functionalisation of N2 under mild conditions is a clear goal. Accessing appropriately reactive phosphorus building blocks presents a different set of challenges based on the sustainability and efficiency of chemical transformations required; from phosphate rock minerals which are mined globally on a 225 million ton scale per year (2013) [11], phosphate fertilisers derived from phosphoric acid are the major products with the remainder used for elemental phosphorus production. Organophosphorus compounds are generally derived from PCl3, obtained by the chlorination of P4, and subsequent multi-step procedures [12,13,14]. Attention has turned to direct and selective activation of elemental phosphorus under mild conditions; this approach is more atom-efficient (which is important given the limited accessible deposits of phosphate rock), avoids the need for large scale production of PCl3 (which is toxic, corrosive and highly reactive), and is both more economically and environmentally sustainable [15].
The area of dinitrogen activation has been reviewed extensively with particular focus on the Haber–Bosch process [16,17,18], and biological nitrogen fixation [5,6,19,20,21,22,23,24,25,26]. There are reviews on transition metal N2 activation which cover N2 binding modes [27,28,29], multimetallic N2 activation [30,31], the relevance of metal hydride complexes to N2 activation [32,33], N2 cleavage and functionalisation [34,35] (including electrochemical [36] and photolytic N2 cleavage [37]), and N2 activation at bare metal atoms [38] and using surface organometallic chemistry [39]. Specific reviews have also focused on activation by group 4 metals [40,41,42], iron [31,43,44], molybdenum [24,45,46,47], and the mid-to-late transition metal centres [48]. In terms of rare earth N2 activation; an account of work from Evans and co-workers to 2004 has been reported [49], and Gardiner more recently reviewed the chemistry of the lanthanides with dinitrogen and reduced derivatives [50]. The area of actinide N2 activation has been discussed in the context of small molecule activation by trivalent uranium complexes [51,52].
Transition metal-mediated white phosphorus activation has been previously reviewed [53,54,55,56], with specific reviews on both early transition metal complexes [57,58] and late transition metal complexes [59]. More broadly, reviews of P4 activation by p-block compounds have also been reported [60,61,62,63,64].
This review seeks to cover all examples of molecular pnictogen activation (dinitrogen, white phosphorus, yellow arsenic) by both rare earth and actinide complexes to date, focusing on synthetic methodology and the structure and bonding of the resulting complexes. Only well-defined homogeneous complexes will be discussed; heterogeneous and surface chemistry lie beyond the scope of this review.

2. Dinitrogen Activation by Rare Earth Complexes

2.1. Complexes Containing a Formal N22− Ligand

The majority of rare earth complexes that activate dinitrogen (N2) result in its formal reduction to the N22− anion and the formation of bimetallic complexes of the general form [A2(thf)xLn]2(μ-η22-N2) where there is side-on binding of N2 (“A is defined as a group that exists as an anion in LnA3 and provides reductive reactivity in combination with an alkali metal or the equivalent”) [65]. The work of Evans and co-workers has led to the development of three key methodologies to access these species: (i) salt metathesis reactions of divalent lanthanide halides with alkali metal salts; (ii) combination of trivalent Ln complexes with alkali metals (LnA3/M or LnA2A′/M method); (iii) photochemical activation of LnA2A′ systems (Figure 1).
Figure 1. Routes to N22− complexes using rare earth metals.
Figure 1. Routes to N22− complexes using rare earth metals.
Inorganics 03 00597 g001
For reference, N–N and M–N(N2) bond lengths obtained from single crystal X-ray diffraction experiments, N–N stretching frequencies (obtained by IR or Raman spectroscopy) and 14/15N-NMR spectroscopic data are summarised in Table 1.
Table 1. Summary of rare earth N22− complexes.
Table 1. Summary of rare earth N22− complexes.
Complex (#) [Reference]N–N Bond Length (Å)Ln–N(N2) Bond Lengths (Å)N–N Frequency (cm−1)14/15N-NMR Spectroscopy (ppm) a
N21.0975 [66]-2331 [67]−75 [68]
[(η5-C5Me4H)2Sc]2(μ-η22-N2) (1) [69]1.239(3)2.216(1)--
2.220(1)
[(η5-C5Me4H)2Sc]2(μ-η22-N2) (1′) [70]1.229(3)2.197(2)-385
2.179(2)
[(η5-C5Me5)2Y]2(μ-η22-N2) (2) [71]1.172(6)2.279(3)-496
2.292(3)
[(η5-1,2,4-tBu–C5H2)2Nd]2(μ-η22-N2) (3) [67]1.226(12)2.495(2)1622 (14N2)-
2.497(2)1569 (15N2)
[(η5-C5Me5)2Sm]2(μ-η22-N2) b (4) [72]1.088(12)2.348(6)-−117 (263 K)
2.367(6)−161 (203 K)
[(η5-C5Me5)2Dy]2(μ-η22-N2) (5) [73]----
[(η5-SiMe3–C5H4)2Dy]2(μ-η22-N2) (6) [74]Connectivity only-
[(η5-C5Me5)2Tm]2(μ-η22-N2) (7) [75]Connectivity only-
[(η5-1,3-SiMe3–C5H3)2Tm]2(μ-η22-N2) (8) [75]1.259(4)2.273(2)--
2.272(2)
[(η5-C5Me5)2Lu]2(μ-η22-N2) (9) [71]Connectivity only527
[(η5-C5Me5)(η5-C5Me4H)Lu]2(μ-η22-N2) (10) [76]1.275(3)2.291(3)1736 (14N2)-
2.295(3)1678 (15N2)
[(η5-C5Me4H)2Y(thf)]2(μ-η22-N2) c (11) [77]1.252(5)2.338(3)-468
2.370(3)
[(η5-SiMe3–C5H4)2Y(thf)]2(μ-η22-N2) c (12) [78]1.244(2)2.3214(14)--
2.3070(14)
[(η5-C5Me5)2La(thf)]2(μ-η22-N2) d (13) [79]1.233(5)2.537(4)-569
2.478(4)
[(η5-C5Me4H)2La(thf)]2(μ-η22-N2) c (14) [79]1.243(4)2.457(2)-495
2.503(2)
[(η5-C5Me5)2Ce(thf)]2(μ-η22-N2) c (15) [68]1.258(9)2.4548(15)-871
2.542(2)
[(η5-C5Me4H)2Ce(thf)]2(μ-η22-N2) d (16) [68]1.235(6)2.428(3)-1001
2.475(3)
[(η5-C5Me5)2Pr(thf)]2(μ-η22-N2) d (17) [68]1.242(9)2.4459(14)-2231
2.512(2)
[(η5-C5Me4H)2Pr(thf)]2(μ-η22-N2) c (18) [68]1.235(7) e2.418(4) e-2383
2.455(3) e
[(η5-C5Me4H)2Nd(thf)]2(μ-η22-N2) c (19) [79]1.241(5) e2.404(3) e--
2.451(2) e
[(η5-SiMe3–C5H4)2Tm(thf)]2(μ-η22-N2) c (20) [75]1.236(8)2.274(4)--
2.302(4)
[(η5-C5Me4H)2Lu(thf)]2(μ-η22-N2) c (21) [80]1.243(12)2.290(6)-521
2.311(6)
[{(Me3Si)2N}2Y(thf)]2(μ-η22-N2) c (22) [81]1.274(3)2.297(2)1425 (14N2)+513 (t)
2.308(2)1377 (15N2)
[{(Me3Si)2N}2La(thf)]2(μ-η22-N2) (23) [81]---516
[{(Me3Si)2N}2Nd(thf)]2(μ-η22-N2) c (24) [81]1.258(3)2.3758(16)--
2.3938(16)
[{(Me3Si)2N}2Gd(thf)]2(μ-η22-N2) c (25) [81]1.278(4)2.326(2)--
2.353(2)
[{(Me3Si)2N}2Tb(thf)]2(μ-η22-N2) c (26) [81]1.271(4)2.301(2)--
2.328(2)
[{(Me3Si)2N}2Dy(thf)]2(μ-η22-N2) c (27) [82]1.305(6)2.287(3)--
2.312(3)
[{(Me3Si)2N}2Ho(thf)]2(μ-η22-N2) c (28) [83]1.264(4)2.296(2)--
2.315(2)
[{(Me3Si)2N}2Er(thf)]2(μ-η22-N2) c (29) [81]1.276(5)2.271(3)--
2.302(3)
[{(Me3Si)2N}2Tm(thf)]2(μ-η22-N2) c (30) [82]1.261(4)2.271(2)--
2.296(2)
[{(Me3Si)2N}2Lu(thf)]2(μ-η22-N2) c (31) [83]1.285(4)2.241(2)1451 (14N2)557
2.272(2)
[{(Me3Si)2N}2Y(PhCN)]2(μ-η22-N2) c (32) [84]1.258(2)2.2848(13)--
2.3092(13)
[{(Me3Si)2N}2Y(C5H5N)]2(μ-η22-N2) c (33) [84]1.255(3)2.2917(16)--
2.3107(17)
[{(Me3Si)2N}2Y(4-NMe2–C5H4N)]2(μ-η22-N2) c (34) [84]1.259(2)2.2979(12)--
2.3132(12)
[{(Me3Si)2N}2Y(Ph3PO)]2(μ-η22-N2) c (35) [84]1.262(2)2.3000(14)--
2.3022(14)
[{(Me3Si)2N}2Y(Me3NO)]2(μ-η22-N2) c (36) [84]1.198(3)2.2925(17)--
2.2941(18)
[(2,6-tBu–C6H3O)2Nd(thf)2]2(μ-η22-N2) (37) [82]1.242(7)2.397(4)--
2.401(3)
[(2,6-tBu–C6H3O)2Dy(thf)2]2(μ-η22-N2) (38) [82]1.257(7) f2.328(4) f1526 (14N2)-
2.340(4) f
1.256(9) g2.336(5) g
2.336(5) g
[Na4(thf)8][(η5151-Et2calix[4]pyrrole)Pr]2(μ-η22-N2) (39) [85]----
[Na4(dme)5][(η5151-Et2calix[4]pyrrole)Pr]2(μ-η22-N2) (40) [85]1.254(7)2.414(5)--
2.457(5)
[Na4(thf)8][(η5151-Et2calix[4]pyrrole)Nd]2(μ-η22-N2) (41) [85]----
[Na4(dioxane)6][(η5151-Et2calix[4]pyrrole)Nd]2(μ-η22-N2) (42) [85]1.234(8)2.511(4)--
2.508(4)
[{HB(3-tBu-5-Me–pz)}Tm{NH(2,5-tBu–C6H3)}]2(μ-η22-N2) (43) [86]1.215(10)2.274(8)--
2.286(9)
a Referenced to CH315NO; b In equilibrium with [(η5-C5Me5)2Sm]2; c trans arrangement of donor solvent; d cis arrangement of donor solvent; e The authors indicate poor data quality or significant disorder in these structures; f thf solvent of crystallisation in unit cell; g toluene solvent of crystallisation in unit cell.

2.1.1. Cyclopentadienyl Ancillary Ligands

The first isolated, structurally characterised dinitrogen complex of an f-element metal was reported by Evans and co-workers [72]. [(η5-C5Me5)2Sm]2(μ-η22-N2) (4) was isolated by slow crystallisation of a toluene solution of the bent metallocene [(η5-C5Me5)2Sm]2 under an N2 atmosphere (Figure 2). 4 exists in dynamic equilibrium with the metallocene starting material involving reversible SmII/SmIII interconversion. In the solid state, 4 displays tetrahedral coordination around each Sm centre with gearing of the [Sm(C5Me5)2] units and the first example of a co-planar M2N2 diamond core for any metal. The bridging, side-on bound N2 has a short N–N distance of 1.088(12) Å (free N2: 1.0975 Å [66]) and does not imply reduction to N22−; however, recent studies by Arnold and co-workers have shown that N–N bond lengths determined using X-ray diffraction experiments can be underestimated and so may not provide the best way of assessing the level of dinitrogen reduction [87,88]. Both the Sm–N/C bond lengths and the 13C-NMR spectral data support formulation of the complex as [SmIII]2(N22−). Maron and co-workers have reported calculations on the interaction of N2 with [(η5-C5Me5)2Ln] (Ln = Sm, Eu, Yb) [89].
Figure 2. Rare earth complexes with cyclopentadienyl ligands resulting from N2 activation to N22−.
Figure 2. Rare earth complexes with cyclopentadienyl ligands resulting from N2 activation to N22−.
Inorganics 03 00597 g002
Since this landmark discovery, the methodology of using reducing divalent rare earth metal complexes to activate N2 has resulted in analogous cyclopentadienyl complexes of Dy (5, 6) [74] and Tm (7, 20) [75]. Structurally, these complexes all demonstrate a common planar Ln2N2 core (Ln–N–N–Ln dihedral angle = 0°), with the arrangement of the cyclopentadienyl ligands being dependent on the metal centre and the nature of the ligand itself (Figure 3).
The number of dinitrogen complexes has been expanded significantly by the report that combination of trivalent lanthanide complexes LnCpR3 or LnCpR2A′ with an alkali metal can also reduce dinitrogen affording side-on bound N2 complexes [(η5-CpR)2Ln]2(μ-η22-N2) (Ln = Sc (1), Y (2), Nd (3), Dy (5, 6), Lu (9)) [67,69,70,71,76] or [(η5-CpR)2Ln(thf)]2(μ-η22-N2) (Ln = Y (11, 12), La (13, 14), Ce (15, 16), Pr (17, 18), Nd (19), Lu (21)) [68,78,79,80,83]. The generality of this method has been demonstrated by the wide range of metals utilised as well as the use of both homo- and heteroleptic trivalent lanthanide starting materials with a variety of cyclopentadienyl, amide, aryloxide, hydride, halide and borohydride ligands.
Figure 3. Structural variations in the solid state structures of [(η5-CpR)2Ln]2(μ-η22-N2) (Ln = Sc (1) and (1′); Nd (3); Sm (4)).
Figure 3. Structural variations in the solid state structures of [(η5-CpR)2Ln]2(μ-η22-N2) (Ln = Sc (1) and (1′); Nd (3); Sm (4)).
Inorganics 03 00597 g003
Complexes 1121 (with donor solvent bound) also have planar Ln2N2 cores. However, only 13 (La), 15 (Ce) and 17 (Pr) have a cis arrangement of thf molecules and an asymmetrically-bound N22− ligand (as a result of crystallographically non-equivalent N atoms). This is most clearly seen though the difference in Ln–N–Ln′ angles (11: 145.77(16) and 157.33(18)°; 13: 156.9(3) and 144.7(3)°; 15: 157.0(3) and 145.1(3)°) (Figure 4).
Figure 4. Structural variations in the solid state structures of [(η5-C5Me4H)2Ln(thf)]2(μ-η22-N2) (11, 14, 16, 18, 19, 21) (11 depicted, left) and [(η5-C5Me5)2Ln(thf)]2(μ-η22-N2) (13, 15, 17) (13 depicted, right). One C5Me4 ring in 11 is disordered.
Figure 4. Structural variations in the solid state structures of [(η5-C5Me4H)2Ln(thf)]2(μ-η22-N2) (11, 14, 16, 18, 19, 21) (11 depicted, left) and [(η5-C5Me5)2Ln(thf)]2(μ-η22-N2) (13, 15, 17) (13 depicted, right). One C5Me4 ring in 11 is disordered.
Inorganics 03 00597 g004
In terms of bonding, calculations were carried out on 1 and the Sc–N (N2) bonding interaction was found to be a polar covalent two-electron four-centre bond resulting from donation from a filled Sc 3d orbital into an empty N2 πg antibonding orbital in the Sc2N2 plane. The lowest unoccupied molecular orbital (LUMO) of 1 is an unperturbed antibonding πg orbital based on N2. This bonding scheme can likely be extended for all compounds 121, with the donor nd orbital varying based on the nature of the metal ion [69].
Complexes 4 and 1518 were characterised by 15N-NMR spectroscopy; the first reported examples of such spectra for paramagnetic N2 complexes [68]. Trivalent Sm, Ce and Pr were chosen due to the low magnetic susceptibility of these ions (4f5 SmIII, μ = 0.84 μB; 4f1 CeIII, μ = 2.54 μB; 4f2 PrIII, μ = 3.58 μB). Broad singlets at high frequency were observed for 15 (871 ppm), 16 (1001 ppm), 17 (2231 ppm) and 18 (2383 ppm). Consistent with the reversible N2 coordination to 4, only a singlet at −75 ppm corresponding to free N2 is observed at 298 K [68]. Cooling resulted in a new resonance at −117 ppm at 263 K which shifted linearly to −161 ppm at 203 K and accounts for bound N2. In the context of pioneering NMR spectroscopic characterisation of organometallic complexes, the solid state 15N- and 139La-NMR spectra of 1415N2 have also been reported [90].
Most recently, it has been demonstrated by Evans and co-workers that photochemical activation of the closed shell LnA2A′ complexes Ln(η5-C5Me5)23-C5Me4H) or Ln(η5-C5Me5)(η5-C5Me4H)(η3-C5Me4H), which feature a novel η3 binding mode of a cyclopentadienyl ligand, yields side-on dinitrogen complexes [(η5-C5Me5)2M]2(μ-η22-N2) (M = Y (2), Dy (5), Lu (9)) and [(η5-C5Me5)(η5-C5Me4H)Lu]2(μ-η22-N2) (10), with concomitant formation of (C5Me4H)2 [73]. These reactions typically take place in under 5 h but, in the absence of photochemical activation, normally require a number of weeks. Full conversion to 2 and 9 is achieved via this methodology; this is notable given that other synthetic methods afforded yields of 26%–51% and 49%–59% respectively. Sterically induced reduction, typified by bulky M(η5-C5Me5)3 complexes, does not account for this process since the less sterically hindered [C5Me4H] ligand acts as the reductant [91]. Calculations support a mechanism involving electron transfer from the [η3-C5Me4H] ligand into an empty 4dz2 orbital on the metal centre. This affords a [C5Me4H]· radical which dimerises, and excited [Cp2M]* nd1 fragment which reduces N2. Similarly, the allyl complexes Ln(C5Me5)23-C3RH4) (R = H or Me) can be photochemically activated to yield 2 and 9; in this case, propene and isobutene are observed as by-products due to H-atom abstraction from solvent rather than radical dimerisation [92].

