Co-reporter:Polly L. Arnold, Michał S. Dutkiewicz, and Olaf Walter
Chemical Reviews September 13, 2017 Volume 117(Issue 17) pp:11460-11460
Publication Date(Web):August 30, 2017
DOI:10.1021/acs.chemrev.7b00192
Fifty years have passed since the foundation of organometallic neptunium chemistry, and yet only a handful of complexes have been reported, and even fewer have been fully characterized. Yet, increasingly, combined synthetic/spectroscopic/computational studies are demonstrating how covalently bonding, soft, carbocyclic organometallic ligands provide an excellent platform for advancing the fundamental understanding of the differences in orbital contributions and covalency in f-block metal–ligand bonding. Understanding the subtleties is the key to the safe handling and separations of the highly radioactive nuclei. This review describes the complexes that have been synthesized to date and presents a critical assessment of the successes and difficulties in their analysis and the bonding information they have provided. Because of increasing recent efforts to start new Np-capable air-sensitive inorganic chemistry laboratories, the importance of radioactivity, the basics of Np decay and its ramifications (including the radiochemical synthesis of one organometallic compound), and the available anhydrous starting materials are also surveyed. The review also highlights a range of instances in which important differences in the chemical behavior between Np and its closest neighbors, uranium and plutonium, are found.
Co-reporter:Polly L. Arnold;Charlotte J. Stevens;Nicola L. Bell;Rianne M. Lord;Jonathan M. Goldberg;Gary S. Nichol;Jason B. Love
Chemical Science (2010-Present) 2017 vol. 8(Issue 5) pp:3609-3617
Publication Date(Web):2017/05/03
DOI:10.1039/C7SC00382J
The first use of a dinuclear UIII/UIII complex in the activation of small molecules is reported. The octadentate Schiff-base pyrrole, anthracene-hinged ‘Pacman’ ligand LA combines two strongly reducing UIII centres and three borohydride ligands in [M(THF)4][{U(BH4)}2(μ-BH4)(LA)(THF)2] 1-M, (M = Li, Na, K). The two borohydride ligands bound to uranium outside the macrocyclic cleft are readily substituted by aryloxide ligands, resulting in a single, weakly-bound, encapsulated endo group 1 metal borohydride bridging the two UIII centres in [{U(OAr)}2(μ-MBH4)(LA)(THF)2] 2-M (OAr = OC6H2tBu3-2,4,6, M = Na, K). X-ray crystallographic analysis shows that, for 2-K, in addition to the endo-BH4 ligand the potassium counter-cation is also incorporated into the cleft through η5-interactions with the pyrrolides instead of extraneous donor solvent. As such, 2-K has a significantly higher solubility in non-polar solvents and a wider U–U separation compared to the ‘ate’ complex 1. The cooperative reducing capability of the two UIII centres now enforced by the large and relatively flexible macrocycle is compared for the two complexes, recognising that the borohydrides can provide additional reducing capability, and that the aryloxide-capped 2-K is constrained to reactions within the cleft. The reaction between 1-Na and S8 affords an insoluble, presumably polymeric paramagnetic complex with bridging uranium sulfides, while that with CS2 results in oxidation of each UIII to the notably high UV oxidation state, forming the unusual trithiocarbonate (CS3)2− as a ligand in [{U(CS3)}2(μ-κ2:κ2-CS3)(LA)] (4). The reaction between 2-K and S8 results in quantitative substitution of the endo-KBH4 by a bridging persulfido (S2)2− group and oxidation of each UIII to UIV, yielding [{U(OAr)}2(μ-κ2:κ2-S2)(LA)] (5). The reaction of 2-K with CS2 affords a thermally unstable adduct which is tentatively assigned as containing a carbon disulfido (CS2)2− ligand bridging the two U centres (6a), but only the mono-bridged sulfido (S)2− complex [{U(OAr)}2(μ-S)(LA)] (6) is isolated. The persulfido complex (5) can also be synthesised from the mono-bridged sulfido complex (6) by the addition of another equivalent of sulfur.
Co-reporter:James R. Pankhurst;Nicola L. Bell;Markus Zegke;Lucy N. Platts;Carlos Alvarez Lamfsus;Laurent Maron;Louise S. Natrajan;Stephen Sproules;Jason B. Love
Chemical Science (2010-Present) 2017 vol. 8(Issue 1) pp:108-116
Publication Date(Web):2016/12/19
DOI:10.1039/C6SC02912D
The uranyl(VI) complex UO2Cl(L) of the redox-active, acyclic diimino-dipyrrin anion, L− is reported and its reaction with inner- and outer-sphere reductants studied. Voltammetric, EPR-spectroscopic and X-ray crystallographic studies show that chemical reduction by the outer-sphere reagent CoCp2 initially reduces the ligand to a dipyrrin radical, and imply that a second equivalent of CoCp2 reduces the U(VI) centre to form U(V). Cyclic voltammetry indicates that further outer-sphere reduction to form the putative U(IV) trianion only occurs at strongly cathodic potentials. The initial reduction of the dipyrrin ligand is supported by emission spectra, X-ray crystallography, and DFT; the latter also shows that these outer-sphere reactions are exergonic and proceed through sequential, one-electron steps. Reduction by the inner-sphere reductant [TiCp2Cl]2 is also likely to result in ligand reduction in the first instance but, in contrast to the outer-sphere case, reduction of the uranium centre becomes much more favoured, allowing the formation of a crystallographically characterised, doubly-titanated U(IV) complex. In the case of inner-sphere reduction only, ligand-to-metal electron-transfer is thermodynamically driven by coordination of Lewis-acidic Ti(IV) to the uranyl oxo, and is energetically preferable over the disproportionation of U(V). Overall, the involvement of the redox-active dipyrrin ligand in the reduction chemistry of UO2Cl(L) is inherent to both inner- and outer-sphere reduction mechanisms, providing a new route to accessing a variety of U(VI), U(V), and U(IV) complexes.
Co-reporter:Fern Sinclair;Johann A. Hlina;Jordann A. L. Wells;Michael P. Shaver
Dalton Transactions 2017 vol. 46(Issue 33) pp:10786-10790
Publication Date(Web):2017/08/22
DOI:10.1039/C7DT02167D
The C3-symmetric uranium(IV) and cerium(IV) complexes Me3SiOM(OArP)3, M = U (1), Ce (2), OArP = OC6H2-6-tBu-4-Me-2-PPh2, have been prepared and the difference between these 4f and 5f congeners as initiators for the ring opening polymerisation (ROP) of L-lactide is compared. The poorly controlled reactivity of the homoleptic analogue U(OArP)4 (3) demonstrates the importance of the M-OSiMe3 initiating group. The incorporation of a nickel atom in 1 to form the U–Ni heterobimetallic complex Me3SiOU(OArP)3Ni (4) may be the first example of the use of the inverse trans influence to switch the reactivity of a complex. This would imply the formation of the U–Ni bond strengthens the U–OSiMe3 bond to such an extent that the ROP catalysis is switched off. Changing the conditions to immortal polymerisation dramatically increases polymerisation rates, and switches the order, with the Ce complex now faster than the U analogue, suggesting ligand protonolysis to afford a more open coordination sphere. For the ROP of rac-lactide, uranium complex 1 promotes heterotacticity at the highest levels of stereocontrol yet reported for an actinide complex.
Co-reporter: Dr. Polly L. Arnold;Dr. Bradley E. Cowie;Markéta Suvova;Dr. Markus Zegke;Dr. Nicola Magnani;Dr. Eric Colineau;Dr. Jean-Christophe Griveau; Dr. Roberto Caciuffo; Dr. Jason B. Love
Angewandte Chemie International Edition 2017 Volume 56(Issue 36) pp:10775-10779
Publication Date(Web):2017/08/28
DOI:10.1002/anie.201705197
AbstractThe reduction of UVI uranyl halides or amides with simple LnII or UIII salts forms highly symmetric, linear, oxo-bridged trinuclear UV/LnIII/UV, LnIII/UIV/LnIII, and UIV/UIV/UIV complexes or linear LnIII/UV polymers depending on the stoichiometry and solvent. The reactions can be tuned to give the products of one- or two-electron uranyl reduction. The reactivity and magnetism of these compounds are discussed in the context of using a series of strongly oxo-coupled homo- and heterometallic poly(f-block) chains to better understand fundamental electronic structure in the f-block.
Co-reporter:Ryan W.F. Kerr, Mark D. Greenhalgh, Alexandra M.Z. Slawin, Polly L. Arnold, Andrew D. Smith
Tetrahedron: Asymmetry 2017 Volume 28, Issue 1(Issue 1) pp:
Publication Date(Web):15 January 2017
DOI:10.1016/j.tetasy.2016.10.012
An enantioselective N-heterocyclic carbene catalysed formal [3+2] cycloaddition has been developed for the synthesis of oxazolindin-4-one products. The reaction of oxaziridines and α-aroyloxyaldehydes under N-heterocyclic carbene catalysis provides the formal cycloaddition products with excellent control of the diastereo- and enantioselectivity (12 examples, up to >95:5 dr, >99:1 er). A matched-mismatched effect between the enantiomer of the catalyst and oxaziridine was identified, and preliminary mechanistic studies have allowed the proposal of a model to explain these observations.Download high-res image (106KB)Download full-size image
Co-reporter:J. A. Hlina;J. A. L. Wells;J. R. Pankhurst;Jason B. Love;P. L. Arnold
Dalton Transactions 2017 vol. 46(Issue 17) pp:5540-5545
Publication Date(Web):2017/05/02
DOI:10.1039/C6DT04570G
The heterotetra- and bimetallic uranium(IV)–rhodium(I) complexes [UIVI2(μ-OArP-1κ1O,2κ1P)2RhI(μ-I)]2 (2) (ArPO− = 2-(diphenylphosphino)-6-tert-butyl-4-methylphenoxide) and UIVI(μ-I)(μ-OArP-1κ1O,2κ1P)3RhI (3) were prepared by treatment of UIVI(OArP-κ2O,P)3 (1) with rhodium(I) iodide olefin complexes. The reaction of 1 with the monodentate cyclooctene (coe) rhodium(I) precursor [(coe)2RhII]2 gives only the bimetallic complex [UIVRhI] 3, and with the diene [(cod)RhII]2 (5) (cod = 1,5-cyclooctadiene), mixtures of [UIVRhI]2 complex 2 and [UIVRhI] 3 along with (cod)RhIOArP-κ2O,P (4), a RhI side-product from the formation of 2. The complexes were characterised by single crystal X-ray diffraction, NMR and UV-vis-NIR spectroscopy, and electrochemistry. The UIV–RhI intermetallic distances in 2 (2.7601(5) Å) and 3 (2.7630(5) Å) are among the shortest between f-elements and transition metals reported to date. Despite almost identical U–Rh bond lengths in the solid state, in solution only weak, and very different interactions between the metal centres are found.
