royal society of chemistry coordination and organometallic

101
Royal Society of Chemistry Coordination and Organometallic Chemistry Discussion Group Meeting St. Anne’s College, Oxford 3 rd –4 th September 2015

Upload: lamcong

Post on 14-Feb-2017

234 views

Category:

Documents


6 download

TRANSCRIPT

Page 1: Royal Society of Chemistry Coordination and Organometallic

Royal Society of Chemistry

Coordination and Organometallic

Chemistry Discussion Group Meeting

St. Anne’s College, Oxford

3rd–4th September 2015

Page 2: Royal Society of Chemistry Coordination and Organometallic

The organisers acknowledge the generous support of

the following sponsors:

Page 3: Royal Society of Chemistry Coordination and Organometallic

Practical Information

Venue

The lecture programme will take place in Mary Ogilvie Lecture Theatre at St Anne’s.

Lunch around St. Annes’s

About 5 minutes walk away …

Taylors Oxford – Sandwich shop – 1 Woodstock Road

Maison Blanc – French baker, patisserie cafe – 3 Woodstock Road

The Royal Oak – Classic British pub – 42-44 Woodstock Road

Will’s Deli – A restaurant offering a range of salads, hot dishes and cheeses – 15

Woodstock Road

About 10 minutes walk away …

The Eagle and Child – Traditional british pub, best known for being a watering hole for

CS Lewis and JRR Tolkien – 49 St. Giles

St Giles cafe – Cafe with simple British menu – 52 St Giles

The Jericho Cafe –Neighourhood cafe with locally sourced blackboard menu – 112

Walton Street

The Victoria – Traditional tavern with real ales plus simple pub food – 90 Walton Street

Taxis

From Oxford city centre to St. Anne’s or from Oxford train station to St. Anne’s, the

cost is approximately £5.

ABC Taxis: 01865 770077 / 01865 775577

Botley Taxis: 01865 423264 / 07866 423264

001 Taxis: 01865 240000

Page 4: Royal Society of Chemistry Coordination and Organometallic
Page 5: Royal Society of Chemistry Coordination and Organometallic

PROGRAMME RSC Coordination and Organometallic Chemistry

Discussion Group Meeting

St. Anne’s College, Oxford; 3rd–4th September 2015

Thursday 3rd of September

10.30–12.00 pm Registration and setting up of posters

Arrival at St. Anne’s and registration with tea and coffee

Lunch (not provided): 12.00–12.50 pm

Session 1: 12.50–3.00 pm

12.50–1.00 pm: Introductory remarks (Jose M. Goicoechea/Simon Aldridge/Andrew

Weller)

1.00–1.45 pm: PL1: Neil Champness, University of Nottingham

“From frameworks to spheres and surfaces – Exploiting the

coordination bond”

1.45–2.20 pm: IL1: Mark Crimmin, Imperial College London

“Transforming carbon–fluorine bonds into carbon–aluminium bonds”

2.20–2.40 pm: George Kostakis, University of Sussex

“Custom-designed heteronuclear 3d/Dy(III) coordination clusters as

catalysts in a domino reaction”

2.40–3.00 pm: Peter Portius, University of Sheffield

“Nitrogen-rich coordination compounds of p-block elements

Coffee break: 3.00–3.30 pm

Session 2: 3.30–5.30 pm

3.30–4.15 pm: PL2: Joshua Figueroa, University of California San Diego

“Structural dynamics and substrate activation in highly congested π-

Acidic ligand fields”

4.15–4.50 pm: IL2: Simon Pope, University of Cardiff

“Fluorophore functionalised metal complexes”

4.50–5.10 pm: Rhiann Andrew, University of Warwick

“Synthesis, reactivity and dynamic behaviour of NHC-based Rhodium

macrocycles”

5.10–5.30 pm: George Britovsek, Imperial College London

“Towards robust alkane oxidation catalysts using non-heme iron(II)

complexes”

Page 6: Royal Society of Chemistry Coordination and Organometallic

5.30–6.00 pm: Meeting of RSC CODG members

Poster session: 5.30–7.00 pm

Pre-dinner drinks and dinner 7.00 pm onwards

Friday 4th of September

Session 3: 9.00–10.40 am

9.00–9.20 am: Tom Hooper, University of Oxford

“Phosphine-borane dehydrocoupling using a {Cp*Rh(PR3)}2+ fragment:

Stoichiometric reactivity and catalysis”

9.20–9.40 am: Rebecca Musgrave, University of Bristol

“Reversible formation of magnetic polynickelocenes”

9.40–10.00 am: Sergey Zlatogorsky, University of Central Lancashire

“Towards metal carbenes as hydrogenase models”

10.00–10.20 am: Nicola Bell, University of Edinburgh

“Taking uranyl Pacman chemistry to the next level”

10.20–10.40 am: Nathan Patmore, University of Huddersfield

“Mechanistic studies of charge transfer in hydrogen bonded ‘dimers of

dimers’”

Coffee break: 10.40–11.10 am

Session 4: 11.10–12.40 pm

11.10–11.20 am: Katie Dryden-Holt, RSC. Presentation on opportunities for younger

members

11.20–11.25 am: Presentation of Sir Edward Frankland Fellowship to Scott Dalgarno

(Richard Walker; RSC)

11.25–12.00 am: IL3: Scott Dalgarno, Heriot-Watt University

“Calixarene-supported clusters”

12.00–12.20 pm: Lee Collins, University of Bath

“eNHChanting new reactivity with Cu and Zn”

12.20–12.40 pm: Phil Dyer, Durham University

“Exploring phosphine-alkene ligand chemistry”

Lunch (not provided): 12.40–1.40 pm

Page 7: Royal Society of Chemistry Coordination and Organometallic

Session 5: 1.40–4.00 pm

1.40–2.00 pm: Introduction by Philip Mountford

Presentation of the RSC Award for Service to Peter Tasker

Introduction of Jason Love for the Dalton Transactions Lecture

2.00–2.45 pm: Dalton Lecture: Jason Love, University of Edinburgh

“An adventure in ligand design: From Pacman to dynamic assemblies

via uranyl”

2.45–3.20 pm: IL4: Kylie Vincent, University of Oxford

“Proton transfer from a metal hydride: Insight from Infrared

spectroscopy applied to nickel-iron hydrogenases under

electrocatalytic turnover”

3.20–3.40 pm: Matthew Blake, University of Oxford

“Group 2- and group 12-metal complexes”

3.40–4.00 pm: Closing remarks

Page 8: Royal Society of Chemistry Coordination and Organometallic

Abstracts for Plenary Lectures

Page 9: Royal Society of Chemistry Coordination and Organometallic

FROM FRAMEWORKS TO SPHERES AND SURFACES –

EXPLOITING THE COORDINATION BOND

N.R. Champness

School of Chemistry, University of Nottingham, University Park, Nottingham, UK.

[email protected]

Non-covalent directional intermolecular interactions provide a pre-determined

recognition pathway which has been widely exploited in supramolecular chemistry to

form functional nanostructures. Our results using coordination bonding interactions to

enable the directed assembly of extended nanostructures will be presented. The lecture

will focus on the development of coordination polymers and metal-organic frameworks

(MOFs) but with an emphasis on the construction of spherical coordination

architectures and their subsequent organisation into larger arrays, notably on surfaces.

The lecture will span three themes: the use of MOFs to host photoactive species

modifying the properties of the incorporated species1,2 (Fig. i); the development of

nanoscale spheres using coordination interactions3 (Fig. ii); and, lastly, the organisation

of magnetic molecules on surfaces4,5 (Fig. iii). In summary the lecture will draw a direct

connection between supramolecular and coordination chemistry and nanostructure

fabrication.

References.

[1] Blake, A.J.; Champness, N.R.; Easun, T.L.; Allan, D.R.; Nowell, H.; George, M.W.;

Jia, J.; Sun, X-Z. Nat. Chem., 2010, 2, 688.

[2] Easun, T.L.; Jia, J.; Reade, T.J.; Sun, X-Z.; Davies, E.S.; Blake, A.J.; George,

M.W.; Champness, N.R. Chem. Sci., 2014, 5, 539.

[3] Argent, S.P.; Greenaway, A.; Gimenez-Lopez, M.C.; Lewis, W.; Nowell, H.;

Khlobystov, A.N.; Blake, A.J.; Champness, N.R.; Schröder, M. J. Am. Chem. Soc.,

2012, 134, 55.

[4] Slater, A.G.; Perdigao, L.M.A.; Beton, P.H.; Champness, N.R. Acc. Chem. Res.,

2014, 47, 3417.

[5] Saywell, A.; Magnano, G.; Satterley, C.J.; Perdigão, L.M.A.; Britton, A.J.; Taleb,

N.; Giménez-López, M.C.; Champness, N.R.; O’Shea, J.N.; Beton, P.H. Nat.

Commun., 2010, 1, 75.

Page 10: Royal Society of Chemistry Coordination and Organometallic

STRUCTURAL DYNAMICS AND SUBSTRATE ACTIVATION IN

HIGHLY CONGESTED -ACIDIC LIGAND FIELDS

Joshua S. Figueroa

Department of Chemistry and Biochemistry

University of California, San Diego

9500 Gilman Drive, MC 0358

La Jolla, CA 92093 (USA)

[email protected]

Recent results concerning the construction and reactivity of low-coordinate transition-

metal complexes featuring sterically encumbering m-terphenyl isocyanide ligands are

presented. Given the isolobal relationship between organoisocyanides and carbon

monoxide, these complexes serve as mimics of the unsaturated binary metal carbonyls.

The latter have traditionally been studied in either the gas phase or by matrix-isolation

techniques and, consequently, their condensed-phase reactivity patterns are significantly

underexplored. Specifically addressed will be synthetic studies that have delivered

homoleptic isocyanide complexes that model binary carbonyls of the mid-to-late

transition metals. In addition, coordinatively unsaturated heteroleptic isocyanide

complexes are presented that mimic the structural and spectroscopic properties of

reactive intermediates in catalytic processes mediated by transition-metal carbonyls. As

expected for mimics such reactive carbonyl complexes, the newly prepared isocyanide

complexes display rich reactivity patterns. Novel aspects of small-molecule

coordination chemistry and activation, especially as a function of structural and ligand

dynamics, by these isocyanide complexes are discussed.

Page 11: Royal Society of Chemistry Coordination and Organometallic

AN ADVENTURE IN LIGAND DESIGN: FROM PACMAN TO

DYNAMIC ASSEMBLIES VIA URANYL

Jason B. Love

EaStCHEM School of Chemistry, University of Edinburgh, Edinburgh EH9 3FJ, UK

Sustainability in the chemical sciences, especially in making best use of our critical

energy and material resources, is a recurring and global challenge. In this contribution,

the use of ligand design, its application to molecular coordination chemistry, catalysis,

and supramolecular chemistry, and the insight gained from these themes regarding

chemical sustainability will be described. This will include (i) the use of Pacman

complexes as platforms for redox catalysis, in particular oxygen reduction and C-H

bond oxygenation,1 (ii) the development of the f-element chemistry of these ligands to

facilitate the reduction and oxo-functionalisation of the uranyl dication,2 and (iii) the

spontaneous assembly of supramolecular ion pairs to form catalysts for oxidation and

reduction reactions under biphasic conditions.3

References

1 A. M. J. Devoille and J. B. Love, "Double-pillared cobalt Pacman complexes:

synthesis, structures and oxygen reduction catalysis." Dalton Trans., 2012, 41, 65-72; J.

R. Pankhurst, T. Cadenbach, D. Betz, C. Finn and J. B. Love, Dalton Trans., "Towards

dipyrrins: oxidation and metalation of acyclic and macrocyclic Schiff-base

dipyrromethanes," Dalton Trans., 2015, 44, 2066-2070.

2 P. L. Arnold, D. Patel, C. Wilson and J. B. Love, "Reduction and selective oxo

group silylation of the uranyl dication." Nature, 2008, 451, 315-317; P. L. Arnold, R.

M. Lord, A-F. Pécharman, G. M. Jones, E. Hollis, G. S. Nichol, L. Maron, J. Fang, T.

Davin, J. B. Love, "Control of oxo-group functionalization and reduction of the uranyl

ion," Inorg. Chem., 2015, 54, 3702-3710

3 M. Cokoja, I. I. E. Markovits, M. H. Anthofer, S. Poplata, A. Pöthig, D. S.

Morris, P. A. Tasker, W. A. Herrmann, F. E. Kühn and J. B. Love, "Catalytic

epoxidation by perrhenate through the formation of organic-phase supramolecular ion

pairs," Chem. Commun., 2015, 51, 3399-3402; A. M. Wilson, P. J. Bailey, P. A. Tasker,

J. R. Turkington, R. A. Grant, J. B. Love, "Solvent extraction: the coordination

chemistry behind extractive metallurgy." Chem. Soc. Rev., 2014, 43, 123-134

Page 12: Royal Society of Chemistry Coordination and Organometallic

Abstracts for Invited Lectures

Page 13: Royal Society of Chemistry Coordination and Organometallic

TRANSFORMING CARBON–FLUORINE BONDS INTO

CARBON–ALUMINIUM BONDS

Olga Ekkert, Adi Nako, Shuhui Yow and Mark R. Crimmin

Department of Chemistry, Imperial College, South Kensington, London, SW7 2AZ, UK

[email protected]

We report the activation of carbon–fluorine bonds using the aluminium dihydride 1. In

the presence of a transition metal catalyst, 1 is an effective stoichiometric reagent for

two competitive reactions that break strong C–F bonds in fluoroarenes. These include

the the transformation of a C–F bond to a C–Al bond (Figure – C–F Alumination) and

conversion of a C–F bond to a C–H bond (Figure – Hydrodefluorination).1 The former

reaction can be conceptualised in terms of a net elimination of H2 from and net addition

of the C–F bond to the group 13 element.

Through investigating the coordination chemistry,2-3 we show that the generation of an

Al(I) fragment from the dehydrogenation of 1 is possible. Control reactions in which an

Al(I) analogue of 1 is reacted with fluoroarenes and fluoroalkanes give the products of

C–F bond cleavage by oxidative addition. Our findings raise the possibility of catalytic

control in the redox chemistry of main group reagents.

[1] (a) Yow S.; Gates, S. J.; White, A. J. P.; Crimmin, M. R. Angew. Chem., Int. Ed.

2012, 51, 12559. (b) Ekkert, O.; Strudley, S. D. A.; Rozenfeld, A.; White, A. J. P.;

Crimmin, M. R. Organometallics 2014, 33, 7027.

[2] Abdalla, J. A. B.; Riddlestone, I. M.; Turner, J.; Kaufman, P. A.; Tirfoin, R.;

Phillips, N.; Aldridge, S. Chem. Eur. J., 2014, 20, 17624.

[3] (a) Ekkert, O.; Toms, H.; White, A. J. P.; Crimmin, M. R. Chem. Sci. 2015, DOI:

10.1039/C5SC01309G. (b) Nako, A. E.; Tan, Q. W.; White, A. J. P.; Crimmin, M. R.

Organometallics, 2014, 33, 2685.

Page 14: Royal Society of Chemistry Coordination and Organometallic

FLUOROPHORE FUNCTIONALISED METAL COMPLEXES

Simon J. A. Pope

School of Chemistry, Cardiff University, Main Building, Park Place, Cardiff

[email protected]

Luminescent metal ion coordination complexes, including organometallic variants, have

shown great utility in their application, from optoelectronic devices to biological

studies. In a biological context, application can vary from cellular imaging agents, to the

development of multimodal imaging agents, to visualising complexes of therapeutic

interest and studying metal-ligand interactions with DNA and/or proteins.

The focus of this presentation will be the development of fluorescent ligands that

can facilitate the optical bioimaging application of the corresponding complexes.

Discussion will include the design and

synthesis of ligands based upon functionalised

anthraquinone and 1,8-naphthalimide species.

Both anthraquinone and 1,8-naphthalimide

species can possess advantageous fluorescent

properties that are tuneable via the differing

substitution patterns of the aromatic cores.

The ligand design and coordination chemistry

with gold(I) [1] and rhenium(I) [2] will be a

primary focus of the presentation.

Our studies encompass the structural, spectroscopic and photophysical

characterisation of all ligands and complexes. These species were also investigated as

potential bioimaging agents, using confocal fluorescence microscopy, with studies

including cancer cell lines, human osteoarthritic cells and the protistan fish parasite

Spironucleus vortens. The results show that both anthraquinone and 1,8-naphthalimide

based complexes can facilitate cell imaging studies. More specifically, the nature of the

1,8-naphthalimide structure serves to influence the cytotoxicity, uptake and intracellular

localization of these imaging agents. In the case of the gold(I) complexes, the

cytotoxicity was significantly enhanced versus the corresponding free ligands.

References

[1] Langdon-Jones, E.E.; Lloyd, D.; Hayes, A.J.; Wainwright, S.D.; Mottram, H.J.;

Coles, S.J.; Horton, P.N.; Pope, S.J.A. Inorg. Chem. 2015, 54, 6606; Langdon-Jones,

E.E.; Pope, S.J.A. Chem. Commun. 2014, 50, 10343; Balasingham, R.G.; Williams,

C.F.; Mottram, H.J.; Coogan, M.P.; Pope, S.J.A. Organometallics 2012, 31, 5835.

[2] Langdon-Jones, E.E.; Symonds, N.O.; Yates, S.E.; Hayes, A.J.; Lloyd, D.; Williams,

R.; Coles, S.J.; Horton, P.N.; Pope, S.J.A. Inorg. Chem. 2014, 53, 3788,

Page 15: Royal Society of Chemistry Coordination and Organometallic

CALIXARENE-SUPPORTED CLUSTERS

Scott J. Dalgarno

Rm 2.04 William Perkin Building, Institute of Chemical Sciences, Heriot-Watt

University, Riccarton, Edinburgh, EH14 4AS

[email protected]

Methylene-bridged calix[n]arenes display markedly different coordination chemistry to

their thia-, sulfonyl- and sulfinyl-bridged analogues, and have recently emerged as

excellent ligand supports for the synthesis of polynuclear transition, lanthanide metal

and 3d-4f clusters. The presentation will cover the formation of a series of cluster motifs

including MnIII2MnII

2(Calix[4])2 SMMs,1 structurally related MnIII2MnIILnIII(Calix[4])2

and MnIII2LnIII

2(Calix[4])2 species (Figure 1),2 MnIII4LnIII

4(Calix[4])4 SMMs and

magnetic refrigerants,3 tri-capped trigonal prismatic enneanuclear CuII9(Calix[4])3 and

octahedral LnIII6(Calix[4])2 clusters. Metal ion binding rules arising from experiments

with calix[4 and 8]arenes will be discussed as they allow one to control metal ion

incorporation depending on the building block employed. New developments in related

bis-calix[4]arene cluster chemistry will also be presented as these a) display structural

characteristics of calix[4]arene and b) allow one to employ cluster formation in directed

assembly.8,9

Figure 1. Series of structurally related calix[4]arene-supported Mn , Mn / Ln and Ln

clusters formed under various conditions.

References

[1] Karotsis, G.; Teat, S. J.; Wernsdorfer, W.; Piligkos, S.; Dalgarno, S. J.; Brechin, E.

K. Angew. Chem. Int. Ed. 2009, 48, 8285.

[2] Palacios, M. A.; McLellan, R.; Taylor, S. M.; Beavers, C. M.; Teat, S. J.;

Wernsdorfer, W.; Piligkos, S.; Dalgarno, S. J.; Brechin, E. K. Chem. Eur. J. 2015, 21,

11212.

[3] Karotsis, G.; Evangelisti, M.; Dalgarno, S. J.; Brechin, E. K. Angew. Chem. Int. Ed.

2009, 48, 9928.

[4] McLellan, R.; Palacios, M. A.; Beavers, C. M.; Teat, S. J.; Piligkos, S.; Brechin, E.

K.; Dalgarno, S. J. Chem. Eur. J. 2015, 21, 2804.

[5] Coletta, M.; McLellan, R.; Murphy, P.; Leube, B. T.; Sanz, S.; Clowes, R.; Gagnon,

K. J.; Teat, S. J.; Cooper, A. I.; Paterson, M. J.; Brechin, E. K.; Dalgarno, S. J.

submitted.