2.1.2. Amide Ancillary Ligands

The simple silylamide ligand [N(SiMe3)2] has also proved suitable to prepare related complexes of the form [{(Me3Si)2N}2Ln(thf)]2(μ-η22-N2) (Ln = Y (22), La (23), Nd (24), Gd (25), Tb (26), Dy (27), Ho (28), Er (29), Tm (30), Lu (31)) (Figure 5, left) [65,81,82,83,93,94,95]. All complexes can be prepared using the LnA3/M or LnA2A′/M method and in addition, 24, 27 and 30 can be prepared directly from the divalent starting materials NdI2, DyI2 and TmI2(thf)3 respectively (though 24 is notable in that it is only isolated in 4% yield using this synthetic approach) [82].
Lewis base coordination of yttrium complex 22 was investigated through a series of substitution reactions to afford [{(Me3Si)2N}2Ln(L)]2(μ-η22-N2) (L = PhCN (32), C5H5N (33), 4-NMe2–C5H4N (34), Ph3PO (35), Me3NO (36)) (Figure 5, right) [84]. Varying the donor ligand had little effect on the planar Ln2N2 structural core of 22 and 3235, though the N–N distance in 36 was unexpectedly short at 1.198(3) Å. Similarly for complexes 121, calculations on [{(Me3Si)2N}2Y(L)]2(μ-η22-N2) (22) indicate that the Y–N (N2) bonding interaction involves donation from a filled Y 4d orbital into an antibonding N2 πg orbital in the Y2N2 plane. The LUMO is an unperturbed antibonding N2 πg orbital [84,93]. UV-Vis spectra of 22 and 3135 all contained a low energy, low intensity absorption around 700 nm which corresponds to the formally electric-dipole forbidden HOMO–LUMO (ag–ag) transition and act as a fingerprint of the electronic structure of the Y2N2 core.
Figure 5. Rare earth complexes with amide ligands resulting from N2 activation to N22−.
Figure 5. Rare earth complexes with amide ligands resulting from N2 activation to N22−.
Inorganics 03 00597 g005
In terms of bonding, the closed shell 4f14 LuIII ion in 31 provides an interesting contrast to yttrium complex 22 with 4f° YIII ions. Calculations support that the bonding is described in analogy with 22, but using higher energy, radially diffuse 5d orbitals which have a good energy match with the N2 antibonding πg orbital and are the correct symmetry for overlap [65]. Hughbanks and co-workers have also reported calculations on the 4f7 GdIII complex 33 to analyse magnetic coupling [96].

2.1.3. Aryloxide Ancillary Ligands

Aryloxide ancillary ligands have been used to prepare side-on N2 complexes of the form [(2,6-tBu–C6H3O)2Ln(thf)2]2(μ-η22-N2) (Ln = Nd (37) and Dy (38)) which now contain two molecules of coordinating solvent per metal (Figure 6) [65,82,93]. Calculations on 38, which contains open shell 4f9 DyIII ions, indicate that Dy–N2 bonding is derived from a 5d–πg interaction in the Dy2N2 plane [65].
Figure 6. Rare earth complexes with aryloxide ligands resulting from N2 activation to N22−.
Figure 6. Rare earth complexes with aryloxide ligands resulting from N2 activation to N22−.
Inorganics 03 00597 g006

2.1.4. Multidentate Ancillary Ligands

Floriani and co-workers reported N2 complexes of PrIII (39, 40) and NdIII (41, 42) using a calix[4]pyrrole ligand; these complexes were obtained as single crystals suitable for X-ray diffraction studies and no isolated yields were reported (Figure 7) [85]. [Na4(thf)8][(η5151-Et2calix[4]pyrrole)Ln]2(μ-η22-N2) (Ln = Pr (39), Nd (41)) were prepared by reduction of [Na(thf)2][(η5151-Et2calix[4]pyrrole)Ln(thf)] (Ln = Pr or Nd) using sodium metal with catalytic napthalene. Addition of dimethoxyethane (dme) to 39 led to solvent exchange to afford [Na4(dme)5][(η5151-Et2calix[4]pyrrole)Ln]2(μ-η22-N2) (40). Similarly, addition of dioxane to 41 affords [Na4(dioxane)6][(η5151-Et2calix[4]pyrrole)Ln]2(μ-η22-N2) (42). The solid state structures of 40 and 42 revealed different coordination modes of the sodium counterions and feature N–N bond distances of 1.254(7) and 1.234(8) Å which are consistent with reduction to N22−. Measured magnetic moments are also consistent with LnIII oxidation state for both metal centres.
Figure 7. Rare earth complexes with multidentate ligands resulting from N2 activation to N22−.
Figure 7. Rare earth complexes with multidentate ligands resulting from N2 activation to N22−.
Inorganics 03 00597 g007
Takats and co-workers reported the scorpionate complex [{HB(3-tBu-5-Me–pz)}Tm{NH(2,5-tBu–C6H3)}]2(μ-η22-N2) (43) (pz = C3HN2 = pyrazolyl) which was prepared by the protonolysis reaction of 2,5-tBu–C6H3NH2 with the heteroleptic TmII hydrocarbyl compound {HB(3-tBu-5-Me–pz)}Tm{CH(SiMe3)2} [86]. The N–N bond distance of 1.215(10) Å is consistent with reduction to N22− and the Ln2N2 core is slightly bent with an Ln–N–N–Ln torsion angle of 0.37°.

2.2. Complexes Containing a Formal N23− Ligand

2.2.1. Amide Ancillary Ligands

For reference, N–N and M–N(N2) bond lengths obtained from single crystal X-ray diffraction experiments and N–N stretching frequencies (obtained by IR or Raman spectroscopy) are summarised in Table 2.
Table 2. Summary of rare earth N23− complexes.
Table 2. Summary of rare earth N23− complexes.
Complex (#) [Reference]N–N Bond Length (Å)Ln–N (N2) Bond Lengths (Å)Ln–N–N–Ln Torsion Angle (°)N–N Frequency (cm−1)
N21.0975 [66]--2331 [67]
[K(thf)6][{(Me3Si)2N}2Y(thf)]2(μ-η22-N2) b (44) [93]1.401(6)2.194(3)01002 (14N2) (calculated)
2.218(3)
1.401(6) a2.190(3) a
2.213(3) a
[K(thf)6][{(Me3Si)2N}2La(thf)]2(μ-η22-N2) (45) [65]--0-
[K(thf)6][{(Me3Si)2N}2Lu(thf)]2(μ-η22-N2) b (46) [65]1.414(8)2.163(4)0979 (14N2)
2.180(4)
[K][{(Me3Si)2N}2Y(thf)]23222-N2) b (47) [93]1.405(3)2.225(2)14.22989 (14N2)
2.242(2)956 (15N2)
[K][{(Me3Si)2N}2Gd(thf)]23222-N2) b (48) [97]1.395(3)2.248(2)13.64-
2.274(2)
[K][{(Me3Si)2N}2Tb(thf)]23222-N2) b (49) [97]1.401(3)2.235(2)16.12-
2.260(2)
[K][{(Me3Si)2N}2Dy(thf)]23222-N2) b (50) [97]1.404(5)2.229(4)15.27-
2.242(4)
[K(18c6)(thf)2][{(Me3Si)2N}2Y(thf)]2(μ-η22-N2) b (51) [97]1.396(3)2.1909(17)0-
2.2136(16)
[K(18c6)(thf)2][{(Me3Si)2N}2Gd(thf)]2(μ-η22-N2) b (52) [94]1.401(4)2.224(2)0-
2.249(2)
[K(18c6)(thf)2][{(Me3Si)2N}2Tb(thf)]2(μ-η22-N2) b (53) [94]1.394(3)2.2056(15)0-
2.2345(15)
[K(18c6)(thf)2][{(Me3Si)2N}2Dy(thf)]2(μ-η22-N2) b (54) [94]1.393(7)2.199(4)0-
2.213(4)
[K(18c6)(thf)2][{(Me3Si)2N}2Ho(thf)]2(μ-η22-N2) b (55) [95]1.404(4)2.188(2)0-
2.210(2)
[K(18c6)(thf)2][{(Me3Si)2N}2Er(thf)]2(μ-η22-N2) b (56) [65]1.409(4)2.178(2)0-
2.204(2)
[Na(thf)6][{(Me3Si)2N}2Y(thf)]2(μ-η22-N2) b (57) [65]1.393(7)2.199(4)0-
2.213(4)
[Na(thf)6][{(Me3Si)2N}2Er(thf)]2(μ-η22-N2) b (58) [65]1.403(4)2.1817(19) 2.2019(19)0-
[K(thf)6][(2,6-tBu–C6H3O)2Dy(thf)]2(μ-η22-N2) (59) [93]1.396(7)2.197(3)0962 (14N2)
2.203(4)
[K(thf)][(2,6-tBu–C6H3O)2Dy(thf)]23222-N2) (60) [93]1.402(7)2.235(5)6.59-
2.209(5)
a Second independent molecule in unit cell; b trans arrangement of thf.
The first definitive evidence for an N23− reduction product of dinitrogen was demonstrated by Evans and co-workers [93]. The LnA3/M system of Y{N(SiMe3)2}3 with KC8 in thf afforded a mixture of [{(Me3Si)2N}2Y(thf)]2(μ-η22-N2) (22), [K(thf)6][{(Me3Si)2N}2Y(thf)]2(μ-η22-N2) (44) and [K][{(Me3Si)2N}2Y(thf)]23222-N2) (47) from which 44 and 47 could be isolated (Figure 8). The EPR spectrum of 4415N2 has a 9-line pattern consistent with a triplet of triplets due to two 15N and two 89Y nuclei and has a hyperfine coupling constant of 8.2 G implying a N-centred radical, while 4715N2 shows extra coupling to potassium; all spectra indicate the presence of the N23− ion. The N–N bond distances are 1.401(6) and 1.405(3) Å respectively and are intermediate between N–N single bonds (1.47 Å in N2H4) and N=N double bonds (1.25 Å in PhN=NPh) [98]. The N–N vibrational stretching frequency in 47 is 989 cm−1, significantly reduced from 1425 cm−1 for 22. Similarly to complexes 2236, the Y–N2 bonding interaction in these complexes can be described by the donation from a filled Y 4d orbital into an antibonding N2 πg orbital (HOMO). However, the orthogonal antibonding N2 πg orbital is now also occupied by a single electron.
Figure 8. Rare earth complexes with amide ligands resulting from N2 activation to N23−.
Figure 8. Rare earth complexes with amide ligands resulting from N2 activation to N23−.
Inorganics 03 00597 g008
Following this remarkable report, these types of complexes have been extended to more rare earth metal centres; [K(thf)6][{(Me3Si)2N}2Ln(thf)]2(μ-η22-N2) (Ln = La (45), Lu (46)) and [K][{(Me3Si)2N}2Ln(thf)]23222-N2) (Ln = Gd (48), Tb (49), Dy (50)). Additionally, the solvated counterion can also be varied to include 18-crown-6 in [K(18c6)(thf)2][{(Me3Si)2N}2Ln(thf)]2(μ-η22-N2) (Ln = Y (51), Gd (52), Tb (53), Dy (54), Ho (55), Er (56)) or can be exchanged for sodium in [Na(thf)6][{(Me3Si)2N}2Ln(thf)]2(μ-η22-N2) (Ln = Y (57), Er (58)) [65,94,95,97].
The solid state structures of 4446 and 5158 which contain outer-sphere counterions have the common Ln2N2 planar core like those of rare earth N22− complexes 142 which have been structurally characterised. However, complexes 4750, which have inner-sphere K+ ions, display bent Ln2N2 cores with torsion angles of 14.22° (47), 13.64° (48), 16.12° (49) and 15.27° (50).
Complexes 4850 (inner-sphere K+) and 5256 (outer-sphere K+) all display interesting magnetic properties; 48 and 52 have the strongest magnetic exchange couplings in a GdIII complex with exchange constants of −27 cm−1, and 49, 50 and 5356 demonstrate single-molecule-magnet behaviour [94,95,99,100]. Combination of rare earth complexes which demonstrate both high anisotropy and strong exchange coupling potentially provides a route to single-molecule magnets with high blocking temperatures. The diffuse nature of the N23− radical facilitates strong coupling in these systems by overlap of the Ln 4f orbitals with the bridging dinitrogen ligand. 53 and 54 exhibit magnetic hysteresis up to record blocking temperatures of 13.9 K (0.9 mTs−1 sweep rate) and 8.3 K (0.08 Ts−1 sweep rate) respectively. Competing LnIII–LnIII antiferromagnetic coupling is observed in complexes at low temperatures in 4850, which have a non-zero Ln–N–N–Ln dihedral angle, demonstrating the importance of geometry of the Ln2N2 unit to magnetic behaviour.