Co-reporter:Michał S. Dutkiewicz;Christos Apostolidis;Olaf Walter
Chemical Science (2010-Present) 2017 vol. 8(Issue 4) pp:2553-2561
Publication Date(Web):2017/03/28
DOI:10.1039/C7SC00034K
Neptunium complexes in the formal oxidation states II, III, and IV supported by cyclopentadienyl ligands are explored, and significant differences between Np and U highlighted as a result. A series of neptunium(III) cyclopentadienyl (Cp) complexes [Np(Cp)3], its bis-acetonitrile adduct [Np(Cp)3(NCMe)2], and its KCp adduct K[Np(Cp)4] and [Np(Cp′)3] (Cp′ = C5H4SiMe3) have been made and characterised providing the first single crystal X-ray analyses of NpIII Cp complexes. In all NpCp3 derivatives there are three Cp rings in η5-coordination around the NpIII centre; additionally in [Np(Cp)3] and K[Np(Cp)4] one Cp ring establishes a μ-η1-interaction to one C atom of a neighbouring Np(Cp)3 unit. The solid state structure of K[Np(Cp)4] is unique in containing two different types of metal–Cp coordination geometries in the same crystal. NpIII(Cp)4 units are found exhibiting four units of η5-coordinated Cp rings like in the known complex [NpIV(Cp)4], the structure of which is now reported. A detailed comparison of the structures gives evidence for the change of ionic radii of ca. −8 pm associated with change in oxidation state between NpIII and NpIV. The rich redox chemistry associated with the syntheses is augmented by the reduction of [Np(Cp′)3] by KC8 in the presence of 2.2.2-cryptand to afford a neptunium(II) complex that is thermally unstable above −10 °C like the UII and ThII complexes K(2.2.2-cryptand)[Th/U(Cp′)3]. Together, these spontaneous and controlled redox reactions of organo-neptunium complexes, along with information from structural characterisation, show the relevance of organometallic Np chemistry to understanding fundamental structure and bonding in the minor actinides.
Co-reporter:Johann A. Hlina; James R. Pankhurst; Nikolas Kaltsoyannis
Journal of the American Chemical Society 2016 Volume 138(Issue 10) pp:3333-3345
Publication Date(Web):March 4, 2016
DOI:10.1021/jacs.5b10698
Heterobimetallic complexes containing short uranium–group 10 metal bonds have been prepared from monometallic IUIV(OArP-κ2O,P)3 (2) {[ArPO]− = 2-tert-butyl-4-methyl-6-(diphenylphosphino)phenolate}. The U–M bond in IUIV(μ-OArP-1κ1O,2κ1P)3M0, M = Ni (3–Ni), Pd (3–Pd), and Pt (3–Pt), has been investigated by experimental and DFT computational methods. Comparisons of 3–Ni with two further U–Ni complexes XUIV(μ-OArP-1κ1O,2κ1P)3Ni0, X = Me3SiO (4) and F (5), was also possible via iodide substitution. All complexes were characterized by variable-temperature NMR spectroscopy, electrochemistry, and single crystal X-ray diffraction. The U–M bonds are significantly shorter than any other crystallographically characterized d–f-block bimetallic, even though the ligand flexes to allow a variable U–M separation. Excellent agreement is found between the experimental and computed structures for 3–Ni and 3–Pd. Natural population analysis and natural localized molecular orbital (NLMO) compositions indicate that U employs both 5f and 6d orbitals in covalent bonding to a significant extent. Quantum theory of atoms-in-molecules analysis reveals U–M bond critical point properties typical of metallic bonding and a larger delocalization index (bond order) for the less polar U–Ni bond than U–Pd. Electrochemical studies agree with the computational analyses and the X-ray structural data for the U–X adducts 3–Ni, 4, and 5. The data show a trend in uranium–metal bond strength that decreases from 3–Ni down to 3–Pt and suggest that exchanging the iodide for a fluoride strengthens the metal–metal bond. Despite short U–TM (transition metal) distances, four other computational approaches also suggest low U–TM bond orders, reflecting highly transition metal localized valence NLMOs. These are more so for 3–Pd than 3–Ni, consistent with slightly larger U–TM bond orders in the latter. Computational studies of the model systems (PH3)3MU(OH)3I (M = Ni, Pd) reveal longer and weaker unsupported U–TM bonds vs 3.
Co-reporter:Charlotte J. Stevens, Alessandro Prescimone, Floriana Tuna, Eric J. L. McInnes, Simon Parsons, Carole A. Morrison, Polly L. Arnold, and Jason B. Love
Inorganic Chemistry 2016 Volume 55(Issue 1) pp:214-220
Publication Date(Web):December 18, 2015
DOI:10.1021/acs.inorgchem.5b02151
The effect of pressure on the intranuclear M···M separation and intermolecular secondary interactions in the dinuclear chromium Pacman complex [Cr2(L)](C6H6) was evaluated because this compound contains both a short Cr···Cr separation and an exogenously bound molecule of benzene in the solid state. The electronic structure of [Cr2(L)] was determined by electron paramagnetic resonance spectroscopy, SQUID magnetometry, and density functional theory calculations and shows a diamagnetic ground state through antiferromagnetic exchange, with no evidence for a Cr–Cr bond. Analysis of the solid-state structures of [Cr2(L)](C6H6) at pressures varying from ambient to 3.0 GPa shows little deformation in the Cr···Cr separation, i.e., no Cr–Cr bond formation, but instead a significantly increased interaction between the exogenous arene and the chromium iminopyrrolide environment. It is therefore apparent from this analysis that [Cr2(L)] would be best exploited as a rigid chemical synthon, with pressure regulation being used to mediate the approach and secondary interactions of possible substrates.
Co-reporter:Nicola L. Bell; Laurent Maron
Journal of the American Chemical Society 2015 Volume 137(Issue 33) pp:10492-10495
Publication Date(Web):August 5, 2015
DOI:10.1021/jacs.5b06630
Molecules containing actinide–nitrogen multiple bonds are of current interest as simple models for new actinide nitride nuclear fuels, and for their potential for the catalytic activation of inert hydrocarbon C–H bonds. Complexes with up to three uranium–nitrogen double bonds are now being widely studied, yet those with one thorium–nitrogen double bond are rare, and those with two are unknown. A new, simple mono(imido) thorium complex and the first bis(imido) thorium complex, K[Th(═NAr)N″3] and K2[Th(═NAr)2N″2], are readily made from insertion reactions (Ar = aryl, N″ = N(SiMe3)2) into the Th–C bond of the cyclometalated thorium amides [ThN″2(N(SiMe3)(SiMe2CH2))] and K[ThN″(N(SiMe3)(SiMe2CH2))2]. X-ray and computational structural analyses show a “transition-metal-like” cis-bis(imido) geometry and polarized Th═N bonds with twice the Wiberg bond order of the formally single Th–N bond in the same molecule.
Co-reporter:Markus Zegke, Gary S. Nichol, Polly L. Arnold and Jason B. Love
Chemical Communications 2015 vol. 51(Issue 27) pp:5876-5879
Publication Date(Web):02 Mar 2015
DOI:10.1039/C5CC00867K
Reactions between the uranyl(VI) Pacman complex [(UO2)(py)(H2L)] of the Schiff-base polypyrrolic macrocycle L and Tebbe's reagent or DIBAL result in the first selective reductive functionalisation of the uranyl oxo by Al to form [(py)(R2AlOUO)(py)(H2L)] (R = Me or iBu). The clean displacement of the oxo-coordinated Al(III) by Group 1 cations has enabled the development of a one-pot, DIBAL-catalysed reduction of the U(VI) uranyl complexes to a series of new, mono-oxo alkali-metal-functionalised uranyl(V) complexes [(py)3(MOUO)(py)(H2L)] (M = Li, Na, K).
Co-reporter:Polly L. Arnold; Anne-Frédérique Pécharman; Rianne M. Lord; Guy M. Jones; Emmalina Hollis; Gary S. Nichol; Laurent Maron; Jian Fang; Thomas Davin;Jason B. Love
Inorganic Chemistry 2015 Volume 54(Issue 7) pp:3702-3710
Publication Date(Web):March 23, 2015
DOI:10.1021/acs.inorgchem.5b00420
Uranyl complexes of a large, compartmental N8-macrocycle adopt a rigid, “Pacman” geometry that stabilizes the UV oxidation state and promotes chemistry at a single uranyl oxo-group. We present here new and straightforward routes to singly reduced and oxo-silylated uranyl Pacman complexes and propose mechanisms that account for the product formation, and the byproduct distributions that are formed using alternative reagents. Uranyl(VI) Pacman complexes in which one oxo-group is functionalized by a single metal cation are activated toward single-electron reduction. As such, the addition of a second equivalent of a Lewis acidic metal complex such as MgN″2 (N″ = N(SiMe3)2) forms a uranyl(V) complex in which both oxo-groups are Mg functionalized as a result of Mg–N bond homolysis. In contrast, reactions with the less Lewis acidic complex [Zn(N″)Cl] favor the formation of weaker U–O–Zn dative interactions, leading to reductive silylation of the uranyl oxo-group in preference to metalation. Spectroscopic, crystallographic, and computational analysis of these reactions and of oxo-metalated products isolated by other routes have allowed us to propose mechanisms that account for pathways to metalation or silylation of the exo-oxo-group.
Co-reporter:Dr. Polly L. Arnold;M.Chem. Max W. McMullon;M.Sc. Julia Rieb;Dr. Fritz E. Kühn
Angewandte Chemie International Edition 2015 Volume 54( Issue 1) pp:82-100
Publication Date(Web):
DOI:10.1002/anie.201404613
Abstract
Most homogeneous catalysis relies on the design of metal complexes to trap and convert substrates or small molecules to value-added products. Organometallic lanthanide compounds first gave a tantalizing glimpse of their potential for catalytic CH bond transformations with the selective cleavage of one CH bond in methane by bis(permethylcyclopentadienyl)lanthanide methyl [(η5-C5Me5)2Ln(CH3)] complexes some 25 years ago. Since then, numerous metal complexes from across the periodic table have been shown to selectively activate hydrocarbon CH bonds, but the challenges of closing catalytic cycles still remain; many f-block complexes show great potential in this important area of chemistry.
Co-reporter:Rami J. Batrice, Jamie McKinven, Polly L. Arnold, and Moris S. Eisen
Organometallics 2015 Volume 34(Issue 16) pp:4039-4050
Publication Date(Web):August 4, 2015
DOI:10.1021/acs.organomet.5b00455
A catalyzed conversion of terminal alkynes into dimers, trimers, and trisubstituted benzenes has been developed using the actinide amides U[N(SiMe3)2]3 (1) and [(Me3Si)2N]2An[κ2-(N,C)-CH2Si(CH3)N(SiMe3)] (An = U (2), Th (3)) as precatalysts. These complexes allow for preferential product formation according to the identity of the metal and the catalyst loading. While these complexes are known as valuable precursors for the preparation of various actinide complexes, this is the first demonstration of their use as catalysts for C–C bond forming reactions. At high uranium catalyst loading, the cycloaddition of the terminal alkyne is generally preferred, whereas at low loadings, linear oligomerization to form enynes is favored. The thorium metallacycle produces only organic enynes, suggesting the importance of the ability of uranium to form stabilizing interactions with arenes and related π-electron-containing intermediates. Kinetic, spectroscopic, and mechanistic data that inform the nature of the activation and catalytic cycle of these reactions are presented.
Co-reporter:Polly L. Arnold, Joy H. Farnaby, Michael G. Gardiner, and Jason B. Love
Organometallics 2015 Volume 34(Issue 11) pp:2114-2117
Publication Date(Web):January 20, 2015
DOI:10.1021/om5012193
U(III) complexes of the conformationally flexible, small-cavity macrocycle trans-calix[2]benzene[2]pyrrolide (L)2–, [U(L)X] (X = O-2,6-tBu2C6H3, N(SiMe3)2), have been synthesized from [U(L)BH4] and structurally characterized. These complexes show binding of the U(III) center in the bis(arene) pocket of the macrocycle, which flexes to accommodate the increase in the steric bulk of X, resulting in long U–X bonds to the ancillary ligands. Oxidation to the cationic U(IV) complex [U(L)X][B(C6F5)4] (X = BH4) results in ligand rearrangement to bind the smaller, harder cation in the bis(pyrrolide) pocket, in a conformation that has not been previously observed for (L)2–, with X located between the two ligand arene rings.