Page 16: Royal Society of Chemistry Coordination and Organometallic

PROTON TRANSFER FROM A METAL HYDRIDE: INSIGHT

FROM INFRARED SPECTROSCOPY APPLIED TO NICKEL-

IRON HYDROGENASES UNDER ELECTROCATALYTIC

TURNOVER

R. Hidalgo, P.A. Ash and K.A. Vincent

University of Oxford, Department of Chemistry, Inorganic Chemistry Laboratory, South

Parks Road, Oxford, OX1 3QR, UK

[email protected]

Rapid oxidation of H2 is catalysed at a NiFe catalytic site in the family of nickel-iron

hydrogenase enzymes. The reaction is controlled readily using electrochemistry, where

the electrode provides a sink for fast uptake of the electrons released during H2

oxidation by a film of immobilised hydrogenase. We have made use of a new approach

to infrared (IR) spectroscopy coupled with electrochemistry to study a NiFe

hydrogenase under catalytic turnover and gain insight into the catalytic cycle for H2

activation.[1] The native CO and CN– ligands on Fe at the active site of the enzyme

provide sensitive spectroscopic probes for identifying redox and coordination changes at

the active site.

It is established that H2 reacts with the NiFe hydrogenase active site at the NiII FeII

level. A NiIII(H–)FeII intermediate is known, but the steps required to remove the

hydrido ligand to restore the starting NiII FeII state have remained obscure. A NiI FeII

state, resulting from initial transfer of a proton from the active site to a nearby base, has

been proposed in computational studies,[2,3] but has not previously been observed

experimentally under conditions relevant to catalysis. Here we provide evidence from

IR spectroelectrochemical studies on hydrogenase I from E. coli, under turnover and

non-turnover conditions, that the NiIII(H–)FeII level does indeed give rise to a detectable

NiI FeII intermediate rather than requiring concerted removal of a proton and

electron.[1,4]

The mechanistic insight gained from these studies of the hydrogenase active site during

fast catalytic turnover should help to inspire development of functional biomimetic

catalysts for H2 activation at non-noble metal sites.

References.

[1] Hidalgo, R.; Ash, P.A.; Healy, A.J.; Vincent, K.A. Angew. Chemie Int. Ed. 2015,

127, 7216.

[2] Pardo, A.; De Lacey, A. L.; Fernández, V. M.; Fan, H.-J.; Fan, Y.; Hall, M. B. J.

Biol. Inorg. Chem. 2006, 11, 286.

[3] Lill, S. O. N.; Siegbahn, P. E. M. Biochemistry 2009, 48, 1056.

[4] Murphy, B.J.; Hidalgo, R.; Roessler, M.M.; Evans, R.M.; Ash, P.A.; Myers, W.K.;

Vincent, K.A.; Armstrong, F.A. J. Am. Chem. Soc. 2015, 137, 8484.

Page 17: Royal Society of Chemistry Coordination and Organometallic

Abstracts for Oral Presentations

Page 18: Royal Society of Chemistry Coordination and Organometallic

CUSTOM-DESIGNED HETERONUCLEAR 3D/DY(III)

COORDINATION CLUSTERS AS CATALYSTS IN A DOMINO

REACTION

Kieran Griffiths, Christopher W. D. Gallop, Alaa Abdul-Sada, Alfredo Vargas, Oscar

Navarro and George E. Kostakis

School of Life Sciences, University of Sussex, Brighton BN1 9QJ, United Kingdom [email protected]

The synthesis of polynuclear Coordination Clusters (CCs) containing 3d and

Dy(III) metal ions is currently extremely popular and a research objective for several

groups due to their aesthetically pleasant structures and potential applications.1

The present work is directed on synthesizing tetranuclear 3d/Dy(III) CCs,

possessing a defect dicubane topology, with Schiff base ligands and their functional

uses, for the first time in catalysis. Thus, three isoskeletal tetranuclear CCs with general

formula [MII

2DyIII

2L4(EtOH)6](ClO4)2·2(EtOH), (M = Co, 1; M= Ni, 2) and

[NiII

2DyIII

2L4Cl2(CH3CN)2]·2(CH3CN) (3), have been synthesized and characterized.

These air-stable compounds, and in particular 3, display efficient homogeneous catalytic

behavior towards the room temperature synthesis of trans-4,5-diaminocyclopent-2-

enones from 2-furaldehyde (Scheme 1) and primary or secondary amines under a non-

inert atmosphere.2

Scheme 1 The proposed domino reaction catalyzed by compounds 1 – 3.

This work signifies a unique contribution to 3d/4f coordination chemistry and

opens new research directions by producing a new generation for 3d/Dy(III) molecules

that can act in tandem in catalysis.

References.

(1) Sessoli, R.; Powell, A. K. Coord. Chem. Rev. 2009, 253, 2328–2341.

(2) Griffiths, K.; Gallop, C. W. D.; Abdul-Sada, A.; Vargas, A.; Navarro, O.;

Kostakis, G. E. Chem. Eur. J. 2015, 21, 6358–6361.

Page 19: Royal Society of Chemistry Coordination and Organometallic

NITROGEN-RICH COORDINATION COMPOUNDS OF P-BLOCK

ELEMENTS

R. M. Campbell, B. Peerless, B. F. Crozier and P. Portius

Department of Chemistry, The University of Sheffield, Brook Hill, S3 7HF, UK [email protected]

N-rich compounds are a fascinating and promising replacement for conventional propel-lants, explosives & pyrotechnics. They generate mainly N2 upon decomposition - ideal for smokeless, CO2-free and “green” energetic materials.1 However, low barriers towards decomposition pose major challenges in their preparation, isolation, and

characterisation. Our research aims at stabilising and controlling the reactivity of N-rich coordination compounds. Specifically, the presentation will focus on E(Y)n, n = 2-6, Y = N-rich ligand; E = p-block coord. centre in low or high ox. states. Applying synthetic approaches novel to energetic chemistry, involving combinations of hypercoordination, bulky counter ions, and ligand exchanges, the preparation and isolation of several new classes of N-rich tetrazolato and azido complexes has been achieved as candidates for efficient & controllable energy storage.5 These include neutral Lewis base adducts and homoleptic complexes (a),2-4, 6,7 covalent, binary E(N3)n azides which were generated in bulk for the first time (b),4 the first homoleptic tetrazolato p-block element complexes

E(CHN4)62 and E(CHN4)3

(E = Si, Ge, Sn, Fig. 1) (c). The new complexes have a unique chemistry – reactions with nitriles and phosphines afford unusual poly(tetrazolato) and poly(phosphiniminato) complexes. Syntheses, reactivity, structures, and thermal and spectroscopic properties of the new species, including application of time-resolved infrared spectroscopy to study photoreactivity, will be described.

[1] Steinhauser, G.; Klapoetke, T. M. Angew. Chem. Int. Ed. 2008, 2. [2] Portius, P.; Fowler, P. W.; et al. Inorg. Chem. 2008, 12004. [4] Portius, P.; Filippou, A. C.; et al. Angew. Chem. Int. Ed. 2010, 8013. [5] Davis, M.; Portius, P. Coord. Chem. Rev. 2013, 1011. [6] Peerless, B.; Keane, T.; et al. Chem. Commun. 2015, 7435. [7] Campbell, R. M.; Davis, M. F.; Fazakerley, M.; Portius, P. paper in preparation 2015.

Fig. First examples of tetrazol-based N-rich coord. networks and complexes with p-block centres.

Page 20: Royal Society of Chemistry Coordination and Organometallic

SYNTHESIS, REACTIVITY AND DYNAMIC BEHAVIOUR OF

NHC-BASED RHODIUM MACROCYCLES

Rhiann E. Andrew,a Dominic W. Ferdani,

a C. André Ohlin

b and Adrian B. Chaplin

a,*

a Department of Chemistry, University of Warwick, Coventry CV4 7AL, UK

b School of Chemistry, Monash University, Clayton, Victoria 3800, Australia

[email protected]

Pincer ligand architectures featuring N-heterocyclic carbene (NHC) donors are

increasingly prominent in organometallic chemistry, combining the strong σ-donor

characteristics of NHCs with the favourable thermal stability and reaction control

possible with a mer-tridentate geometry.1 With a view to exploiting their unique steric

profile, we have recently become engaged in the investigation of macrocyclic variants

of these NHC pincers.2 In this report we present the synthesis, fluxional behaviour and

reactivity of rhodium complexes containing macrocyclic pincer ligands.3 A general

synthetic procedure involving ring-closing olefin metathesis enabled access to rhodium

systems incorporating both lutidine (1a, n = 1) and pyridine (1b, n = 0) central donor

groups. For comparison, a bidentate xylene bridged macrocyclic system (1c) was also

prepared.

In the presence of CO, 1a exhibits coordination-induced atropisomerisation akin to that

reported for its acyclic analogues.4 Through computational modelling and simulation-

aided analysis of variable temperature 1H NMR spectra, we deduced that

interconversion occurs through a symmetric intermediate similar to 1c. The absence of

bridging methylene groups makes for a more rigid architecture (1b) that cannot

coordinate excess CO or dissociate pyridine, but more readily undergoes oxidative

addition – in comparison to 1a – to provide isolable rhodium(III) complexes.

References

1. Poyatos, M.; Mata, J. A.; Peris, E. Chem. Rev. 2009, 109, 3677– 3707.

2. Andrew, R. E.; Chaplin, A. B. Dalton Trans. 2014, 43, 1413– 1423.

3. a) Andrew, R. E.; Chaplin, A. B. Inorg. Chem. 2015, 54, 312– 322; b) Andrew, R.

E.; Ferdani, D. W.; Ohlin, C. A.; Chaplin, A. B. Organometallics 2015, 34, 913–917.

4. Miecznikowski, J. R.; Grundemann, S.; Albrecht, M; Megret, C; Clot, E.; Faller, J.

W.; Eisenstein, O.; Crabtree, R. H. Dalton Trans. 2003, 5, 831 – 838.

Page 21: Royal Society of Chemistry Coordination and Organometallic

TOWARDS ROBUST ALKANE OXIDATION CATALYSTS USING NON-HEME IRON(II) COMPLEXES

Michaela Grau and George Britovsek

Department of Chemistry, Imperial College London, Exhibition Road, London, SW72AY

[email protected]

The selective oxidation of organic molecules under energy efficient, cheap, non toxic

and environmentally friendly conditions is an important target for the chemical industry

and organic synthesis on laboratory scale. In nature there are metalloenzymes such as

cytochrome P450 or methane monooxygenase, which feature iron complexes at the

active site and perform excellent regio- and stereoselective oxidation catalysis.[1]

Inspired by the reactivity and selectivity of metalloenzymes, a large variety of non-

heme iron(II) complexes featuring linear, tripodal or cyclic tetradentate N-donor ligands

have been developed.[2]

A key finding has been that catalysts with tetradentate ligands

tend to give the most active catalysts. However, rapid degradation of most catalysts

under the operating conditions presents a major problem.[3]

N

N

N

N N

N

N N

N N

N

N N

N N

X

L3 X = H

L4 X = Cl

L1 L2

To establish whether catalyst stability is affected by an increase in ligand denticity,

we present here our results on mononuclear non-heme iron(II) complexes containing

linear pentadentate ligands (L1-4). The reactivity of these complexes as catalysts for the

oxidation of cyclohexane with hydrogen peroxide has been investigated. Analysis of

the different complexes revealed an interesting and unusual change in the coordination

geometry from octahedral to pentagonal bipyramidal. Furthermore, in solution it was

found that the pentagonal bipyramidal complex changes its coordination geometry with

temperature resulting in a novel temperature-dependent coordination equilibrium.[4]

References [1] B. Meunier, S. P. de Visser, S. Shaik, Chem. Rev. 2004, 104, 3947-3980.

[2] M. Grau, G. J. P. Britovsek, Top. Organomet. Chem. 2015, 50, 145-172.

[3] a) J. England, C. R. Davis, M. Banaru, A. J. P. White, G. J. P. Britovsek, Adv.

Synth. Catal. 2008, 350, 883-897; b) J. England, R. Gondhia, L. Bigorra-Lopez,

A. R. Petersen, A. J. P. White, G. J. P. Britovsek, Dalton Trans. 2009, 5319-

5334; c) M. Grau, A. Kyriacou, F. Cabedo Martinez, I. de Wispelaere, A. J. P.

White, G. J. P. Britovsek, Dalton Trans. 2014, 43, 17108-17119.

[4] M. Grau, J. England, R. Torres Martin de Rosales, H. S. Rzepa, A. J. P. White,

G. J. P. Britovsek, Inorg. Chem. 2013, 52, 11867-11874.

Page 22: Royal Society of Chemistry Coordination and Organometallic

PHOSPHINE-BORANE DEHYDROCOUPLING USING A

{Cp*Rh(PR3)}2+

FRAGMENT: STOICHIOMETRIC REACTIVITY

AND CATALYSIS

Thomas N. Hooper and Andrew S. Weller*

University of Oxford, Department of Chemistry, Chemistry Research Laboratory, 12 Mansfield Road, Oxford, OX1 3TA

[email protected]

Phosphinoborane polymers are valence isoelectronic with polyolefins and have been

shown to have applications in lithography and as preceramic polymers for

semiconductors.1 Manners et al. reported the homogeneous catalytic synthesis of

polyphosphinoboranes using Rh(I) and Rh(III) precatalysts, but the nature of the active

species and the mechanism was not determined.1 We present a new system based on a

{Cp*Rh(PR3)}2+

fragment which allows for analysis of intermediates and mechanism

through stoichiometric reactivity and catalytic production of dehydrocoupled products.

H3B·PH2Ph was catalytically dehydrocoupled in toluene solution using

[Cp*RhMe(PMe3)(CH2Cl2)][BArF

4] (100 °C, 1 mol%) to produce well-defined

polymeric material. Under similar conditions at 5 mol%, the secondary phosphine-

borane H3B·PHPh2 was dehydrocoupled to form a linear dimer and temporal

interrogation allowed the organometallic species to be identified.

Stoichiometric reaction of secondary phosphine-boranes with

[Cp*RhMe(PMe3)(CH2Cl2)][BArF

4] formed products dependent on the phosphine-

borane substituents (see figure 1).Reaction intermediates were studied to probe the

mechanism by which dehydrocoupling could occur.

Figure 1. Reaction of [Cp*RhMe(PMe3)(CH2Cl2)][BArF

4] with H3B·PHR2.

References

[1] Clark, T.J.; Lee, K.; Manners, I. Chem. Eur. J., 2006, 12, 8634.

Page 23: Royal Society of Chemistry Coordination and Organometallic

REVERSIBLE FORMATION OF MAGNETIC

POLYNICKELOCENES

R. A. Musgrave, A. D. Russell, G. R. Whittell and Ian Manners

School of Chemistry, University of Bristol, Bristol, BS8 1TS, UK

[email protected]

The field of [n]metallocenophane chemistry has been expanding rapidly since the

1960s; in particular, the propensity of these strained molecules to undergo ring-opening

polymerization (ROP). Recently this reaction has been extended to

[n]nickelocenophanes, yielding polynickelocenes, a class of magnetic S = 1

metallopolymers.1 Interestingly, this ROP reaction shows reversibility at ambient

temperatures (Figure 1), and thermodynamic propensity for this unusual reactivity is

explored.2 The synthesis and characterization of a number of other novel

[n]nickelocenophane monomers and their respective magnetic polymers will also be

presented.

Figure 1. Reversible formation of poly(nickelocenylpropylene).

[1] Baljak, S.; Russell, A. D.; Binding, S. C.; Haddow, M. F.; O’Hare, D.; Manners, I. J. Am. Chem. Soc. 2014, 136, 5864-5867.

[2] Musgrave, R. A.; Russell, A. D.; Whittell, G. R.; Manners, I. Unpublished Results

Page 24: Royal Society of Chemistry Coordination and Organometallic

TOWARDS METAL CARBENES AS HYDROGENASE MODELS

D. Britton, A. Boak, S. McDougall, Z. Moinuddin, N. Bramah, A. Panchal, and

S. Zlatogorsky

School of Physical Sciences and Computing, UCLan, Preston, UK [email protected]

Structural and functional models of hydrogenases are important for understanding redox

reactions involving hydrogen in nature, modelling enzymatic behaviour, and are the key

to new materials for hydrogen storage.

N-Heterocyclic carbenes (NHCs) are known to support iron in a range of oxidation

states. Electronic properties of NHCs (σ-donors capable of backbonding) are similar to

those of CN- and CO found in native hydrogenases and, unlike carbenes, already tried in

their modelling.1 The basicity of a free carbene coupled with its lability when

coordinated to Fe2,3

makes Fe-NHCs ideal unexplored candidates for hydrogenase-like

behaviour (e.g. heterolytic H2 cleavage across the Fe-NHC bond). The validity of this

idea is strongly supported by the finding that Fe acts as an electrophile to activate H2 in

an Fe-NHC system.4 However, due to their lability, Fe-NHC complexes lack stability.

2,3

We propose to introduce sulphur-based “Fe-anchoring” function(s) into an NHC ligand.

Metallation of the latter, given the well-known stability of sulphur-bridged Fe clusters,

will give stable Fe-S-NHC complexes.

We suggest that sulphur-containing functional groups will include thioketone(s),

thiolate(s), thiocarboxylate(s), thiazole(s), etc. (Scheme 1)

N

E

E'

N

N

N

E'

E

D DS

N

E

E'

N

N

N

E'

E

DD

S

S

x x yy

SPACER = o-/m-xylylidene, hexamethylenebenzene, oligomethylene

D = donor group, e.g.

E, E' = C, N (imidazole/triazole)

SR, R = H, Alk/Ar

SPACERSPACER

Scheme 1

N

S

N

S

SPACERx x

The preparation of imidazolium, triazolium and (benzo)thiazolium pro-ligands shown

above, as well as initial attempts at their metallation, will be reported.

[1] Tard, C.; Pickett, C. J. Chem. Rev. 2009, 109, 2245.

[2] Zlatogorsky, S.; Muryn, C. A.; Tuna, F.; Evans, D. J.; Ingleson, M. J.

Organometallics 2011, 30, 4974.

[3] Zlatogorsky, S.; Ingleson, M. J. Dalton Trans. 2012, 41, 2685.

[4] Runyon, J. W.; Steinhof, O.; Rasika Dias, H. V.; Calabrese, J. C.; Marchall, W. J.;

Arduengo III, A. J. Aust. J. Chem., 2011, 64, 1165.

Page 25: Royal Society of Chemistry Coordination and Organometallic

TAKING URANYL PACMAN CHEMISTRY TO THE NEXT

LEVEL

Nicola L. Bell1, Jason Love

1 and Polly L. Arnold

1

1 EastChem, University of Edinburgh, The Kings Buildings, Edinburgh

[email protected]

The strong and covalent O=U=O multiple bonds in the linear uranyl dication [UO2]2+

are relatively inert with most reactivity of the U(VI) metal occurring around the

equatorial plane.1 However, we have previously shown that encapsulation of [UO2]

2+

within the orthophenylene-linked polypyrrolic Schiff base ‘Pacman’ macrocycle can result in facile reduction of the U(VI) to U(V) and allow silylation and metalation

reactivity.2 Due to steric constraints two linear uranyl moieties cannot be accommodated

within a single orthophenylene-linked ‘Pacman’ ligand; instead a doubly reduced ‘butterfly’ compound with a ‘cis-uranyl’ motif is accessed (Scheme 1).3

Scheme 1: Synthesis of bis-uranyl ‘Pacman’ complexes

Working with our extended anthracenylene-linked macrocycle we have recently

developed a synthetic route to both mono- and bis-uranyl complexes the latter of which

contains two U(VI) linear trans-dioxo uranyl cations in contrast to the reduced

U(V)/U(V) ‘butterfly’ species observed with the orthophenylene-linked macrocycle.4

Herein, we will discuss their synthesis, properties and reactivity.

References. [1] Denning, R. G.; Green, J. C.; Hutchings, T. E.; Dallera, C.; Tagliaferri, A.; Giarda, K.; Brookes,

N. B.; Braicovich, L. J. Chem. Phys. 2002, 117, 8008; Vallet, V.; Wahlgren, U.; Grenthe, I. J.

Phys. Chem. A 2012, 116, 12373-12380.