2.2.2. Aryloxide Ancillary Ligands

As for yttrium complexes 44 and 47, the dysprosium aryloxide complexes [K(thf)6][(2,6-tBu–C6H3O)2Dy(thf)]2(μ-η22-N2) (59) and [K(thf)][(2,6-tBu–C6H3O)2Dy(thf)]23222-N2) (60) were first isolated from reaction of DyI2 with KO-2,6-tBu–C6H3 (Figure 9) [65,93]. Reoxidation of 59 with AgBPh4 affords the N22− complex [(2,6-tBu–C6H3O)2Dy(thf)2]2(μ-η22-N2) (38). From the solid state structures the N–N bond lengths are 1.396(7) (59) and 1.402(7) Å (60), which fall in the range of reported N23− complexes (Table 2). As anticipated, 59 has a planar Ln2N2 core in the solid state whereas it is bent in complex 60 with a Ln–N–N–Ln torsion angle of 6.59°.
Figure 9. Rare earth complexes with aryloxide ligands resulting from N2 activation to N23−.
Figure 9. Rare earth complexes with aryloxide ligands resulting from N2 activation to N23−.
Inorganics 03 00597 g009

2.3. Complexes Containing a Formally N24− Ligand

For reference, N–N and M–N(N2) bond lengths obtained from single crystal X-ray diffraction experiments are summarised in Table 3.
Table 3. Summary of rare earth N24− complexes.
Table 3. Summary of rare earth N24− complexes.
Complex (#) [Reference]N–N Bond Length (Å)Ln–N (N2) Bond Lengths (Side-on) (Å)Ln–N (N2) Bond Lengths (End-on) (Å)
N21.0975 [66]--
[Li(thf)2]2[(Et2calix[4]pyrrole)Sm]2(N2Li4) (61) [101]1.525(4)2.357(2)-
2.342(2)
[{Ph2C(C4H3N)2}Sm(thf)]441122-N2) (62) [102]1.412(17)2.327(3)2.177(8)
2.327(3)2.177(8)
[{CyC(C4H3N)2}Sm(thf)]441122-N2) (63) [103]1.392(16)2.339(3)2.160(8)
2.339(3)2.160(8)
[{Et2C(C4H3N)2}Sm(thf)]441122-N2) (64) [104]1.415(4)2.328(3)2.145(3)
2.342(3)
[{Ph(Me)C(C4H3N)2}Sm(dme)]441122-N2) (65) [104]1.42(2)2.316(13)2.149(11)
2.316(12)
[Na(thf)]2[{CyC(C4H3N)2}Sm(thf)]46111122-N2) (66) [103]1.371(16)2.332(11)2.178(10)
2.324(11)
[{Li(thf)}33-Cl)][(Cycalix[4]pyrrole)Sm]2(μ-η22-N2) (67) [105]1.08(3)2.880(18)-
2.974(18)
[(Li(thf)2][(Cycalix[4]pyrrole)2Sm3Li2](μ511222-N2) (68) [105]1.502(5)2.249(4) (Sm(1)–N(1))-
2.253(4) (Sm(1)–N(1))
2.355(4) (Sm(2)–N(1))
2.370(4) (Sm(2)–N(1))
2.398(4) (Sm(3)–N(1))
2.376(4) (Sm(3)–N(1))
To date, all examples of complexes containing an N24− ligand, derived from dinitrogen activation by a rare earth metal centre involve multidentate ligands (Figure 10) [101,102,103,104,105,106]. The first example of a structurally characterised rare earth complex containing an N24− ligand was reported by Gambarotta and co-workers [101]. The octametallic complex [Li(thf)2]2[(Et2calix[4]pyrrole)Sm]2(N2Li4) (61) was prepared by reaction of SmCl3(thf)3 with [Li(thf)]4[Et2calix[4]pyrrole] and subsequent reduction with Li metal under an argon atmosphere, followed by exposure to N2. The reaction by-products were not determined. In the solid state structure, there is an octahedron of metal ions coordinating to the N24− ligand which is η2-bound to the two SmIII in the apical sites, and both η2- and η1-bound to the opposite pairs of the four Li+ ions in the equatorial plane. The N–N distance of 1.525(4) Å, combined with a total magnetic moment of 2.72 μB, is consistent with formulation of the complex as [SmIII]2(N24−).
Figure 10. Lanthanide complexes with multidentate donors resulting from N2 activation to N24−.
Figure 10. Lanthanide complexes with multidentate donors resulting from N2 activation to N24−.
Inorganics 03 00597 g010
A family of tetrametallic Sm dipyrrolide complexes [{R2C(C4H3N)2}Sm(L)]441122-N2) (L = thf, CR2 = CPh2 (62), CCy (63), CEt2 (64); L = dme, CR2 = C(Me)Ph (65)) was prepared by reaction of SmI2(thf)2 with the corresponding alkali metal dipyrrolide salt, and display both side-on and end-on N2 coordination [102,103,104]. 6265 were stable both thermally and in vacuo. All compounds were structurally characterised using X-ray diffraction experiments; the N–N bond lengths range from 1.392(16) to 1.42(2) Å, and variable coordination modes of the pyrrolide ligands are observed across all complexes. Reduction of 63 with Na afforded [Na(thf)]2[{R2C(C4H3N)2}Sm(thf)]46111122-N2) (66). The N24− ligand is bound end-on to two Sm ions and side-on to the other two, as well as being end-on bound to two Na+ ions. Assignment of the reduction level of dinitrogen to N24− leads to a formal oxidation state of +2.5 for each samarium centre. This is proposed on the basis of the N–N bond distance (1.371(16) Å; slightly shorter than that expected for an N–N single bond), magnetic moment (4.05 μB; lower than the analogous divalent complex 63) and short Sm–Sm contacts in the solid state which may promote magnetic couplings.
Tetra-calix-pyrrole complexes [{Li(thf)}33-Cl)][(Cycalix[4]pyrrole)Sm]2(μ-η22-N2) (67) and the unusual trimetallic [Li(thf)2][(Cycalix[4]pyrrole)2Sm3Li2](μ511222-N2) (68) contain formal N24− ligands, on the basis of charge neutrality implied from the presence of SmIII centres but have disparate N–N bond lengths of 1.08(3) Å and 1.502(5) Å respectively [105]. It should be noted that the crystallographic data for 67 was only of a good enough quality to obtain structural connectivity.

3. Dinitrogen Activation by Actinide Complexes

3.1. Complexes Containing an Activated N2 Ligand

Perhaps surprisingly, given the number of examples of dinitrogen activation by rare earth complexes, there are very few examples of N2 activation with actinide complexes despite the presence of uranium in early catalysts for the Haber–Bosch process (Figure 11) [107]. Actinide-element bonding in the model system [X3An]2(μ-η22-N2) (An = Th–Pu, X = F, Cl, Br, Me, H, OPh) has recently been studied using relativistic DFT calculations [108].
Figure 11. Actinide complexes resulting from N2 activation (no N–N bond cleavage).
Figure 11. Actinide complexes resulting from N2 activation (no N–N bond cleavage).
Inorganics 03 00597 g011
For reference, N–N bond lengths obtained from single crystal X-ray diffraction experiments, N–N stretching frequencies (obtained by IR or Raman spectroscopy) and 14/15N-NMR spectroscopic data of actinide N2 complexes are summarised in Table 4.
Table 4. Summary of actinide N2 complexes.
Table 4. Summary of actinide N2 complexes.
Complex (#) [Reference]StabilityN–N Bond Length (Å)N–N Frequency (cm−1)14/15N-NMR Spectroscopy (ppm) a
N2-1.0975 [66]2331 [67]−75 [68]
[{N(CH2CH2NSiMe2tBu)3}U]2(μ-η22-N2) (69) [109]Stable under N2 (1 atm)1.109(7)--
N2 dissociation in vacuo
{Ph(tBu)N}3Mo(μ211-N2)U{N(tBu)(3,5-Me–C6H3)}3 (70) [110]Stable in vacuo at 25 °C1.232(11)- (14N2)-
“Thermally stable”1547 (15N2)
{(3,5-Me-C6H3)(Ad)N}3Mo(μ211-N2)U{N(tBu)(3,5-Me–C6H3)}3 (71) [110]Stable in vacuo at 25 °C1.23(2)1568 (14N2)-
“Thermally stable”1527 (15N2)
[(η5-C5Me5)(η8-1,4-SiiPr3–C8H4)U]2(μ-η22-N2) (72) [111]75% conversion to 72 at 50 psi N21.232(10)--
N2 dissociation in vacuo, in solution and solid state
5-C5Me5)3U(η1-N2) (73) [112]Crystallisation at 80 psi N21.120(14)2207 (14N2)-
N2 dissociation in vacuo or in solution under N2 (1 atm)2134 (15N2)
[(2,6-tBu–C6H3O)3U]2(μ-η22-N2) (74) [88]N2 dissociation in vacuo and in solution at 25 °C1.163(19)--
1.204 (17)
1.201(19)
[(2,4,6-tBu–C6H2O)3U]2(μ-η22-N2) (75) [88]Stable in vacuo at 25 °C1.236(5)1451 (14N2)-
N2 dissociation at 80 °C in solution1404 (15N2)
[{(Mes)3SiO}3U]2(μ-η22-N2) (76) [87]Stable in vacuo at 25 °C1.124(12) (eclipsed)1437 (14N2)4213.5
Slowly forms U{OSi(Mes)3}4 at 100 °C in solution1.080(11) (staggered)1372 (15N2)
a Referenced to CH315NO.
The first example of dinitrogen activation by an actinide complex was reported by Scott and co-workers; the trivalent uranium complex {N(CH2CH2NSiMe2tBu)3}U reacts with N2 (1 atm) to yield [{N(CH2CH2NSiMe2tBu)3}U]2(μ-η22-N2) (69) [109]. In solution, the reaction is reversible and 69 converts back to the trivalent uranium starting material when freeze-pump-thaw degassed (Scheme 1). The solid state structure of 69 illustrates the side-on binding mode of N2 and features an N–N bond length of 1.109(7) Å. Alongside solution magnetic susceptibility measurements of 3.22 μB per uranium centre, these data agree with the dimer being formulated as [UIII]2(N20).
Scheme 1. Reversible N2 binding in 69.
Scheme 1. Reversible N2 binding in 69.
Inorganics 03 00597 g017
Cummins and co-workers reported thermally stable heterobimetallic U–Mo N2 complexes featuring the end-on binding mode of N2; the UIII tris(amide) U{N(tBu)Ar}3(thf) (Ar = 3,5-Me–C6H3) reacts with Mo{N(R)Ar′}3 under an N2 atmosphere (1 atm) to yield {Ar′(R)N}3Mo(μ211-N2)U{N(tBu)Ar}3 (R = tBu, Ar′ = Ph (70); R = Ad, Ar′ = 3,5-Me–C6H3 (71)) [110]. The solid state molecular structure of 70 shows an N–N bond distance of 1.232(11) Å, consistent with N22− and a formal oxidation state of UIV. In principal, UIII and MoIII metal centres can provide the 6 electrons required for N2 bond cleavage and the stability of 70 and 71 should be noted with reference to Mo{N(tBu)Ar′}3 which cleaves N2 and forms a terminal nitride (Mo≡N) under mild conditions [113,114,115].
Reversible side-on binding and N2 activation was demonstrated by Cloke et al., using a mixed sandwich UIII pentalene complex. (η5-C5Me5)(η8-1,4-SiiPr3–C8H4)U reacts with N2 to yield [(η5-C5Me5)(η8-1,4-SiiPr3–C8H4)U]2(μ-η22-N2) (72) [111]. The solid state structure features an N–N bond length of 1.232(10) Å which is consistent with reduction to N22−; regardless of the formal reduction level, the relief of steric crowding in 72 likely drives the facile loss of N2 when there is no overpressure.
To date, the only example of a monometallic f-element complex of N2 was reported by Evans et al.; sterically crowded U(η5-C5Me5)3 reacts with N2 (80 psi) to afford (η5-C5Me5)3U(η1-N2) (73) [112]. N2 binding is reversible and lowering the pressure results in N2 dissociation. The solid state structure of 73 shows the N2 ligand is bound end-on and linearly (U–N–N = 180°) and the N–N distance is 1.120(14) Å which is statistically equivalent with free N2.
Arnold and co-workers reported that the trivalent uranium aryloxides U(OAr)3 (Ar = 2,6-tBu–C6H3 or 2,4,6-tBu–C6H2) bind N2 (1 atm) to form the side-on bound N2 adducts [(ArO)3U]2(μ-η22-N2) (74 and 75 respectively) [88]. Though 74 was obtained as a minor product, the more sterically hindered 75 was formed in quantitative yield and was stable under dynamic vacuum and in the presence of coordinating solvents and polar small molecules (CO and CO2) under ambient conditions. N2 loss was observed when 75 was heated to 80 °C in a toluene solution. The solid state N–N bond lengths are 1.163(19), 1.204(17), 1.201(19) Å (74) and 1.236(5) Å (75) which indicate significant N2 reduction by the electron rich UIII metal centres. Consistent with this, Raman spectroscopy performed on 75 showed a strong band at 1451 cm−1 for the N–N stretch (1404 cm−1 in the 7515N2) which is significantly lower than in free N2 (2331 cm−1) [67]. DFT calculations indicate a 5Ag ground state which agrees with the [UIV]2(N22−) description of 75. Significantly, the U–N (N2) interaction derives from two occupied MOs showing π backbonding from uranium f orbitals into an N2 antibonding πg orbitals and that the interaction is strongly polarised. The bonding description is very similar to that in previously calculated models for 69 (formally N20) [116,117] and 72 (N22−) [118] which display very different N–N bond lengths. While experimental N–N bond distances determined by X-ray diffraction experiments are undeniably useful for quick comparisons, the bond length is likely underestimated since the data is based on electron density rather than atomic positions and thus may not reflect the level of dinitrogen reduction. In these studies, N–N stretching wavenumbers were more accurately reproduced by calculation than bond length and it is proposed that this would be a more suitable measurement for probing N2 reduction.
The most robust actinide N2 complex prepared to date is [{(Mes)3SiO}3U]2(μ-η22-N2) (76) (Mes = 2,4,6-Me–C6H2) and is stable both in vacuo and in toluene solution up to 100 °C, at which point U{OSi(Mes)3}4 is slowly formed as the major product (52% conversion after 18 h) [87]. 76 is isolated from the reaction of U{N(SiMe3)2}3 with 3 equivalents of HOSi(Mes)3 under an N2 atmosphere (1 atm). Raman spectroscopy shows a peak at 1437 cm−1 assigned to the N–N stretching mode, indicating a significant level of reduction with respect to free dinitrogen (2331 cm−1) and comparing well with 1451 cm−1 recorded for 75 where reduction to N22− was assigned. The N–N distances in the solid state are 1.124(12) Å (76-Eclipsed) and 1.080(11) Å (76-Staggered), which are statistically equivalent to that of free N2; the disparity in implied reduction of N2 from Raman spectroscopy and X-ray diffraction experiments again highlighting that the latter may not be best suited for assigning reduction in these systems.