Co-reporter: Polly L. Arnold;Dr. Alessro Prescimone;Dr. Joy H. Farnaby;Dr. Stephen M. Mansell; Simon Parsons; Nikolas Kaltsoyannis
Angewandte Chemie International Edition 2015 Volume 54( Issue 23) pp:6735-6739
Publication Date(Web):
DOI:10.1002/anie.201411250
Abstract
The diuranium(III) compound [UN′′2]2(μ-η6:η6-C6H6) (N′′=N(SiMe3)2) has been studied using variable, high-pressure single-crystal X-ray crystallography, and density functional theory . In this compound, the low-coordinate metal cations are coupled through π- and δ-symmetric arene overlap and show close metalCH contacts with the flexible methyl CH groups of the sterically encumbered amido ligands. The metal–metal separation decreases with increasing pressure, but the most significant structural changes are to the close contacts between ligand CH bonds and the U centers. Although the interatomic distances are suggestive of agostic-type interactions between the U and ligand peripheral CH groups, QTAIM (quantum theory of atoms-in-molecules) computational analysis suggests that there is no such interaction at ambient pressure. However, QTAIM and NBO analyses indicate that the interaction becomes agostic at 3.2 GPa.
Co-reporter:Dr. Polly L. Arnold;M.Chem. Max W. McMullon;M.Sc. Julia Rieb;Dr. Fritz E. Kühn
Angewandte Chemie 2015 Volume 127( Issue 1) pp:84-103
Publication Date(Web):
DOI:10.1002/ange.201404613
Abstract
Homogene Katalyse hängt zu erheblichen Teilen von der Synthese von Metallkomplexen ab, die in der Lage sind, (niedermolekulare) Substrate zu binden und zu wertvolleren Produkten umzusetzen. Einen ersten Vorgeschmack der katalytischen C-H-Aktivierung mit metallorganischen Lanthanoidkomplexen gab es vor etwa 25 Jahren, als es gelang, mit Bis(pentamethylcyclopentadenyl)lanthanoid-Methylkomplexen [(η5-(C5Me5)2Ln(CH3)] eine C-H-Bindung von Methan zu spalten. In der Zwischenzeit wurden zahlreiche Metallkomplexe aus dem ganzen Periodensystem gefunden, die selektiv C-H-Bindungen aktivieren können, aber die Herausforderung, einen geschlossenen Katalysezyklus zu etablieren, bleibt bestehen. Viele f-Block-Komplexe scheinen aber über großes Potenzial auf diesem wichtigen Gebiet zu verfügen.
Co-reporter: Polly L. Arnold;Dr. Alessro Prescimone;Dr. Joy H. Farnaby;Dr. Stephen M. Mansell; Simon Parsons; Nikolas Kaltsoyannis
Angewandte Chemie 2015 Volume 127( Issue 23) pp:6839-6843
Publication Date(Web):
DOI:10.1002/ange.201411250
Abstract
The diuranium(III) compound [UN′′2]2(μ-η6:η6-C6H6) (N′′=N(SiMe3)2) has been studied using variable, high-pressure single-crystal X-ray crystallography, and density functional theory . In this compound, the low-coordinate metal cations are coupled through π- and δ-symmetric arene overlap and show close metalCH contacts with the flexible methyl CH groups of the sterically encumbered amido ligands. The metal–metal separation decreases with increasing pressure, but the most significant structural changes are to the close contacts between ligand CH bonds and the U centers. Although the interatomic distances are suggestive of agostic-type interactions between the U and ligand peripheral CH groups, QTAIM (quantum theory of atoms-in-molecules) computational analysis suggests that there is no such interaction at ambient pressure. However, QTAIM and NBO analyses indicate that the interaction becomes agostic at 3.2 GPa.
Co-reporter:Polly L. Arnold, Charlotte J. Stevens, Joy H. Farnaby, Michael G. Gardiner, Gary S. Nichol, and Jason B. Love
Journal of the American Chemical Society 2014 Volume 136(Issue 29) pp:10218-10221
Publication Date(Web):July 8, 2014
DOI:10.1021/ja504835a
A new robust and high-yielding synthesis of the valuable UIII synthon [U(BH4)3(THF)2] is reported. Reactivity in ligand exchange reactions is found to contrast significantly to that of uranium triiodide. This is exemplified by the synthesis and characterization of azamacrocyclic UIII complexes, including mononuclear [U(BH4)(L)] and dinuclear [Li(THF)4][{U(BH4)}2(μ-BH4)(LMe)] and [Na(THF)4][{U(BH4)}2(μ-BH4)(LA)(THF)2]. The structures of all complexes have been determined by single-crystal X-ray diffraction and display two new UIII2(BH4)3 motifs.
Co-reporter:Polly L. Arnold, Joy H. Farnaby, Rebecca C. White, Nikolas Kaltsoyannis, Michael G. Gardiner and Jason B. Love
Chemical Science 2014 vol. 5(Issue 2) pp:756-765
Publication Date(Web):02 Dec 2013
DOI:10.1039/C3SC52072B
New, conformationally restricted ThIV and UIV complexes, [ThCl2(L)] and [UI2(L)], of the small-cavity, dipyrrolide, dianionic macrocycle trans-calix[2]benzene[2]pyrrolide (L)2− are reported and are shown to have unusual κ5:κ5 binding in a bent metallocene-type structure. Single-electron reduction of [UI2(L)] affords [UI(THF)(L)] and results in a switch in ligand binding from κ5-pyrrolide to η6-arene sandwich coordination, demonstrating the preference for arene binding by the electron-rich UIII ion. Facile loss of THF from [UI(THF)(L)] further increases the amount of U–arene back donation. [UI(L)] can incorporate a further UIII equivalent, UI3, to form the very unusual dinuclear complex [U2I4(L)] in which the single macrocycle adopts both κ5:κ5 and η6:κ1:η6:κ1 binding modes in the same complex. Hybrid density functional theory calculations carried out to compare the electronic structures and bonding of [UIIII(L)] and [UIII2I4(L)] indicate increased contributions to the covalent bonding in [U2I4(L)] than in [UI(L)], and similar U–arene interactions in both. MO analysis and QTAIM calculations find minimal U–U interaction in [U2I4(L)]. In contrast to the reducible U complex, treatment of [ThCl2(L)] with either a reductant or non-nucleophilic base results in metallation of the aryl rings of the macrocycle to form the (L−2H)4− tetraanion and two new and robust Th–C bonds in the –ate complexes [K(THF)2ThIV(μ-Cl)(L−2H)]2 and K[ThIV{N(SiMe3)2}(L−2H)].
Co-reporter:Polly L. Arnold, Isobel A. Marr, Sergey Zlatogorsky, Ronan Bellabarba and Robert P. Tooze
Dalton Transactions 2014 vol. 43(Issue 1) pp:34-37
Publication Date(Web):28 Oct 2013
DOI:10.1039/C3DT52762J
A Sc NHC complex readily activates three equivalents of CO2 showing ‘Frustrated Lewis Pair’ type reactivity with each metal–carbene bond, but whilst CS2 is also activated by the labile carbenes, no metal involvement is observed.
Co-reporter:Polly L. Arnold, Thomas Cadenbach, Isobel H. Marr, Andrew A. Fyfe, Nicola L. Bell, Ronan Bellabarba, Robert P. Tooze and Jason B. Love
Dalton Transactions 2014 vol. 43(Issue 38) pp:14346-14358
Publication Date(Web):11 Jun 2014
DOI:10.1039/C4DT01442A
The reactivity of a series of organometallic rare earth and actinide complexes with hemilabile NHC-ligands towards substrates with acidic C–H and N–H bonds is described. The synthesis, characterisation and X-ray structures of the new heteroleptic mono- and bis(NHC) cyclopentadienyl complexes LnCp2(L) 1 (Ln = Sc, Y, Ce; L = alkoxy-tethered carbene [OCMe2CH2(1-C{NCHCHNiPr})]), LnCp(L)2 (Ln = Y) 2, and the homoleptic tetrakis(NHC) complex Th(L)44 are described. The reactivity of these complexes, and of the homoleptic complexes Ln(L)3 (Ln = Sc 3, Ce), with E–H substrates is described, where EH = pyrrole C4H4NH, indole C8H6NH, diphenylacetone Ph2CC(O)Me, terminal alkynes RCCH (R = Me3Si, Ph), and cyclopentadiene C5H6. Complex 1-Y heterolytically cleaves and adds pyrrole and indole N–H across the metal carbene bond, whereas 1-Ce does not, although 3 and 4 form H-bonded adducts. Complexes 1-Y and 1-Sc form adducts with CpH without cleaving the acidic C–H bond, 1-Ce cleaves the Cp–H bond, but 2 reacts to form the very rare H+–[C5H5]−–H+ motif. Complex 1-Ce cleaves alkyne C–H bonds but the products rearrange upon formation, while complex 1-Y cleaves the C–H bond in diphenylacetone forming a product which rearranges to the Y–O bonded enolate product.
Co-reporter:Polly L. Arnold ; Emmalina Hollis ; Gary S. Nichol ; Jason B. Love ; Jean-Christophe Griveau ; Roberto Caciuffo ; Nicola Magnani ; Laurent Maron ; Ludovic Castro ; Ahmed Yahia ; Samuel O. Odoh ;Georg Schreckenbach
Journal of the American Chemical Society 2013 Volume 135(Issue 10) pp:3841-3854
Publication Date(Web):March 4, 2013
DOI:10.1021/ja308993g
The heterobimetallic complexes [{UO2Ln(py)2(L)}2], combining a singly reduced uranyl cation and a rare-earth trication in a binucleating polypyrrole Schiff-base macrocycle (Pacman) and bridged through a uranyl oxo-group, have been prepared for Ln = Sc, Y, Ce, Sm, Eu, Gd, Dy, Er, Yb, and Lu. These compounds are formed by the single-electron reduction of the Pacman uranyl complex [UO2(py)(H2L)] by the rare-earth complexes LnIII(A)3 (A = N(SiMe3)2, OC6H3But2-2,6) via homolysis of a Ln–A bond. The complexes are dimeric through mutual uranyl exo-oxo coordination but can be cleaved to form the trimetallic, monouranyl “ate” complexes [(py)3LiOUO(μ-X)Ln(py)(L)] by the addition of lithium halides. X-ray crystallographic structural characterization of many examples reveals very similar features for monomeric and dimeric series, the dimers containing an asymmetric U2O2 diamond core with shorter uranyl U═O distances than in the monomeric complexes. The synthesis by LnIII–A homolysis allows [5f1-4fn]2 and Li[5f1-4fn] complexes with oxo-bridged metal cations to be made for all possible 4fn configurations. Variable-temperature SQUID magnetometry and IR, NIR, and EPR spectroscopies on the complexes are utilized to provide a basis for the better understanding of the electronic structure of f-block complexes and their f-electron exchange interactions. Furthermore, the structures, calculated by restricted-core or all-electron methods, are compared along with the proposed mechanism of formation of the complexes. A strong antiferromagnetic coupling between the metal centers, mediated by the oxo groups, exists in the UVSmIII monomer, whereas the dimeric UVDyIII complex was found to show magnetic bistability at 3 K, a property required for the development of single-molecule magnets.