[2] Arnold, P. L.; Blake, A. J.; Wilson, C.; Love, J. B. Inorg. Chem. 2004, 43, 8206-8208; Arnold,

P. L.; Patel, D.; Wilson, C.; Love, J. B. Nature 2008, 451, 315-U313; Arnold, P. L.;

Pecharman, A. F.; Hollis, E.; Yahia, A.; Maron, L.; Parsons, S.; Love, J. B. Nat. Chem. 2010, 2,

1056-1061; Arnold, P. L.; Hollis, E.; Nichol, G. S.; Love, J. B.; Griveau, J.-C.; Caciuffo, R.;

Magnani, N.; Maron, L.; Castro, L.; Yahia, A.; Odoh, S. O.; Schreckenbach, G. J. Am. Chem.

Soc. 2013, 135, 3841-3854.

[3] Arnold, P. L.; Jones, G. M.; Odoh, S. O.; Schreckenbach, G.; Magnani, N.; Love, J. B. Nat Chem

2012, 4, 221-227.

[4] Arnold, P. L.; Jones, G. M.; Pan, Q.-J.; Schreckenbach, G.; Love, J. B. Dalton Trans. 2012, 41,

6595-6597.

Page 26: Royal Society of Chemistry Coordination and Organometallic

MECHANISTIC STUDIES OF CHARGE TRANSFER IN

HYDROGEN BONDED ‘DIMERS OF DIMERS’

L. A. Wilkinson, Kevin B. Vincent, Anthony J. H. M Meijer and N. J. Patmore

Department of Chemical Sciences, University of Huddersfield, Huddersfield HD1 3DH

[email protected]

Small molecule models are vital tools in improving our understanding of electron

transfer processes that occur in nature and chemistry. Mixed-valence (MV) compounds

consist of an electron donor and acceptor bridged by a conjugated spacer. As well as

covalently linked systems, it is also conceivable that hydrogen bond bridges could be

used to link redox centers in MV systems. However, examples remain rare, with little

detail about possible mechanisms by which the MV state is stabilised.1

In this talk we discuss charge transfer in hydrogen bonded ‘dimers of dimers’ of form

[Mo2(TiPB)3(HDON)]2 (TiPB = 2,4,6-triisopropylbenzoate; H2DON = 2,7-

dihydroxynaph-thyridine) and [Mo2(TiPB)3(HDOP)]2 (H2DOP = 3,6-

dihydroxypyridazine). The cyclic voltammograms of these compounds show two one-

electron processes indicating stabilization of the MV state.2 The magnitude of

stabilization is affected by addition of electron donating and withdrawing groups to the

ancillary and bridging ligands, providing insight into possible mechanisms. We have

recently probed the timescale and mechanism of charge transfer in these compounds by

cyclic voltammetry, UV/vis/NIR and EPR spectroscopy, and theoretical studies.

Stabilisation of the MV state in these compounds is proposed to be related to the proton

coordinate of the bridging ligand, and is termed proton-coupled mixed valency (Figure

1). These are the best models to date for studying this process, and a possible

mechanism will be discussed.

Figure 1. Proton-coupled mixed valency in [Mo2(TiPB)3(HDOP)]2

+.

[1] a) Goeltz, J. C.; Kubiak, C. P. J. Am. Chem. Soc., 2010, 132, 17390. b) Sun, H.:

Steeb, J.; Kaifer, A. E. J. Am. Chem. Soc., 2006, 128, 2820. c) Tadokoro, M.; Inoue,

T.; Tamaki, S.; Fujii, K.; Isogai, K.; Nakazawa, H.; Takeda, S.; Isobe, K.; Koga, N.;

Ichimura, N.; Nakasuji, K. Angew. Chem. Int. Ed., 2007, 46, 5938.

[2] a) Wilkinson, L. A.; McNeill, L.; Meijer, A. J. H. M.; Patmore, N. J. J. Am. Chem.

Soc., 2013, 135, 1723. b) Wilkinson, L. A.; McNeill, L.; Scattergood, P. A.;

Patmore, N. J. Inorg. Chem. 2013, 52, 9683.

Page 27: Royal Society of Chemistry Coordination and Organometallic

eNHChanting NEW REACTIVITY WITH Cu AND Zn

Lee R. Collins, Mary F. Mahon, Ian M. Riddlestone and Michael K. Whittlesey

Department of Chemistry, University of Bath, Claverton Down, Bath, BA2 7AY, UK

[email protected]

N-heterocyclic carbenes (NHCs) are increasingly employed as ligands in

organometallic chemistry. Their easily tuneable electronic and steric demands make

them ideal for varying the reactivity of metal centres.

A diverse range of NHC structures are now available, with enhanced

nucleophilic or electrophilic character through altering their HOMO-LUMO energies.1

Di-amido carbenes (DACs) exhibit enhanced electrophilicity at the C2 position due to

electron withdrawing carbonyls α to the carbene’s nitrogen substituents.2

We have previously reported on the remarkable electrophilicity of DACs, and

the facile intramolecular migratory insertion reactions of [(6-MesDAC)MR2] (M = Zn,

Cd; R = Et, Me) complexes.3 We now report that this migratory insertion chemistry

also occurs for catalytically relevant copper hydride complexes, and even for di-amino

carbene complexes. The difference in susceptibility to migratory insertion is used to

explain the stark difference in activity of di-amino and di-amido carbene complexes for

catalytic reductions.

References

[1] Melaimi, M.; Soleilhavoup, M.; Bertrand, G. Angew. Chem. Int. Ed. 2010, 49, 8810–8849

[2] Verlinden, K.; Buhl, H.; Frank, W.; Ganter, C. Eur. J. Inorg. Chem 2015, 14, 2416-

2425

[3] Collins, L. R.; Hierlmeier, G.; Mahon, M. F.; Riddlestone, I. M.; Whittlesey, M. K.

Chem. Eur. J. 2015, 21, 3215–3218

Page 28: Royal Society of Chemistry Coordination and Organometallic

EXPLORING PHOSPHINE-ALKENE LIGAND CHEMISTRY

L. W. Tuxworth,1 L. Estévez,

2 J-M. Sotiropoulos,

2 K. Miqueu,

2 and P. W. Dyer

1

1

Centre for Sustainable Chemical Processes, Department of Chemistry, Durham

University, South Road, Durham, DH1 3LE, UK. 2

Institut des Sciences Analytiques et de Physico-Chimie pour l’Environnement et les Matériaux, Université de Pau et des Pays de l’Adour, Hélioparc, 64053 Pau, France.

[email protected]

Phosphine-alkene (P^=) ligands have received much recent attention, finding use in a

range of metal-mediated reactions, e.g. 1,4-additions and imine hydrogenations [1,2].

P^= systems provide electronic asymmetry that results from the combination of a

weakly π-acidic phosphine and a strongly π-accepting alkene within the same scaffold.

When bound in a bidentate fashion, this unusual electronic character can accelerate

reductive elimination reactions, something often the rate-limiting step in many Pd-

mediated cross-couplings [3].

P

Pd

L1

P

Pd

P(L1)

X

Cl

Me

L1 P

Pd

P(L1)

Me

P(L1)

PPd

P

–[L1Me]Cl

Assisted TS

1

Me

X : Cl

L1

P

Pd

P(L1)

Me

Me

X : Me

direct reductive elimination

pseudo-reductive elimination

#

P

Pd

P(L1)

Me

Me

#

Cl

–C2H6

N

Ph2P

L1 :

5-coordinate intermediate

P^= L1 promotes reductive elimination of ethane from [PdMe2(TMEDA)] forming

pseudo-homoleptic [Pd0(κ2

-P,C-L1)2] (1) [4]. Reaction of L1 with

[PdCl(Me)(TMEDA)] also leads to the formation of complex 1, but via an unusual

ligand-assisted methyl abstraction pathway, which has been probed by computation and

by variable temperature NMR studies.

References.

[1] Maire, P.; Deblon, S.; Breher, F.; Geier, J.; Bohler, C.; Ruegger, H.; Schonberg, H.;

Grützmacher, H. Chem. Eur. J., 2004, 10, 4198.

[2] Shintani, R.; Duan, W.L.; Nagano, T.; Okada, A.; Hayashi, T. Angew. Chem. Int.

Ed., 2005, 44, 4611.

[3] Zhang, H.; Luo, X.C.; Wongkhan, K.; Duan, H.; Li, Q.; Zhu, L.Z.; Wang, J.;

Batsanov, A.S.; Howard, J.A.K.; Marder, T.B.; Lei, A.W. Chem. Eur. J., 2009, 15,

3823.

[4] Estévez, L.; Tuxworth, L.W.; Sotiropoulos, J.-M.; Dyer, P.W.; Miqueu, K. Dalton

Trans, 2014, 43, 11165.

Page 29: Royal Society of Chemistry Coordination and Organometallic

GROUP 2– AND GROUP 12–METAL COMPLEXES

M. P. Blakea, N. Kaltsoyannis

b and P. Mountford

a

aDepartment of Chemistry, University of Oxford, OX1 3TA, UK

bDepartment of Chemistry, University College London, WC1H 0AJ, UK

[email protected]

The structure, bonding and reactivity of molecular compounds containing unusual

metal–metal bonds continues to be a topic of considerable interest.[1]

Reports of the first

Zn–Zn[2]

and Mg–Mg single bonds,[3]

and the emergence of a growing body of

lanthanide– and actinide–transition metal bonds highlight the activity in this area.[4]

Despite this, there remains relatively little experimental or theoretical information

regarding bonds between the Group 2 elements and the transition metals. We recently

showed that reductive cleavage of Fp2 (Fp = CpFe(CO)2) with Ca/Hg or Yb/Hg

amalgams gave the Ca–Fe and Yb–Fe compounds 1 and 2, with 1 possessing the first

bond between calcium and a transition metal.[5]

Analogous Mg–Fe (e.g. 3) and related

compounds were also readily formed, with or without supporting ligand sets,[6]

and this

led us to the first insertion reactions of Group 2–transition metal bonds. Recent related

studies in our laboratory gave the first compound with a bond between two different

Group 12 metals, namely 4 which contains a formal ZnI–Hg

0–ZnI linkage.

[7] In this

presentation we report further on some of these recently communicated results, along

with additional new work from our laboratories in this area.

[1] Liddle, S. T. (Ed.), Molecular Metal-Metal Bonds: Compounds, Synthesis,

Properties, Wiley-VCH: Weinheim, Germany, 2015. [2] Resa, I.; Carmona, E.;

Gutierrez-Puebla, E.; Monge, A. Science 2004, 305, 1136. [3] Green, S. P.; Jones, C.;

Stasch, A. Science 2007, 318, 1754. [4] Liddle, S. T.; Mills, D. P. Dalton Trans. 2009,

5592; Roesky, P. W. Dalton Trans. 2009, 1887; Bauer, J.; Braunschweig, H.; Dewhurst,

R. D. Chem. Rev. 2012, 112, 4329; Oelkers, B.; Butovskii, M. V.; Kempe, R. Chem.

Eur. J. 2012, 18, 13566. [5] Blake, M. P.; Kaltsoyannis, N.; Mountford, P. J. Am. Chem.

Soc. 2011, 113, 15358. [6] Blake, M. P.; Kaltsoyannis, N.; Mountford, P. Chem.

Commun. 2013, 49, 3315. [7] Blake, M. P.; Kaltsoyannis, N.; Mountford, P. Chem.

Commun. 2015, 51, 5743.

Page 30: Royal Society of Chemistry Coordination and Organometallic

Abstracts for Poster Presentations

Page 31: Royal Society of Chemistry Coordination and Organometallic

SILVER N-HETEROCYCLIC CARBENES DERIVED FROM

NATURAL PRODUCTS AS ANTICANCER AGENTS

Dr. Charlotte E. Willans,1 Prof. Roger M. Phillips2 and H. Abdelgawad1

1School of Chemistry, University of Leeds, Woodhouse Lane, Leeds, LS2 9JT, UK, 2School of Applied Sciences, University of Huddersfield, HD1 3DH, UK.

[email protected]

The discovery of the platinum-containing compound cisplatin as a chemotherapeutic compound opened a new area of research in organometallic chemistry for the treatment of cancer.1 Some of this research is focused on silver with silver-N-heterocyclic carbene (NHC) complexes showing significant promise as chemotherapeutic agents.2-4 This work describes the synthesis of novel silver-NHC complexes derived from xanthine precursors such as theophylline, theobromine and caffeine which are found naturally in cocoa beans. These complexes were assessed for their cytotoxicity against eight cancerous cell lines and results were compared to those obtained for cisplatin. Hydrophobicity studies and structure activity relationships indicate that both stability and ligand sterics play a part in their activity as anticancer agents.5

[1] Rosenber.B; Vancamp, L.; Krigas, T. Nature 1965, 205, 698. [2] Monteiro, D. C. F.; Phillips, R. M.; Crossley, B. D.; Fielden, J.; Willans, C. E. Dalton Trans. 2012, 41, 3720. [3] Medvetz, D. A.; Hindi, K. M.; Panzner, M. J.; Ditto, A. J.; Yun, Y. H.; Youngs, W. J. Metal-based Drugs 2008, 2008. [4] Mohamed, H. A.; Willans, C. E. RSC Specialist Periodical Reports in Organometallic Chemistry 2014, 39, 26. [5] Mohamed, H. A.; Lake, B. R. M.; Laing, T.; Phillips, R. M.; Willans, C. E. Dalton Trans. 2015, 44, 7563.

E.g.1. R1 = Me, R2 = Benzyl

E.g.2. R1 = Me, R2 = nBu

Page 32: Royal Society of Chemistry Coordination and Organometallic

A STUDY OF NEW {Rh(XANTPHOS)}+ COMPLEXES

FOR THE DEHYDROPOLYMERISATION OF AMINE-BORANES

G. M. Adams and A. S. Weller*

Department of Chemistry, University of Oxford, UK

[email protected]

Polyaminoboranes, [H2BNRH]n, are an underdeveloped class of materials that are

isoelectronic with technologically and societally ubiquitous polyolefins. They are

formed by the controlled dehydropolymerisation of a parent amine-borane, and the

fundamental understanding of this mechanism has only recently started to be

investigated.[1]

Herein, we report the synthesis of new dehydropolymerisation catalysts

based upon a previously reported {Rh(Xantphos)}+ fragment,

[2] one example being the

straightforwardly prepared mer-[Rh(κ3-POP-Xantphos)(η2

-PhCCPh)][BArF

4] complex

(1).[3]

This system produces polymer of low PDI at 0.2 mol% catalyst loading, and by

tuning the reaction conditions the molecular weight of the resulting polymeric material

can be controlled. Modifying the Xantphos ligand backbone is also explored with regard

to the mechanism of dehydropolymerisation.

Scheme 1: Dehydropolymerisation of an amine-borane by complex 1

References

[1] Staubitz, A.; Sloan, M. E.; Robertson, A. P. M.; Friedrich, A.; Schneider, S.; Gates,

P. J.; Gunne, J; Manners, I. J. Am. Chem. Soc. 2010, 132, 13332

[2] Johnson, H. C.; Leitao, E. M.; Whittell, G. R.; Manners, I.; Lloyd-Jones, G. C.;

Weller, A. S. J. Am. Chem. Soc. 2014, 136, 9078

[3] Ren, P.; Pike, S. D.; Pernik. I; Weller, A. S.; Willis, M. C. Organometallics 2015,

34, 711

Page 33: Royal Society of Chemistry Coordination and Organometallic

UPGRADING ETHANOL; MECHANISTIC INSIGHTS

H. Aitchison, R. L. Wingad, and D. F. Wass

School of Chemistry, University of Bristol, Cantock's Close, Bristol, BS8 1TS [email protected]

The search for renewable and sustainable sources of energy is a matter of exceptional importance and demands substantial attention and research effort. Numerous studies have shown the damaging and detrimental effects increased carbon dioxide can have on the Earth’s fragile climate; a direct consequence of utilising fossil-based fuels. All alternative fuels presently being used have both distinct advantages and inadequacies, such as the current plant-based fuel bioethanol.

Research within the Wass group has focused on the synthesis of n-butanol and iso-butanol from ethanol and ethanol/methanol blends respectively. Both butanol isomers have energy densities much more similar to that of gasoline (>86%), and their more alkane-like features assist in reducing the disadvantages bioethanol presents as a sustainable biofuel. The proposed feedstock is waste plant-matter, hence combustion of biobutanol results in a complete carbon cycle and does not adversely affect the environment.

Specifically novel transition metal catalysed routes are currently under investigation. The reaction is postulated to proceed via a well-known process known as the Guerbet reaction. We have developed ruthenium based catalysts that achieve unprecedented selectivity in the catalytic upgrading of ethanol and ethanol/methanol blends. The catalytic capability of these systems has been studied, and several important and intriguing observations highlight the need to investigate the specifics of the mechanism.

Subtle changes in the ligand sets allow for n-butanol, iso-butanol (catalysts 1

and 2) and sec-butanol (catalyst 3) formation. Unambiguously defining the origins of the selectivity of these ruthenium systems will enable selectivity tuning. This is a mechanistic feature that will undoubtedly provide information necessary for scale-up, as well as future design of new catalysts for the transformation of alcoholic substrates. [1] Dowson, G. R. M.; Haddow, M. F.; Lee J.; Wingad R. L.; Wass D. F., Angew.

Chemie – Int. Ed., 2013, 52, 9005-9008.

Page 34: Royal Society of Chemistry Coordination and Organometallic

DEVELOPMENT OF FLUORINE-18 RADIOLABELLED THIA-

FATTY ACID TRACERS AND A COORDINATION CHEMISTRY

DELIVERY MECHANISM FOR IMAGING CARDIAC

METABOLISM BY POSITRON EMISSION TOMOGRAPHY (PET)

Zainab Al-Alia,b

, Juozas Domarkasa,b

, Benjamin Burkea,b

, Torsten Ruestb, Faisal Nuhu

d,

Robert Atkinsond, Chris Cawthorne

bc, Anne-Marie Seymour

bc and Stephen J.

Archibaldab

aDepartment of Chemistry, bPositron Emission Tomography Research Centre, cSchool of Biological, Biomedical and Environmental Sciences, University of Hull, Cottingham

Road, Hull, HU6 7RX, UK.

[email protected]

Clinical need. Cardiac diseases are the leading cause of death worldwide, thus, an early

detection and better understanding of different pathologies is needed. Positron emission

tomography1 is the most sensitive noninvasive medical imaging technique capable of

detecting metabolic changes during or preceding the establishment of a pathology.

Biochemical rational. Cardiac retention of thia-fatty acids (TFAs) is expected to reflect

the extent of fatty acid metabolism.2 Long chain TFAs are actively taken up by

myocites, but show poor blood solubility and fast liver-based elimination due to their

high lipophilicity.2 There is no active transport of short TFAs which may result in more

favorable pharmacokinetics. Project aims. Develop a novel TFA delivery mechanism

to the cardiac cells, based on their coordination to mitochondria targeted transition

metal complexes, allowing us to take advantage of the pharmacokinetics of various

short fatty acids and enable their use as tracers of cardiac metabolism.3 Results. 16-

[18

F]-fluoro-4-thia-palmitic acid (18

F-FTP) which should be actively transported into

cells and 10-[18

F]-fluoro-4-thia-capric acid (18

F-FTC2) which should enter only by

passive diffusion, were successfully radiolabelled from in house made precursors. 18

F-

FTP cardiac uptake was confirmed by PET imaging. Non mitochondria targeted zinc-

cross bridged cyclam complex was prepared as a model for coordination. Future work.

Free 18

F-FTC2 cardiac uptake will be investigated by PET and compared with that of 18

F-FTP. FTC2 will be coordinated with prepared Zn-cross bridged cyclam complex and

its in vitro and in vivo behavior will be studied by LC/MS and PET.

References.

[1] S. Vallabhajosula, Molecular Imaging: Radiopharmaceuticals for PET and SPECT,

2009, 299.

[2] A. P. Belanger, M. K. Pandey and T. R. DeGrado, Nucl. Med. Biol., 2011, 38, 435.

[3] A. Khan et al. J. Am. Chem. Soc. 2009, 131, 3416–3417

Page 35: Royal Society of Chemistry Coordination and Organometallic

SUPPORTED INDENYL METALLOCENES FOR ETHYLENE

POLYMERISATION

P. Angpanitcharoen, G. Hay, J.-C. Buffet and D. O’Hare

Chemistry Research Laboratory, University of Oxford, 12 Mansfield Road, Oxford OX1 3TA.