3.2. Complexes Resulting from N2 Cleavage

Tetra-calix-pyrrole ligands bound to SmII centres have been demonstrated to activate N2 by Gambarotta and co-workers [101]. With UIII, an unprecedented example of N–N bond cleavage using an molecular f-element complex was observed; when [K(dme)][(Et2calix[4]pyrrole)U(dme)] is treated with potassium naphthalenide under an atmosphere of N2, N–N bond cleavage occurs to afford [K(dme)4][{K(dme)(Et2calix[4]pyrrole)U}2(μ-NK)2] (77) (Scheme 2) [119]. 77 contains two bridging nitrides (U–N: 2.076(6) and 2.099(5) Å) which have contacts with potassium ions (N–K: 2.554(6) Å) that bridge two pyrrolide units on separate ligands. It was postulated that 77 is a Class 1 UIV–UV mixed valence complex on the basis of an absorption at 1247 nm in the near-IR spectrum which is characteristic of UV. The paramagnetism of 77 resulted in NMR silence in both 15N- and 14N-NMR spectra.
Scheme 2. N–N bond cleavage using a UIII complex to form 71.
Scheme 2. N–N bond cleavage using a UIII complex to form 71.
Inorganics 03 00597 g018
Reduction of the thorium bisphenolate complex [K(dme)2][{(2-tBu-4-Me–C6H2O)2-6-CH2}2ThCl(dme)] (A) with potassium naphthalenide under an atmosphere of dinitrogen unexpectedly resulted in the amide complex [K(dme)4][{(2-tBu-4-Me–C6H2O)2-6-CH2}2Th(NH2)(dme)] 78 which is the first example of N2 functionalisation using an f-element complex (Scheme 3). The parent [NH2] amide ligand is confirmed through 15N-NMR spectroscopy which shows a triplet at 155.01 ppm (1JNH = 57.2 Hz) [120]. The proposed mechanism of this transformation involves formation of a formally zero-valent thorium intermediate which contains two bound [C10H8K(18c6)] fragments (identified through a single crystal X-ray diffraction experiment). This intermediate can then react with the starting material A leading to N2 activation, cleavage and hydrogenation as a result of H atom abstraction from solvent molecules.
Scheme 3. N2 cleavage and hydrogenation by a thorium complex.
Scheme 3. N2 cleavage and hydrogenation by a thorium complex.
Inorganics 03 00597 g019

4. White Phosphorus Activation by Rare Earth Complexes

Rare earth complexes resulting from P4 activation are illustrated in Figure 12. For reference, average P–P and Ln–P bond lengths obtained from single crystal X-ray diffraction experiments, and 31P-NMR spectroscopic resonances are summarised in Table 5. The structural cores of complexes 7984 are shown in Figure 13 for clarity.
Figure 12. Rare earth complexes resulting from P4 activation.
Figure 12. Rare earth complexes resulting from P4 activation.
Inorganics 03 00597 g012
Figure 13. Overview of the LnxPn structural cores resulting from P4 activation by rare earth complexes.
Figure 13. Overview of the LnxPn structural cores resulting from P4 activation by rare earth complexes.
Inorganics 03 00597 g013
Table 5. Summary of rare earth P4 activation complexes.
Table 5. Summary of rare earth P4 activation complexes.
Complex (#) [Reference]Average P–P Bond Lengths (Å)Average M–P Bond Length (Å)31P-NMR Spectroscopy (298 K) (ppm) a
P42.21 [121]-−488 to −527 [122]
[(η5-C5Me5)2Sm]442222-P8) (79) [123]2.195 (Pcorner–Pinner)3.047-
2.291 (Pinner–Pinner)
[{Fe(1-NSitBuMe2–C5H4)2}Sc]442222-P8) (80) [124]2.204 (Pcorner–Pinner)2.768+45.7
2.308 (Pinner–Pinner)+96.2
[{Fe(1-NSitBuMe2–C5H4)2}Sc]33222-P7) (81) [124]2.229 (Pbottom–Pbottom)2.750+23.1
2.197 (Pedge–Pbottom)−118.9
2.201 (Papex–Pedge)−131.4
[{Fe(1-NSitBuMe2–C5H4)2}Y(thf)]33222-P7) (82) [124]2.238 (Pbottom–Pbottom)2.950−21.1
2.176 (Pedge–Pbottom)−82.4
2.188 (Papex–Pedge)−130.3
[{Fe(1-NSitBuMe2–C5H4)2}La(thf)]33222-P7) (83) [125]2.258 (Pbottom–Pbottom)3.120−75
2.161 (Pedge–Pbottom)
2.191 (Papex–Pedge)
[{Fe(1-NSitBuMe2–C5H4)2}Lu(thf)]33222-P7) (84) [125]2.233 (Pbottom–Pbottom)2.893+0.8
2.181 (Pedge–Pbottom)−96.8
2.183 (Papex–Pedge)−133.3
a Referenced to 85% H3PO4.
Roesky and co-workers reported the first example of a molecular polyphosphide of the rare earth elements, [(η5-C5Me5)2Sm]442222-P8) (79) [123]. The samarocene (η5-C5Me5)2Sm activates P4 to yield a P84− fragment with a realgar-type structure, a process proposed to be driven by the one-electron oxidation of the divalent samarium metal centre. 79 has molecular D2d symmetry and the [Cp*2Sm] units bridge the P84− cage with Sm–P distances in the range of 2.997(2) to 3.100(2) Å. DFT calculations support the strongly ionic character of the Sm–P bonds.
Activation of P4 by group 3 metal centres was first reported by Diaconescu and co-workers [124]. Reaction of the scandium arene inverse-sandwich complexes [{Fe(1-NSitBuMe2–C5H4)2}Sc]2(μ-arene) (arene = C10H8 or C14H10) with P4 resulted in displacement of the neutral arene and formation of a mixture of the tetrametallic [{Fe(1-NSitBuMe2–C5H4)2}Sc]442222-P8) (80) and trimetallic [{Fe(1-NSitBuMe2–C5H4)2}Sc]33222-P7) (81). The mixtures were readily separated and the product distribution could be controlled by the stoichiometry of P4 and the nature of the arene starting material. 80 possesses a realgar-type P84− unit whereas 81 contains a Zintl-type P73− unit [126], the first example of its formation in the absence of strong alkali metal reducing agents. Solution phase 31P-NMR spectroscopy demonstrates a diagnostic AA′A′′MM′M′′X spin system. The analogous yttrium arene inverse-sandwich complex [{Fe(1-NSitBuMe2–C5H4)2}Y(thf)]2(μ-C10H8) activates P4 to yield [{Fe(1-NSitBuMe2–C5H4)2}Y(thf)]33222-P7) (82) as the sole product where the larger coordination sphere of yttrium is saturated with an additional thf molecule [127]. Importantly, in the context of functionalisation of white phosphorus to organophosphorus compounds, both 81 and 82 were shown to react with 3 equivalents of Me3SiI to yield P7(SiMe3)3 and {Fe(1-NSitBuMe2–C5H4)2}MI (Scheme 4).
Scheme 4. P73− functionalisation with Me3SiI.
Scheme 4. P73− functionalisation with Me3SiI.
Inorganics 03 00597 g020
This chemistry was later extended to lanthanum and lutetium using the same methodology, forming [{Fe(1-NSitBuMe2–C5H4)2}Ln(thf)]33222-P7) (Ln = La (83), Lu (84)) [125]. The valence tautomerisation of P73− in 83, which occurs at a similar temperature to Li3P7 but does not require donor solvents, was proposed to take place by a lanthanum-assisted mechanism involving simultaneous formation and breaking of 4 La–P bonds.

5. White Phosphorus Activation by Actinide Complexes

Actinide complexes resulting from P4 activation are illustrated in Figure 14. For reference, average P–P and An–P bond lengths obtained from single crystal X-ray diffraction experiments, and 31P-NMR spectroscopic resonances are summarised in Table 6. The structural cores of complexes 8591 are shown in Figure 15 for clarity.
Table 6. Summary of actinide P4 activation complexes.
Table 6. Summary of actinide P4 activation complexes.
Complex (#) [Reference]Average P–P Bond Lengths (Å)Average An–P Bond Length (Å)31P-NMR Spectroscopy (ppm) a
P42.21 [121]-−488 to −527 [122]
[(η5-1,3-tBu–C5H3)2Th](μ-η33-cyclo-P3)[(η5-1,3-tBu–C5H3)2ThCl] (85) [128]2.1852.913−75.7 (293 K)
[(η5-1,3-tBu–C5H3)2Th]2(μ-η33-P6) (86) [128]2.2342.904 (Th-η2-P)+125.4 (293 K)
2.844 (Th-η1-P)−41.9 (293 K)
[{(3,5-Me–C6H3)(tBu)N}3U](μ-η44-cyclo-P4) (87) [129]2.1603.127+794
[{(3,5-Me–C6H3)(Ad)N}3U](μ-η44-cyclo-P4) (88) [129]2.1593.124+803
[(η5-C5Me5)(η8-1,4-SiiPr3–C8H6)U]2(μ-η22-cyclo-P4) (89) [130]2.1502.977+718
[HC(SiMe2N-4-Me–C6H4)3U]33222-P7) (90) [130]2.249 (Pbottom–Pbottom)2.990-
2.187 (Pedge–Pbottom)
2.209 (Papex–Pedge)
[{N(CH2CH2NSiiPr3)3}U]2(μ-η55-cyclo-P5) (91) [131]2.0063.280-
a Referenced to 85% H3PO4.
Figure 14. Actinide complexes resulting from P4 activation.
Figure 14. Actinide complexes resulting from P4 activation.
Inorganics 03 00597 g014
Figure 15. Overview of the AnxPn structural cores resulting from P4 activation by actinide complexes.
Figure 15. Overview of the AnxPn structural cores resulting from P4 activation by actinide complexes.
Inorganics 03 00597 g015
The first report of P4 activation by an actinide complex came from Scherer et al., using thorium [128]. The butadiene complex (η5-1,3-tBu–C5H3)2Th(η4-C4H6) reacts with P4 at 100 °C in the presence of MgCl2(OEt2) to yield [(η5-1,3-tBu–C5H3)2Th](μ-η33-cyclo-P3)[(η5-1,3-tBu–C5H3)2ThCl] (85). In the absence of MgCl2(OEt2), [(η5-1,3-tBu–C5H3)2Th]2(μ-η33-P6) (86) was afforded as a consequence of P4 fragmentation and subsequent catenation. 85 features a cyclo-P33− unit and formal ThIV centres. In the solid state structure, the coordination environment about the thorium centres is trigonal planar and tetrahedral (Cl bound) with Th–P distances ranging from 2.809(6) to 2.974(8) Å. The P–P distances are 2.171(9), 2.192(9) and 2.192(8) Å. Only a single broad resonance is observed in the 31P-NMR spectrum at 293 K but cooling to 193 K leads to splitting and the observation of an A2B system; the barrier to rotation of the cyclo-P3 unit was estimated to be ca. 44 kJ·mol−1. 86 contains a P64− bicycle with thorium metal centres capping the five-membered rings. Th–P bond lengths in the solid state structure range from 2.840(7) to 2.919(7) Å.
The following report of P4 activation came over a decade later and was the first using a uranium complex [129]. The UIII tris(amide) U{N(R)Ar}3(thf) (R = tBu or Ad, Ar = 3,5-Me–C6H3) reacts with 0.5 equivalents of P4 to yield [{Ar(R)N}3U](μ-η44-cyclo-P4) (R = tBu (87), Ad (88)) which contains a cyclo-P42− unit and where the metal centres have been formally oxidised to UIV. In the solid state structures, the average P–P bond distance is 2.159 Å and the P–P–P angle is 90°; both statistically equivalent across the two structures. Resonances at 794 and 803 ppm were observed for 87 and 88 respectively in the 31P-NMR spectrum. Computational studies implied that the U–P bonding character is largely ionic with the presence of a weak δ-bonding interaction between filled U df hybrid orbitals and the P42− LUMO.
Cloke and co-workers described the related cyclo-P4 example [(η5-C5Me5)(η8-1,4-SiiPr3–C8H6)U]2(μ-η22-cyclo-P4) (89) which was prepared from (η5-C5Me5)(η8-1,4-SiiPr3–C8H6)U(thf) and 0.5 equivalents of P4 [130]. This was the first example of the μ-η22-P4 coordination mode [132,133] and DFT studies on a model system [(η5-C5H5)(η8-C8H8)U](μ-η22-cyclo-P4) support the formulation of the dimer with P42− and UIV oxidation states. The tilted cyclo-P42− unit leads to U–P bonding interactions involving both σ and π orbitals. The wedge shaped nature of the sterically demanding (η5-C5Me5)(η8-1,4-SiiPr3–C8H6)U fragment likely results in the slipped μ-η22 coordination mode.
Following on from the rare earth inverse sandwich complexes that resulted in P84− and P73− clusters, and P73− functionalisation [124,125], Liddle and co-workers reported the reaction of [HC(SiMe2NAr)3U]2(μ-η66-C6H5CH3) (Ar = 4-Me–C6H4) with 1.1 equivalents of P4 which afforded the first actinide Zintl complex [HC(SiMe2NAr)3U]33222-P7) (90) [134]. U–P bonding was determined to be essentially ionic. Interestingly, reaction of 90 with a number of electrophiles under ambient conditions led to functionalisation of the P73− unit and liberation of P7R3 (R = SiMe3, Me, Ph, Li(tmeda)) after P–Si, P–C or P–Li bond formation. Though not catalytic, 90 could be regenerated from this reaction mixture and two turnovers achieved demonstrating a significant step towards controlled P4 activation under mild conditions.
Very recently, Liddle and co-workers described the first example of a cyclo-P5 complex resulting from activation of P4 by an f-block complex [131]. [{N(CH2CH2NSiiPr3)3}U]2(μ-η55-cyclo-P5) (91) was prepared by reaction of {N(CH2CH2NSiiPr3)3}U with 0.25 equivalents of P4. Spectroscopic and magnetic measurements support oxidation to afford UIV centres and charge transfer resulting in a formal P52− ligand in this inverse sandwich complex. Despite the isolobal analogy of cyclo-P5 with the cyclopentadienyl anion, which bonds to metal centres using primarily σ- and π-bonding, calculations on 91 suggest that the principal U–P interactions involve polarised δ-bonding and this can be attributed to the energetically available uranium 5f orbitals of correct δ-symmetry.
Compared to both rare earth and actinide complexes, the activation of P4 by transition metal complexes have proven to result in a wide variety of activation products [58,59]; with notable examples including fragmentation resulting in terminal and bridging P1 ligands [135,136,137], P2 ligands [138], cyclo-P3 ligands [56,139], fragmentation to other P4 ligands [140], coordination of P4 tetrahedra [141,142], and expansion to Pn (n = 5–14) ligands [143,144,145,146]. More significantly, functionalisation of these phosphorus units has also been observed.