Co-reporter:Ayodele Oladeji, Polly L. Arnold, Mohammad I. Ali, Slawomir Sujecki, Andrew Phillips, Igor V. Sazanovich and Julia A. Weinstein
Journal of Materials Chemistry A 2013 vol. 1(Issue 48) pp:8075-8085
Publication Date(Web):21 Oct 2013
DOI:10.1039/C3TC31601G
Er-doped SiO2 powders were prepared by a simple and straightforward sol–gel method using tetraethoxysilane and erbium triflate as sources of SiO2 and Er2O3. Excitation with near-infrared 800 and 980 nm light initiated an energy up-conversion process in these materials, which resulted in the emission in the predominantly red and green parts of the spectrum. Laser power dependence of the intensity of the up-converted emission was studied both experimentally and by using a numerical model to understand the up-conversion mechanism. Excited state absorption (ESA) and five unique ion–ion energy transfer up-conversion (ETU) processes between the relevant levels are identified as possible mechanisms for the observed luminescence. The efficient NIR-visible upconversion was observed even at comparatively low powers of 200 mW cm−2.
Co-reporter:Stephen M. Mansell, Fanny Bonnet, Marc Visseaux and Polly L. Arnold
Dalton Transactions 2013 vol. 42(Issue 25) pp:9033-9039
Publication Date(Web):10 Apr 2013
DOI:10.1039/C3DT50131K
The synthesis and structural characterisation of the uranium(IV) amido-borohydrides (N′′)2U{κ2-N(SiMe3)SiMe2CH2BBN-H} and U{κ2-N(SiMe3)SiMe2CH2BBN-H}2, and their activity as pre-catalysts for the polymerisation of isoprene are described.
Co-reporter:Polly L. Arnold, Zöe R. Turner, Anne I. Germeroth, Ian J. Casely, Gary S. Nichol, Ronan Bellabarba and Robert P. Tooze
Dalton Transactions 2013 vol. 42(Issue 5) pp:1333-1337
Publication Date(Web):27 Nov 2012
DOI:10.1039/C2DT31698F
The reactions of f-block silylamido N-heterocyclic carbene (NHC) complexes ([M(L)(N{SiMe3}2)2], M = Y, Ce, and U, L = bidentate alkoxy-tethered NHC ligand) with CO and CO2 have been studied and compared to each other, to those of selected [M(L)2(N{SiMe3}2)] complexes, and to those of [M(N{SiMe3}2)3] to identify the effect of the labile NHC group on the small molecule activation chemistry. The small molecules COS and N2CPh2 have also been studied.
Co-reporter:Guy M. Jones; Polly L. Arnold;Dr. Jason B. Love
Chemistry - A European Journal 2013 Volume 19( Issue 31) pp:10287-10294
Publication Date(Web):
DOI:10.1002/chem.201301067
Abstract
Simple and versatile routes to the functionalization of uranyl-derived UV–oxo groups are presented. The oxo-lithiated, binuclear uranium(V)–oxo complexes [{(py)3LiOUO}2(L)] and [{(py)3LiOUO}(OUOSiMe3)(L)] were prepared by the direct combination of the uranyl(VI) silylamide “ate” complex [Li(py)2][(OUO)(N”)3] (N”=N(SiMe3)2) with the polypyrrolic macrocycle H4L or the mononuclear uranyl (VI) Pacman complex [UO2(py)(H2L)], respectively. These oxo-metalated complexes display distinct UO single and multiple bonding patterns and an axial/equatorial arrangement of oxo ligands. Their ready availability allows the direct functionalization of the uranyl oxo group leading to the binuclear uranium(V) oxo–stannylated complexes [{(R3Sn)OUO}2(L)] (R=nBu, Ph), which represent rare examples of mixed uranium/tin complexes. Also, uranium–oxo-group exchange occurred in reactions with [TiCl(OiPr)3] to form U-OC bonds [{(py)3LiOUO}(OUOiPr)(L)] and [(iPrOUO)2(L)]. Overall, these represent the first family of uranium(V) complexes that are oxo-functionalised by Group 14 elements.
Co-reporter:Stephen M. Mansell, Joy H. Farnaby, Anne I. Germeroth, and Polly L. Arnold
Organometallics 2013 Volume 32(Issue 15) pp:4214-4222
Publication Date(Web):July 12, 2013
DOI:10.1021/om4003957
A new dinitrogen adduct of a homoleptic uranium tris(siloxide) complex, [U{OSi(Mes)3}3]2(μ-η2:η2-N2), is reported. Synthesis of the 15N-labeled isotopomer and Raman spectroscopy confirm the reductive activation of N2 to a (N2)2– dianion. The 15N NMR shift of the 15N2-labeled isotopomer is also reported. Crystallographic characterization shows a side-on (N2)2– coordinated in either an eclipsed or staggered conformation in different crystals. The U–N2–U complex is stable to vacuum and shows high thermal stability, retaining the formally reduced dinitrogen at 100 °C. The parent three-coordinate uranium(III) [U{OSi(Mes)3}3] could not be isolated in our hands, with N2-free syntheses affording only uranium(IV) compounds. The rational synthesis and full characterization of two such U(IV) byproducts, [U{OSi(Mes)3}{N(SiMe3)2}3] and [U{OSi(Mes)3}4], is also reported.
Co-reporter:Polly L. Arnold, Guy M. Jones, Qing-Jiang Pan, Georg Schreckenbach and Jason B. Love
Dalton Transactions 2012 vol. 41(Issue 22) pp:6595-6597
Publication Date(Web):30 Apr 2012
DOI:10.1039/C2DT30658A
Expansion of a Schiff-base polypyrrolic macrocycle allows the formation of a binuclear uranyl complex with co-linear uranyl ions and a very short oxo–oxo distance.
Co-reporter:Guy M. Jones; Polly L. Arnold;Dr. Jason B. Love
Angewandte Chemie International Edition 2012 Volume 51( Issue 50) pp:12584-12587
Publication Date(Web):
DOI:10.1002/anie.201207609
Co-reporter:Guy M. Jones; Polly L. Arnold;Dr. Jason B. Love
Angewandte Chemie 2012 Volume 124( Issue 50) pp:12752-12755
Publication Date(Web):
DOI:10.1002/ange.201207609
Co-reporter:Stephen M. Mansell ; Nikolas Kaltsoyannis
Journal of the American Chemical Society 2011 Volume 133(Issue 23) pp:9036-9051
Publication Date(Web):May 17, 2011
DOI:10.1021/ja2019492
Previously unanticipated dinitrogen activation is exhibited by the well-known uranium tris(aryloxide) U(ODtbp)3, U(OC6H3-But2-2,6)3, and the tri-tert-butyl analogue U(OTtbp)3, U(OC6H2-But3-2,4,6)3, in the form of bridging, side-on dinitrogen complexes [U(OAr)3]2(μ-η2:η2-N2), for which the tri-tert-butyl N2 complex is the most robust U2(N2) complex isolated to date. Attempted reduction of the tris(aryloxide) complex under N2 gave only the potassium salt of the uranium(III) tetra(aryloxide) anion, K[U(OAr)4], as a result of ligand redistribution. The solid-state structure is a polymeric chain formed by each potassium cation bridging two arenes of adjacent anions in an η6 fashion. The same uranium tris(aryloxides) were also found to couple carbon monoxide under ambient conditions to give exclusively the ynediolate [OCCO]2– dianion in [U(OAr)3]2(μ-η1:η1-C2O2), in direct analogy with the reductive coupling recently shown to afford [U{N(SiMe3)2}3]2(μ-η1:η1-C2O2). The related UIII complexes U{N(SiPhMe2)2}3 and U{CH(SiMe3)2}3 however do not show CO coupling chemistry in our hands. Of the aryloxide complexes, only the U(OC6H2-But3-2,4,6)3 reacts with CO2 to give an insertion product containing bridging oxo and aryl carbonate moieties, U2(OTtbp)4(μ-O)(μ-η1:η1-O2COC6H2-But3-2,4,6)2, which has been structurally characterized. The presence of coordinated N2 in [U(OTtbp)3]2(N2) prevents the occurrence of any reaction with CO2, underscoring the remarkable stability of the N2 complex. The di-tert-butyl aryloxide does not insert CO2, and only U(ODtbp)4 was isolated. The silylamide also reacts with carbon dioxide to afford U(OSiMe3)4 as the only uranium-containing material. GGA and hybrid DFT calculations, in conjunction with topological analysis of the electron density, suggest that the U–N2 bond is strongly polar, and that the only covalent U→N2 interaction is π backbonding, leading to a formal (UIV)2(N2)2– description of the electronic structure. The N–N stretching wavenumber is preferred as a metric of N2 reduction to the N–N bond length, as there is excellent agreement between theory and experiment for the former but poorer agreement for the latter due to X-ray crystallographic underestimation of r(N–N). Possible intermediates on the CO coupling pathway to [U(OAr)3]2(μ-C2O2) are identified, and potential energy surface scans indicate that the ynediolate fragment is more weakly bound than the ancillary ligands, which may have implications in the development of low-temperature and pressure catalytic CO chemistry.
Co-reporter:Polly L. Arnold ; Zoë R. Turner ; Ronan Bellabarba ;Robert P. Tooze
Journal of the American Chemical Society 2011 Volume 133(Issue 30) pp:11744-11756
Publication Date(Web):June 9, 2011
DOI:10.1021/ja204209t
Two functional groups can be delivered at once to organo-rare earth complexes, (L)MR2 and (L)2MR (M = Sc, Y; L = ({1-C(NDippCH2CH2N)}CH2CMe2O), Dipp = 2,6-iPr2-C6H3; R = CH2SiMe3, CH2CMe3), via the addition of E–X across the metal–carbene bond to form a zwitterionic imidazolinium–metal complex, (LE)MR2X, where LE = {1-EC(NDippCH2CH2N)}CH2CMe2O, E is a p-block functional group such as SiR3, PR2, or SnR3, and X is a halide. The “ate” complex (LLi)ScR3 is readily accessible and is best described as a Li carbene adduct, ({1-Li(THF)C(NDippCH2CH2N)}CH2CMe2O)Sc(CH2SiMe3)3, since structural characterization shows the alkoxide ligand bridging the two metals and the carbene Li-bound with the shortest yet recorded Li–C bond distance. This can be converted via lithium halide-eliminating salt metathesis reactions to alkylated or silylated imidazolinium derivatives, (LE)ScR3 (E = SiMe3 or CPh3). All the E-functionalized imidazolinium complexes spontaneously eliminate functionalized hydrocarbyl compounds upon warming to room temperature or slightly above, forming new organic products ER, i.e., forming C–Si, C–P, and C–Sn bonds, and re-forming the inorganic metal carbene (L)MR(X) or (L)2MX complex, respectively. Warming the tris(alkyl) complexes (LE)MR3 forms organic products arising from C–C or C–Si bond formation, which appears to proceed via the same elimination route. Treatment of (L)2Sc(CH2SiMe3) with iodopentafluorobenzene results in the “reverse sense” addition, which upon thermolysis forms the metal aryl complex (L)2Sc(C6F5) and releases the iodoalkane Me3SiCH2I, again facilitated by the reversible functionalization of the N-heterocyclic carbene group in these tethered systems.