[email protected]

With a global production volume of 150 million metric tonnes per annum, mostly by

heterogeneous Ziegler-Natta catalysts, polyolefins have become an essential material for

the modern world.[1]

The use of metallocenes in polyolefins production has enabled an

unprecedented ability to fine-tune polymer microstructure by ligand design.[2]

Supported

metallocene catalysts capture the best of both worlds by, to an extent, retaining high

polymerisation activity and control of metallocenes while providing templates for

polymer growth.[3]

A new substituted indenyl family have recently shown very high

activity towards solution-phase polymerisation of ethylene.[4,5]

Here, we will present our study on the effects of ligand modification. (Me2SB(

3-

EtI*))ZrCl2, its unbridged analogues and alkyl analogues have been synthesised,

characterised and tested for ethylene polymerisation activity.

Synthesis of dimethylsilylbis(3-ethyl-2,4,5,6,7-tetramethylindenyl)zirconium

dichloride, (Me2SB(

3-EtI*))ZrCl2.

References:

[1] Gahleitner, M.; Resconi, L.; Doshev, P. MRS Bull. 2013, 38, 229.

[2] Welborn, H. C.; Ewen, J. A. US5324800 1991.

[3] Severn, J. R.; Chadwick, J. C.; Duchateau, R.; Friederichs, N. Chem. Rev. 2005,

105, 4073.

[4] Ransom, P.; Ashley, A. E.; Brown, N. D.; Thompson, A. L.; O’Hare, D. Organometallics 2011, 30, 800.

[5] Arnold, T. A. Q.; Buffet, J.-C.; Turner, Z. R.; O’Hare, D. J. Organomet. Chem. 2015. http://dx.doi.org/10.1016/j.jorganchem.2015.01.019

Page 36: Royal Society of Chemistry Coordination and Organometallic

PALLADIUM-STIBINE COMPLEXES: UNPRECEDENTED TRIPLY BRIDGING COORDINATION OF

TRIMETHYLANTIMONY

S. Benjamina, T. Kraemer

b, W. Levason

c, S. Macgregor

b and G. Reid

c.

aNottingham Trent University, NG11 8NS

bHeriot-Watt University, EH14 4AS

cUniversity of Southampton, SO17 1BJ

[email protected]

Palladium phosphine complexes are widely used as catalysts in organic cross coupling reactions, and the electronic and steric characteristics of the phosphine ligands can be

varied in order to tune catalyst activity and selectivity.1 Palladium complexes with

stibines, SbR3, the heavier congeners of phosphines, have been much less explored. Stibine ligands exhibit a number of unusual coordination behaviours which are distinct from their lighter analogues, such as redox activity and formation of acceptor

interactions with anions and other electron donors both intra- and intermolecularly.2,3

The first (and still rare) example of a bridging organopnictine was a µ 2-SbiPr3 bridged

Rh2 dimer, and though the phosphine analogue could not be prepared directly, ligand metathesis with this stibine complex allowed a bridging phosphine to be isolated for the

first time.4

Investigation of the coordination chemistry of stibine

ligands with Pd(II) has led to the isolation of a number

of dimeric or cluster complexes, featuring short Pd-Pd Pd Sb

distances and several unusual ligand behaviours. The

Lewis acidity of coordinated halostibines is

demonstrated by the formation of intermolecular

secondary interactions between stibine and halide

ligands. A [Pd(0)]4 tetrahedron incorporating µ3-SbMe3

ligands has been prepared. The triply-bridging behaviour

of a monodentate organopnictine is unprecedented, and

the electronic structure and bonding in this complex has

been examined using ADF calculations.

[1] Gillespie, J. A.; Zuidema, E.; van Leeuwen, P. W. N. M.; Kamer, P. C.

J. In Phosphorus(III) Ligands in Homogeneous Catalysis: Design and Synthesis; John

Wiley & Sons, Ltd. 2012, ch. 1, 1-26.

[2] Benjamin, S. L.; Reid, G. Coord. Chem. Rev. 2015, 297–298, 168-180.

[3] Wade, C. R.; Ke, I.; Gabbaï, F. P. Angew. Chem. Int. Ed. 2012, 51, 478-481.

[4] Werner, H. Angew. Chem. Int. Ed. 2004, 43, 938-954.

Page 37: Royal Society of Chemistry Coordination and Organometallic

DISCERNING THE MECHANISM(S) OF IRON-CATALYZED

CROSS-COUPLING REACTIONS.

L. Brown, J. R. Birkett, J. B. Sweeney and N. J. Patmore

Department of Chemical Sciences, University of Huddersfield, HD1 3DH [email protected]

Formation of new carbon-carbon bonds via metal-catalyzed cross-coupling reactions are

amongst the most indispensable tools in the repertoire of synthetic chemists. Precious

metals are routinely used and chief amongst these is palladium, unrivaled in the

versatility of its application. The inherent cost of the metal leads to costly syntheses

driving prices of commodities and pharmaceuticals. While the use of catalysis is

essential in the drive towards greener chemistry the catalysts themselves should be both

environmentally and biologically benign. As a result the use of the more abundant first

row transition metals has been investigated. Iron in particular has proven a promising

candidate.1 Being less well explored however, there remain many questions regarding

the exact functioning mechanism(s) in already established catalytic schemes.2–4

This work explores the mechanism(s) operating in iron-catalyzed Kumada type

couplings; the reaction between halogenated hydrocarbons and Grignard reagents using

simple iron salt catalysts.5 A number of catalytically relevant species, including novel

hetero-bimetallic species are isolated by stoichiometric reaction of Fe(acac)3 with aryl

Grignards. These complexes have been characterized by a variety of methods including 1H NMR and IR spectroscopy, X-ray crystallography and magnetic susceptibility

measurements. Furthermore, the isolation of Fe(acac)2 as a side product and the

revelation that each of these species has equal catalytic competence in selected cross-

coupling reactions suggests the presence of competing catalytic pathways.

(1) Tamura, M.; Kochi, J. J. Am. Chem. Soc. 1971, 93, 1487–1489.

(2) Smith, R. S.; Kochi, J. J. Org. Chem. 1976, 41, 502–509.

(3) Fürstner, A.; Martin, R.; Krause, H.; Seidel, G.; Goddard, R.; Lehmann, C. W. J. Am. Chem. Soc. 2008, 130, 8773–8787.

(4) Hatakeyama, T.; Yoshimoto, Y.; Gabriel, T.; Nakamura, M. Org. Lett. 2008, 10, 5431–5344.

(5) Tamao, K.; Sumitani, K.; Kumada, M.; Kiso, Y. J. Am. Chem. Soc. 1972, 94, 9268–9269.

Page 38: Royal Society of Chemistry Coordination and Organometallic

POLYETHYLENE SYNTHESIS USING PERMETHYLINDENYL

METALLOCENE COMPLEXES

J.-C. Buffet, Z. R. Turner and D. O'Hare

Chemistry Research Laboratory, 12 Mansfield Road, OX1 3TA, Oxford, UK

[email protected]

Polyethylene (PE) is the most widely used polyolefin with an annual production of

above 75 million tons per year and a global demand poised for growth. In contrast to

heterogeneous Ziegler-Natta systems, homogeneous metallocene-based catalysts

produce polyethylene with a narrow molecular weight distribution due to their single-

site nature; polymer properties can also be fine-tuned through careful ligand design.[1]

By immobilising metallocene catalysts onto support materials, the processability

problem is solved and the advantages of metallocene catalysts are preserved to a great

extent.[2]

We recently reported the synthesis of permethylindenyl complexes and their use in

the solution polymerisation of ethylene,[3] and the development of solid catalysts using

layered double hydroxides as a support.[4]

Here, we present the synthesis of unsymmetrical permethylindenyl complexes,

Figure 1a, and their use in the polymerisation of ethylene in both solution and slurry,

Figure 1b.

a) b)

Figure 1. a) Solid-state molecular structure of a permethylindenyl zirconocene Me

2SB(tBu2Flu,I*)ZrCl2 and b) SEM images of PE synthesised in the slurry process.

References. [1]Hlatky, G. G. Chem. Rev. 2000, 100, 1347. [2]Severn, J. R.; Chadwick, J. C.;

Duchateau, R.; Friederichs, N. Chem. Rev. 2005, 105, 4073. [3]Arnold, T.A.Q.; Buffet,

J.-C.; Turner, Z. R.; O’Hare, D. J. Organomet. Chem., 2015, DOI:

10.1016/j.jorganchem.2015.01.019. [5]Buffet, J.-C.; Wanna, N.; Arnold, T. A. Q.;

Gibson, E. K.; Wells, P. P.; Wang, Q.; Tantirungrotechai, J.; O'Hare, D. Chem Mater.

2015, 27, 1495.

Page 39: Royal Society of Chemistry Coordination and Organometallic

DEVELOPING THE COORDINATION CHEMISTRY OF LEAD(II)

– NEW DIPHOSPHINE COMPLEXES AND REAGENTS FOR

SUPERCRITICAL FLUID ELECTRODEPOSITION

W. Levason, G. Reid and J. Burt

Chemistry, University of Southampton, Highfield, Southampton, SO17 1BJ, UK

[email protected]

Owing to the unique properties of supercritical fluids, the technique of supercritical

fluid electrodeposition (SCFED) has the potential to produce technologically important

materials on the nanoscale with atomic layer control. SCFED has already been

successfully used to deposit a range of reactive materials, including Ge films and 3 nm

diameter Cu nanowires.1

The development of this approach to enable deposition of other p-block materials is of

particular interest, and recently work has focussed on identifying, synthesising and

characterising suitable tin(II) and lead(II) precursors. Here we present voltammetric and

deposition studies of several Sn(II) and Pb(II) reagents in CH2Cl2, together with the

growth of 13 nm tin nanowires from supercritical CH2F2.

While numerous phosphine complexes of most heavy p-block elements (e.g. Bi(III),

Sn(II) and Sn(IV)) are known,2 surprisingly the only well characterised examples with

Pb(II) are two recently reported adducts with lead thiolates.3 In order to further explore

the coordination chemistry of lead(II) complexes with neutral diphosphine ligands we

have synthesised and characterised (by X-ray crystallography, IR spectroscopy,

microanalysis and NMR spectroscopy where possible) several diphosphine complexes

of Pb(NO3)2 and Pb(SiF6).2H2O, finding evidence for anion coordination in all cases.

Despite its size and ability to achieve large coordination numbers (up to 12) only one

diphosphine ligand is observed to coordinate to Pb(II).4

[1] Bartlett, P.N.; Cook, D.C.; George, M.W.; Hector, A.L.; Ke, J.; Levason, W.; Reid,

G.; Smith, D.C.; Zhang, W. Phys. Chem. Chem. Phys. 2014, 16, 9202. The SCFED

Project (www.scfed.net) is a multidisciplinary collaboration of British universities

investigating the fundamental and applied aspects of supercritical fluids.

[2] Burt, J.; Levason, W.; Reid, G. Coord. Chem. Rev. 2014, 260, 65.

[3] Rossini, A.J.; Macgregor, A.W.; Smith, A.S.; Schatte, G.; Schurko, R.W.; Briand,

G.G. Dalton Trans. 2013, 9533.

[4] Burt, J.; Grantham, W.; Levason, W.; Reid, G. Dalton Trans. 2015, 11533.

Page 40: Royal Society of Chemistry Coordination and Organometallic

(E)-SELECTIVE ALKENE ISOMERISATION CATALYSED BY A

HETEROBIMETALLIC Mg---Zr TRIHYDRIDE

M. J. Butler and M. R. Crimmin

Department of Chemistry, Imperial College, South Kensington, London, SW7 2AZ, UK

[email protected]

We recently reported the hydride-bridged heterobimetallic complex Al•Zr (Figure -

M•Zr), thought to play a role in the activation of a group 4 pre-catalyst in two separate

C–X (X = F, O) inert bond cleavages.1 Here we describe the preparation of Mg•Zr and

Zn•Zr: well-defined heterobimetallic hydrides of M(II) and Zr(IV). The Mg•Zr complex is a selective, single-site catalyst for the isomerisation of terminal alkenes,

controlling the stereoselectivity of the reaction.

We show that terminal alkenes are isomerised to internal alkenes across one position

with a strong bias for (E)-stereochemistry in the products.2

Absence of a M(II) fragment

– i.e. use of [Cp2ZrH(-H)]2 – results in loss of activity, selectivity, and competitive

substrate hydrogenation. The labile Al•Zr is also vulnerable to these pitfalls. In-situ pre-

catalyst activation is possible through mixing of the main group hydride and [Cp2ZrCl2].

Finally, we have probed the mechanism with a simple stoichiometric reaction of

(Dipp

BDI)Mg-alkyl complexes, which give Mg•Zr when reacted with [Cp2ZrH(-H)]2.

References

[1] a) Yow, S.; Gates, S. J.; White, A. P.; Crimmin, M. R. Angew. Chem. Int. Ed. 2012,

51, 12559-12563 b) Yow, S.; Nako, A. E.; Neveu, L.; White, A. J. P.; Crimmin, M. R.

Organometallics 2013, 32, 5260-5262.

[2] Examples of other group 4 systems: a) Averbuj, C.; Eisen, M. J. Am. Chem. Soc.

1999, 121, 8755-8759 b) Petrisor, C. E.; Frutos, L. M.; Castaño, O.; Mosquera, M. E.

G.; Royo, E. ; Cuenca, T. Dalton Trans. 2008, 2670-2673

Page 41: Royal Society of Chemistry Coordination and Organometallic

SYNTHETIC APPLICATIONS OF BORYLZINC COMPOUNDS:

BORYL TRANSFER CHEMISTRY AND CATALYTIC

BORYLATION

Jesus Campos, Simon Aldridge*

Chemistry Research Laboratory, University of Oxford, Mansfield Road, OX1 3TA,

Oxford

[email protected]

Since the groundbreaking isolation of the first boryllithium species by Yamashita and

Nozaki,[1]

the chemistry of boryl nucleophiles has rapidly expanded. Interestingly, the

reactivity of the boryl anion is highly dependent on its metallic counterion, as

exemplified by the dissimilar reactivity between a diaminosubstituted boryllithium and

its magnesium analogue when added to benzaldehyde.[2]

Considering the rich chemistry

of alkyl zinc reagents in synthetic applications and the growing interest in zinc as an

inexpensive and environmentally benign metal in catalysis, we decided to explore the

still undeveloped chemistry of borylzinc compounds (A in Figure 1).

Figure 1. Synthetic applications of borylzinc compounds.

Transmetalation reactions using borylzinc species allowed us to synthesize a number of

Transition Metal and Main Group metal boryl complexes with considerably higher

yields and increased selectivity when compared to the commonly used boryllithium

analogues. More interestingly, the direct use of borylzinc precursors for cross-coupling

processes resulted in the borylation of a plethora of organic halides and pseudohalides,

including unactivated alkyl halides, in high yields, with high functional group tolerance

and under mild conditions.

[1] Segawa, Y.; Yamashita, M.; Nozaki, K. Science 2006, 314, 113.

[2] Yamashita, M.; Suzuki, Y.; Segawa, Y.; Nozaki, K. J. Am. Chem. Soc. 2007, 129,

9570.

Page 42: Royal Society of Chemistry Coordination and Organometallic

SOFT DONOR COMPLEXES WITH TRI- AND TETRA-VALENT NIOBIUM AND TANTALUM

Yao-Pang Chang; William Levason; Gill Reid.

School of Chemistry, University of Southampton, Southampton, SO17 1BJ, UK

[email protected]

The chemistry of early transition metal halide complexes with soft, neutral donor

ligands such as phosphines, thio- or selenoethers is often different from that with hard

O- or N-donor ligands.1 Very recently, we have reported a series of group V metal

halide complexes with chalcogenoethers in their highest oxidation state (+5), and

showed their utility in the chemical vapour deposition of NbS2 and NbSe2 thin films.2

However, complexes of the lower oxidation states, such as Nb(III), Ta(III), Nb(IV) or

Ta(IV) with soft donor ligands are unusual and their chemistry is not well developed.

Mononuclear Nb(III) or Ta(III) halides (NbCl3 or TaCl3) do not exist, instead, dinuclear

complexes of the form [M2Cl6(L)3] (M = Nb, Ta; L = Me2S, THT) are known to form

from Na/Hg or Mg/Et2O reduction of MCl5 in the presence of the S ligand.3 Complexes

obtained via these starting materials often form edge-sharing dimers and contain a

M=M double bond.4 Nb(IV) halides, in contrast, are made by aluminium reduction of

MCl5 over a temperature gradient.5,6

Here we report a series of Nb(III) and Ta(III) dimer complexes obtained by reaction of

[M2Cl6(Me2S)3] with soft, neutral dithioether or diselenoether ligands. In addition,

NbCl4 has been prepared and directly reacted with neutral ligands. The chemistry

associated with these complexes is described, together with their spectroscopic and

structural characterisation.

References:

[1] S. L. Benjamin, W. Levason and G. Reid, Chem. Soc. Rev., 2013, 42, 1460-

1499.

[2] S. L. Benjamin, Y. P. Chang, C. Gurnani, A. L. Hector, M. Huggon, W. Levason

and G. Reid, Dalton Trans., 2014, 43, 16640-16648.

[3] T. Waters, A. G. Wedd, M. Ziolek and I. Nowak, in Comprehensive

Coordination Chemistry II, eds. J. A. McCleverty and T. J. Meyer, Elsevier B.V.,

Toronto, Editon edn., 2003, vol. 4.

[4] E. Babaian-Kibala, F. A. Cotton and P. A. Kibala, Inorg. Chem., 1990, 29, 4002-

4005.

[5] McCarley, R. E.; Torp, B. A., Inorg. Chem. 1963, 2, 540

[6] Fowles, G. W. A.; Tidmarsh, D. J.; Walton, R. A., Inorg. Chem. 1969, 8, 631.

Page 43: Royal Society of Chemistry Coordination and Organometallic

AN ELECTROCHEMICAL FLOW-REACTOR FOR THE

SYNTHESIS OF ORGANOMETALLIC COMPLEXES

Dr C. E. Willans, Dr B. N. Nguyen and M. R. Chapman

Institute of Process Research and Development (iPRD), School of Chemistry, University

of Leeds, LS2 9JT, UK

[email protected]

Developed by Arduengo over two decades ago, the notion of employing N-heterocyclic

carbenes (NHCs) as ancillary ligands for transition metal-based catalysts has been

refined such that they now present all required attributes for broad application –

typically offering high return over their phosphine rivals (e.g. enhanced thermal stability

and greater tunability).[1] Despite these advances, a number of drawbacks currently

exist with traditional methods of metal-NHC preparation, notably when considering

such complexes for industrial use. The pre-requisite for strongly basic/strictly inert

conditions within current syntheses largely limits the scope of suitable substrate for

metal-NHC complexation.[2] Complementary routes include transmetallation of a

carbenic moiety from a basic metal oxide (e.g. Ag2O),[3] leading to the accumulation of

metal salt byproducts.

In light of these challenges, the design, construction and optimisation of an innovative

electrochemical flow-reactor has been developed which circumvents such issues. The

electrochemical approach, which has been published in Chemical Communications, [4,

5] provides a clean and atom economical route to metal-NHCs as a result of: (i) no

external reagents are required, (ii) a simple evaporative work-up and (iii) only H2 gas

byproduct.

[1] Arduengo, A. J.; Harlow, R. L.; Kline, M. J. Am. Chem. Soc. 1991, 113, 361-

363.

[2] Bourissou, D.; Guerret, O.; Gabbai, F. P.; Bertrand, G. Chem. Rev. 2000, 100,

39-91.

[3] Lin, I. J. B.; Vasam, C. S. Coord. Chem. Rev. 2007, 251 (5+6), 642-670.

[4] Lake, B. R. M.; Bullough, E. K.; Williams, T. J.; Whitwood, A. C.; Little, M.

A.; Willans, C. E. Chem. Commun. 2012, 48 (40), 4887-4889.

[5] Chapman, M. R.; Shafi, Y. M.; Kapur, N.; Nguyen, B. N.; Willans, C. E. Chem.

Commun. 2015, 51 (7), 1282-1284.

Page 44: Royal Society of Chemistry Coordination and Organometallic

SYNTHESIS AND REACTIVITY OF TITANIUM BORYLIMIDO

COMPOUNDS

Benjamin A. Clough, Andrey V. Protchenko, Simon Aldridge and Philip Mountford*

Chemistry Research Laboratory, University of Oxford, Mansfield Road, Oxford, OX1 3TA, U.K.