6. Arsenic, Antimony and Bismuth Activation by Rare Earth and Actinide Complexes

There is only a single example of molecular arsenic activation by a rare earth or actinide metal complex. Reaction of the thorium butadiene complex (η5-1,3-tBu–C5H3)2Th(η4-C4H6) with As4 (yellow arsenic) in boiling xylene affords [(η5-1,3-tBu–C5H3)2Th]2(μ-η33-As6) in analogy to the previously reported P4 chemistry of Scherer et al. (Figure 16) [147]. In the solid state molecular structure; the average Th–(η2-As6) bonds are 3.027 Å, average Th–(η1-As) bonds are 2.922 Å and average As–As bonds are 2.459 Å.
Figure 16. Actinide complexes resulting from As4 activation.
Figure 16. Actinide complexes resulting from As4 activation.
Inorganics 03 00597 g016
In contrast, activation of molecular arsenic by homogeneous transition metal complexes is considerably more diverse; formation of cyclo-Asn ligands (n = 3–6, 8) [54,148], metal arsenic clusters [149], coordination of intact As4 tetrahedra to metal ions [133,150,151], fragmentation into As2 and other As4 ligands [152,153,154], catenation to As10 and As12 ligands [155], reactions to form PnAsm ligands [156], and full As4 fragmentation resulting in terminal M≡As arsenide bonds [157] have all been reported.
Beyond arsenic in group 15 are antimony and bismuth. While activation of molecular forms of these elements is unlikely, it is worth noting that Scheer and co-workers reported a tungsten terminal stibido complex {N(CH2CH2NSiMe3)3}3W≡Sb prepared from reaction of {N(CH2CH2NSiMe3)3}3WCl with LiSb(H){CH(SiMe3)2} [158], while Breunig et al., have reported [(η5-C5Me5)2Mo(CO)2](cyclo-Sb3) and [(η5-C5H5)2Mo(CO)2](cyclo-Sb3), which are a result of reaction of [(η5-CpR)2Mo(CO)3]2 with (tBuSb)4 [2]. In terms of rare earth complexes; [(η5-C5Me5)2Sm]23122-Sb3){(η5-C5Me5)2(thf)Sm} and [(η5-C5Me5)2Sm]222-Bi2) were prepared by reaction of [(η5-C5Me5)2Sm]2 with SbPh3 and BiPh3 respectively [3,4].

7. Conclusions and Perspectives

To date, a wide range of rare earth dinitrogen complexes have been prepared (168), including group 3 metal ions and 4f elements at both ends of the periodic table, despite the limited radial extension of the 4f orbitals and the trivalent oxidation state being the most prevalent. In fact, apart from the very first f-element dinitrogen complex (4), all of the other complexes are air-sensitive but stable to N2 dissociation in vacuo. Reduction of N2 to N22− has most commonly been achieved with the [A2(thf)xLn]2(μ-η22-N2) structural motif (138) whereas examples of reduction to N24− have all involved more complex multidentate ligands (6168). The nature of the bonding in complexes of the form [A2(thf)xLn]2(μ-η22-N2) has been extensively studied for both group 3, and closed and open shell 4fn metal ions demonstrating that the Ln–N (N2) bonding is based on a Ln nd–N2 π* interaction. These systems have allowed for the first definitive characterisation of the N23− radical reduction product of dinitrogen (44, 47) and this has now been extended to many of the rare earth elements (45, 46, 4860). Isolation of the N23− radical in homogeneous complexes is significant since it is likely to be a transient species in other transition metal systems and may also have a role in biological N2 fixation.
In terms of reactivity, the rare earth N2 complexes prepared thus far react with N2 dissociation rather than N2 cleavage and functionalisation [77,159,160,161]. Related to this, the understanding of the nature of bonding of Ln–N multiple bonds is of fundamental interest and the isolation of terminal imido (Ln=NR) complexes has only recently been reported [162,163,164]. It is also significant to consider the reactivity of Ln–N2 complexes in the context of other low electron count early transition metal complexes; here, N2 cleavage or functionalisation can be achieved through ligand induced reductive cleavage [165,166,167,168,169,170,171,172,173], and more generally, N–N bond scission of reduced N2 derivatives can be attained through metal-ligand cooperativity [174,175].
Actinide dinitrogen complexes are much rarer, with just 8 examples of well-defined molecular uranium complexes (6976). Of these, only the heterobimetallic U–Mo end-on N2 complexes (70, 71) and recently prepared UIV aryloxide and siloxide side-on N2 complexes (75, 76) are thermally robust and stable when exposed to vacuum. Key steps forward have been made in the understanding of U–N (N2) bonding in the side-on N2 complexes as a polar covalent U 5f-N2 π* interaction, and in the rationalisation of the differences between solid state N–N bond lengths and the overall electronic structure of these complexes. This understanding, in combination with well-designed ligand sets may lead the way in the preparation of other isolable actinide dinitrogen complexes for further study.
Though the isolated actinide N2 complexes, like the rare earth N2 complexes, tend to react with N2 loss rather than N2 functionalisation or cleavage, there are two reports of such reactivity. Importantly, in both cases, the putative An(N2) complex was not observed. Cleavage of N2 by a uranium tetra-calix-pyrrole complex results in a bimetallic complex with bridging nitrides (78) whereas a thorium bisphenolate complex activates and functionalises N2 to a parent amide ligand [NH2] (79) through an unknown mechanism. Both reactions occur in the presence of an external reductant. Examples of isolable terminal uranium nitrides (U≡N), derived from NaN3, have only recently been reported [176,177,178], but it has already been demonstrated that these systems are capable of nitride functionalisation. 78 and 79 remain standout examples in demonstrating that actinide complexes can both cleave and functionalise N2, but also highlight how much more remains to be understood in this field.
P4 activation by rare earth complexes has led to both P84− ions with realgar-type structures (79, 80), and P73− ions (8184) using cyclopentadienyl and amido ancillary ligands. Promisingly, functionalisation of the P73− unit in 81 and 82 was found possible using Me3SiI to afford P7(SiMe3)3. This is an interesting prospect for the synthesis of organophosphorus compounds from a P4 building block. Actinide P4 activation results in a more diverse array of phosphorus ligands; P33− (85), P64−, (86), cyclo-P42− (8789), P73− (90), and cyclo-P52− (91). Similar to rare earth chemistry, the P73− ions in 90 could be functionalised by P–Si, P–C or P–Li bond formation to afford P7R3 units and this reaction cycle could be completed with two turnovers. 91 is the first example of an f-element being able to fragment and catenate P4 to cyclo-P52−. Despite the parallels with the cyclopentadienyl ligand, calculations suggest the U–P (cyclo-P5) interaction to be based on polarised δ-bonding and electronic structure in these systems can be described as [UIV]2(P52−). Putting this area into perspective, transition metals have already been shown to activate molecular phosphorus and, in limited examples, to result in further functionalisation.
There remains only a lone example of arsenic activation by a thorium butadiene complex leading to a P6 cage (92) and no examples using rare earth metals. The first examples of crystallographically characterised uranium arsenide (U–AsH2), arsenidene (U=AsH) and arsenido (U≡AsK) complexes have only recently been reported, using KAsH2 as a source of arsenic [179]. These compounds raise the question of the diverse reactivity that actinide complexes could be expected to show with molecular arsenic and whether the formation of an unsupported, terminal actinide arsenide bond (M≡As) is accessible.
It is clear that molecular pnictogen activation by rare earth and actinide metal complexes is an exciting field of study which remains underdeveloped with respect to transition metals and main group elements. These unique metals offer the potential of new reactivity and functionalisation chemistry with the pnictogen elements, while the fundamental study of M–pnictogen bonds remains important.