Co-reporter:Polly L. Arnold, Zoë R. Turner, Ronan M. Bellabarba and Robert P. Tooze
Chemical Science 2011 vol. 2(Issue 1) pp:77-79
Publication Date(Web):28 Oct 2010
DOI:10.1039/C0SC00452A
A simple coordination complex of uranium(III), a uranium tris(amide), can selectively couple gaseous CO to the linear ynediolate [OCCO]2− dianion, at room temperature and pressure, regardless of the reagent stoichiometry. This product exhibits further reactivity upon warming in the form of the addition of a C–H bond of a methyl group across the CC triple bond, this second carbon–carbon bond forming reaction generating a functionalised enediolate dianion.
Co-reporter:Polly L. Arnold
Chemical Communications 2011 vol. 47(Issue 32) pp:9005-9010
Publication Date(Web):25 May 2011
DOI:10.1039/C1CC10834D
Molecular complexes of uranium are capable of activating a range of industrially and economically important small molecules such as CO, CO2, and N2; new and often unexpected reactions provide insight into an element that needs to be well-understood if future clean-energy solutions are to involve nuclear power.
Co-reporter: Polly L. Arnold;Dr. Emmalina Hollis;Dr. Fraser J. White;Dr. Nicola Magnani; Roberto Caciuffo;Dr. Jason B. Love
Angewandte Chemie International Edition 2011 Volume 50( Issue 4) pp:887-890
Publication Date(Web):
DOI:10.1002/anie.201005511
Co-reporter: Polly L. Arnold;Anne-Frédérique Pécharman ;Dr. Jason B. Love
Angewandte Chemie International Edition 2011 Volume 50( Issue 40) pp:9456-9458
Publication Date(Web):
DOI:10.1002/anie.201104359
Co-reporter: Polly L. Arnold;Dr. Emmalina Hollis;Dr. Fraser J. White;Dr. Nicola Magnani; Roberto Caciuffo;Dr. Jason B. Love
Angewandte Chemie 2011 Volume 123( Issue 4) pp:917-920
Publication Date(Web):
DOI:10.1002/ange.201005511
Co-reporter: Polly L. Arnold;Anne-Frédérique Pécharman ;Dr. Jason B. Love
Angewandte Chemie 2011 Volume 123( Issue 40) pp:9628-9630
Publication Date(Web):
DOI:10.1002/ange.201104359
Co-reporter:Zoë R. Turner ; Ronan Bellabarba ; Robert P. Tooze
Journal of the American Chemical Society 2010 Volume 132(Issue 12) pp:4050-4051
Publication Date(Web):March 4, 2010
DOI:10.1021/ja910673q
Silyl, phosphinyl, stannyl, and boryl reagents can be added across the neutral metal−carbon dative bond in d0 f-block metal N-heterocyclic carbene complexes in a reversible manner, allowing additional functional groups to be incorporated into redox-inactive organo-f-block compounds.
Co-reporter:Polly L. Arnold, Natalie A. Potter (née Jones), Christopher D. Carmichael, Alexandra M. Z. Slawin, Paul Roussel and Jason B. Love
Chemical Communications 2010 vol. 46(Issue 11) pp:1833-1835
Publication Date(Web):12 Feb 2010
DOI:10.1039/B921132B
Trinuclear, supramolecular wheel structures are formed spontaneously from the metallation of a Schiff-base-pyrrole macrocycle by Ce3+ cations, while the related actinide U3+ cation is instead oxidised to U4+ and encapsulated by the macrocyclic framework.
Co-reporter:Polly L. Arnold ; Natalie A. Potter (née Jones) ; Nicola Magnani ; Christos Apostolidis ; Jean-Christophe Griveau ; Eric Colineau ; Alfred Morgenstern ; Roberto Caciuffo ;Jason B. Love
Inorganic Chemistry 2010 Volume 49(Issue 12) pp:5341-5343
Publication Date(Web):May 26, 2010
DOI:10.1021/ic100374j
Syntheses of the bimetallic uranium(III) and neptunium(III) complexes [(UI)2(L)], [(NpI)2(L)], and [{U(BH4)}2(L)] of the Schiff-base pyrrole macrocycles L are described. In the absence of single-crystal structural data, fitting of the variable-temperature solid-state magnetic data allows the prediction of polymeric structures for these compounds in the solid state.
Co-reporter:Jean-Charles Buffet ; Jun Okuda
Inorganic Chemistry 2010 Volume 49(Issue 2) pp:419-426
Publication Date(Web):December 14, 2009
DOI:10.1021/ic900740n
The indium complex InL2N′′ has been prepared from the reaction of 2 equiv of (tBu)2P(O)CH2CH(tBu)OH (HL) with InN′′3 (N′′ = N(SiMe3)2). This complex reacts with a further equivalent of 2,6-di-tert-butylphenol or HL to afford the adducts InL2(OAr) and InL3, respectively. Confirmation that the anion L− exhibits “ligand self-recognition” in the formation of predominantly homochiral complexes RR-InL2N′′ and SS-InL2N′′ is obtained from 1H and 31P NMR spectroscopic data. However, the self-recognition is less effective at the indium cation, and mixtures of InL3 complexes with different configurations are observed. Single-crystal X-ray diffraction data confirm the five-coordinate, distorted bipyramidal In center in InL2N′′ and InL2(OAr) as anticipated. Selected crystals of InL3 show two of the possible configurations: one is the fac-RRR-InL3 complex, analogous to the lanthanide complexes LnL3 reported previously (Ln = Y, Eu, Er, Yb); another is the alternative, homochiral mer form RRR′-InL3. All three complexes are efficient single-component initiators for the ring-opening polymerization of rac-lactide over a wide range of temperatures and monomer-to-initiator ratios, exhibiting reasonable control over the synthesis of isotactic polylactide. Despite its poorly defined structure, InL3 is the fastest initiator among the three complexes for the polymerization of rac-lactide, and shows the best tacticity control. The polylactide samples have high molecular weights Mn,exp (between 44 000 and 270 000 g/mol at completion) and narrow polydispersities (as low as 1.25 at completion).
Co-reporter:Polly L. Arnold, Dipti Patel, Anne-Frédérique Pécharman, Claire Wilson and Jason B. Love
Dalton Transactions 2010 vol. 39(Issue 14) pp:3501-3508
Publication Date(Web):02 Mar 2010
DOI:10.1039/B922115H
The synthesis of the mono-uranyl complex [UO2(THF)(H2LMe)] of a ditopic Schiff-base pyrrole macrocycle is described and is shown to adopt a Pacman wedge-shaped structure in which the uranyl dication is desymmetrised and sits solely in one N4-donor compartment to leave the other vacant. While investigating the mechanism of the previously reported, base-initiated, reductive silylation chemistry of [UO2(THF)(H2LMe)], we found that uranyl hydroxide complexes could be isolated. As such, the reaction between [UO2(THF)(H2LMe)] and KH in THF generated the dimeric cation-cation hydroxide [{UO2(OH)K(C6H6)(H2LMe)}2] when crystallised from C6H6, or alternatively, when crystallised from THF, the monomeric THF-adducted cation-cation complex [UO2(OH)K(THF)2(H2LMe)] was isolated. These compounds result formally from the substitution of the equatorial THF molecule by hydroxide, and it was also shown that the reaction between dry KOH and [UO2(THF)(H2LMe)] generated [{UO2(OH)K(C6H6)(H2LMe)}2].
Co-reporter:PollyL. Arnold ;ZoëR. Turner;Nikolas Kaltsoyannis ;Panagiota Pelekanaki;RonanM. Bellabarba Dr.;RobertP. Tooze Dr.
Chemistry - A European Journal 2010 Volume 16( Issue 31) pp:9623-9629
Publication Date(Web):
DOI:10.1002/chem.201001471
Abstract
Oxidative halogenation with trityl chloride provides convenient access to CeIV and UIV chloroamides [M(N{SiMe3}2)3Cl] and their N-heterocyclic carbene derivatives, [M(L)(N{SiMe3}2)2Cl] (L=OCMe2CH2(CNCH2CH2NDipp) Dipp=2,6-iPr2C6H3). Computational analysis of the bonding in these and a fluoro analogue, [U(L)(N{SiMe3}2)2F], provides new information on the covalency in this relative rare oxidation state for molecular cerium complexes. Computational studies reveal increased Mayer bond orders in the actinide carbene bond compared with the lanthanide carbene bond, and natural and atoms-in-molecules analyses suggest greater overall ionicity in the cerium complexes than in the uranium analogues.
Co-reporter:Stephen M. Mansell, Bernabé Fernandez Perandones, Polly L. Arnold
Journal of Organometallic Chemistry 2010 695(25–26) pp: 2814-2821
Publication Date(Web):
DOI:10.1016/j.jorganchem.2010.08.019
Co-reporter:Polly L. Arnold and Ian J. Casely
Chemical Reviews 2009 Volume 109(Issue 8) pp:3599
Publication Date(Web):April 9, 2009
DOI:10.1021/cr8005203
Co-reporter:Polly L. Arnold ; Melanie S. Sanford ;Stephen M. Pearson
Journal of the American Chemical Society 2009 Volume 131(Issue 39) pp:13912-13913
Publication Date(Web):September 10, 2009
DOI:10.1021/ja905713t
A PdIV complex that represents a viable catalytic intermediate in Pd-catalyzed C−H bond halogenation reactions has been isolated and structurally characterized. It contains the first examples of both a PdIV NHC bond and a PdIV alkoxide bond and serves as a precatalyst for C−H bond halogenation. As such, this represents a new class of tunable supporting ligand systems in PdIV catalysis.
Co-reporter:Polly L. Arnold, Jason B. Love, Dipti Patel
Coordination Chemistry Reviews 2009 Volume 253(15–16) pp:1973-1978
Publication Date(Web):August 2009
DOI:10.1016/j.ccr.2009.03.014
The uranyl dication, [UO2]2+, is the most prevalent and most thermodynamically stable form of uranium and is a soluble and problematic environmental contaminant. It is also extraordinarily chemically robust due to the strongly covalent trans-UO2 bonding. In contrast, the pentavalent uranyl cation [UO2]+ is unstable in an aqueous environment with respect to disproportionation into tetravalent uranium species and [UO2]2+. Aside from fundamental interest, an understanding of the pentavalent [UO2]+ cation is desirable since it is important environmentally as a key intermediate in the precipitation of uranium from groundwater.In the last 2 years, the use of anaerobic coordination chemistry techniques and organometallic reagents has allowed the isolation of a few kinetically inert complexes containing the f1 [UO2]+ cation. The synthesis and characterisation of these, and the insight they give into subsequent reactivity of the trans-UO2 unit, is discussed in this review.
Co-reporter:Polly L. Arnold, Jonathan McMaster and Stephen T. Liddle
Chemical Communications 2009 (Issue 7) pp:818-820
Publication Date(Web):17 Dec 2008
DOI:10.1039/B819072K
A salt-elimination reaction between the neodymium monoiodide [Nd(L′)(N″)(I)]2 [L′ = ButNCH2CH2{C(NCSiMe3CHNBut)}; N″ = N(SiMe3)2] and K[FeCp(CO)2] affords the first complex with an unsupported 4f–3d metal–metal bond that is sufficiently stable to be isolated; the bond is identified as principally ionic in nature by DFT calculations.