[email protected]

The chemistry of metal–nitrogen multiple bonds has been of interest to researchers for

more than thirty years.1 In particular, the synthesis and small molecule reactivity of

Group 4 imido and hydrazido complexes, (L)M=NR and (L)M=NNR2, have been

extensively studied within this research group and by others,2

and more recently, the

first titanium alkoxyimido complexes, (L)Ti=NOR, were reported.3

In contrast, only five examples of the related borylimides, (L)M=NBR2, have been

reported across all of the transition metals to date,4 and we here present the first

systematic pursuit of borylimido complexes, namely of titanium, using the borylamine

H2NB{N(Ar')CH}2 (Ar' = 2,6-C6H3iPr2).

To this end, we report two novel routes into titanium borylimido chemistry, and the

subsequent installation of a varied range of supporting ligand sets (see Figure 1), which

have allowed us to begin the first reactivity studies of this functional group.

Figure 1: Examples of new titanium borylimido complexes.

References.

[1] For example: Acc. Chem. Res., 2005, 38, 955; Acc. Chem. Res., 2005, 38, 839; J.

Am. Chem. Soc., 2001, 123, 2923.

[2] For example: Organometallics, 2011, 30, 1182; J. Am. Chem. Soc., 2004, 126, 1794;

Organometallics, 2009, 28, 4747.

[3] Chem. Commun., 2011, 47, 4926; Inorg. Chem. 2011, 50, 12155; Organometallics,

2013, 32, 7520.

[4] Polyhedron, 1993, 12, 1061; Z. Anorg. Allg. Chem., 2003, 629, 744; Angew. Chem.

Int. Ed. Engl., 2002, 41, 3709-3712; J. Am. Chem. Soc., 2014, 136, 8197.

Page 45: Royal Society of Chemistry Coordination and Organometallic

BIMETALLIC GROUP 4 COMPLEXES AS OLEFIN

POLYMERISATION CATALYSTS

R. A. Collins,a A. F. Russell,

a A. Berthoud,

b R. T. W. Scott

b and P. Mountford*

,a

a Chemistry Research Laboratory, Department of Chemistry, University of Oxford,

Mansfield Road, Oxford, OX1 3TA, UK. [email protected]

b LANXESS Elastomers B.V., Global R&D, P.O. Box 1130, 6160 BC Geleen, The Netherlands

LANXESS Elastomers have introduced (as Keltan ACETM) a new class of к1

-amidinate

complexes of the type (η-C5R5)M{NC(Ar)NR'2}X2 (X = Me or Cl; M = Ti, Zr or Hf),

which are extremely active pre-catalysts for the commercial homo- and co-

polymerization of olefins.[1,2]

Bimetallic catalysts can have beneficial cooperative effects in olefin polymerisation,

which arise from the complementary interaction of the polymeric chain between the

adjacent metal centres. This often results in an increased molecular weight, chain

branching and incorporation of alternate monomeric species. [3]

Here we report new bimetallic catalysts with different bridging moieties and compare

their catalytic activity and the resultant polymer properties to that of their mononuclear

analogues.

Figure 1. Solid state structures of representative mononuclear and bimetallic cyclopentadienyl-amidinate

complexes

[1] Ijpeij, E. G.; Coussens, B.; Zuideveld, M. A.; van Doremaele, G. H. J.; Mountford,

P.; Lutz, M.; Spek, A. L. Chem. Commun. 2010, 46, 3339.

[2] Ijpeij, E. G.; Windmuller, P. J. H.; Arts, H. J.; Van, d. B. F.; Van, D. G. H. J.;

Zuideveld, M. A. WO2005090418A1 2005.

[3] Delferro, M.; Marks, T. J. Chem. Rev. 2011, 111, 2450.

Page 46: Royal Society of Chemistry Coordination and Organometallic

ON THE ROAD TO TETRAZOLATE BASED NITROGEN RICH

MATERIALS

Peter Portius and Ben Crozier

Dainton Building, Brook Hill, Sheffield, S3 7HF

[email protected]

My research is focussed on the synthesis of homoleptic tetrazolato- main group

complexes of the type [E(N4CH)6]2-

(where E= Si, Ge and Sn). Such complexes are

unknown for p-block elements, while only certain transition metal examples exist, such

as [ML6](BF4)2 where M = Mn, Co, Cu, Zn and L = 1-methyltetrazole, 1-ethyltetrazole

and 1-propyltetrazole. 1 The appeal of the tetrazolate ligand is its high nitrogen content

which imparts this class of compounds with a high nitrogen content and a high energy

density. This characteristic originates within the strength of the dinitrogen triple bond.

Compounds that contains the weaker N-N and N=N bonds strain against an enthalpic

drive to exist as dinitrogen. Therefore, thermally activated decomposition of these

compounds is accompanied by a large release of heat and the formation of mostly N2

gas. It is for this reason that nitrogen-rich energetics are seen as a more environmentally

friendly alternative to conventional energetic materials. 2

The novel homoleptic anionic polytetrazolate complexes [Si(N4CH)6]2-

and

[Ge(N4CH)6]2-

were prepared recently in our laboratory. 3 The next step was an

extension of the reaction scheme to the related heavier analog [Sn(N4CH)6]2-

.

Employing ligand exchange reactions, the reaction was found to stop at the stage of

(PPN)2[Sn(N4CH)3Cl3]. For this reason, alternative tetrazoles transfer reagents were

investigated, amongst which trimethylsilyltetrazole, TMS-N4CH, 4

was found to be

highly promising. TMS-N4CH is readily accessible. It is synthetically highly versatile as

it relies on the formation of TMS-F in the exchange reaction with the relative strength

of the Si-F bond providing the driving force. In a one-pot synthesis SnF4, TMS-N4CH

and PPN(N4CH) are combined and react to produce the fully exchanged

(PPN)2[Sn(N4CH)6]. TMS-N4CH also holds promise for generating low-valent

tetrazolates. Combination with SnF2 in dry pyridine has yielded a co-ordination polymer

with a repeating unit of {Sn(N4CH)2(py)}. 5

This is the first example of a tetrazolate-

bridge main group co-ordination polymer. From each 2D planar sheet pyridine rings

extend out of the plane and pi-pi stack with pyridine bound to an opposing sheet,

producing a bilayer in the extended structure. This is a highly interesting structure as it

suggests a 3D metal organic framework (MOF) utilising a tin-tetrazolate core could be

synthesised. Such a system could preferentially adsorb CO2.6 In addition with a suitably

nitrogen rich organic linker, it would be a rare example of an energetic MOF. 7

[1] J. Kusz, H. Spiering, and P. Gütlich, J. Appl. Crystallogr., 2004, 37, 589.

[2] G. Steinhauser and T. M. Klapötke, Angew. Chem. Int. Ed., 2008, 47, 3330.

[3] P. Portius, B. F. Crozier, In preparation, 2015.

[4] L. Birkofer et al., Chem. Ber., 1963, 96, 2750.

[5] P. Portius, B. F. Crozier, In preparation, 2015.

[6] R. Banerjee et al., Cryst. Growth Des., 2011, 11, 5176.

[7] M Shreeve et al., Angew. Chem. Int. Ed., 2014, 53, 2540.

Page 47: Royal Society of Chemistry Coordination and Organometallic

STOICHIOMETRIC AND CATALYTIC C-F BOND ACTIVATION

BY TRANS-DIHYDRIDE RUTHENIUM NHC COMPLEXES

Mateusz K. Cybulski, Mary F. Mahon, Ian M. Riddlestone and Michael K. Whittlesey

Department of Chemistry, University of Bath, Claverton Down, Bath BA2 7AY, U.K.

[email protected]

In light of our previous work on catalytic hydrodefluorination (HDF) of fluorinated

aromatics using ruthenium N-heterocyclic (NHC) dihydride precursors and a well

understood mechanism involving nucleophilic attack by metal-bound hydride ligands

investigated with DFT calculations [1], we now present our recent findings on

stoichiometric and catalytic C-F activation by Ru NHC complexes containing

increasingly more nucleophilic hydride ligands. The mixed phosphine-carbene species,

Ru(IEt2Me2)2(PPh3)2H2 (1), proved to be a viable precursor for HDF, converting C6F6 to

a mixture of tri, di and monofluorobenzenes over several days at 90°C with 10 mol%

catalyst and Et3SiH as the reductant [2]. The employment of the tetracarbene species,

Ru(IMe4)4H2 (2), allowed for milder reaction conditions, affording selective formation

of 1,2,4,5-tetrafluorobenzene within minutes at room temperature.

References.

[1] Reade, S. P.; Mahon, M. F.; Whittlesey, M. K. J. Am. Chem. Soc. 2009, 131, 1847;

Panetier, J. A.; Macgregor, S. A.; Whittlesey, M. K. Angew. Chem. Int. Ed. 2011, 50,

2783; Macgregor, S. A.; McKay D.; Panetier, J. A.; Whittlesey, M. K. Dalton Trans.

2013, 42, 7386

[2] Cybulski, M. K.; Riddlestone, I. M.; Mahon, M. F.; Woodman, T. J.; Whittlesey, M.

K. Dalton Trans. 2015, DOI: 10.1039/c5dt01996f

Page 48: Royal Society of Chemistry Coordination and Organometallic

ALKALINE EARTH ORGANOHYDROBORATE COMPLEXES FOR THE RING-OPENING POLYMERISATION OF RAC-

LACTIDE

Nichabhat Diteepeng, Junjuda Unruangsri, Insun Yu and Philip Mountford*

Chemistry Research Laboratory, Department of Chemistry, University of Oxford

Mansfield Road Oxford OX1 3TA UK. [email protected]

Polymers from renewable resources such as polylactide (PLA) have gained great

attraction from both commercial and academic perspectives because of their

environmental advantages and various applications ranging from drug delivery to

packaging materials. The preparation of well-characterised metal complexes employed

as ring-opening polymerisation (ROP) catalysts has been a focus in both industrial and

academic research.1

The hydride transfer and reducing potential of metal organohydroborate complexes

allows them to be candidates for rac-LA ROP initiators. To date, no alkaline earth

organohydroborate complexes have been established as ROP initiators.2 Here we report

mechanistic investigations and polymerisation studies of the ROP of rac-LA using

alkaline earth organohydroborate complexes (HBEt3) supported by 3-methyl,5-tert-

butyl tris(pyrazolyl)hydroborate ligands (TptBu,Me

) (1-2, Figure 1)3 and the synthesis and

characterisation of a new calcium alkoxide complex derived from the hydride transfer

between TpCaHBEt3(THF) and benzophenone (3, Figure 1).

Figure 1: New alkaline earth initiators (1 – 3) for the ROP of rac-LA.

References

[1] Recent reviews: Wheaton, C. A.; Hayes, P. G.; Ireland, B. J. Dalton Trans. 2009, 25, 4832, Platel, R. H.; Hodgson, L. M.; Williams, C. K. Polym. Rev. 2008, 48,

11, Dechy-Cabaret, O.; Martin-Vaca, B.; Bourissou, D. Chem. Rev. 2004, 104,

6147, O'Keefe, B. J.; Hillmyer, M. A.; Tolman, W. B. J.Chem. Soc., Dalton Trans.

2001, 2215.

[2] See for example: Harvey, M. J.; Hanusa, T. P.; Pink, M. Chem. Commun. 2000, 6,

489, Pillai Sarish, S.; Jana, A.; Roesky, H. W.; Schulz, T.; John, M.; Stalke, D.

Inorg. Chem. 2010, 49, 3816, Krieck, S.; Görls, H.; Westerhausen, M. Inorg.

Chem. Commun. 2010, 13, 1466.

[3] Unruangsri, J.; Mountford, P. unpublished results.

Page 49: Royal Society of Chemistry Coordination and Organometallic

EXPLOITING THE POLYAMINE TRANSPORT SYSTEM TO

DELIVER THERANOSTIC RHENIUM(I) COMPLEXES

P. M. Cullis, M. P. Lowe, H. L. Parker and V. L. Emms.

Department of Chemistry, University of Leicester, University Road, Leicester, LE1 7RH

[email protected]

Polyamines are crucial to cell viability and play a key role in proliferation. Cancer cells

frequently exhibit an upregulated polyamine transport system; as such polyamine-drug

conjugates have been designed to selectively target cancer cells.1 Polyamines have been

shown to enter the cell via endocytosis, and be sequestered into acidic vesicles.2

Targeted Re(I) complexes have been developed for use as luminescent probes for

biological imaging.3 We have extended this approach to use the Re(I) tricarbonyl core

as a scaffold upon which to build a theranostic complex, incorporating cancer cell

specific delivery and controlled release of cytotoxic agents.

Bioconjugation of polyamines to the Re(I) tricarbonyl core is achieved using

diazotransfer4 and Cu(I) catalyzed azide-alkyne cycloaddition

5 to generate

pyridyltriazole (pyta) chelating ligands. Luminescent Re(I) tricarbonyl pyta complexes

have long emission lifetimes and a large Stokes shift, which makes them suitable for use

as biological probes. However, these complexes show weak emission when excited at

405 nm, a wavelength commonly used in confocal microscopy. The electronic structure

of the ligand has been modified to shift the absorption wavelength to generate more

emissive biological probes.

References. [1] Leblond, P.; Boulet, E.; Bal-Mahieu, C.; Pillon, A., Kruczynski, A.; Guilbaud, N.; Bailly, C.; Sarrazin,

T.; Lartigau, E.; Lansiaux A. and Meignan, S. Invest. New Drugs, 2014, 32, 883.

[2] Belting, M.; Mani, K.; Jönsson, M.; Cheng, F.; Sandgren, S.; Jonsson, S.; Ding, K. ; Delcros J.-G. and

Fransson, L.-A. J. Biol. Chem., 2003, 278, 47181.

[3] Coogan M. P. and Fernández-Moreira, V. Chem. Commun., 2014, 50, 384.

[4] Ethan D. Goddard-Borger, E. D. and Stick, R. V. Org. Lett., 2007, 9, 3797.

[5] Rostovtsev, V. V.; Green, L. G.; Fokin, V. V. and Sharpless, K. B. Angew. Chem. 2002, 114, 2708.

Page 50: Royal Society of Chemistry Coordination and Organometallic

GROUP IV PERMETHYLPENTALENE COMPLEXES AS

ETHYLENE POLYMERISATION CATALYSTS

D. A. Fraser, F. Mark Chadwick, J.-C. Buffet and D. O'Hare*

Chemistry Research Laboratory, University of Oxford, 12 Mansfield Road, Oxford.

OX1 3TA [email protected]

Polyethylene accounts for nearly 65% of total annual polyolefin output, with global production over 70 million tonnes per year.[1,2] Although heterogeneous Ziegler-Natta systems define the industry standard, large molecular weight distributions and uneven monomer incorporation restricts application of the resultant polymer. Homogeneous catalysts meanwhile often give far more desireable polymer properties. Following the synthesis of the permethylpentalenyl dianion (Li2[Pn*], Pn* = C8Me6)

[3] a series of new titanium and zirconium complexes were synthesised and their activity for the polymerisation of ethylene in solution were tested.[4] They were considered “Very active” on the Gibson scale[5]

Here, we present a more detailed study of the solution phase polymerisation of ethylene including mechanistic studies. Variable temperature NMR studies were carried out to probe the early stages of activation following addition of AlMe3. References.

[1] Severn, J. R.; Chadwick, J. C.; Duchateau, R.; Friederichs, N., Chem. Rev., 2005, 105, 4073 [2] Piringer, O. G.; Baner, A. L. Plastic Packaging: Interactions with Food and Phatmaceuticals;

2nd Ed.; Wiley, 2008 [3] a) A. E. Ashley, A. R. Cowley, D. O’Hare, Chem. Comm., 2007, 1512 b) A. E, Ashley, a. R.

Cowley, D. O’Hare, Eur. J. Org. Chem., 2007, 26, 2239 [4] F. M. Chadwick; R. T. Cooper; A. E. Ashely; J.-C. Buffet; D. M. O’Hare, Organometallics,

2014, 33(14), 3775 – 3785 [5] G. J. P. Britovsek, V. C. Gibson, D. F. Wass, Angew. Chem. Int. Ed. Engl., 1999, 38(4), 428-447

Fig. 1: Representation and solid state molecular structure of Pn*ZrClCp. Ellipsoids at 50%

Page 51: Royal Society of Chemistry Coordination and Organometallic

HYDROPHOSPHINATION USING AN AIR-STABLE IRON PRE-

CATALYST

K. J. Gallagher and R. L. Webster*

Department of Chemistry, University of Bath, Claverton Down, Bath, BA2 7AY, UK [email protected]

The synthesis of phosphines using a 100% atom-economic transformation is an attractive process, not least because phosphorus compounds in the +3 oxidation state are important ligands for transition metal catalysis. They also make excellent organocatalysts and reagents in organic synthesis.

Hydrophosphination reactions are most commonly catalysed by Group 10 (Ni, Pd, Pt), Group 3 (Y) and lanthanide (La, Sm, Yb) metal salts.1 More recently, iron salts have also been used, a particularly appealing approach due to the vast abundance and low toxicity of the metal.2

Although the use of iron salts is operationally simple, catalyst design is needed to further advance iron catalysed hydrophosphination. By introducing an easily accessible salen ligand we have significantly improved functional group tolerance, lowered catalyst loading and have shown that the reaction can proceed efficiently under milder reaction conditions than those reported using iron chloride salts.2,3 These simple phosphine products have also been shown to be active for iron-catalysed Negishi coupling.

This is the first time a designed iron catalyst has been used for hydrophosphination, these results, along with comprehensive mechanistic studies, will be presented.

[1] L. Rosenberg, ACS Catal. 2013, 3, 2845-2855. [2] L. Routaboul, F. Toulgoat, J. Gatignol, J.-F. Lohier, B. Norah, O. Delacroix, C. Alayrac, M. Taillefer and A.-C. Gaumont, Chem. Eur. J. 2013, 19, 8760-8764. [3] K. J. Gallagher, R. L. Webster, Chem. Commun. 2014, 50, 12109-12111

Page 52: Royal Society of Chemistry Coordination and Organometallic

TANTALUM HYDRAZIDE COMPLEXES WITH A TRIDENTATE

TRIANIONIC LIGAND

Karen Y. Gamero-Vega, Richard A. Collins and Philip Mountford*

Department of Chemistry, University of Oxford, OX1 3TA, UK

[email protected]

The chemistry of metal-nitrogen multiple bond has been an ongoing area of extensive

interest. The study of imido and hydrazido complexes of Group 4 and 6 metals is well

established and many applications have been reported.[1]

The synthesis, properties and

applications of the tridentate trianionic pincer have been explored with high valent

metals over the past decade.[2]

The aim of this project is to synthesise the first Group 5 hydrazido complex with a

tridentate trianionic ligand and explore any reactivity exhibited with small molecules.

Three different synthetic routes to this compound, using 2,2’-bis(trimethylsilylamino)diphenylamine

[3,4] as a ligand backbone, are reported and the

chemistry of the synthetic intermediates explored.

[1] Gade, L. H.; Mountford, P. Coordination Chemistry Reviews 2001, 216-217, 65-

97. Pickett, C. J. J. Biol. Inorg. Chem., 1996, 1, 601. Hazari, N.; Mountford, P.

Acc. Chem. Res. 2005, 38, 839. Haak, E.; Bytschkov, I.; Doye, S. Angew. Chem.,

Int. Ed., 1999, 38, 3389.

[2] O’Reilly, M. E.; Veige, A. S. Chem. Soc. Rev. 2014, 43, 6325.

[3] Ren, P.; Vechorkin, O.; von Allmen, K.; Scopelliti, R.; Hu, X. J. Am. Chem. Soc.

2011, 133, 7084–7095.

[4] Schrock, R. R.; Lee, J.; Liang, L.; Davis, W. M. Inorg. Chim. Acta. 1998, 270,

353-362.

Page 53: Royal Society of Chemistry Coordination and Organometallic

SYNTHESIS OF IRIDIUM COMPLEXES CONTAINING N-

HETEROCYCLIC CARBENE-BASED PINCER LIGANDS

Lucero González-Sebastin and Adrian B. Chaplin*

Department of Chemistry, University of Warwick, Coventry CV4 7AL, UK

[email protected]

N-heterocyclic carbenes (NHCs) are of growing interest in organometallic chemistry

and catalysis as alternatives to widely used phosphine ligands.1 In addition to their

strong donor properties, inclusion into rigid mer-tridentate ‘pincer’ ligand architectures

is of great interest for chemically challenging applications requiring high thermal

stability, such as alkane dehydrogenation.2 Our research has focused on developing a

new series of NHC-based pincer ligands containing imidazolinylidene donor groups

(see below). These CCC-pincer ligands were prepared from 1,3-diaminobenzene

through standard multistep synthetic routes.3 The coordination chemistry of these

ligands with various iridium precursors is currently being pursued and will be

discussed. Ongoing and future research aims to evaluate iridium derivatives in

hydrocarbon activation reactions.