Acknowledgments

Zoë R. Turner thanks Prof. Dermot O’Hare (University of Oxford), SCG Chemicals for financial support and a SCG Research Fellowship, and Trinity College for a Junior Research Fellowship.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Kaltsoyannis, N. Does covalency increase or decrease across the actinide series? Implications for minor actinide partitioning. Inorg. Chem. 2013, 52, 3407–3413. [Google Scholar] [CrossRef] [PubMed]
  2. Breunig, H.J.; Rösler, R.; Lork, E. Complexes with Sb2 and cyclo-Sb3 ligands: The tetrahedranes [{C5H5(CO)2Mo}2Sb2], [C5H5(CO)2MoSb3], and [C5Me5(CO)2MoSb3]. Angew. Chem. Int. Ed. 1997, 36, 2819–2821. [Google Scholar] [CrossRef]
  3. Evans, W.J.; Gonzales, S.L.; Ziller, J.W. The utility of (C5Me5)2Sm in isolating crystallographically characterizable zintl ions. X-ray crystal structure of a samarium complex of (Sb3)3−. J. Chem. Soc. Chem. Commun. 1992, 1138–1139. [Google Scholar] [CrossRef]
  4. Evans, W.J.; Gonzales, S.L.; Ziller, J.W. Organosamarium-mediated synthesis of bismuth–bismuth bonds: X-ray crystal structure of the first dibismuth complex containing a planar M2(μ-η22-Bi2) unit. J. Am. Chem. Soc. 1991, 113, 9880–9882. [Google Scholar] [CrossRef]
  5. Hoffman, B.M.; Dean, D.R.; Seefeldt, L.C. Climbing nitrogenase: Toward a mechanism of enzymatic nitrogen fixation. Acc. Chem. Res. 2009, 42, 609–619. [Google Scholar] [CrossRef] [PubMed]
  6. Hu, Y.; Ribbe, M.W. Decoding the nitrogenase mechanism: The homologue approach. Acc. Chem. Res. 2010, 43, 475–484. [Google Scholar] [CrossRef] [PubMed]
  7. Hellman, A.; Baerends, E.J.; Biczysko, M.; Bligaard, T.; Christensen, C.H.; Clary, D.C.; Dahl, S.; van Harrevelt, R.; Honkala, K.; Jonsson, H.; et al. Predicting catalysis: Understanding ammonia synthesis from first-principles calculations. J. Phys. Chem. B 2006, 110, 17719–17735. [Google Scholar] [CrossRef] [PubMed]
  8. Schrock, R.R. Reduction of dinitrogen. Proc. Natl. Acad. Sci. USA 2006, 103, 17087. [Google Scholar] [CrossRef] [PubMed]
  9. Smil, V. Enriching the Earth: Fritz Haber, Carl Bosch and the Transformation of World Food Production; Massachusetts Institute of Technology Press: Cambridge, MA, USA, 2001. [Google Scholar]
  10. Erisman, J.W.; Sutton, M.A.; Galloway, J.; Klimont, Z.; Winiwarter, W. How a century of ammonia synthesis changed the world. Nat. GeoSci. 2008, 1, 636–639. [Google Scholar] [CrossRef]
  11. U.S. Geographical Survey, Mineral Commodity Surveys; U.S. Department of the Interior: Washington, DC, USA, 2015.
  12. Corbridge, D. Phosphorus: An Outline of its Chemistry, Biochemistry, and Technology, 5th ed.; Elsevier: New York, NY, USA, 1994. [Google Scholar]
  13. Quin, L.D. A Guide to Organophosphorus Chemistry; Wiley: New York, NY, USA, 2000. [Google Scholar]
  14. Engel, R. Synthesis of Carbon Phosphorus Bonds, 2nd ed.; CRC Press: Boca Raton, FL, USA, 2004. [Google Scholar]
  15. Withers, P.J.A.; Elser, J.J.; Hilton, J.; Ohtake, H.; Schipper, W.J.; van Dijk, K.C. Greening the global phosphorus cycle: How green chemistry can help achieve planetary P sustainability. Green Chem. 2015, 17, 2087–2099. [Google Scholar] [CrossRef]
  16. Liu, H. Catalytic Ammonia Synthesis; Plenum Press: New York, NY, USA, 1991. [Google Scholar]
  17. Schlögl, R. Catalytic synthesis of ammonia—A “Never-Ending Story”? Angew. Chem. Int. Ed. 2003, 42, 2004–2008. [Google Scholar] [CrossRef] [PubMed]
  18. Leigh, G.J. Haber–Bosch and Other Industrial Processes. In Catalysts for Nitrogen Fixation; Smith, B., Richards, R., Newton, W., Eds.; Springer: Houten, The Netherlands, 2004; Volume 1, pp. 33–54. [Google Scholar]
  19. Studt, F.; Tuczek, F. Theoretical, spectroscopic, and mechanistic studies on transition-metal dinitrogen complexes: Implications to reactivity and relevance to the nitrogenase problem. J. Comput. Chem. 2006, 27, 1278–1291. [Google Scholar] [CrossRef] [PubMed]
  20. MacKay, B.A.; Fryzuk, M.D. Dinitrogen coordination chemistry: On the biomimetic borderlands. Chem. Rev. 2004, 104, 385–401. [Google Scholar] [CrossRef] [PubMed]
  21. Barrière, F. Model Complexes of the Active Site of Nitrogenases: Recent Advances. In Bioinspired Catalysis; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2014; pp. 225–248. [Google Scholar]
  22. Rolff, M.; Tuczek, F. Nitrogenase and Nitrogen Activation. In Comprehensive Inorganic Chemistry II, 2nd ed.; Poeppelmeier, J.R., Ed.; Elsevier: Amsterdam, The Netherlands, 2013; pp. 593–618. [Google Scholar]
  23. Ribbe, M.W. Nitrogen Fixation: Methods and Protocols; Humana Press: New York, NY, USA, 2011. [Google Scholar]
  24. Barrière, F. Modeling of the molybdenum center in the nitrogenase FeMo-cofactor. Coord. Chem. Rev. 2003, 236, 71–89. [Google Scholar] [CrossRef]
  25. Smith, B.E.; Durrant, M.C.; Fairhurst, S.A.; Gormal, C.A.; Grönberg, K.L.C.; Henderson, R.A.; Ibrahim, S.K.; le Gall, T.; Pickett, C.J. Exploring the reactivity of the isolated iron-molybdenum cofactor of nitrogenase. Coord. Chem. Rev. 1999, 185–186, 669–687. [Google Scholar] [CrossRef]
  26. Rehder, D. Vanadium nitrogenase. J. Inorg. Biochem. 2000, 80, 133–136. [Google Scholar] [CrossRef]
  27. MacLachlan, E.A.; Fryzuk, M.D. Synthesis and reactivity of side-on-bound dinitrogen metal complexes. Organometallics 2006, 25, 1530–1543. [Google Scholar] [CrossRef]
  28. Fryzuk, M.D. Side-on end-on bound dinitrogen: An activated bonding mode that facilitates functionalizing molecular nitrogen. Acc. Chem. Res. 2009, 42, 127–133. [Google Scholar] [CrossRef] [PubMed]
  29. Poveda, A.; Perilla, I.C.; Pérez, C.R. Some considerations about coordination compounds with end-on dinitrogen. J. Coord. Chem. 2001, 54, 427–440. [Google Scholar] [CrossRef]
  30. Gambarotta, S.; Scott, J. Multimetallic cooperative activation of N2. Angew. Chem. Int. Ed. 2004, 43, 5298–5308. [Google Scholar] [CrossRef] [PubMed]
  31. McWilliams, S.F.; Holland, P.L. Dinitrogen binding and cleavage by multinuclear iron complexes. Acc. Chem. Res. 2015, 48, 2059–2065. [Google Scholar] [CrossRef] [PubMed]
  32. Ballmann, J.; Munha, R.F.; Fryzuk, M.D. The hydride route to the preparation of dinitrogen complexes. Chem. Commun. 2010, 46, 1013–1025. [Google Scholar] [CrossRef] [PubMed]
  33. Jia, H.-P.; Quadrelli, E.A. Mechanistic aspects of dinitrogen cleavage and hydrogenation to produce ammonia in catalysis and organometallic chemistry: Relevance of metal hydride bonds and dihydrogen. Chem. Soc. Rev. 2014, 43, 547–564. [Google Scholar] [CrossRef] [PubMed]
  34. Sivasankar, C.; Baskaran, S.; Tamizmani, M.; Ramakrishna, K. Lessons learned and lessons to be learned for developing homogeneous transition metal complexes catalyzed reduction of N2 to ammonia. J. Organomet. Chem. 2014, 752, 44–58. [Google Scholar] [CrossRef]
  35. Tanabe, Y.; Nishibayashi, Y. Developing more sustainable processes for ammonia synthesis. Coord. Chem. Rev. 2013, 257, 2551–2564. [Google Scholar] [CrossRef]
  36. Van der Ham, C.J.M.; Koper, M.T.M.; Hetterscheid, D.G.H. Challenges in reduction of dinitrogen by proton and electron transfer. Chem. Soc. Rev. 2014, 43, 5183–5191. [Google Scholar] [CrossRef] [PubMed]
  37. Rebreyend, C.; de Bruin, B. Photolytic N2 splitting: A road to sustainable NH3 production? Angew. Chem. Int. Ed. 2015, 54, 42–44. [Google Scholar] [CrossRef] [PubMed]
  38. Himmel, H.-J.; Reiher, M. Intrinsic dinitrogen activation at bare metal atoms. Angew. Chem. Int. Ed. 2006, 45, 6264–6288. [Google Scholar] [CrossRef] [PubMed]
  39. Chow, C.; Taoufik, M.; Quadrelli, E.A. Ammonia and dinitrogen activation by surface organometallic chemistry on silica-grafted tantalum hydrides. Eur. J. Inorg. Chem. 2011, 2011, 1349–1359. [Google Scholar] [CrossRef]
  40. Ohki, Y.; Fryzuk, M.D. Dinitrogen activation by group 4 metal complexes. Angew. Chem. Int. Ed. 2007, 46, 3180–3183. [Google Scholar] [CrossRef] [PubMed]
  41. Chirik, P.J. Dinitrogen functionalization with bis(cyclopentadienyl) complexes of zirconium and hafnium. Dalton Trans. 2007, 1, 16–25. [Google Scholar] [CrossRef] [PubMed]
  42. Kuganathan, N.; Green, J.C.; Himmel, H.-J. Dinitrogen fixation and activation by Ti and Zr atoms, clusters and complexes. New J. Chem. 2006, 30, 1253–1262. [Google Scholar] [CrossRef]
  43. Crossland, J.L.; Tyler, D.R. Iron–dinitrogen coordination chemistry: Dinitrogen activation and reactivity. Coord. Chem. Rev. 2010, 254, 1883–1894. [Google Scholar] [CrossRef]
  44. Hazari, N. Homogeneous iron complexes for the conversion of dinitrogen into ammonia and hydrazine. Chem. Soc. Rev. 2010, 39, 4044–4056. [Google Scholar] [CrossRef] [PubMed]
  45. Schrock, R.R. Catalytic reduction of dinitrogen to ammonia at a single molybdenum center. Acc. Chem. Res. 2005, 38, 955–962. [Google Scholar] [CrossRef] [PubMed]
  46. Nishibayashi, Y. Molybdenum-catalyzed reduction of molecular dinitrogen into ammonia under ambient reaction conditions. Comptes Rendus Chim. 2015, 18, 776–784. [Google Scholar] [CrossRef]
  47. Schrock, R.R. Catalytic Reduction of Dinitrogen to Ammonia by Molybdenum. In Catalysis without Precious Metals; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2010; pp. 25–50. [Google Scholar]
  48. Khoenkhoen, N.; de Bruin, B.; Reek, J.N.H.; Dzik, W.I. Reactivity of dinitrogen bound to mid- and late-transition-metal centers. Eur. J. Inorg. Chem. 2015, 2015, 567–598. [Google Scholar] [CrossRef]
  49. Evans, W.J.; Lee, D.S. Early developments in lanthanide-based dinitrogen reduction chemistry. Can. J. Chem. 2005, 83, 375–384. [Google Scholar] [CrossRef]
  50. Gardiner, M.G.; Stringer, D.N. Dinitrogen and related chemistry of the lanthanides: A review of the reductive capture of dinitrogen, as well as mono- and di-aza containing ligand chemistry of relevance to known and postulated metal mediated dinitrogen derivatives. Materials 2010, 3, 841–862. [Google Scholar] [CrossRef]
  51. Gardner, B.M.; Liddle, S.T. Small-molecule activation at uranium(III). Eur. J. Inorg. Chem. 2013, 2013, 3753–3770. [Google Scholar] [CrossRef]
  52. Liddle, S.T. The renaissance of non-aqueous uranium chemistry. Angew. Chem. Int. Ed. 2015, 54, 8604–8641. [Google Scholar] [CrossRef] [PubMed]
  53. Scherer, O.J. Complexes with substituent-free acyclic and cyclic phosphorus, arsenic, antimony, and bismuth ligands. Angew. Chem. Int. Ed. 1990, 29, 1104–1122. [Google Scholar] [CrossRef]
  54. Scherer, O.J. Pn and Asn ligands: A novel chapter in the chemistry of phosphorus and arsenic. Acc. Chem. Res. 1999, 32, 751–762. [Google Scholar] [CrossRef]
  55. Peruzzini, M.; Gonsalvi, L.; Romerosa, A. Coordination chemistry and functionalization of white phosphorus via transition metal complexes. Chem. Soc. Rev. 2005, 34, 1038–1047. [Google Scholar] [CrossRef] [PubMed]
  56. Vaira, M.D.; Sacconi, L. Transition metal complexes with cyclo-triphosphorus (η3-P3) and tetrahedro-tetraphosphorus (η1-P4) ligands. Angew. Chem. Int. Ed. 1982, 21, 330–342. [Google Scholar] [CrossRef]
  57. Figueroa, J.S.; Cummins, C.C. A niobaziridine hydride system for white phosphorus or dinitrogen activation and N- or P-atom transfer. Dalton Trans. 2006, 2161–2168. [Google Scholar] [CrossRef] [PubMed]
  58. Cossairt, B.M.; Piro, N.A.; Cummins, C.C. Early-transition-metal-mediated activation and transformation of white phosphorus. Chem. Rev. 2010, 110, 4164–4177. [Google Scholar] [CrossRef] [PubMed]
  59. Caporali, M.; Gonsalvi, L.; Rossin, A.; Peruzzini, M. P4 activation by late-transition metal complexes. Chem. Rev. 2010, 110, 4178–4235. [Google Scholar] [CrossRef] [PubMed]
  60. Giffin, N.A.; Masuda, J.D. Reactivity of white phosphorus with compounds of the p-block. Coord. Chem. Rev. 2011, 255, 1342–1359. [Google Scholar] [CrossRef]
  61. Balázs, G.; Seitz, A.; Scheer, M. Activation of White Phosphorus (P4) by Main Group Elements and Compounds. In Comprehensive Inorganic Chemistry II, 2nd ed.; Poeppelmeier, J.R., Ed.; Elsevier: Amsterdam, The Netherlands, 2013; pp. 1105–1132. [Google Scholar]
  62. Khan, S.; Sen, S.S.; Roesky, H.W. Activation of phosphorus by group 14 elements in low oxidation states. Chem. Commun. 2012, 48, 2169–2179. [Google Scholar] [CrossRef] [PubMed]
  63. Johnson, B.P.; Balázs, G.; Scheer, M. Low-coordinate E1 ligand complexes of group 15 elements—A developing area. Coord. Chem. Rev. 2006, 250, 1178–1195. [Google Scholar] [CrossRef]
  64. Scheer, M.; Balázs, G.; Seitz, A. P4 activation by main group elements and compounds. Chem. Rev. 2010, 110, 4236–4256. [Google Scholar] [CrossRef] [PubMed]
  65. Fang, M.; Bates, J.E.; Lorenz, S.E.; Lee, D.S.; Rego, D.B.; Ziller, J.W.; Furche, F.; Evans, W.J. (N2)3− radical chemistry via trivalent lanthanide salt/alkali metal reduction of dinitrogen: New syntheses and examples of (N2)2− and (N2)3− complexes and density functional theory comparisons of closed shell Sc3+, Y3+, and Lu3+ versus 4f9 Dy3+. Inorg. Chem. 2011, 50, 1459–1469. [Google Scholar] [CrossRef] [PubMed]
  66. Fryzuk, M.D.; Johnson, S.A. The continuing story of dinitrogen activation. Coord. Chem. Rev. 2000, 200–202, 379–409. [Google Scholar] [CrossRef]
  67. Jaroschik, F.; Momin, A.; Nief, F.; LeGoff, X.-F.; Deacon, G.B.; Junk, P.C. Dinitrogen reduction and C–H activation by the divalent organoneodymium complex [(C5H2tBu3)2Nd(μ-I)K([18]crown-6)]. Angew. Chem. Int. Ed. 2009, 48, 1117–1121. [Google Scholar] [CrossRef] [PubMed]