Co-reporter:Polly L. Arnold, Ian J. Casely, Zoë R. Turner, Ronan Bellabarba and Robert B. Tooze
Dalton Transactions 2009 (Issue 35) pp:7236-7247
Publication Date(Web):29 Jul 2009
DOI:10.1039/B907034F
The synthesis of magnesium and zinc complexes of bidentate anionic alkoxide ligands with saturated-backbone carbene groups is reported. Mono(ligand) and bis(ligand) complexes [M(LR)N″]2 and [M(LR)2] (M = Mg, Zn, N″ = N(SiMe3)2, LR = [OCMe2CH2{CNCH2CH2NR}] R = iPr, Mes, Dipp) have been isolated, and some structurally characterised and compared with the new unsaturated carbene complex [Mg(L)2]. Reactions with silyl halides show either addition across the metal carbene bond, or across the metal alkoxide bond, in accordance with the metals’ electronegativity difference: the metal alkoxide bonds are stronger for MgII complexes, for which the carbene is silylated to form zwitterionic [MgI(Me3SiLR)N″] (Me3SiLR = OCMe2CH2{Me3SiCNCH2CH2NR}) while the metal-bound alkoxide group is silylated in the ZnII complexes forming [ZnI(Me3SiOLR)N″] (Me3SiOLR = Me3SiOCMe2CH2{CNCH2CH2NR}). The proligand [HLR] is silylated at the alcohol group, forming the iodide salt [Me3SiOCMe2CH2{HCNCH2CH2NR}]I.Preliminary results on the use of these complexes as initiators for the polymerisation of rac-lactide are reported, and suggest different initiation mechanisms are occurring for the two metals, in agreement with the different silylation reactivity observed. The polymerisation reactions are facile at room temperature even without an initiator, and yield polymers of reasonable molecular weight and heterotacticity and with good PDI. These are the first magnesium NHC complexes demonstrated to effect lactide polymerisation.Also, an adduct instead of the anticipated potassium alkoxycarbene is generated from the reaction of the proligand [HLR] with potassium amide KN″; this has been structurally characterised.
Co-reporter:PollyL. Arnold Dr.;Jean-Charles Buffet;Robert Blaudeck;Slawomir Sujecki Dr.;Claire Wilson Dr.
Chemistry - A European Journal 2009 Volume 15( Issue 33) pp:8241-8250
Publication Date(Web):
DOI:10.1002/chem.200900522
Abstract
The reaction of a chiral racemic bidentate ligand HL1 (tBu2P(O)CH2CH(tBu)OH) with mid to late trivalent lanthanide cations affords predominantly homochiral lanthanide complexes (RRR)-[Ln(L1)3] and (SSS)-[Ln(L1)3]. A series of reactions are reported that demonstrate that the syntheses are under thermodynamic control, and driven by a ligand ‘self-recognition’ process, in which the large asymmetric bidentate L1 ligands pack most favourably in a C3 geometry around the lanthanide cation. The synthesis of bis(L1) adducts [Ln(L1)2X] (X=N(SiMe3)2, OC6H3tBu-2,6) is also reported. Analysis of the diastereomer mixtures shows that homochiral (L1)2 complexes are favoured but to a lesser extent. The complexes Ln(L1)3 and [Ln(L1)2(OC6H3tBu-2,6)] have been studied as initiators for the polymerization of ε-caprolactone and its copolymer with lactide, glycolide and its copolymer with lactide, and ε-caprolactam.
Co-reporter:Christopher D. Carmichael ; Natalie A. Jones
Inorganic Chemistry 2008 Volume 47(Issue 19) pp:8577-8579
Publication Date(Web):September 3, 2008
DOI:10.1021/ic801138e
Uranium turnings react with elemental iodine in diethyl ether at room temperature, with sonication and/or stirring, over a period of days to afford UI3, UI4(OEt2)2, or UI4(OBun2) depending on the stoichiometry or ether solvent. This is the first room temperature, and thus safe and convenient, synthesis of UI3.
Co-reporter:Polly L. Arnold ; Sergey Zlatogorsky ; Natalie A. Jones ; Christopher D. Carmichael ; Stephen T. Liddle ; Alexander J. Blake ;Claire Wilson
Inorganic Chemistry 2008 Volume 47(Issue 19) pp:9042-9049
Publication Date(Web):September 4, 2008
DOI:10.1021/ic801046u
The d0 yttrium N-heterocyclic carbene compound YL3 (L = OCMe2CH2[C{N(CHCH)NPri}]) has been made and structurally characterized. It adopts a mer configuration of the three bidentate ligands. A comparison of this with the isostructural d1 titanium complex TiL3 is made in order to seek experimental evidence of a π-bonding contribution to the M−C bond. This has been augmented by DFT calculations. Experimentally, the metal radius-corrected Ti−C distance is shorter than the Y−C distance, suggesting a π-bonding contribution in the d1 complex, but the computational data suggest that a shorter σ bond might simply be formed by the more strongly polarizing titanium cation. From the potassium reduction of TiL(OPri)3, only a byproduct arising from silicone grease activation was isolable, identified as a mixed-valent, multinuclear, d0/d1 cluster [TiIIIL2{PriOSiMe2O}K2OTiIV(OPri)4]2 in which the carbene ligands are bound to the TiIII centers in preference to TiIV, with longer Ti−C distances than those found in TiL3.
Co-reporter:PollyL. Arnold Dr.;Jean-Charles Buffet;RobertP. Blaudeck;Slawomir Sujecki Dr.;AlexerJ. Blake ;Claire Wilson Dr.
Angewandte Chemie International Edition 2008 Volume 47( Issue 32) pp:6033-6036
Publication Date(Web):
DOI:10.1002/anie.200801279
Co-reporter:PollyL. Arnold Dr.;IanJ. Casely;ZoëR. Turner ;ChristopherD. Carmichael Dr.
Chemistry - A European Journal 2008 Volume 14( Issue 33) pp:10415-10422
Publication Date(Web):
DOI:10.1002/chem.200801321
Abstract
A new and modular route to bidentate ligands that combines an alkoxide with a saturated backbone N-heterocyclic carbene (NHC) is presented. The bi(heterocyclic) compounds are formally the addition product of a saturated NHC and the alcohol group of the N-functionalised arm. Using these compounds, the synthesis and structural characterisation of the first electropositive metal complexes of saturated N-heterocyclic carbenes has been achieved, and examples structurally characterised for the yttrium(III) and the uranyl [UO2]2+ cations.
Co-reporter:PollyL. Arnold Dr.;Jean-Charles Buffet;RobertP. Blaudeck;Slawomir Sujecki Dr.;AlexerJ. Blake ;Claire Wilson Dr.
Angewandte Chemie 2008 Volume 120( Issue 32) pp:6122-6125
Publication Date(Web):
DOI:10.1002/ange.200801279
Co-reporter:Stephen T. Liddle, Ian S. Edworthy and Polly L. Arnold
Chemical Society Reviews 2007 vol. 36(Issue 11) pp:1732-1744
Publication Date(Web):03 Jul 2007
DOI:10.1039/B611548A
Since the discovery of a stable “bottleable” N-heterocyclic carbene (NHC), there has been a spectacular explosion of interest in this ligand class. This interest stems from a desire to understand the fundamentals of the structure and bonding of these systems, but also because of their numerous and emerging applications in small molecule activation, homogeneous catalysis and Lewis acid-catalysed reactions. In this Tutorial Review, we introduce the reader to NHCs, cover general synthetic methods to prepare anionic tethered NHCs and their metal complexes, and discuss emerging applications in reactivity and catalytic studies.
Co-reporter:Ian J. Casely, Stephen T. Liddle, Alexander J. Blake, Claire Wilson and Polly L. Arnold
Chemical Communications 2007 (Issue 47) pp:5037-5039
Publication Date(Web):09 Oct 2007
DOI:10.1039/B713041D
The tetravalent organometallic cerium complex [CeL4] is readily accessible from the oxidation of the trivalent [CeL3], L = a bidentate N-heterocyclic carbene alkoxide ligand, [C{(NPri)CHCHN}CH2CMe2O]. The [CeL4] complex should behave like the [UL4] analogue, but the two complexes show significantly different structures, highlighting the differences between 4f and 5f metals.
Co-reporter:Stephen M. Mansell, Polly L. Arnold
Polyhedron (25 September 2016) Volume 116() pp:
Publication Date(Web):25 September 2016
DOI:10.1016/j.poly.2016.03.048
[U(ODtbp)3] (ODtbp = O-2,6-tBu2C6H3) reacts in a 1:1 ratio with the isocyanide CN-Xyl (Xyl = 2,6-Me2C6H3) to form the pseudo-tetrahedral 4-coordinate adduct [U(CNXyl)(ODtbp)3] with νCN 24 cm−1 higher compared to the free isocyanide. Uranium(III) complexes with bulky ligands UX3 (X: ODtbp, N″ = N(SiMe3)2) react with cyclooctatetraene (COT) in a 2:1 U:COT ratio to generate the half-sandwich UIV [U(COT)X2] and [UX4] (which for X = N″ spontaneously converts into the more stable metallacycle [U(N″)2{κ2-N(SiMe3)SiMe2CH2}] and HN″), as opposed to the other potential product, the inverse COT-sandwich [(UX2)2(μ-COT)]. The heteroleptic UIII amido-iodide [{U(N″)2(thf)(μ-I)}2] can be isolated in a low yield (14%) from the 2:1 reaction of KN″ and [UI3(thf)4] in thf, and its molecular structure was shown to be dimeric with iodine atoms bridging the U centres.No coupling, no sandwiches: reactions of simple, homoleptic U(III) complexes with bulky ligands show the potential limits at which small π-acceptor molecules can be reductively coupled, and at which cyclic δ-acceptor organics can be sandwiched by reductive activation.
Co-reporter:Polly L. Arnold, Jonathan McMaster and Stephen T. Liddle
Chemical Communications 2009(Issue 7) pp:NaN820-820
Publication Date(Web):2008/12/17
DOI:10.1039/B819072K
A salt-elimination reaction between the neodymium monoiodide [Nd(L′)(N″)(I)]2 [L′ = ButNCH2CH2{C(NCSiMe3CHNBut)}; N″ = N(SiMe3)2] and K[FeCp(CO)2] affords the first complex with an unsupported 4f–3d metal–metal bond that is sufficiently stable to be isolated; the bond is identified as principally ionic in nature by DFT calculations.
Co-reporter:Polly L. Arnold, Ian J. Casely, Zoë R. Turner, Ronan Bellabarba and Robert B. Tooze
Dalton Transactions 2009(Issue 35) pp:NaN7247-7247
Publication Date(Web):2009/07/29
DOI:10.1039/B907034F
The synthesis of magnesium and zinc complexes of bidentate anionic alkoxide ligands with saturated-backbone carbene groups is reported. Mono(ligand) and bis(ligand) complexes [M(LR)N″]2 and [M(LR)2] (M = Mg, Zn, N″ = N(SiMe3)2, LR = [OCMe2CH2{CNCH2CH2NR}] R = iPr, Mes, Dipp) have been isolated, and some structurally characterised and compared with the new unsaturated carbene complex [Mg(L)2]. Reactions with silyl halides show either addition across the metal carbene bond, or across the metal alkoxide bond, in accordance with the metals’ electronegativity difference: the metal alkoxide bonds are stronger for MgII complexes, for which the carbene is silylated to form zwitterionic [MgI(Me3SiLR)N″] (Me3SiLR = OCMe2CH2{Me3SiCNCH2CH2NR}) while the metal-bound alkoxide group is silylated in the ZnII complexes forming [ZnI(Me3SiOLR)N″] (Me3SiOLR = Me3SiOCMe2CH2{CNCH2CH2NR}). The proligand [HLR] is silylated at the alcohol group, forming the iodide salt [Me3SiOCMe2CH2{HCNCH2CH2NR}]I.Preliminary results on the use of these complexes as initiators for the polymerisation of rac-lactide are reported, and suggest different initiation mechanisms are occurring for the two metals, in agreement with the different silylation reactivity observed. The polymerisation reactions are facile at room temperature even without an initiator, and yield polymers of reasonable molecular weight and heterotacticity and with good PDI. These are the first magnesium NHC complexes demonstrated to effect lactide polymerisation.Also, an adduct instead of the anticipated potassium alkoxycarbene is generated from the reaction of the proligand [HLR] with potassium amide KN″; this has been structurally characterised.