[Ir]

R =

[Ir]N

N

N

N

R

R

2(Cl)

N

N

N

N

R

R

References 1. Pugh, D.; Danopoulos, A. A., Coord. Chem. Rev. 2007, 251, 610-641.

2. Chianese, A. R.; Drance, M. J.; Jensen, K. H.; McCollom, S. P.; Yusufova, N.;

Shaner, S. E.; Shopov, D. Y.; Tendler, J. A., Organometallics 2014, 33, 457-464.

3. Rubio, R. J.; Andavan, G. T. S.; Bauer, E. B.; Hollis, T. K.; Cho, J.; Tham, F. S.;

Donnadieu, B., J. Organomet. Chem. 2005, 690, 5353-5364.

Page 54: Royal Society of Chemistry Coordination and Organometallic

ALKALINE EARTH-TRANSITION METAL COMPLEXES

Ross Green, Alicia Walker, Matthew P. Blake and Philip Mountford*

Chemistry Research Laboratory, University of Oxford, 12 Mansfield Road,

Oxford OX1 3TA, UK.

[email protected]

Molecular compounds containing metal-metal bonds have long been of interest in

inorganic chemistry.1 Although there are numerous examples of compounds with bonds

between transition metal (TM) elements, there exist relatively few between transition

metals and the alkaline earth (Ae) elements.2

Our group has recently reported several compounds containing Ae–TM bonds. The

reaction of the β-diketiminate magnesium species Mg(NacNac)I(THF) with KFp (Fp =

CpFe(CO)2) afforded Mg(NacNac)Fp(THF) (Mg–Fe = 2.6326(4) Å).3 This was shown

to react with TolNCNTol to afford Mg(NacNac){(NTol)2CFp}, the first net insertion of

an unsaturated substrate into an alkaline earth-transition metal bond.

We have recently developed amidinate species of the type Mg{RC(NR')2}M'(THF)

(M' = CpFe(CO)2, CpRu(CO)2 or Co(CO)3(PCy3), R = Mes, R' = iPr, Dipp or Mes),

which feature fewer carbon atoms in the ligand backbone and yield a more sterically

open system which may have the potential for increased reactivity with respect to the

analogous system. In addition, the guanidinate species [Mg{Me2NC(NDipp)2}Fp]2 has

been synthesised which features two Mg–Fe bonds (Mg–Fe = 2.5279(4) Å). These are

the shortest known Mg–Fe bonds to date.

Figure 1. Mg–TM bonded complexes featuring β-diketiminate, amidinate and

guanidinate ligands.

[1] (a) F. A. Cotton, C. A. Murillo and R. A. Walton, Multiple Bonds Between Metal Atoms; 3rd

ed.;

Springer: New York, 2005. (b) S. T. Liddle (ed.), Molecular Metal-Metal Bonds; Wiley:

Weinheim, Germany, 2015. (c) L. H. Gade, Angew. Chem. Int. Ed., 2000, 39, 2658.

[2] For example: (a) H. Braunschweig, K. Gruss and K. Radacki, Angew. Chem. Int. Ed., 2009, 48,

4239. (b) H. Felkin, P. J. Knowles, B. Meunier, A. Mitschler, L. Ricard and R. Weiss, J. Chem.

Soc. Chem. Commun., 1974, 44. (c) K. Jonas, G. Koepe and C. Krüger, Angew. Chem. Int. Ed.,

1986, 25, 923. (d) H. Deng and S. G. Shore, J. Am. Chem. Soc., 1991, 113, 8538.

[3] M. P. Blake, N. Kaltsoyannis, P. Mountford, Chem. Commun., 2013, 49, 3315.

Page 55: Royal Society of Chemistry Coordination and Organometallic

MECHANISTIC STUDY OF OF ANHYDRIDE FORMATION

FROM RHODIUM(III) ACETYL COMPLEXES AND

CARBOXYLATES

D. Griffin and A. Haynes

Department of Chemistry, University of Sheffield, Sheffield, S3 7HF, UK,

[email protected]

Product formation in rhodium-catalysed methanol carbonylation is conventionally

proposed to proceed via reductive elimination of acetyl iodide from

[Rh(COMe)(CO)2I3]- and subsequent hydrolysis to form acetic acid.

1 Evidence has

recently been reported for an alternative mechanism involving replacement of an I-

ligand by acetate, formed in situ, and subsequent reductive elimination of acetic

anhydride which is then hydrolysed.2

The reactions of chelate complexes [Rh(COMe)(L-L)I2] (1a-d) with carboxylates have

been studied to elucidate the mechanism of anhydride formation. Rapid substitution of I-

by acetate occurs to give 2a-d, followed by relatively slow reductive elimination of

acetic anhydride. Intermediates 2a-c were observed in situ by IR and NMR

spectroscopy and 2d has been isolated and characterised by X-ray crystallography.

References

[1] Haynes, A. Adv. Catal. 2010, 53, 1–45 and references therein.

[2] (a) Lassauque, N.; Davin, T.; Nguyen, D. H.; Adcock, R. J.; Coppel, Y.; Le

Berre, C.; Serp, P.; Maron, L.; Kalck, P. Inorg. Chem. 2012, 51, 4–6.

(b) Nguyen, D. H.; Lassauque, N.; Vendier, L.; Mallet-Ladeira, S.; Le Berre, C.;

Serp, P.; Kalck, P. Eur. J. Inorg. Chem. 2014, 326–336.

Page 56: Royal Society of Chemistry Coordination and Organometallic

FUNCTIONALISATION OF METAL-ORGANIC FRAMEWORK

UiO-66-NH2 VIA POST-SYNTHETIC MODIFICATION

A. D. Burrows, H. Amer Hamzah

Department of Chemistry, University of Bath, Claverton Down, Bath BA2 7AY, UK.

[email protected], [email protected]

Metal-organic frameworks (MOFs) are a relatively new class of porous materials which

find applications in gas storage, catalysis, drug delivery and sensors. MOFs can be post-

synthetically modified to introduce complex functionalities in the pores thus altering

their physical and chemical properties. In this context, post-synthetic modification

(PSM) is defined as a process in which a pre-formed MOF is transformed into a new

MOF via crystal-to-crystal transformation.1 PSM is of interest as it can allow for the

formation of functionalised MOFs which cannot be formed via direct synthesis.

We report herein the functionalisation of an amine-containing MOF [Zr6O4(OH)4(BDC-

NH2)6] (BDC-NH2 = 2-amino-1,4-benzenedicarboxylate), UiO-66-NH2 via PSM. An α-

amino phosphonate moiety has been partially incorporated into the pores of UiO-66 via

a one pot, three component reaction (Scheme 1). Tandem PSM was achieved by further

reaction of the unreacted amino groups with acetic anhydride. Crystallinity and porosity

were maintained during the transformations. The physical and chemical properties (e.g.

thermal stability, BET surface area and CO2/N2 selectivity) of the modified MOF have

been compared to the pristine UiO-66-NH2.

Scheme 1. PSM of UiO-66-NH2 via Kabachnik-Fields reaction.

Reference

[1] Wang, Z.; Tanabe, K. K.; Cohen, S. M. Inorg. Chem. 2009, 48, 296-306.

Page 57: Royal Society of Chemistry Coordination and Organometallic

NOVEL SINGLE SOURCE PRECURSORS FOR THE DEPOSITION

OF TiSe2 FILMS Samantha L. Hawken,

a Andrew L. Hector,

a Marek Jura,

b William Levason,

a Gillian

Reida and Gavin Stenning

b

aSchool of Chemistry, University of Southampton, Southampton SO17 1BJ, United

Kingdom

Email: [email protected]

bSTFC, ISIS, Harwell Innovation Campus, Didcot, Oxfordshire OX11 0QX, United

Kingdom

Properties including superconductivity, exhibiting charge density waves and a tuneable

band gap make transition metal dichalcogenide materials exciting prospects for many

applications.1 The pseudo octahedral complexes [TiCl4{

nBuSe(CH2)nSe

nBu}] (n= 2,3)

have been synthesized by the reaction of TiCl4 with nBuSe(CH2)nSe

nBu (1:1) and

characterised by IR, multinuclear 1H,

13C{

1H} and

77Se{

1H} NMR spectroscopy and

microanalysis. Low pressure chemical vapour deposition (LPCVD) using the complexes

show that they function effectively as single source precursors (SSP) for the deposition

of crystalline TiSe2 thin films on SiO2 substrates. The hexagonal 1T-TiSe2 films,

characterised by X-ray diffraction patterns, show a high degree of stability in air when

prepared from these precursors. The analysis revealed that both precursors showed

equal promise as LPCVD SSP, hence only the C2 linked complex was used in further

experiments. The dependence on temperature and precursor quantity of the orientation,

film thickness and roughness is shown through a number of deposition experiments.

1 M. Chhowalla, H. S. Shin, G. Eda, L.-J. Li, K. P. Loh and H. Zhang, Nat. Chem., 2013, 5,

263–75.

Page 58: Royal Society of Chemistry Coordination and Organometallic

ONE-STEP, MODULAR ROUTE TO OPTICALLY ACTIVE

DIPHOS LIGANDS

E. Louise Hazeland and Paul G. Pringle

School of Chemistry, University of Bristol, Cantock’s Close, Bristol BS8 1TS, U.K. [email protected]

Efficient asymmetric hydrogenation using metal-phosphine catalysts is one of the

success stories of modern chemistry with multiple industrial applications.1 The field is

over 40 years old but challenges still remain. For example, there are potentially

commercially valuable substrates for which highly enantioselective catalysts have not

yet been discovered. One-step automated routes to chiral monophos ligands have been a

major development since 2000 and led to the application by DSM of high-throughput

ligand development.2 Chiral diphos ligands have advantages over monophos ligands for

several catalytic processes, but their synthesis is often multi-step and laborious.

Previously, single atom-backbone chiral diphos ligands3,4

with C2- or C1-symmetry

have been shown to be very effective for asymmetric hydrogenation. We have exploited

our discovery of the one-step reaction (Scheme 1) to synthesise a range of chiral diphos

ligands.5

Specifically, this presentation will focus on the exploitation of this route to synthesise

a range of novel C1-symmetric, C1-backbone, diphos ligands such as those shown in

Scheme 2, their application in asymmetric hydrogenation and their potential for the

application of high-throughput catalyst discovery methods.

[1] de Vries, J. G.; Elsevier C. J., The Handbook of Homogeneous Hydrogenation,

2008, Wiley-VCH. [2] Lefort, L.; Boogers, J. F.; deVries, A. M.; deVries, J., Top. Catal. 2006, 40, 185-191. [3] Hoge, G.; Wu, H.-P.; Kissel, W. S.; Pflum, D. A.; Greene,

D. J.; Bao, J., J. Am. Chem. Soc. 2004, 126, 5966-5967. [4] Gridnev, I. D.; Imamoto, T.;

Hoge, G.; Kouchi, M.; Takahashi, H., J. Am. Chem. Soc. 2008, 130, 2560-2572. [5]

Hazeland, E. L.; Chapman, A. M.; Pringle, P. G.; Sparkes, H. A., Chem. Commun.

2015, 51, 10206-10209.

= optically active

phosphacycle

Scheme 1: General one-step route where X = C, N; R = Ph, tBu,

iPr, Cy and R’ = Me, Ph.

Scheme 2: One-step route to diphos ligands

Page 59: Royal Society of Chemistry Coordination and Organometallic

-COMPLEXES OF COPPER(I) WITH MAIN GROUP HYDRIDES

(M = Mg, Zn)

A. Hicken and M. R. Crimmin

Department of Chemistry, Imperial College, South Kensington, London, SW7 2AZ, UK

[email protected]

We have recently documented the synthesis of a series of σ–complexes of copper in

which M–H (M = Al, Zn) and E–H (E = B) σ–bonds coordinate to a copper(I)

fragment.1,2

Here we report a series of fluorinated copper(I) complexes (1-3), prepared

by the reaction of copper mesityl with nac-nac and ac-nac proligands. We show that the

hexafluorinated complex 22toluene crystallises as an inverse-sandwich.3 In solution,

displacement of the arene from 1-3 with molecular magnesium and zinc hydrides has

allowed the generation of six new σ–complexes of copper(I), including the first

examples that incorporate a Mg–H bond.

References

(1) Nako, A. E.; White, A. J. P.; Crimmin, M. R. Dalton Trans. 2015, 44, 12530.

(2) Nako, A. E.; Tan, Q. W.; White, A. J. P.; Crimmin, M. R. Organometallics 2014,

33, 2685.

(3) Badiei, Y. M.; Dinescu, A.; Dai, X.; Palomino, R. M.; Heinemann, F. W.; Cundari,

T. R.; Warren, T. H. Angew. Chem., Int. Ed. 2008, 120, 10109.

Page 60: Royal Society of Chemistry Coordination and Organometallic

A TETHERED NHC-CARBORANE LIGAND: SELECTIVE C-H

ACTIVATION AND VERSITILE COORDINATION

Dr. C. E. Willans and J. Holmes

School of Chemistry, University of Leeds, LS2 9JT [email protected]

The development of new catalysts and catalytic processes is essential for a

sustainable and green future. Ligand design is central to these developments as they

control the activity and reactivity of a metal centre. Whilst modifying current ligands

may induce incremental changes in the outcome of a reaction, brand new ligand

architectures can lead to more diverse pathways and processes. The fusion of two very

different ligand families namely N-Heterocyclic carbenes (NHCs)[1]

and carboranes[2]

,

gives rise to a brand new ligand architecture; dicarba-dodecaboranes bearing

imidazolium tethers. This ligand system exhibits unprecedented C-H activation with

divergent reactivity towards cyclometalation leading to versatile coordination to a RhI

centre.

Figure 1. Synthesis of a tethered NHC-carborane ligand and its subsequent coordination to Rh

I

[1] Hopkinson, M. N.; Richter, C.; Schedler, M.; Glorius, F. Nature 2014, 510, 485.

[2] Grimes, R. N. Coord. Chem. Rev 2000, 200, 773.

CCH N N tBu CCHBr N NtBu + CCHN NtBu +N N tBu H-

Br-xs.

NucleophilicSubstitution Deboronation

CC

NNtBu

HRh ClC C

NNtBu

RhC

CHNN

tBuRh NCMe

RhAg2OMeCN

Ag2O DCM[Rh(COD)Cl]2i) NaH, THFii) [Rh(COD)Cl]2, MeCN

Page 61: Royal Society of Chemistry Coordination and Organometallic

SYNTHESIS AND CATALYTIC ACTIVITY OF OXO-

FUNCTIONALISED TRIAZOLYLIDENE RUTHENIUM(II)

COMPLEXES

C. Johnson and M. Albrecht

Departement für Chemie und Biochemie, Universität Bern, Freiestrasse 3, CH-3012

Bern, Switzerland

[email protected]

In recent years, mesoionic 1,2,3-triazolylidenes have emerged as a highly versatile

subclass of N-heterocyclic carbene (NHC) ligands.1 This NHC scaffold can be

effectively tailored to specific functions as a consequence of the flexibility of the [3 + 2]

cycloaddition of alkynes with azides. This feature, coupled with the ligands’ stong σ-

donor abilities have led to their diverse application in catalytic transformations.2

Here we present the synthesis of a range of oxo-functionalised triazolylidene

ruthenium(II) complexes. The implications of this C,O-bidentate motif is discussed in

terms of catalytic activity.

N N

NMes

ORuCl

References:

[1] Mathew, P.; Neels, A.; Albrecht, M. J. Am. Chem. Soc. 2008, 130, 13534.

[2] Donnelly, K. F.; Petronilho, A.; Albrecht, M. Chem. Commun. 2013, 49, 1145.

Page 62: Royal Society of Chemistry Coordination and Organometallic

DEVELOPMENT OF CHELATORS FOR 89

Zr IN

POSITRON EMISSION TOMOGRAPHY IMAGING

Boon-Uma Jowanaridhi a,b

, Johanna Seemannb, Benjamin P. Burke

a,b and Stephen J.

Archibald a,b

* aDepartment of Chemistry,

bPositron Emission Tomography Research Centre

University of Hull, Cottingham Road, Hull, HU6 7RX, UK.

[email protected]

Positron emission tomography (PET) is used for the non-invasive study of dynamic

processes.1 Immuno-PET is a relatively young discipline used to improve visualisation

by combining the high sensitivity of PET with the specificity of monoclonal antibodies.

It allows images to be collected after the highly specific accumulation of radioactive

compound in the target organ and the clearance of unbound tracer from the blood pool,

which increases target-to-background activity ratio.2 Zirconium-89 is an ideal immuno-

PET radionuclide as it decays with a half-life of 78.41 h via positron emission and

therefore allows tracking of the relatively slow accumulating antibodies over several

days.3 DFO is the most common ligand used as

89Zr-chelate. However, it has been

reported that 89

Zr-DFO-mAb conjugates show significant uptake of radioactivity in

bone of mice which indicates release of the radiometal.4

Recently, the development of a more suitable chelate for 89

Zr has been gaining

attention. The rigid structure of macrocycles, such as cyclen or cyclam can be utilised as

stabilising platform offering preorientation for attached donating groups. In a first

instance, cyclen-based ligands were synthesised and functionalised with various types

of pendant arms containing hydroxamate, phosphonic acid, phosphinic acid, picolinic

acid, kojic acid and acrylamide as donating moieties. Radiolabelling with 89

Zr was

evaluated using variations in temperature, pH, type and concentration of buffer, as well

as concentration of ligands. 6,6',6'',6'''-((1,4,7,10-Tetraazacyclododecane-1,4,7,10-

tetrayl)tetrakis(methylene))-tetrapicolinic acid could be labelled to 39.8 % RCY using a

temperature of 90 oC and 0.2 M NaOAc at pH 4 with a ligand concentration of 20 µM.

In an additional experiment the ligand was labelled with another PET radioisotope, 68

Ga, to 91.2 % RCY in 3 min and the stability tested against apotransferrin with only 4

% of tracer decomposed over 2 h.

References. 1. Van Der Veldt, A.; Smit, E.; Lammertsma, A. A., Positron emission tomography as a method for

measuring drug delivery to tumors in vivo: the example of [11C]docetaxel. Frontiers in

Oncology 2013, 3.

2. Deri, M. A.; Ponnala, S.; Zeglis, B. M.; Pohl, G.; Dannenberg, J. J.; Lewis, J. S.; Francesconi, L.

C., Alternative Chelator for 89Zr Radiopharmaceuticals: Radiolabeling and Evaluation of 3,4,3-

(LI-1,2-HOPO). Journal of Medicinal Chemistry 2014, 57 (11), 4849-4860.

3. van de Watering, F. C. J.; Rijpkema, M.; Perk, L.; Brinkmann, U.; Oyen, W. J. G.; Boerman, O.

C., Zirconium-89 Labeled Antibodies: A New Tool for Molecular Imaging in Cancer Patients.

BioMed Research International 2014, 2014, 13.

4. Holland, J. P.; Divilov, V.; Bander, N. H.; Smith-Jones, P. M.; Larson, S. M.; Lewis, J. S., 89Zr-

DFO-J591 for ImmunoPET of Prostate-Specific Membrane Antigen Expression In Vivo. Journal

of Nuclear Medicine 2010, 51 (8), 1293-1300.

Page 63: Royal Society of Chemistry Coordination and Organometallic

REACTIVITY OF THE 2-PHOSPHAETHYNOLATE ANION

AND PHOSPHINECARBOXAMIDE

Jose M. Goicoechea and Andrew R. Jupp

Department of Chemistry, University of Oxford,

Chemistry Research Laboratory, 12 Mansfield Road, Oxford, OX1 3TA, U.K

[email protected]

We recently reported the synthesis of the 2-phosphaethynolate anion, PCO–, by an

unprecedented activation of the group 15 Zintl cluster, P73–

.[1]

This phosphorus-

containing analogue of the ubiquitous cyanate anion has been the subject of a number of

recent articles by our group and those of Grützmacher[2]

and Cummins.[3]

We have

systematically explored the cycloaddition chemistry[1,4]

and ligand properties of this

remarkable anion compared to its isoelectronic counterparts.[5]

Furthermore, by analogy with Wöhler’s paradigm-shifting synthesis of urea in 1828, the

reaction of PCO– with ammonium salts yields the unprecedented

phosphinecarboxamide.[6]

This inorganic analogue of urea is a rare example of an air-

stable primary phosphine, and its ligand properties have been explored.[7]

We have also

explored its Brønsted acidity to afford a series of novel phosphides, secondary and

tertiary phosphines, and phosphine oxides.[8]

References

[1] Jupp, A. R.; Goicoechea, J. M. Angew. Chem., Int. Ed. 2013. 52, 10064.