  68. Evans, W.J.; Rego, D.B.; Ziller, J.W. Synthesis, structure, and 15N-NMR studies of paramagnetic lanthanide complexes obtained by reduction of dinitrogen. Inorg. Chem. 2006, 45, 10790–10798. [Google Scholar] [CrossRef] [PubMed]
  69. Demir, S.; Lorenz, S.E.; Fang, M.; Furche, F.; Meyer, G.; Ziller, J.W.; Evans, W.J. Synthesis, structure, and density functional theory analysis of a scandium dinitrogen complex, [(C5Me4H)2Sc]2(μ-η22-N2). J. Am. Chem. Soc. 2010, 132, 11151–11158. [Google Scholar] [CrossRef] [PubMed]
  70. Demir, S.; Siladke, N.A.; Ziller, J.W.; Evans, W.J. Scandium and yttrium metallocene borohydride complexes: Comparisons of (BH4)1− vs. (BPh4)1− coordination and reactivity. Dalton Trans. 2012, 41, 9659–9666. [Google Scholar] [CrossRef] [PubMed]
  71. Schmiege, B.M.; Ziller, J.W.; Evans, W.J. Reduction of dinitrogen with an yttrium metallocene hydride precursor, [(C5Me5)2YH]2. Inorg. Chem. 2010, 49, 10506–10511. [Google Scholar] [CrossRef] [PubMed]
  72. Evans, W.J.; Ulibarri, T.A.; Ziller, J.W. Isolation and X-ray crystal structure of the first dinitrogen complex of an f-element metal, [(C5Me5)2Sm]2N2. J. Am. Chem. Soc. 1988, 110, 6877–6879. [Google Scholar] [CrossRef]
  73. Fieser, M.E.; Bates, J.E.; Ziller, J.W.; Furche, F.; Evans, W.J. Dinitrogen reduction via photochemical activation of heteroleptic tris(cyclopentadienyl) rare-earth complexes. J. Am. Chem. Soc. 2013, 135, 3804–3807. [Google Scholar] [CrossRef] [PubMed]
  74. Evans, W.J.; Allen, N.T.; Ziller, J.W. Expanding divalent organolanthanide chemistry: The first organothulium(II) complex and the in situ organodysprosium(II) reduction of dinitrogen. Angew. Chem. Int. Ed. 2002, 41, 359–361. [Google Scholar] [CrossRef]
  75. 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] [PubMed]
  76. Mueller, T.J.; Fieser, M.E.; Ziller, J.W.; Evans, W.J. (C5Me4H)1−-based reduction of dinitrogen by the mixed ligand tris(polyalkylcyclopentadienyl) lutetium and yttrium complexes, (C5Me5)3−x(C5Me4H)xLn. Chem. Sci. 2011, 2, 1992–1996. [Google Scholar] [CrossRef]
  77. Lorenz, S.E.; Schmiege, B.M.; Lee, D.S.; Ziller, J.W.; Evans, W.J. Synthesis and reactivity of bis(tetramethylcyclopentadienyl) yttrium metallocenes including the reduction of Me3SiN3 to [(Me3Si)2N] with [(C5Me4H)2Y(THF)]2(μ-η22-N2). Inorg. Chem. 2010, 49, 6655–6663. [Google Scholar] [CrossRef] [PubMed]
  78. MacDonald, M.R.; Ziller, J.W.; Evans, W.J. Synthesis of a crystalline molecular complex of Y2+, [(18-crown-6)K][(C5H4SiMe3)3Y]. J. Am. Chem. Soc. 2011, 133, 15914–15917. [Google Scholar] [CrossRef] [PubMed]
  79. Evans, W.J.; Lee, D.S.; Lie, C.; Ziller, J.W. Expanding the LnZ3/alkali-metal reduction system to organometallic and heteroleptic precursors: Formation of dinitrogen derivatives of lanthanum. Angew. Chem. Int. Ed. 2004, 43, 5517–5519. [Google Scholar] [CrossRef] [PubMed]
  80. Evans, W.J.; Lee, D.S.; Johnston, M.A.; Ziller, J.W. The elusive (C5Me4H)3Lu: Its synthesis and LnZ3/K/N2 reactivity. Organometallics 2005, 24, 6393–6397. [Google Scholar] [CrossRef]
  81. Evans, W.J.; Lee, D.S.; Rego, D.B.; Perotti, J.M.; Kozimor, S.A.; Moore, E.K.; Ziller, J.W. Expanding dinitrogen reduction chemistry to trivalent lanthanides via the LnZ3/Alkali metal reduction system: Evaluation of the generality of forming Ln2(μ-η22-N2) complexes via LnZ3/K. J. Am. Chem. Soc. 2004, 126, 14574–14582. [Google Scholar] [CrossRef] [PubMed]
  82. Evans, W.J.; Zucchi, G.; Ziller, J.W. Dinitrogen reduction by Tm(II), Dy(II), and Nd(II) with simple amide and aryloxide ligands. J. Am. Chem. Soc. 2003, 125, 10–11. [Google Scholar] [CrossRef] [PubMed]
  83. Evans, W.J.; Lee, D.S.; Ziller, J.W. Reduction of dinitrogen to planar bimetallic M2(μ-η22-N2) complexes of Y, Ho, Tm, and Lu using the K/Ln[N(SiMe3)2]3 reduction system. J. Am. Chem. Soc. 2004, 126, 454–455. [Google Scholar] [CrossRef] [PubMed]
  84. Corbey, J.F.; Farnaby, J.H.; Bates, J.E.; Ziller, J.W.; Furche, F.; Evans, W.J. Varying the Lewis base coordination of the Y2N2 core in the reduced dinitrogen complexes {[(Me3Si)2N]2(L)Y}2(μ-η22-N2) (L = benzonitrile, pyridines, triphenylphosphine oxide, and trimethylamine N-oxide). Inorg. Chem. 2012, 51, 7867–7874. [Google Scholar] [CrossRef] [PubMed]
  85. Campazzi, E.; Solari, E.; Floriani, C.; Scopelliti, R. The fixation and reduction of dinitrogen using lanthanides: Praseodymium and neodymium meso-octaethylporphyrinogen-dinitrogen complexes. Chem. Commun. 1998, 2603–2604. [Google Scholar] [CrossRef]
  86. Cheng, J.; Takats, J.; Ferguson, M.J.; McDonald, R. Heteroleptic Tm(II) complexes: One more success for Trofimenko’s scorpionates. J. Am. Chem. Soc. 2008, 130, 1544–1545. [Google Scholar] [CrossRef] [PubMed]
  87. Mansell, S.M.; Farnaby, J.H.; Germeroth, A.I.; Arnold, P.L. Thermally stable uranium dinitrogen complex with siloxide supporting ligands. Organometallics 2013, 32, 4214–4222. [Google Scholar] [CrossRef]
  88. Mansell, S.M.; Kaltsoyannis, N.; Arnold, P.L. Small molecule activation by uranium tris(aryloxides): Experimental and computational studies of binding of N2, coupling of CO, and deoxygenation insertion of CO2 under ambient conditions. J. Am. Chem. Soc. 2011, 133, 9036–9051. [Google Scholar] [CrossRef] [PubMed]
  89. Perrin, L.; Maron, L.; Eisenstein, O.; Schwartz, D.J.; Burns, C.J.; Andersen, R.A. Bonding of H2, N2, ethylene, and acetylene to bivalent lanthanide metallocenes: Trends from DFT calculations on Cp2M and Cp*2M (M = Sm, Eu, Yb) and experiments with Cp*2Yb. Organometallics 2003, 22, 5447–5453. [Google Scholar] [CrossRef]
  90. Hamaed, H.; Lo, A.Y.; Lee, D.S.; Evans, W.J.; Schurko, R.W. Solid-state 139La- and 15N-NMR spectroscopy of lanthanum-containing metallocenes. J. Am. Chem. Soc. 2006, 128, 12638–12639. [Google Scholar] [CrossRef] [PubMed]
  91. Evans, W.J.; Davis, B.L. Chemistry of tris(pentamethylcyclopentadienyl) f-element complexes, (C5Me5)3M. Chem. Rev. 2002, 102, 2119–2136. [Google Scholar] [CrossRef] [PubMed]
  92. Fieser, M.E.; Johnson, C.W.; Bates, J.E.; Ziller, J.W.; Furche, F.; Evans, W.J. Dinitrogen reduction, sulfur reduction, and isoprene polymerization via photochemical activation of trivalent bis(cyclopentadienyl) rare-earth-metal allyl complexes. Organometallics 2015, 34, 4387–4393. [Google Scholar] [CrossRef]
  93. Evans, W.J.; Fang, M.; Zucchi, G.; Furche, F.; Ziller, J.W.; Hoekstra, R.M.; Zink, J.I. Isolation of dysprosium and yttrium complexes of a three-electron reduction product in the activation of dinitrogen, the (N2)3− radical. J. Am. Chem. Soc. 2009, 131, 11195–11202. [Google Scholar] [CrossRef] [PubMed]
  94. Rinehart, J.D.; Fang, M.; Evans, W.J.; Long, J.R. Strong exchange and magnetic blocking in N23−-radical-bridged lanthanide complexes. Nat. Chem. 2011, 3, 538–542. [Google Scholar] [CrossRef] [PubMed]
  95. Rinehart, J.D.; Fang, M.; Evans, W.J.; Long, J.R. A N23− radical-bridged terbium complex exhibiting magnetic hysteresis at 14 K. J. Am. Chem. Soc. 2011, 133, 14236–14239. [Google Scholar] [CrossRef] [PubMed]
  96. Roy, L.E.; Hughbanks, T. Magnetic coupling in dinuclear Gd complexes. J. Am. Chem. Soc. 2006, 128, 568–575. [Google Scholar] [CrossRef] [PubMed]
  97. Meihaus, K.R.; Corbey, J.F.; Fang, M.; Ziller, J.W.; Long, J.R.; Evans, W.J. Influence of an inner-sphere K+ ion on the magnetic behavior of N23− radical-bridged dilanthanide complexes isolated using an external magnetic field. Inorg. Chem. 2014, 53, 3099–3107. [Google Scholar] [CrossRef] [PubMed]
  98. Allen, F.H.; Kennard, O.; Watson, D.G.; Brammer, L.; Orpen, A.G.; Taylor, R. Tables of bond lengths determined by X-ray and neutron diffraction. Part 1. Bond lengths in organic compounds. J. Chem. Soc. Perkin Trans. 1987, 2, S1–S19. [Google Scholar] [CrossRef]
  99. Zhang, Y.-Q.; Luo, C.-L.; Wang, B.-W.; Gao, S. Understanding the magnetic anisotropy in a family of N23− radical-bridged lanthanide complexes: Density functional theory and ab initio calculations. J. Phys. Chem. A 2013, 117, 10873–10880. [Google Scholar] [CrossRef] [PubMed]
  100. Rajeshkumar, T.; Rajaraman, G. Is a radical bridge a route to strong exchange interactions in lanthanide complexes? A computational examination. Chem. Commun. 2012, 48, 7856–7858. [Google Scholar] [CrossRef] [PubMed]
  101. Jubb, J.; Gambarotta, S. Dinitrogen reduction operated by a samarium macrocyclic complex. Encapsulation of dinitrogen into a Sm2Li4 metallic cage. J. Am. Chem. Soc. 1994, 116, 4477–4478. [Google Scholar] [CrossRef]
  102. Dubé, T.; Conoci, S.; Gambarotta, S.; Yap, G.P.A.; Vasapollo, G. Tetrametallic reduction of dinitrogen: Formation of a tetranuclear samarium dinitrogen complex. Angew. Chem. Int. Ed. 1999, 38, 3657–3659. [Google Scholar] [CrossRef]
  103. Dubé, T.; Ganesan, M.; Conoci, S.; Gambarotta, S.; Yap, G.P.A. Tetrametallic divalent samarium cluster hydride and dinitrogen complexes. Organometallics 2000, 19, 3716–3721. [Google Scholar] [CrossRef]
  104. Bérubé, C.D.; Yazdanbakhsh, M.; Gambarotta, S.; Yap, G.P.A. Serendipitous isolation of the first example of a mixed-valence samarium tripyrrole complex. Organometallics 2003, 22, 3742–3747. [Google Scholar] [CrossRef]
  105. Guan, J.; Dubé, T.; Gambarotta, S.; Yap, G.P.A. Dinitrogen labile coordination versus four-electron reduction, THF cleavage, and fragmentation promoted by a (calix-tetrapyrrole)Sm(II) complex. Organometallics 2000, 19, 4820–4827. [Google Scholar] [CrossRef]
  106. Ganesan, M.; Gambarotta, S.; Yap, G.P.A. Highly reactive SmII macrocyclic clusters: Precursors to N2 reduction. Angew. Chem. Int. Ed. 2001, 40, 766–769. [Google Scholar] [CrossRef]
  107. Fox, A.R.; Bart, S.C.; Meyer, K.; Cummins, C.C. Towards uranium catalysts. Nature 2008, 455, 341–349. [Google Scholar] [CrossRef] [PubMed]
  108. Huang, Q.-R.; Kingham, J.R.; Kaltsoyannis, N. The strength of actinide-element bonds from the quantum theory of atoms-in-molecules. Dalton Trans. 2015, 44, 2554–2566. [Google Scholar] [CrossRef] [PubMed]
  109. Roussel, P.; Scott, P. Complex of dinitrogen with trivalent uranium. J. Am. Chem. Soc. 1998, 120, 1070–1071. [Google Scholar] [CrossRef]
  110. Odom, A.L.; Arnold, P.L.; Cummins, C.C. Heterodinuclear uranium/molybdenum dinitrogen complexes. J. Am. Chem. Soc. 1998, 120, 5836–5837. [Google Scholar] [CrossRef]
  111. Cloke, F.G.N.; Hitchcock, P.B. Reversible binding and reduction of dinitrogen by a uranium(III) pentalene complex. J. Am. Chem. Soc. 2002, 124, 9352–9353. [Google Scholar] [CrossRef] [PubMed]
  112. Evans, W.J.; Kozimor, S.A.; Ziller, J.W. A monometallic f element complex of dinitrogen: (C5Me5)3U(η1-N2). J. Am. Chem. Soc. 2003, 125, 14264–14265. [Google Scholar] [CrossRef] [PubMed]
  113. Laplaza, C.E.; Cummins, C.C. Dinitrogen cleavage by a three-coordinate molybdenum(III) complex. Science 1995, 268, 861–863. [Google Scholar] [CrossRef] [PubMed]
  114. Mindiola, D.J.; Meyer, K.; Cherry, J.-P.F.; Baker, T.A.; Cummins, C.C. Dinitrogen cleavage stemming from a heterodinuclear niobium/molybdenum N2 complex: New nitridoniobium systems including a niobazene cyclic trimer. Organometallics 2000, 19, 1622–1624. [Google Scholar] [CrossRef]
  115. Curley, J.J.; Cook, T.R.; Reece, S.Y.; Müller, P.; Cummins, C.C. Shining light on dinitrogen cleavage: Structural features, redox chemistry, and photochemistry of the key intermediate bridging dinitrogen complex. J. Am. Chem. Soc. 2008, 130, 9394–9405. [Google Scholar] [CrossRef] [PubMed]
  116. Kaltsoyannis, N.; Scott, P. Evidence for actinide metal to ligand π backbonding. Density functional investigations of the electronic structure of [{(NH2)3(NH3)U}2222-N2)]. Chem. Commun. 1998, 1665–1666. [Google Scholar] [CrossRef]
  117. Roussel, P.; Errington, W.; Kaltsoyannis, N.; Scott, P. Back bonding without σ-bonding: A unique π-complex of dinitrogen with uranium. J. Organomet. Chem. 2001, 635, 69–74. [Google Scholar] [CrossRef]
  118. Cloke, F.G.N.; Green, J.C.; Kaltsoyannis, N. Electronic structure of [U22-N2)(η5-C5Me5)28-C8H4(SiPri3)2)2]. Organometallics 2004, 23, 832–835. [Google Scholar] [CrossRef]
  119. Korobkov, I.; Gambarotta, S.; Yap, G.P.A. A highly reactive uranium complex supported by the calix[4]tetrapyrrole tetraanion affording dinitrogen cleavage, solvent deoxygenation, and polysilanol depolymerization. Angew. Chem. Int. Ed. 2002, 41, 3433–3436. [Google Scholar] [CrossRef]
  120. Korobkov, I.; Gambarotta, S.; Yap, G.P.A. Amide from dinitrogen by in situ cleavage and partial hydrogenation promoted by a transient zero-valent thorium synthon. Angew. Chem. Int. Ed. 2003, 42, 4958–4961. [Google Scholar] [CrossRef] [PubMed]
  121. Corbridge, D.E.C.; Lowe, E.J. Structure of white phosphorus: Single crystal X-ray examination. Nature 1952, 170, 629–629. [Google Scholar] [CrossRef]
  122. Kühl, O. Phosphorus-31 NMR Spectroscopy: A Concise Introduction for the Synthetic Organic and Organometallic Chemist; Springer-Verlag GmbH: Berlin, Germany; Heidelberg, Germany, 2008. [Google Scholar]
  123. Konchenko, S.N.; Pushkarevsky, N.A.; Gamer, M.T.; Köppe, R.; Schnöckel, H.; Roesky, P.W. [{(η5-C5Me5)2Sm}4P8]: A molecular polyphosphide of the rare-earth elements. J. Am. Chem. Soc. 2009, 131, 5740–5741. [Google Scholar] [CrossRef] [PubMed]