Co-reporter:J. A. Hlina, J. A. L. Wells, J. R. Pankhurst, Jason B. Love and P. L. Arnold
Dalton Transactions 2017 - vol. 46(Issue 17) pp:NaN5545-5545
Publication Date(Web):2017/02/03
DOI:10.1039/C6DT04570G
The heterotetra- and bimetallic uranium(IV)–rhodium(I) complexes [UIVI2(μ-OArP-1κ1O,2κ1P)2RhI(μ-I)]2 (2) (ArPO− = 2-(diphenylphosphino)-6-tert-butyl-4-methylphenoxide) and UIVI(μ-I)(μ-OArP-1κ1O,2κ1P)3RhI (3) were prepared by treatment of UIVI(OArP-κ2O,P)3 (1) with rhodium(I) iodide olefin complexes. The reaction of 1 with the monodentate cyclooctene (coe) rhodium(I) precursor [(coe)2RhII]2 gives only the bimetallic complex [UIVRhI] 3, and with the diene [(cod)RhII]2 (5) (cod = 1,5-cyclooctadiene), mixtures of [UIVRhI]2 complex 2 and [UIVRhI] 3 along with (cod)RhIOArP-κ2O,P (4), a RhI side-product from the formation of 2. The complexes were characterised by single crystal X-ray diffraction, NMR and UV-vis-NIR spectroscopy, and electrochemistry. The UIV–RhI intermetallic distances in 2 (2.7601(5) Å) and 3 (2.7630(5) Å) are among the shortest between f-elements and transition metals reported to date. Despite almost identical U–Rh bond lengths in the solid state, in solution only weak, and very different interactions between the metal centres are found.
Co-reporter:Polly L. Arnold, Natalie A. Potter (née Jones), Christopher D. Carmichael, Alexandra M. Z. Slawin, Paul Roussel and Jason B. Love
Chemical Communications 2010 - vol. 46(Issue 11) pp:NaN1835-1835
Publication Date(Web):2010/02/12
DOI:10.1039/B921132B
Trinuclear, supramolecular wheel structures are formed spontaneously from the metallation of a Schiff-base-pyrrole macrocycle by Ce3+ cations, while the related actinide U3+ cation is instead oxidised to U4+ and encapsulated by the macrocyclic framework.
Co-reporter:Polly L. Arnold
Chemical Communications 2011 - vol. 47(Issue 32) pp:NaN9010-9010
Publication Date(Web):2011/05/25
DOI:10.1039/C1CC10834D
Molecular complexes of uranium are capable of activating a range of industrially and economically important small molecules such as CO, CO2, and N2; new and often unexpected reactions provide insight into an element that needs to be well-understood if future clean-energy solutions are to involve nuclear power.
Co-reporter:Polly L. Arnold, Isobel A. Marr, Sergey Zlatogorsky, Ronan Bellabarba and Robert P. Tooze
Dalton Transactions 2014 - vol. 43(Issue 1) pp:NaN37-37
Publication Date(Web):2013/10/28
DOI:10.1039/C3DT52762J
A Sc NHC complex readily activates three equivalents of CO2 showing ‘Frustrated Lewis Pair’ type reactivity with each metal–carbene bond, but whilst CS2 is also activated by the labile carbenes, no metal involvement is observed.
Co-reporter:Ian J. Casely, Stephen T. Liddle, Alexander J. Blake, Claire Wilson and Polly L. Arnold
Chemical Communications 2007(Issue 47) pp:NaN5039-5039
Publication Date(Web):2007/10/09
DOI:10.1039/B713041D
The tetravalent organometallic cerium complex [CeL4] is readily accessible from the oxidation of the trivalent [CeL3], L = a bidentate N-heterocyclic carbene alkoxide ligand, [C{(NPri)CHCHN}CH2CMe2O]. The [CeL4] complex should behave like the [UL4] analogue, but the two complexes show significantly different structures, highlighting the differences between 4f and 5f metals.
Co-reporter:Polly L. Arnold, Joy H. Farnaby, Rebecca C. White, Nikolas Kaltsoyannis, Michael G. Gardiner and Jason B. Love
Chemical Science (2010-Present) 2014 - vol. 5(Issue 2) pp:NaN765-765
Publication Date(Web):2013/12/02
DOI:10.1039/C3SC52072B
New, conformationally restricted ThIV and UIV complexes, [ThCl2(L)] and [UI2(L)], of the small-cavity, dipyrrolide, dianionic macrocycle trans-calix[2]benzene[2]pyrrolide (L)2− are reported and are shown to have unusual κ5:κ5 binding in a bent metallocene-type structure. Single-electron reduction of [UI2(L)] affords [UI(THF)(L)] and results in a switch in ligand binding from κ5-pyrrolide to η6-arene sandwich coordination, demonstrating the preference for arene binding by the electron-rich UIII ion. Facile loss of THF from [UI(THF)(L)] further increases the amount of U–arene back donation. [UI(L)] can incorporate a further UIII equivalent, UI3, to form the very unusual dinuclear complex [U2I4(L)] in which the single macrocycle adopts both κ5:κ5 and η6:κ1:η6:κ1 binding modes in the same complex. Hybrid density functional theory calculations carried out to compare the electronic structures and bonding of [UIIII(L)] and [UIII2I4(L)] indicate increased contributions to the covalent bonding in [U2I4(L)] than in [UI(L)], and similar U–arene interactions in both. MO analysis and QTAIM calculations find minimal U–U interaction in [U2I4(L)]. In contrast to the reducible U complex, treatment of [ThCl2(L)] with either a reductant or non-nucleophilic base results in metallation of the aryl rings of the macrocycle to form the (L−2H)4− tetraanion and two new and robust Th–C bonds in the –ate complexes [K(THF)2ThIV(μ-Cl)(L−2H)]2 and K[ThIV{N(SiMe3)2}(L−2H)].
Co-reporter:Polly L. Arnold, Zoë R. Turner, Ronan M. Bellabarba and Robert P. Tooze
Chemical Science (2010-Present) 2011 - vol. 2(Issue 1) pp:NaN79-79
Publication Date(Web):2010/10/28
DOI:10.1039/C0SC00452A
A simple coordination complex of uranium(III), a uranium tris(amide), can selectively couple gaseous CO to the linear ynediolate [OCCO]2− dianion, at room temperature and pressure, regardless of the reagent stoichiometry. This product exhibits further reactivity upon warming in the form of the addition of a C–H bond of a methyl group across the CC triple bond, this second carbon–carbon bond forming reaction generating a functionalised enediolate dianion.
Co-reporter:Stephen T. Liddle, Ian S. Edworthy and Polly L. Arnold
Chemical Society Reviews 2007 - vol. 36(Issue 11) pp:NaN1744-1744
Publication Date(Web):2007/07/03
DOI:10.1039/B611548A
Since the discovery of a stable “bottleable” N-heterocyclic carbene (NHC), there has been a spectacular explosion of interest in this ligand class. This interest stems from a desire to understand the fundamentals of the structure and bonding of these systems, but also because of their numerous and emerging applications in small molecule activation, homogeneous catalysis and Lewis acid-catalysed reactions. In this Tutorial Review, we introduce the reader to NHCs, cover general synthetic methods to prepare anionic tethered NHCs and their metal complexes, and discuss emerging applications in reactivity and catalytic studies.
Co-reporter:Jamie McKinven, Gary S. Nichol and Polly L. Arnold
Dalton Transactions 2014 - vol. 43(Issue 46) pp:NaN17421-17421
Publication Date(Web):2014/10/22
DOI:10.1039/C4DT02995J
Bulky terphenolate ligands allow the synthesis of rare heteroleptic thorium chloride, and borohydride complexes; in the absence of donor solvents, the terphenolate ligands protect the metal ions through neutral Th–η6-arene interactions in a thorium bis(arene) sandwich motif.
Co-reporter:Polly L. Arnold, Nicola L. Bell, Isobel H. Marr, Siyi She, Jonathan Hamilton, Craig Fraser and Kai Wang
Dalton Transactions 2014 - vol. 43(Issue 41) pp:NaN15428-15428
Publication Date(Web):2014/09/04
DOI:10.1039/C4DT01464B
A set of β-ketoimidazolium and β-ketoimidazolinium salts of the general formula [R1C(O)CH2{CH[NCR3CR3N(R2)]}]X (R1 = tBu, naphth; R2 = iPr, Mes, tBu; R3 = H, Me, (H)2; X = Cl, Br) show contrasting reactivity with superhydride bases MHBEt3; two are reduced to chiral β-alcohol carbene–boranes R1CH(OH)CH2{C(BEt3)[NCR3CR3N(R2)]} 2 (R1 = tBu; R2 = iPr, Mes; R3 = H), two with bulky R2 substituents are reduced to chiral β-borate imidazolium salts [R1CH(OBEt3)CH2{CH[NCR3CR3N(R2)]}]X 3 (R1 = tBu, naphth; R2 = Mes, tBu; R3 = H, Me; X = Cl, Br), and the two saturated heterocycle derivatives remain unreduced but form carbene–borane adducts R1C(O)CH2{C(BEt3)[NCR3CR3N(R2)]} 4 (R1 = tBu, naphth; R2 = Mes; R3 = (H)2). Heating solutions of the imidazolium borates 3 results in the elimination of ethane, in the first example of organic borates functioning as Brønsted bases and forming carbene boranes R1CH(OBEt2)CH2{C[NCR3CR3N(R2)]} 5 (R1 = naphth; R2 = Mes; R3 = Me). The ‘abnormal’ carbene borane of the form 2 R1CH(OH)CH2{CH[NC(BEt3)CR3N(R2)]} (R1 = tBu; R2 = tBu; R3 = H), is also accessible by thermolysis of 3, suggesting that the carbene–borane alcohol is a more thermodynamically stable combination than the zwitterionic imidazolium borate. High-temperature thermolysis also can result in complete cleavage of the alcohol arm, eliminating tert-butyloxirane and forming the B–N bound imidazolium borate 7. The strong dependence of reaction products on the steric and electronic properties of each imidazole precursor molecule is discussed.
Co-reporter:Ayodele Oladeji, Polly L. Arnold, Mohammad I. Ali, Slawomir Sujecki, Andrew Phillips, Igor V. Sazanovich and Julia A. Weinstein
Journal of Materials Chemistry A 2013 - vol. 1(Issue 48) pp:NaN8085-8085
Publication Date(Web):2013/10/21
DOI:10.1039/C3TC31601G
Er-doped SiO2 powders were prepared by a simple and straightforward sol–gel method using tetraethoxysilane and erbium triflate as sources of SiO2 and Er2O3. Excitation with near-infrared 800 and 980 nm light initiated an energy up-conversion process in these materials, which resulted in the emission in the predominantly red and green parts of the spectrum. Laser power dependence of the intensity of the up-converted emission was studied both experimentally and by using a numerical model to understand the up-conversion mechanism. Excited state absorption (ESA) and five unique ion–ion energy transfer up-conversion (ETU) processes between the relevant levels are identified as possible mechanisms for the observed luminescence. The efficient NIR-visible upconversion was observed even at comparatively low powers of 200 mW cm−2.