[2] Puschmann, F. F.; Stein, D.; Heift, D.; Hendriksen, C.; Gal, Z. A.; Grützmacher, H.-

F.; Grützmacher, H. Angew. Chem. Int. Ed. 2011, 50, 8420.

[3] Krummenacher, I.; Cummins, C. C. Polyhedron, 2012, 32, 10.

[4] Heift, D.; Benkő, Z.; Grützmacher, H.; Jupp, A. R.; Goicoechea, J. M. Chem. Sci.

2015, 6, 4017.

[5] Jupp, A. R.; Geeson, M. B.; McGrady, J. E.; Goicoechea, J. M. manuscript

submitted.

[6] Jupp, A. R.; Goicoechea, J. M. J. Am. Chem. Soc. 2013, 135, 19131.

[7] Jupp, A. R.; Trott, G.; Payen de la Garanderie, É.; Holl, J. D. G.; Carmichael, D.;

Goicoechea, J. M. Chem. Eur. J. 2015, 21, 8015.

[8] Geeson, M. B.; Jupp, A. R.; McGrady, J. E.; Goicoechea, J. M. Chem. Commun.

2014, 50, 12281.

Page 64: Royal Society of Chemistry Coordination and Organometallic

TITANIUM ‘DOUBLE-SANDWICH’ COMPLEXES FOR THE REDUCTIVE ACTIVATION OF CO2

A. F. R. Kilpatrick, J. C. Green and F. G. N. Cloke

Chemistry Research Laboratory, Department of Chemistry, University of Oxford, Mansfield Road, Oxford OX1 3TA, UK [email protected]

Metalmetal bonds have been the subject of intense research, given that many

bimetallic compounds exhibit unusual electronic properties and novel reactivity. The

pentalene ligand, Pn = C8H6, displays a variety of coordination modes in its

organometallic complexes and can bind two metal centres with η5- hapticity in so-called

'double-sandwich' complexes.

We have recently reported the synthesis of a di-titanium bis(pentalene) complex,

Ti2(Pn†)2, Pn

† = 1,4-{Si

iPr3}2C8H4, which features a rare example of a TiTi double

bond.[1]

Ti2(Pn†)2 reductively splits CO2 under very mild conditions into O

2- and CO

complexes,[2]

and shows exceptional reactivity with heteroallenes CS2 and COS.[3,4]

Synthetic, structural and computational studies into the CO2 reaction mechanism are

presented, in addition to recent results exploring the reductive activation of other

unsaturated compounds by titanium double-sandwiches.

[1] Kilpatrick, A. F. R.; Green, J. C.; Cloke, F. G. N.; Tsoureas, N. Chem. Commun. 2013, 9434–9436.

[2] Kilpatrick, A. F. R.; Cloke, F. G. N. Chem. Commun. 2014, 50, 2769–2771.

[3] Kilpatrick, A. F. R.; Green, J. C.; Cloke, F. G. N. Organometallics 2015, DOI:

10.1021/acs.organomet.5b00315.

[4] Kilpatrick, A. F. R.; Green, J. C.; Cloke, F. G. N. Organometallics 2015, DOI:

10.1021/acs.organomet.5b00363.

Page 65: Royal Society of Chemistry Coordination and Organometallic

A NEW SMALL MOLECULE GELATOR AND 3D FRAMEWORK

LIGATOR OF LEAD(II)

Jane V. Knichal, William J. Gee, Andrew D. Burrows, Paul R. Raithby and Chick C.

Wilson

Univeristy of Bath Bath BA2 7AY

[email protected]

Supramolecular gels formed by low molecular weight gelators are a class of material in

which interest is rapidly expanding to provide a diverse range of applications to

commerce, industry and medicine1. Our continued work on uncommonly substituted

benzene dicarboxylate ligands in metal frameworks2 has lead us to one of the first lead

based metallogels to be reported. The solvothermal reaction of a new diacid allene,

(H2abd) with hydrated lead(II) acetate has yielded a metallogel with a critical gelation

point of 1% w/v in DMF. SEM studies confirm a worm-like morphology composed of

nanoscale fibres. Under non-solvothermal conditions, combining H2abd and hydrated

lead(II) acetate resulted in formation of single crystals, which were identified as a 3D

coordination polymer by X-ray diffraction. Structural features observed within this

framework provide the basis for assigning the molecular structure to the fibrils present

within the corresponding metallogel. This is a useful addition to the literature focusing

on understanding metal-based gel materials and their applications3.

References.

[1] A. Y.-Y. Tam and V. W.-W. Yam, Chem. Soc. Rev., 2013, 42, 1540

[2] (a) J. V. Knichal, W. J. Gee, A. D. Burrows, P. R. Raithby and C. C. Wilson, Cryst. Growth & Des., 2015, 15, 465-474 (b) J. V. Knichal, W. J. Gee, A. D. Burrows, P. R.

Raithby, S. J. Teat and C. C. Wilson, Chem. Commun., 2014, 50, 14436

[3] (a) A. Mallick, E.-M. Schön, T. Panda, K. Sreenivas, D. D. Díaz and R. Banerjee,

Journal of Materials Chemistry, 2012, 22, 14951 (b) S. Sengupta and R. Mondal, J. Mater. Chem. A, 2014, 2, 16373-16377

Page 66: Royal Society of Chemistry Coordination and Organometallic

COORDINATION-DRIVEN FORMATION OF INCLUSION

COMPLEXES OF RESORCIN[4]ARENE CAVITANDS

Richard C. Knighton and Adrian B. Chaplin*

Department of Chemistry, University of Warwick, Coventry CV4 7AL, UK

[email protected]

Since the inception of supramolecular chemistry, resorcin[4]arene cavitands have been

synonymous with host-guest chemistry.1 Guests have ranged from polar substrates such

as charged2 and hydrogen-bonded species

3 to apolar substrates which are included in the

cavity due to solvophobic effects.4 While non-covalent interactions are important in the

formation of host-guest complexes, they can be augmented through ligation of the guest

to a metal centre - thereby increasing the strength of association.5

In this work we describe the use of tetrasubstituted asymmetric resorcin[4]arene

cavitands as bidentate ligands for metal complexes to obtain 'metal-strapped' capsules

(Figure 1).

Figure 1: Small-molecule binding by an asymmetric resorcin[4]arene cavitand

References

1. (a) Koblenz, T. S.; Wassenaar, J.; Reek, J. N. H. Chem. Soc. Rev. 2008, 37, 247-262;

(b) Pochorovski, I.; Ebert, M.-O.; Gisselbrecht, J.-P.; Boudon, C.; Schweizer, W. B.;

Diederich, F. J. Am. Chem. Soc. 2012, 134, 14702-14705; (c) Pochorovski, I.;

Diederich, F. Acc. Chem. Res. 2014, 47, 2096-2105.

2. Carnegie, R. S.; Gibb, C. L. D.; Gibb, B. C. Angew. Chem. Int. Ed. 2014, 126,

11682-11684.

3. Cacciarini, M.; Azov, V. A.; Seiler, P.; Kunzer, H.; Diederich, F. Chem. Commun.

2005, 5269-5271.

4. Sullivan, M. R.; Gibb, B. C. Org. Biomol. Chem 2015, 13, 1869-1877.

5. (a) Iwamoto, H.; Nishi, S.; Haino, T. Chem. Commun. 2011, 47, 12670-12672; (b)

Wieser-Jeunesse, C.; Matt, D.; De Cian, A. Angew. Chem. Int. Ed. 1998, 37, 2861-

2864.

Page 67: Royal Society of Chemistry Coordination and Organometallic

CHELATING BIS(DIAZABORYL) LIGANDS FOR PREPARATION

OF CYCLIC BISBORYL COMPLEXES

Eugene L. Kolychev, Rémi Tirfoin, Simon Aldridge*

Chemistry Research Laboratory, University of Oxford, Mansfield Road, OX1 3TA,

Oxford

[email protected]

Recent reports by Yamashita and Nozaki et al. on the isolation[1]

and reactivity[2]

of the

first stable boryllithium reagent have opened new possibilities for the application of

diazaboryl fragments as strong -donor anionic ligands. Preliminary studies conducted

in our laboratory showed that diazaboryl complexes of Main Group elements[3]

are

encouraging candidates as transition metal free bond activation reagents. Further

investigation is currently focused on the preparation of bifunctional boryl ligands, with

the aims (i) of increasing the stability of complexes by the chelate effect, and (ii) of

facilitating re-reduction to lower oxidation states due to the presence of constrained ring

systems with appropriately narrow bite angles.

Synthetically, key precursors for such chelate systems are bis(diazaboryl) dibromides

which can be prepared by high yielding routes starting from commercially available

starting materials. In subsequent steps, a range of approaches including metathesis using

alkali metal boryl complexes and oxidative insertion of low valent metal centres into B-

H bonds of the analogous boryl hydrides have been examined to establish reliable

synthetic routes to chelating boryl complexes.

Figure 1. Cyclic bis(diazaboryl) complexes.

[1] Segawa, Y.; Yamashita, M.; Nozaki, K. Science 2006, 314, 113.

[2] Segawa, Y.; Suzuki, Y.; Yamashita, M.; Nozaki, K. J Am Chem Soc, 2008, 130,

16069.

[3] e.g. Protchenko, A.V. ; Dange, D.; Schwarz, A.D.; Blake, M.P.; Jones, C.;

Mountford, P.; Aldridge, S. J Am Chem Soc. 2014, 136, 10902.

Page 68: Royal Society of Chemistry Coordination and Organometallic

UNEXPECTED DIMER FORMATION DURING

{Rh(PiPR2(CH2)3P

iPR2)}

+ CATALYSED DEHYDROCOUPLING OF

H3B·NH3

A. Kumar and A. S. Weller

Department of Chemistry, University of Oxford, Oxford, U.K

[email protected].

Transition metal catalysed dehydrocoupling of H3B·NH3 is of significant interest

because of its potential use in chemical hydrogen storage and the synthesis of new

polymers, ceramics and piezoelectric materials. Recently, the Weller group has reported

the use of {Rh(PPh2(CH2)3PPh2)}+ fragment in 1,2-C6H4F2 for the fast dehydrocoupling

of H3B·NMe2H.1 Although complete mechanistic details are yet to be resolved,

formation of a dimeric species [{Rh(Ph2P(CH2)3PPh2)}2(H)2(H2BNMe2)]+ was observed

during the dehydrocoupling process through ESI-MS (as [M-H]+). However, no further

role of the dimeric species in the catalysis was explored.

We here report the use of a {Rh(PiPr2(CH2)3P

iPr2)}

+ fragment for the dehydrocoupling

of H3B·NH3 in Et2O. Interestingly, we observe the formation of

[Et2O·H2B·NH3][BArF

4] and [{Rh(PiPr2(CH2)3P

iPr2)}2(H)(BH2NH2)][BAr

F4]

intermediates during the dehydrocoupling process (Scheme 1). Formation of the dimeric

species is consistent with the previous report by the Weller group as mentioned earlier.1

[{Rh(PiPr2(CH2)3P

iPr2)}2(H)(BH2NH2)][BAr

F4] has also been independently

synthesized and to the best of our knowledge represents the first example of bridging

amino-borane bound to two metal centres. The role of the dimeric species in the

catalytic dehydrocoupling is currently being studied.

Scheme 1: Reaction of {Rh(PiPr2(CH2)3P

iPr2)}

+ fragment with H3B·NH3, [BAr

F4]

-

anion not shown.

References:

1. Dallanegra, R.; Robertson, A. P. M.; Chaplin, A. B.; Manners I.; Weller, A. S.

Chem. Commun. 2011, 47, 3763.

Page 69: Royal Society of Chemistry Coordination and Organometallic

GEOMETRY CONSTRAINED MAIN-GROUP COMPLEXES FOR

SMALL MOLECULES ACTIVATION

J. M. Goicoechea and S. K. Lo

Department of Chemistry, Chemistry Research Laboratory, University of Oxford, 12

Mansfield Road, Oxford, OX1 3TA, U.K.

[email protected]

Precious transition metals are well known for undergoing reversible redox processes in

the activation of small molecules and subsequent catalytic synthesis of many

compounds. However these metals are rare and expensive, therefore creating a lot of

interest in abundant main-group alternatives. In line with this, recent studies have

highlighted a number of main-group systems that are able to activate challenging small

molecules.1

The Arduengo group has previously shown that by distorting a P(III)

compound into a planar “T-shaped” geometry it provides an orbital arrangement akin to

transition metal, which allow for the cleavage of polarised E–H bonds and preliminary

evidence of catalytic activity.2 Radosevich and his group has successfully developed

catalytic hydrogenations using the Arduengo’s compound.3

Our group is interested in tuning the steric and electronic properties of a variety of novel

geometry constrained ligands of group 15 elements. Our under study includes redox

active ligands as well as saturated and unsaturated ligands (see figure 1). Our

investigation expands to arsenic analogous of described ligands. The +3 oxidation state

of arsenic is lowered in energy which should enhance the reactivity of the reductive

elimination step. We are currently investigating the reactivity of these compounds

towards small molecules.

N

N

H

N

H

Dipp DippP

N

O OAs

N

O OPn

MeO OMe

tBu

tBu

tBu

tBu

Figure 1

Pn = P, As

References:

[1] (a) Power, P.P. Nature 2010, 463, 171 (b) Stephan, D.W.; Erker, G. Angew. Chem.

Int. Ed. 2010, 49, 46

[2] Cully, S. A.; Arduengo, A. J. Am. Chem. Soc. 1984, 106, 1164

[3] (a) Dunn, N. L. ; Radosevich, A. T. J. Am. Chem. Soc. 2012, 13, 11330 (b)

McCarthy, S. M. ; Lin, Y.-C. ; Devarajan, D. ; Chang, J. W. ; Yennaway, H. P. ; Rioux,

R. M. ; Ess, D. H. ; Radosevich, A. T. J. Am. Chem. Soc. 2014, 136, 4640

Page 70: Royal Society of Chemistry Coordination and Organometallic

B-H ACTIVATION REACTIONS OF TITANIUM COMPOUNDS

Simona Mellino

a, Eric Clotb and Philip Mountforda

aChemistry Research Laboratory, University of Oxford, Mansfield Road, Oxford OX1

3TA, U.K., bInstitut Charles Gerhardt, Université Montpellier 2 and CNRS, cc 1501,

Place Eugène Bataillon, 34095 Montpellier cedex 5, France. [email protected]

Reactions of Group 4 (L)M=NR imido complexes have been extensively reported in the literature [1]. More recently the corresponding Group 4 hydrazide (L)M=NNR2 chemistry has been developed [2]. The reaction of both imido and hydrazido compounds with small molecules is an area of ongoing interest and many reaction products have been reported. Recently the activation of Si-H bonds of RSiH3 (R= Ph or Bu) has been demonstrated by the hydrazide compound Cp*Ti{MeC(NiPr)2}(NNMe2) which was found to undergo reversible 1,2-addition across Ti=N bond to yield titanium hydride species [3]. Following this line of research we decided to explore the reactivity of Cp*Ti{MeC(NiPr)2}(NNR2) compounds, where R2 = Me2, Ph2, CPh2, with boranes. For for the alkylidene hydrazide (R2 = CPh2) reaction with both pinacolborane (HBPin) and 9-BBN gives B-H addition across the N=C double bond, yielding compounds 1 and 2. The reactivity of the analogues hydrazido compounds (R2 = Me2) is different, resulting in either the product of 1,2-addition across the Ti=N bond with HBPin (3) or the borylimido compound (4) with one equivalent of 9-BBN dimer. [1] R. R. Schrock, Acc. Chem. Res., 2005, 38, 955; N. Hazari and P. Mountford, Acc. Chem.

Res., 2005, 38, 839.[2] P. J. Tiong, A. Nova, L. R. Groom, A. D. Schwarz, J. D. Selby, A. D Schofield, E. Clot and P. Mountford, Organometallics, 2011, 30, 1182. A. D. Schofield, A. Nova, J. D. Selby, A. D. Schwarz, E. Clot and P. Mountford, Chem. Eur J., 2011, 17, 265-285. [3] P. J. Tiong, A. Nova, E. Clot and P. Mountford, Chem. Commun., 2011, 47, 3147-3149. P. J. Tiong, A. Nova, A. D. Schwarz, J. D. Selby, E. Clot and P. Mountford, Dalton Trans., 2012, 41, 2277-2288.

Page 71: Royal Society of Chemistry Coordination and Organometallic

ADVANCES IN TRANSITION METAL FRUSTRATED LEWIS

PAIR (FLP) CHEMISTRY

Owen J. Metters, Stephanie R. Flynn, Duncan F. Wass, Ian Manners

School of Chemistry, University of Bristol, Cantock’s Close, Bristol, UK, BS8 1TS

[email protected]

Recent work in the Wass group has focussed on the development of novel

intramolecular frustrated Lewis pair (FLP) systems containing transition metal (Zr(IV),

Ti(III)) fragments. These systems have been shown to be highly reactive towards small

molecules (e.g. CO2, H2, ethers), and also competent catalysts for the dehydrocoupling

of amine-boranes.1 Modifications to these current systems have led us to synthesise a

range of intermolecular FLPs containing the analogous transition metal fragments (A).

These intermolecular systems (A) demonstrate reactivity towards small molecules

similar to that of their intramolecular counterparts, however the absence of the tether

provides easy access to a wider range of Lewis bases, allowing us to explore the effect

of varying phosphine sterics and electronics on FLP reactivity.

Zirconium (IV) cations of the type [(CpR)2ZrOMes][B(C6F5)4] are also found to

be competent catalysts in the catalytic hydrogenation of imines. The mechanism is

found to be FLP-mediated with the formation of a Zr+/N FLP between catalyst and

substrate. Such reactions require low pressures of H2 (1 bar), ambient temperatures (25

ºC) and short reaction times (90 mins).

1 (a) Chapman, A. M.; Haddow, M. F.; Wass D. F., J. Am. Chem. Soc., 2011, 133,

18463, (b) Chapman A. M.; Haddow, M. F.; Wass D. F., J. Am. Chem. Soc., 2011, 133,

8826 (c) Chapman A. M.; D. F; Dalton Trans., 2012, 41, 9067

Page 72: Royal Society of Chemistry Coordination and Organometallic

QUANTITATIVE UNDERSTANDING OF MULTIMETALLIC CLUSTER GROWTH

Stefan Mitzinger,1 Lies Broekaert,

1,2 Werner Massa,

1 Florian Weigend,

2

Stefanie Dehnen1

1Department of Chemistry and Scientific Centre for Material Science, Philipps

University of Marburg, Hans-Meerwein-Straße, D-35043 Marburg, Germany.

2Institute of Nanotechnology, Karlsruhe Institute of Technology,

Hermann-von- Helmholtz Platz 1, D-76344 Eggenstein-Leopoldshafen, Germany.

The chemistry of homoatomic and binary intermetalloid Zintl anions has been

investigated for several decades.1 In most cases, the structural features of Zintl anions

can be explained with Wade-Mingos rules or the Zintl-Klemm concept.2 By extraction

of quaternary metallic phases containing K/Ge/As/MV (MV = V, Nb, Ta) with

ethylenediamine (en), a new class of multimetallic Zintl anions was synthesized under

encapsulation of electron-poor transition metals.3

In case of the extraction of a K/Ge/As/Ta phase five new compounds were fully

characterized within one reaction allowing the elucidation of the stepwise formation of

multimetallic clusters, based on the crystal structures and complementary quantum

chemical studies of the involved clusters (Ge2As2)2–

, (Ge7As2)2–

, [Ta@Ge6As4]3–

,

[Ta@Ge8As4]3–

, and [Ta@Ge8As6]3–

. The results can even be generalized for an entire

family of multimetallic clusters. Furthermore first Zintl anions of the Ge/P and Si/P

system will be presented.