  124. Huang, W.; Diaconescu, P.L. P4 activation by group 3 metal arene complexes. Chem. Commun. 2012, 48, 2216–2218. [Google Scholar] [CrossRef] [PubMed]
  125. Huang, W.; Diaconescu, P.L. P4 Activation by lanthanum and lutetium naphthalene complexes supported by a ferrocene diamide ligand. Eur. J. Inorg. Chem. 2013, 2013, 4090–4096. [Google Scholar] [CrossRef]
  126. Turbervill, R.S.P.; Goicoechea, J.M. From clusters to unorthodox pnictogen sources: Solution-phase reactivity of [E7]3− (E = P–Sb) anions. Chem. Rev. 2014, 114, 10807–10828. [Google Scholar] [CrossRef] [PubMed]
  127. Shannon, R. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. Sect. A Found. Crystallogr. 1976, 32, 751–767. [Google Scholar] [CrossRef]
  128. Scherer, O.J.; Werner, B.; Heckmann, G.; Wolmershäuser, G. Bicyclic P6 as complex ligand. Angew. Chem. Int. Ed. 1991, 30, 553–555. [Google Scholar] [CrossRef]
  129. Stephens, F.H. Activation of White Phosphorus by Molybdenum and Uranium tris-Amides. Ph.D. Thesis, Massachusetts Institute of Technology, Cambridge, MA, USA, May 2004. [Google Scholar]
  130. Frey, A.S.P.; Cloke, F.G.N.; Hitchcock, P.B.; Green, J.C. Activation of P4 by U(η5-C5Me5)(η8-C8H6(SiiPr3)2-1,4)(THF); the X-ray structure of [U(η5-C5Me5)(η8-C8H6(SiiPr3)2-1,4)]2(μ-η22-P4). New J. Chem. 2011, 35, 2022–2026. [Google Scholar] [CrossRef]
  131. Gardner, B.M.; Tuna, F.; McInnes, E.J.L.; McMaster, J.; Lewis, W.; Blake, A.J.; Liddle, S.T. An inverted-sandwich diuranium μ-η55-cyclo-P5 complex supported by U–P5 δ-bonding. Angew. Chem. Int. Ed. 2015, 54, 7068–7072. [Google Scholar] [CrossRef] [PubMed]
  132. Forfar, L.C.; Clark, T.J.; Green, M.; Mansell, S.M.; Russell, C.A.; Sanguramath, R.A.; Slattery, J.M. White phosphorus as a ligand for the coinage metals. Chem. Commun. 2012, 48, 1970–1972. [Google Scholar] [CrossRef] [PubMed]
  133. Spitzer, F.; Sierka, M.; Latronico, M.; Mastrorilli, P.; Virovets, A.V.; Scheer, M. Fixation and release of intact E4 tetrahedra (E = P, As). Angew. Chem. Int. Ed. 2015, 54, 4392–4396. [Google Scholar] [CrossRef] [PubMed]
  134. Patel, D.; Tuna, F.; McInnes, E.J.L.; Lewis, W.; Blake, A.J.; Liddle, S.T. An actinide zintl cluster: A tris(triamidouranium)μ3222-heptaphosphanortricyclane and its diverse synthetic utility. Angew. Chem. Int. Ed. 2013, 52, 13334–13337. [Google Scholar] [CrossRef] [PubMed]
  135. Laplaza, C.E.; Davis, W.M.; Cummins, C.C. A molybdenum–phosphorus triple bond: Synthesis, structure, and reactivity of the terminal phosphido (P3−) complex [Mo(P)(NRAr)3]. Angew. Chem. Int. Ed. 1995, 34, 2042–2044. [Google Scholar] [CrossRef]
  136. Cherry, J.-P.F.; Stephens, F.H.; Johnson, M.J.A.; Diaconescu, P.L.; Cummins, C.C. Terminal phosphide and dinitrogen molybdenum compounds obtained from pnictide-bridged precursors. Inorg. Chem. 2001, 40, 6860–6862. [Google Scholar] [CrossRef] [PubMed]
  137. Chisholm, M.H.; Folting, K.; Pasterczyk, J.W. A phosphido-capped tritungsten alkoxide cluster: W33-P)(μ-OCH2-t-Bu)3(OCH2-t-Bu)6 and speculation upon the existence of a reactive (t-BuCH2O)3W≡P intermediate. Inorg. Chem. 1988, 27, 3057–3058. [Google Scholar] [CrossRef]
  138. Scherer, O.J.; Sitzmann, H.; Wolmershäuser, G. Umsetzung von P4 mit (η5-C5H5)(CO)2Mo≡Mo(CO)25-C5H5) zu den tetraedrischen molybdänkomplexen Pn[Mo(CO)25-C5H5)]4−n (n = 2,3). J. Organomet. Chem. 1984, 268, C9–C12. [Google Scholar] [CrossRef]
  139. Di Vaira, M.; Ghilardi, C.A.; Midollini, S.; Sacconi, L. cyclo-Triphosphorus (δ-P3) as a ligand in cobalt and nickel complexes with 1,1,1-tris(diphenylphosphinomethyl)ethane. Formation and structures. J. Am. Chem. Soc. 1978, 100, 2550–2551. [Google Scholar] [CrossRef]
  140. Chirik, P.J.; Pool, J.A.; Lobkovsky, E. Functionalization of elemental phosphorus with [Zr(η5-C5Me5)(η5-C5H4tBu)H2]2. Angew. Chem. Int. Ed. 2002, 41, 3463–3465. [Google Scholar] [CrossRef]
  141. Gröer, T.; Baum, G.; Scheer, M. Complexes with a monohapto bound phosphorus tetrahedron and phosphaalkyne. Organometallics 1998, 17, 5916–5919. [Google Scholar] [CrossRef]
  142. Peruzzini, M.; Marvelli, L.; Romerosa, A.; Rossi, R.; Vizza, F.; Zanobini, F. Synthesis and characterisation of tetrahedro-tetraphosphorus complexes of rhenium—Evidence for the first bridging complex of white phosphorus. Eur. J. Inorg. Chem. 1999, 1999, 931–933. [Google Scholar] [CrossRef]
  143. Urnėžius, E.; Brennessel, W.W.; Cramer, C.J.; Ellis, J.E.; Schleyer, P.V.R. A carbon-free sandwich complex [(P5)2Ti]2−. Science 2002, 295, 832–834. [Google Scholar]
  144. Scherer, O.J.; Sitzmann, H.; Wolmershäuser, G. Hexaphosphabenzene as complex ligand. Angew. Chem. Int. Ed. 1985, 24, 351–353. [Google Scholar] [CrossRef]
  145. Scherer, O.J.; Berg, G.; Wolmershäuser, G. P8 and P12 as complex ligands. Chem. Ber. 1996, 129, 53–58. [Google Scholar] [CrossRef]
  146. Scheer, M.; Deng, S.; Scherer, O.J.; Sierka, M. Tetraphosphacyclopentadienyl and triphosphaallyl ligands in iron complexes. Angew. Chem. Int. Ed. 2005, 44, 3755–3758. [Google Scholar] [CrossRef] [PubMed]
  147. Scherer, O.J.; Schulze, J.; Wolmershäuser, G. Bicyclisches As6 als komplexligand. J. Organomet. Chem. 1994, 484, C5–C7. [Google Scholar] [CrossRef]
  148. Di Vaira, M.; Midollini, S.; Sacconi, L.; Zanobini, F. cyclo-Triarsenic as μ,η-ligand in transition-metal complexes. Angew. Chem. Int. Ed. 1978, 17, 676–677. [Google Scholar] [CrossRef]
  149. Scherer, O.J.; Kemény, G.; Wolmershäuser, G. [Cp4Fe4(E2)2] clusters with triangulated dodecahedral Fe4E4 skeletons (E = P, As). Chem. Ber. 1995, 128, 1145–1148. [Google Scholar] [CrossRef]
  150. Schwarzmaier, C.; Timoshkin, A.Y.; Scheer, M. An end-on-coordinated As4 tetrahedron. Angew. Chem. Int. Ed. 2013, 52, 7600–7603. [Google Scholar] [CrossRef] [PubMed]
  151. Schwarzmaier, C.; Sierka, M.; Scheer, M. Intact As4 tetrahedra coordinated side-on to metal cations. Angew. Chem. Int. Ed. 2013, 52, 858–861. [Google Scholar] [CrossRef] [PubMed]
  152. Scherer, O.J.; Sitzmann, H.; Wolmershäuser, G. (E2)2-einheiten (E = P, As) als clusterbausteine. J. Organomet. Chem. 1986, 309, 77–86. [Google Scholar] [CrossRef]
  153. Spinney, H.A.; Piro, N.A.; Cummins, C.C. Triple-bond reactivity of an AsP complex intermediate: Synthesis stemming from molecular arsenic, As4. J. Am. Chem. Soc. 2009, 131, 16233–16243. [Google Scholar] [CrossRef] [PubMed]
  154. Heinl, S.; Scheer, M. Activation of group 15 based cage compounds by [CpBIGFe(CO)2] radicals. Chem. Sci. 2014, 5, 3221–3225. [Google Scholar] [CrossRef]
  155. Gra; Bodensteiner, M.; Zabel, M.; Scheer, M. Synthesis of arsenic-rich Asn ligand complexes from yellow arsenic. Chem. Sci. 2015, 6, 1379–1382. [Google Scholar]
  156. Schwarzmaier, C.; Bodensteiner, M.; Timoshkin, A.Y.; Scheer, M. An approach to mixed PnAsm ligand complexes. Angew. Chem. Int. Ed. 2014, 53, 290–293. [Google Scholar] [CrossRef] [PubMed]
  157. Curley, J.J.; Piro, N.A.; Cummins, C.C. A terminal molybdenum arsenide complex synthesized from yellow arsenic. Inorg. Chem. 2009, 48, 9599–9601. [Google Scholar] [CrossRef] [PubMed]
  158. Balázs, G.; Sierka, M.; Scheer, M. Antimony–tungsten triple bond: A stable complex with a terminal antimony ligand. Angew. Chem. Int. Ed. 2005, 44, 4920–4924. [Google Scholar] [CrossRef] [PubMed]
  159. Evans, W.J.; Fang, M.; Bates, J.E.; Furche, F.; Ziller, J.W.; Kiesz, M.D.; Zink, J.I. Isolation of a radical dianion of nitrogen oxide (NO)2−. Nat. Chem. 2010, 2, 644–647. [Google Scholar] [CrossRef] [PubMed]
  160. Corbey, J.F.; Fang, M.; Ziller, J.W.; Evans, W.J. Cocrystallization of (μ-S2)2− and (μ-S)2− and formation of an [η2-S3N(SiMe3)2] ligand from chalcogen reduction by (N2)2− in a bimetallic yttrium amide complex. Inorg. Chem. 2015, 54, 801–807. [Google Scholar] [CrossRef] [PubMed]
  161. Evans, W.J.; Lee, D.S.; Ziller, J.W.; Kaltsoyannis, N. Trivalent [(C5Me5)2(THF)Ln]2(μ-η22-N2) complexes as reducing agents including the reductive homologation of CO to a ketene carboxylate, (μ-η4-O2CCCO)2. J. Am. Chem. Soc. 2006, 128, 14176–14184. [Google Scholar] [CrossRef] [PubMed]
  162. Lu, E.; Li, Y.; Chen, Y. A scandium terminal imido complex: Synthesis, structure and DFT studies. Chem. Commun. 2010, 46, 4469–4471. [Google Scholar] [CrossRef] [PubMed]
  163. Rong, W.; Cheng, J.; Mou, Z.; Xie, H.; Cui, D. Facile preparation of a scandium terminal imido complex supported by a phosphazene ligand. Organometallics 2013, 32, 5523–5529. [Google Scholar] [CrossRef]
  164. Schädle, D.; Meermann-Zimmermann, M.; Schädle, C.; Maichle-Mössmer, C.; Anwander, R. Rare-earth metal complexes with terminal imido ligands. Eur. J. Inorg. Chem. 2015, 2015, 1334–1339. [Google Scholar] [CrossRef]
  165. Sobota, P.; Janas, Z. Formation of a nitrogen–carbon bond from N2 and CO. Influence of MgCl2 on the N2 reduction process in the system TiCl4/Mg. J. Organomet. Chem. 1984, 276, 171–176. [Google Scholar] [CrossRef]
  166. Semproni, S.P.; Margulieux, G.W.; Chirik, P.J. Di- and tetrametallic hafnocene oxamidides prepared from CO-induced N2 bond cleavage and thermal rearrangement to hafnocene cyanide derivatives. Organometallics 2012, 31, 6278–6287. [Google Scholar] [CrossRef]
  167. Knobloch, D.J.; Lobkovsky, E.; Chirik, P.J. Carbon monoxide-induced dinitrogen cleavage with group 4 metallocenes: Reaction scope and coupling to N–H bond formation and CO deoxygenation. J. Am. Chem. Soc. 2010, 132, 10553–10564. [Google Scholar] [CrossRef] [PubMed]
  168. Semproni, S.P.; Milsmann, C.; Chirik, P.J. Structure and reactivity of a hafnocene μ-nitrido prepared from dinitrogen cleavage. Angew. Chem. Int. Ed. 2012, 51, 5213–5216. [Google Scholar] [CrossRef] [PubMed]
  169. Semproni, S.P.; Chirik, P.J. Synthesis of a base-free hafnium nitride from N2 cleavage: A versatile platform for dinitrogen functionalization. J. Am. Chem. Soc. 2013, 135, 11373–11383. [Google Scholar] [CrossRef] [PubMed]
  170. MacKay, B.A.; Johnson, S.A.; Patrick, B.O.; Fryzuk, M.D. Functionalization and cleavage of coordinated dinitrogen via hydroboration using primary and secondary boranes. Can. J. Chem. 2005, 83, 315–323. [Google Scholar] [CrossRef]
  171. Fryzuk, M.D.; MacKay, B.A.; Johnson, S.A.; Patrick, B.O. Hydroboration of coordinated dinitrogen: A new reaction for the N2 ligand that results in its functionalization and cleavage. Angew. Chem. Int. Ed. 2002, 41, 3709–3712. [Google Scholar] [CrossRef]
  172. Fryzuk, M.D.; MacKay, B.A.; Patrick, B.O. Hydrosilylation of a dinuclear tantalum dinitrogen complex: Cleavage of N2 and functionalization of both nitrogen atoms. J. Am. Chem. Soc. 2003, 125, 3234–3235. [Google Scholar] [CrossRef] [PubMed]
  173. Spencer, L.P.; MacKay, B.A.; Patrick, B.O.; Fryzuk, M.D. Inner-sphere two-electron reduction leads to cleavage and functionalization of coordinated dinitrogen. Proc. Natl. Acad. Sci. USA 2006, 103, 17094–17098. [Google Scholar] [CrossRef] [PubMed]
  174. Margulieux, G.W.; Turner, Z.R.; Chirik, P.J. Synthesis and ligand modification chemistry of a molybdenum dinitrogen complex: Redox and chemical activity of a bis(imino)pyridine ligand. Angew. Chem. Int. Ed. 2014, 53, 14211–14215. [Google Scholar] [CrossRef] [PubMed]
  175. Milsmann, C.; Turner, Z.R.; Semproni, S.P.; Chirik, P.J. Azo N=N bond cleavage with a redox-active vanadium compound involving metal–ligand cooperativity. Angew. Chem. Int. Ed. 2012, 51, 5386–5390. [Google Scholar] [CrossRef] [PubMed]
  176. King, D.M.; Tuna, F.; McInnes, E.J.L.; McMaster, J.; Lewis, W.; Blake, A.J.; Liddle, S.T. Synthesis and structure of a terminal uranium nitride complex. Science 2012, 337, 717–720. [Google Scholar] [CrossRef] [PubMed]
  177. King, D.M.; Tuna, F.; McInnes, E.J.L.; McMaster, J.; Lewis, W.; Blake, A.J.; Liddle, S.T. Isolation and characterization of a uranium(VI)–nitride triple bond. Nat. Chem. 2013, 5, 482–488. [Google Scholar] [CrossRef] [PubMed]
  178. Cleaves, P.A.; King, D.M.; Kefalidis, C.E.; Maron, L.; Tuna, F.; McInnes, E.J.L.; McMaster, J.; Lewis, W.; Blake, A.J.; Liddle, S.T. Two-electron reductive carbonylation of terminal uranium(V) and uranium(VI) nitrides to cyanate by carbon monoxide. Angew. Chem. Int. Ed. 2014, 53, 10412–10415. [Google Scholar] [CrossRef] [PubMed]
  179. Gardner, B.M.; Balázs, G.; Scheer, M.; Tuna, F.; McInnes, E.J.L.; McMaster, J.; Lewis, W.; Blake, A.J.; Liddle, S.T. Triamidoamine uranium(IV)–arsenic complexes containing one-, two- and threefold U–As bonding interactions. Nat. Chem. 2015, 7, 582–590. [Google Scholar] [CrossRef] [PubMed]

Share and Cite

MDPI and ACS Style

Turner, Z.R. Molecular Pnictogen Activation by Rare Earth and Actinide Complexes. Inorganics 2015, 3, 597-635. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics3040597

AMA Style

Turner ZR. Molecular Pnictogen Activation by Rare Earth and Actinide Complexes. Inorganics. 2015; 3(4):597-635. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics3040597

Chicago/Turabian Style

Turner, Zoë R. 2015. "Molecular Pnictogen Activation by Rare Earth and Actinide Complexes" Inorganics 3, no. 4: 597-635. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics3040597

Article Metrics

Back to TopTop