Co-reporter:Michał S. Dutkiewicz, Christos Apostolidis, Olaf Walter and Polly L. Arnold
Chemical Science (2010-Present) 2017 - vol. 8(Issue 4) pp:
Publication Date(Web):
DOI:10.1039/C7SC00034K
Co-reporter:Jordann A. L. Wells, Megan L. Seymour, Markéta Suvova and Polly L. Arnold
Dalton Transactions 2016 - vol. 45(Issue 40) pp:NaN16032-16032
Publication Date(Web):2016/08/26
DOI:10.1039/C6DT02630C
Two lower-oxidation state uranium cations can be readily combined in a robust, yet flexible and derivatisable, tetraaryloxide ligand framework, affording a new platform at which to use the multi-electron reductive capacity of the two actinide centres.
Co-reporter:James R. Pankhurst, Nicola L. Bell, Markus Zegke, Lucy N. Platts, Carlos Alvarez Lamfsus, Laurent Maron, Louise S. Natrajan, Stephen Sproules, Polly L. Arnold and Jason B. Love
Chemical Science (2010-Present) 2017 - vol. 8(Issue 1) pp:NaN116-116
Publication Date(Web):2016/10/28
DOI:10.1039/C6SC02912D
The uranyl(VI) complex UO2Cl(L) of the redox-active, acyclic diimino-dipyrrin anion, L− is reported and its reaction with inner- and outer-sphere reductants studied. Voltammetric, EPR-spectroscopic and X-ray crystallographic studies show that chemical reduction by the outer-sphere reagent CoCp2 initially reduces the ligand to a dipyrrin radical, and imply that a second equivalent of CoCp2 reduces the U(VI) centre to form U(V). Cyclic voltammetry indicates that further outer-sphere reduction to form the putative U(IV) trianion only occurs at strongly cathodic potentials. The initial reduction of the dipyrrin ligand is supported by emission spectra, X-ray crystallography, and DFT; the latter also shows that these outer-sphere reactions are exergonic and proceed through sequential, one-electron steps. Reduction by the inner-sphere reductant [TiCp2Cl]2 is also likely to result in ligand reduction in the first instance but, in contrast to the outer-sphere case, reduction of the uranium centre becomes much more favoured, allowing the formation of a crystallographically characterised, doubly-titanated U(IV) complex. In the case of inner-sphere reduction only, ligand-to-metal electron-transfer is thermodynamically driven by coordination of Lewis-acidic Ti(IV) to the uranyl oxo, and is energetically preferable over the disproportionation of U(V). Overall, the involvement of the redox-active dipyrrin ligand in the reduction chemistry of UO2Cl(L) is inherent to both inner- and outer-sphere reduction mechanisms, providing a new route to accessing a variety of U(VI), U(V), and U(IV) complexes.
Co-reporter:Polly L. Arnold, Charlotte J. Stevens, Nicola L. Bell, Rianne M. Lord, Jonathan M. Goldberg, Gary S. Nichol and Jason B. Love
Chemical Science (2010-Present) 2017 - vol. 8(Issue 5) pp:NaN3617-3617
Publication Date(Web):2017/03/10
DOI:10.1039/C7SC00382J
The first use of a dinuclear UIII/UIII complex in the activation of small molecules is reported. The octadentate Schiff-base pyrrole, anthracene-hinged ‘Pacman’ ligand LA combines two strongly reducing UIII centres and three borohydride ligands in [M(THF)4][{U(BH4)}2(μ-BH4)(LA)(THF)2] 1-M, (M = Li, Na, K). The two borohydride ligands bound to uranium outside the macrocyclic cleft are readily substituted by aryloxide ligands, resulting in a single, weakly-bound, encapsulated endo group 1 metal borohydride bridging the two UIII centres in [{U(OAr)}2(μ-MBH4)(LA)(THF)2] 2-M (OAr = OC6H2tBu3-2,4,6, M = Na, K). X-ray crystallographic analysis shows that, for 2-K, in addition to the endo-BH4 ligand the potassium counter-cation is also incorporated into the cleft through η5-interactions with the pyrrolides instead of extraneous donor solvent. As such, 2-K has a significantly higher solubility in non-polar solvents and a wider U–U separation compared to the ‘ate’ complex 1. The cooperative reducing capability of the two UIII centres now enforced by the large and relatively flexible macrocycle is compared for the two complexes, recognising that the borohydrides can provide additional reducing capability, and that the aryloxide-capped 2-K is constrained to reactions within the cleft. The reaction between 1-Na and S8 affords an insoluble, presumably polymeric paramagnetic complex with bridging uranium sulfides, while that with CS2 results in oxidation of each UIII to the notably high UV oxidation state, forming the unusual trithiocarbonate (CS3)2− as a ligand in [{U(CS3)}2(μ-κ2:κ2-CS3)(LA)] (4). The reaction between 2-K and S8 results in quantitative substitution of the endo-KBH4 by a bridging persulfido (S2)2− group and oxidation of each UIII to UIV, yielding [{U(OAr)}2(μ-κ2:κ2-S2)(LA)] (5). The reaction of 2-K with CS2 affords a thermally unstable adduct which is tentatively assigned as containing a carbon disulfido (CS2)2− ligand bridging the two U centres (6a), but only the mono-bridged sulfido (S)2− complex [{U(OAr)}2(μ-S)(LA)] (6) is isolated. The persulfido complex (5) can also be synthesised from the mono-bridged sulfido complex (6) by the addition of another equivalent of sulfur.
Co-reporter:Polly L. Arnold, Thomas Cadenbach, Isobel H. Marr, Andrew A. Fyfe, Nicola L. Bell, Ronan Bellabarba, Robert P. Tooze and Jason B. Love
Dalton Transactions 2014 - vol. 43(Issue 38) pp:NaN14358-14358
Publication Date(Web):2014/06/11
DOI:10.1039/C4DT01442A
The reactivity of a series of organometallic rare earth and actinide complexes with hemilabile NHC-ligands towards substrates with acidic C–H and N–H bonds is described. The synthesis, characterisation and X-ray structures of the new heteroleptic mono- and bis(NHC) cyclopentadienyl complexes LnCp2(L) 1 (Ln = Sc, Y, Ce; L = alkoxy-tethered carbene [OCMe2CH2(1-C{NCHCHNiPr})]), LnCp(L)2 (Ln = Y) 2, and the homoleptic tetrakis(NHC) complex Th(L)44 are described. The reactivity of these complexes, and of the homoleptic complexes Ln(L)3 (Ln = Sc 3, Ce), with E–H substrates is described, where EH = pyrrole C4H4NH, indole C8H6NH, diphenylacetone Ph2CC(O)Me, terminal alkynes RCCH (R = Me3Si, Ph), and cyclopentadiene C5H6. Complex 1-Y heterolytically cleaves and adds pyrrole and indole N–H across the metal carbene bond, whereas 1-Ce does not, although 3 and 4 form H-bonded adducts. Complexes 1-Y and 1-Sc form adducts with CpH without cleaving the acidic C–H bond, 1-Ce cleaves the Cp–H bond, but 2 reacts to form the very rare H+–[C5H5]−–H+ motif. Complex 1-Ce cleaves alkyne C–H bonds but the products rearrange upon formation, while complex 1-Y cleaves the C–H bond in diphenylacetone forming a product which rearranges to the Y–O bonded enolate product.
Co-reporter:Polly L. Arnold, Dipti Patel, Anne-Frédérique Pécharman, Claire Wilson and Jason B. Love
Dalton Transactions 2010 - vol. 39(Issue 14) pp:NaN3508-3508
Publication Date(Web):2010/03/02
DOI:10.1039/B922115H
The synthesis of the mono-uranyl complex [UO2(THF)(H2LMe)] of a ditopic Schiff-base pyrrole macrocycle is described and is shown to adopt a Pacman wedge-shaped structure in which the uranyl dication is desymmetrised and sits solely in one N4-donor compartment to leave the other vacant. While investigating the mechanism of the previously reported, base-initiated, reductive silylation chemistry of [UO2(THF)(H2LMe)], we found that uranyl hydroxide complexes could be isolated. As such, the reaction between [UO2(THF)(H2LMe)] and KH in THF generated the dimeric cation-cation hydroxide [{UO2(OH)K(C6H6)(H2LMe)}2] when crystallised from C6H6, or alternatively, when crystallised from THF, the monomeric THF-adducted cation-cation complex [UO2(OH)K(THF)2(H2LMe)] was isolated. These compounds result formally from the substitution of the equatorial THF molecule by hydroxide, and it was also shown that the reaction between dry KOH and [UO2(THF)(H2LMe)] generated [{UO2(OH)K(C6H6)(H2LMe)}2].
Co-reporter:Markus Zegke, Gary S. Nichol, Polly L. Arnold and Jason B. Love
Chemical Communications 2015 - vol. 51(Issue 27) pp:NaN5879-5879
Publication Date(Web):2015/03/02
DOI:10.1039/C5CC00867K
Reactions between the uranyl(VI) Pacman complex [(UO2)(py)(H2L)] of the Schiff-base polypyrrolic macrocycle L and Tebbe's reagent or DIBAL result in the first selective reductive functionalisation of the uranyl oxo by Al to form [(py)(R2AlOUO)(py)(H2L)] (R = Me or iBu). The clean displacement of the oxo-coordinated Al(III) by Group 1 cations has enabled the development of a one-pot, DIBAL-catalysed reduction of the U(VI) uranyl complexes to a series of new, mono-oxo alkali-metal-functionalised uranyl(V) complexes [(py)3(MOUO)(py)(H2L)] (M = Li, Na, K).
Co-reporter:Polly L. Arnold, Guy M. Jones, Qing-Jiang Pan, Georg Schreckenbach and Jason B. Love
Dalton Transactions 2012 - vol. 41(Issue 22) pp:NaN6597-6597
Publication Date(Web):2012/04/30
DOI:10.1039/C2DT30658A
Expansion of a Schiff-base polypyrrolic macrocycle allows the formation of a binuclear uranyl complex with co-linear uranyl ions and a very short oxo–oxo distance.
Co-reporter:Stephen M. Mansell, Fanny Bonnet, Marc Visseaux and Polly L. Arnold
Dalton Transactions 2013 - vol. 42(Issue 25) pp:NaN9039-9039
Publication Date(Web):2013/04/10
DOI:10.1039/C3DT50131K
The synthesis and structural characterisation of the uranium(IV) amido-borohydrides (N′′)2U{κ2-N(SiMe3)SiMe2CH2BBN-H} and U{κ2-N(SiMe3)SiMe2CH2BBN-H}2, and their activity as pre-catalysts for the polymerisation of isoprene are described.
Co-reporter:Polly L. Arnold, Zöe R. Turner, Anne I. Germeroth, Ian J. Casely, Gary S. Nichol, Ronan Bellabarba and Robert P. Tooze
Dalton Transactions 2013 - vol. 42(Issue 5) pp:NaN1337-1337
Publication Date(Web):2012/11/27
DOI:10.1039/C2DT31698F
The reactions of f-block silylamido N-heterocyclic carbene (NHC) complexes ([M(L)(N{SiMe3}2)2], M = Y, Ce, and U, L = bidentate alkoxy-tethered NHC ligand) with CO and CO2 have been studied and compared to each other, to those of selected [M(L)2(N{SiMe3}2)] complexes, and to those of [M(N{SiMe3}2)3] to identify the effect of the labile NHC group on the small molecule activation chemistry. The small molecules COS and N2CPh2 have also been studied.