References:

[1] S. Scharfe, F. Kraus, S. Stegmaier, A. Schier, T. F. Fässler, Angew. Chem. Int. Ed.

2011, 50, 3630–3670. [2] W. Klemm, in Festkörperprobleme 3 (Ed.: F. Sauter),

Springer, Heidelberg, 1964, pp. 233–251.; K. Wade, Advances in Inorganic Chemistry

and Radiochemistry 1976, 18, 1–66. [3] S. Mitzinger, L. Broeckaert, W. Massa, F.

Weigend, S. Dehnen, Chem. Commun. 2015, 51, 3866–3869.

Page 73: Royal Society of Chemistry Coordination and Organometallic

NEW DEVELOPMENTS IN THE COORDINATION CHEMISTRY

OF GROUP 13 METAL FLUORIDE

Rajiv Bhallaa, William Levason

b, Graeme McRobbie

c, Francesco M. Monzittu

b, Gill

Reidb.

a Centre for advanced imaging, University of Queensland, Brisbane, Queensland 4072,

AUS b School of Chemistry, University of Southampton, Southampton SO17, 1BJ, UK c GE Healthcare, The Grove Centre, White Lion Road, Amersham HP7 9LL, UK

[email protected]

Although the hydrates of the Group 13 metal fluorides, MF3·3H2O (M = Ga, Al, In),

have very poor solubility in organic solvents or water, several new complexes have been

studied in recent years.1,2

Hydrothermal synthesis, or in a few examples refluxing in

polar solvents (such as alcohols)3, have been exploited to overcome the poor solubility

of these fluorides. However, the forcing conditions of these methods can lead to

decomposition of the ligand. For these reasons, there is a need to find a pathway into

this chemistry which goes through milder reaction conditions.

Moreover, the Group 13 metal fluorides complexes have attracted much attention for

their possible application as PET imaging agents4,5

in the last few years. 18

F is the most

commonly used PET-imaging isotope due to its high abundance, its low-energy positron

emission, and its short half-life of 109.8 min. Usually, 18

F is attached to targeted

peptides by binding it to a carbon atom, this process however is challenging and often

involves multistep syntheses, so M18

F-complexes have arisen as attractive, possible,

future alternatives.

In this study, we report a new entry

into the chemistry of the hydrates

Group 13 metal fluorides, which

enables reactions with different

ligands, such as pyridine oxide

(Figure 1) or N,N,N’,N’,N’’-pentamethyldiethylenetriamine

(pmdta) at room temperature, further

investigating their coordination

chemistry.

[1]. Bhalla R., Darby C., Levason W., Luthra S. K., McRobbie G., Reid G., Sanderson G., Zhang W.,

Chem. Sci. 2013, 5, 381.

[2]. Bhalla R., Levason W., Luthra S. K., McRobbie G., Monzittu F. M., Palmer J., Reid G.,

Sanderson G., Zhang W., Dalton Trans. 2015, 44, 9569.

[3]. Penkert F. N., Weyhermüller T., Wieghardt K., Chem. Commun. 1998, 5, 557.

[4]. Laverman P., McBride W. J., Sharkey R. M., Eek A., Joosten L., Oyen W. J. G., Goldenberg D.

M., Boerman O. C., J. Nucl. Med. 2010, 51, 454.

[5]. Bhalla R., Levason W., Luthra S. K., McRobbie G., Sanderson G., Reid G., Chem. Eur. J. 2015,

21, 4688.

Figure 1. Structure of the H-bonded dimer of

[GaF3(OH2)2(PyNO)] showing the H-bonding contacts

(blue) and the two geometric isomers of Ga(III).

Page 74: Royal Society of Chemistry Coordination and Organometallic

MACROCYCLIC TRANSITION METAL COMPLEXES FOR

ELECTROCHEMICAL REDUCTION OF CARBON DIOXIDE

James R. Pankhurst, Thomas Cadenbach and Jason B. Love*

EaStCHEM School of Chemistry, University of Edinburgh, UK

[email protected], *[email protected]

In March 2015, the concentration of carbon dioxide in the Earth’s atmosphere exceeded

400 ppm,1 having increased steadily from 280 ppm since the Industrial Revolution due

to anthropogenic activity.2 The electrochemical reduction of CO2 to useful products is a

viable method to be deployed in carbon-capture and utilisation (CCU) strategies,

however the single-electron reduction of CO2, that occurs at -1.90 V vs NHE, is

kinetically disfavoured.3 There is therefore a requirement to design homogeneous

systems that can function as electrocatalysts, to lower the kinetic barriers and required

overpotential for CO2 reduction.

We recently reported a pair of new polypyrrolic Schiff-base macrocycles that act as

binucleating ligands for the late transition metals.4,5

These complexes fold into

structures that offer a reactive cleft between the two metal centres. Their

electrochemically reversible redox features, both in the reductive and oxidative

directions from the +2 oxidation state, make their application in catalysis viable. Cyclic

voltammetry has indicated that some of these complexes react with CO2 at moderate

overpotentials of around -1.2 V vs. NHE, using H2O as a proton source.

[1]

P. Tans and R. Keeling, http://www.esrl.noaa.gov/gmd/ccgg/trends/ 30/05/2015

[2] A. Neftel, E. Moor, H. Oeschger and B. Stauffer, Nature, 1985, 315, 45

[3] E.E. Benson, C.P. Kubiak, A.J. Sathrum and J.M. Smieja, Chem. Soc. Rev., 2009, 38, 89

[4] J.R. Pankhurst, T. Cadenbach, D. Betz, C. Finn and J.B. Love, Dalton Trans., 2015, 44,

2066

[5] T. Cadenbach, J.R. Pankhurst, T.A. Hofmann, M. Curcio, P.L. Arnold and J.B. Love,

Organometallics, 2015

Page 75: Royal Society of Chemistry Coordination and Organometallic

COORDINATION CHEMISTRY OF BIDENTATE

CALIX[4]ARENE-BASED NHC LIGANDS

Ruth Patchett and Adrian B. Chaplin*

Department of Chemistry, University of Warwick, Coventry CV4 7AL, UK

[email protected]

Calix[4]arenes are attractive complexing agents and enzyme mimics due to their

characteristic “basket-like” shape. These macromolecules have been observed to

encapsulate guests and can provide flexible trans-disposed ligand systems.1 In

particular, functionalisation of the upper rim with donor groups enables the possibility

to alter the reactivity of bound transition metals, through their close proximity to the

calix[4]arene cavity.

As part of our work investigating the coordination chemistry of calix[4]arene-based

coordination complexes, we describe and contrast the coordination chemistry of a N-

Heterocyclic Carbene (NHC) functionalised calix[4]arene ligand with a simple mono-

dentate NHC analogue.

References

[1] (a) Wieser-Jeunesse, C.; Matt, D.; De Cian, A. Angew. Chem. Int. Ed. 1998, 37,

2861–2864; (b) Dinarès, I.; Garcia de Miguel, C.; Mesquida, N.; Alcalde, E. J. Org.

Chem., 2009, 74, 482–485; (c) Fang, X. ; Scott, B.L.; Watkin, J.G.; Carter, C.A.G.;

Kubas, G.J. Inorg. Chim. Acta, 2001, 317, 276–281.

Page 76: Royal Society of Chemistry Coordination and Organometallic

NOVEL MAIN GROUP POLYAZIDES: HOMOLEPTIC LOW

VALENT GROUP 14 POLYAZIDES AND BASE STABILISED

GROUP 13 AND 15 POLYAZIDES

P.Portius and B. Peerless

University of Sheffield, Dainton Building, Brook Hill, S3 7HF

[email protected]

Abstract text: Main Group binary azides represent an interesting class of nitrogen rich

compounds owing to their potential as energetic materials, however, due to their

expressed shock and thermal sensitivities binary azides are an experimental challenge to

prepare and characterize. This has led to a number of methods to stabilize these

sensitive compounds including using Lewis bases [1]

or using bulky cations in the

preparation of anionic hypercoordinate azido complexes [2]

. These methods have been

used to prepare low valent germanium and tin azides of the form L’Ge(N3) [3]

,

L”Ge(N3)2 [4]

and L’Sn(N3) [3]

(L’ = N-(n-propyl)-2-(n-propylamino)-troponiminate, L” = N-heterocylcic carbene), however, no Si(II) azido complex or homoleptic polyazido

low valent complex has been prepared previously. Treating GeCl2(1,4-dioxane) or

SnCl2 with PPh4N3 and NaN3 we have shown it is possible to prepare and fully

characterise the new species (PPh4)Ge(N3)3 and (PPh4)Sn(N3)3, representing the first

low valent Main Group homoleptic polyazido complexes [5]

. Attempts to reduce base

stabilised Si(IV) polyazides to Si(II) using a Mg dimer has led to the isolation of a

previously known Si(I) complex and a novel base stabilised magnesium azido

compound. Group 13 and 15 binary azides are scarcer in the literature than Group 14,

Al(N3)3 has only been observed in argon matrices [6]

and P(N3)5 is currently unknown.

Using an N-heterocylic carbene (NHC) we have been able to isolate and structurally

characterise the first Lewis base adducts of Al(N3)3 and P(N3)5 in the form

(NHC)Al(N3)3 and (NHC)P(N3)5.

References.

[1] Filippou, A. C.; Portius, P.; Neumann, D. U.; Wehrstedt, K.-D. Angew. Chem. 2000,

112, 4524; Angew. Chem., Int. Ed. 2000, 39, 4333.

[2] Filippou, A. C.; Portius, P.; Schnakenburg, G. J. Am. Chem. Soc. 2002, 124, 12396

[3] Ayers, A. E.; Marynick,D. S.; Dias, R. H.V. Inorg. Chem. 2000, 39, 4147

[4] Lyhs, B.; Blaser, D.; Wolper, C.; Schulz, S.; Haack, R.; Jansen, G. Inorg. Chem.

2013, 52, 7236−7241

[5] Peerless, B.; Keane, T.; Meijer, A. J. H. M.; Portius, P. Chem. Comm. 2015, 51,

7435

[6] Linnen, C. J.; Macks, D. E.; Coombe, R. D. J. of Phys. Chem. B, 1997, 101, 1602.

Page 77: Royal Society of Chemistry Coordination and Organometallic

NEW CATALYSTS FOR THE UPGRADING OF ETHANOL TO

ADVANCED BIOFUELS

K. J. Pellow and D. F. Wass

School of Chemistry, University of Bristol, Cantock’s Close, Bristol, UK, BS8 1TS

[email protected]

www.wassresearchgroup.com

The need to find alternatives to fossil fuels is crucial both from an environmental perspective, as well as to ensure energy security and sustainability.1 Higher alcohols, termed advanced biofuels, such as butanol are ideal “drop in” gasoline alternatives, outperforming bioethanol both in terms of energy content and ease of use.2

Figure 1. NHC ruthenium complexes employed as catalysts for the upgrading of ethanol to

butanol.

The Wass group have previously employed ruthenium catalysts for the upgrading of ethanol to butanol by way of the Guerbet reaction.3,4 We have extended the known library of catalysts able to carry out this transformation and have developed a series of N-heterocyclic carbene (NHC) ruthenium complexes (figure 1) that successfully catalyse ethanol to the advanced fuels n-butanol and iso-butanol with up to 93 % selectivity. [1] Ragauskas, A. J.; Williams, C. K.; Davison, B. H.; Britovsek, G.; Cairney, J.; Eckert, C. A.; Frederick, W. J.; Hallett, J. P.; Leak, D. J.; Liotta, C. L.; Mielenz, J. R.; Murphy, R.; Templer, R.; Tschaplinski, T. Science 2006, 311, 484. [2] Peralta-Yahya, P. P.; Zhang, F.; del Cardayre, S. B.; Keasling, J. D. Nature 2012, 488, 320. [3] Guerbet, M. C. R. Hebd. Seances Acad. Sci. 1909, 149, 129. [4] Dowson, G. R. M.; Haddow, M. F.; Lee, J.; Wingad, R. L.; Wass, D. F. Angew.

Chem. Int. Ed. 2013, 52, 9005.

Page 78: Royal Society of Chemistry Coordination and Organometallic

COORDINATION CHEMISTRY OF S-BLOCK CATIONS WITH

SOFT DONOR MACROCYCLES

M. J. D. Champion, M. Everett, A. Jolleys, W. Levason, D. Pugh, J. Purkis and G. Reid

Chemistry, University of Southampton, Highfield, Southampton SO17 1BJ

[email protected]

The s-block elements form hard metal cations which have a high affinity for O-donor

ligands, especially water. Consequently the vast majority of their reported coordination

chemistry is with hard O-donor ligands. Coordination complexes with neutral soft donor

ligands are rare, which may in part be due to the lack of suitable precursors. Most s-

block salts have limited solubility in organic media and solvents which solubilise the s-

block cations usually contain competitive O-donor groups (e.g. thf, MeOH, dmso).

We recently discovered that NaBArF has a high affinity for aza macrocycles, resulting

in the isolation of several unusual coordination compounds of Na+, including the

sandwich complex [Na(Me3tacn)2][BArF] (left; Me3tacn = 1,4,7-trimethyl-1,4,7-

triazacyclononane).1

Recently we extended the aza macrocycle chemistry to other Group 1 cations, isolating

compounds such as a sandwich complex of K+.2 We have also reacted even softer

neutral donor ligands such as [24]aneS8 (1,4,7,10,13,16,19,22-octathiacyclotetracosane)

to Na+, forming [Na([24]aneS8)][BAr

F] with homoleptic thioether coordination at Na

+

(right).3 Our methodology is also applicable to Group 2, allowing the isolation of

complexes which contain bonds between soft thio- and selenoether moieties and hard

Group 2 dications.

This chemistry is possible because the BArF salts are highly soluble, allowing reactions

to take place in non-competitive solvents, thus revealing a rich new area of the

coordination chemistry of s-block cations with soft donor ligands.

[1] Everett, M., Jolleys, A., Levason, W., Pugh, D., Reid, G., Chem. Commun., 2014, 50, 5843.

[2] Dyke, J. M. et al., Dalton. Trans., 2015, 44, DOI: 10.1039/c5dt01865j.

[3] Champion, M. J. D. et al., Inorg. Chem., 2015, 54, 2497.

[4] The SCFED project (www.scfed.net) is a multidisciplinary collaboration of British

universities investigating the fundamental and applied aspects of supercritical fluids.

NaBArF

[24]aneS8 2(Me

3tacn)

Page 79: Royal Society of Chemistry Coordination and Organometallic

CATALYSIS IN SERVICE OF MAIN GROUP CHEMISTRY:

SYNTHESIS OF P-BLOCK MATERIALS

D. A. Resendiz–Lara, T. Jurca, J. Turner, A. Schäfer, G. R. Whittell, Ian Manners.

School of Chemistry, University of Bristol, Bristol, BS8 1TS, UK

[email protected]

Catalytic dehydrocoupling chemistry is a versatile approach for the formation of p block

element-element bonds, and is becoming a promising methodology for the formation of

inorganic polymers under mild conditions with high-yielding reactions compared to

analogous routes (e. g. condensation reactions, reductive coupling, etc.).1

Recently, iron-based complexes have been reported to catalyse the polymerisation of

phosphine-boranes,2 and the molecular weight of the resulting polymeric material was

found to vary as a function of the catalyst loading. The synthesis and characterization of

two novel phosphine-borane monomers and their respective polymers will be

presented.3

Scheme 1. Dehydrocoupling of arylphosphine-boranes by [Fe] catalyst.

[1] a) Leitao, E. M.; Jurca T.; Manners, I. Nature Chem., 2013, 5, 817-829. b) Clark,

T. J. Lee, K. Manners, I. Chem. Eur. J., 2006, 12, 8634-8648.

[2] Schäfer A.; Jurca, T.; Turner, J.; Vance, J. R.; Lee, K.; Du, V. A.; Haddow, M. F.;

Whitell, G. R.; Manners, I. Angew. Chem. Int. Ed., 2015, 54, 4836-4841.

[3] Jurca, T.; Turner, J.; Resendiz-Lara, D. A.; Manners, I. Unpublished Results

Page 80: Royal Society of Chemistry Coordination and Organometallic

SYNTHESIS AND REACTIVITY OF ALKALINE EARTH

GALLYL COMPLEXES

J. Adan Reyes-Sanchez, Matthew P. Blake and Philip Mountford*

Chemistry Research Laboratory, University of Oxford, 12 Mansfield Road, Oxford OX1 3TA, UK.

[email protected]

Discrete molecular entities containing metal–metal bonds have enabled a better

understanding of chemical bonding and the reactivity in intermetallic and more complex

species. Their practical applications are of interest in the development of new materials1

and the solution to current energy problems,2 amongst others.

Mountford et al. recently reported the synthesis of complexes containing magnesium–transition metal

3 and lanthanide–gallium and boron

4 bonds. Here, we report the salt

metahesis reactions of the gallium(I) N-heterocyclic carbene (NHC) analogue

[(Dipp

DAB)GaK(Et2O)]2 (DAB = {N(Dipp)C(H)}2, Dipp = 2,6-diisopropylphenyl) with

Group 2 iodide complexes to form compounds containing alkaline earth–gallium bonds.

Group 2 metals in these species are supported by Dipp

NacNac or CzOxMe

ligands (Fig.

1). These compounds have been reacted with small molecules such as carbodiimides,

isocyanates, thioisocyanates and epoxides.

Figure 1. Alkaline earth–gallium complexes.

References

[1] Goll, D.; Kronmller, H. Naturwissenschaften 2000, 87, 423-438.

[2] David, E., J. Mat. Proc. Techn. 2005, 162-163, 169-177.

[3] Blake, M. P.; Kaltsoyannis, N.; Mountford, P. J. Am. Chem. Soc. 2011, 133 (39),

15358-15361.

[4] Saleh, L. M. A.; Birjkumar, K. H.; Protchenko, A. V.; Schwarz, A. D.; Aldridge,

S.; Jones, C.; Kaltsoyannis, N.; Mountford, P. J. Am. Chem. Soc. 2011, 133 (11),

3836-3839.

Page 81: Royal Society of Chemistry Coordination and Organometallic

SMALL MOLECULE COORDINATION AND ACTIVATION BY

[Ru(IPr)2(CO)H]+ (IPr = 1,3-BIS(2,6-

DIISOPROPYLPHENYL)IMIDAZOL-2-YLIDENE)

Ian M. Riddlestone,a David McKay,

b Stuart A. Macgregor,

b Mary F. Mahon,

a and

Michael K. Whittleseya

aDepartment of Chemistry, University of Bath, Claverton Down, Bath, BA2 7AY

bSchool of Engineering and Physical Sciences, Heriott-Watt University, Edinburgh,

EH14 4AS

[email protected]

Very low coordinate transition metal fragments represent long standing synthetic targets

due to their potential for both novel and high reactivity with organic substrates. The

synthesis and characterization of four-coordinate Ru(II)L4 species represents a

challenging target due to the high coordinative unsaturation and corresponding high

Lewis acidity. Although rare, cationic species of the type [Ru(L)2(CO)R]+ (L =

PtBu2Me, R = Ph (1) or H (2)

[1,2]; L = IMes, R = H (3)

[3]) can be accessed via halide

abstraction reactions from the appropriate chloride containing precursors. All feature a

saw-horse geometry at ruthenium and 1 and 2 are stabilized by double agostic

interactions from the supporting phosphine ligand.[1,2]

By contrast, 3 was reported free

of agostic bonding, based on a combination of computational studies and trapping

experiments.[3]

Employment of the sterically demanding IPr

ligand allowed isolation and structural

characterization of the agostically stabilised

16-electron cation [Ru(IPr)2(CO)H]+ (4). 4

readily coordinates (H2, CO and

H3B.NMe2H), but undergoes hydride

abstraction with HBcat and Et2Zn,

eliminating H2 and ethane, forming rare

Ru-boryl and bimetallic Ru-Zn species

respectively.

References

[1] Huang, D.; Streib, W. E.; Eisenstein, O.; Caulton, K. G., Angew. Chem. Int. Ed.

Engl. 1997, 36, 2004.

[2] Huang, D.; Bollinger, J. C.; Streib, W. E.; Folting, K.; Young, J. V.; Eisenstein, O.;

Caulton, K. G., Organometallics 2000, 19, 2281

[3] Lee, J. P.; Ke, Z.; Ramírez, M. A.; Gunnoe, T. B.; Cundari, T. R.; Boyle, P. D.;

Petersen, J. L., Organometallics 2009, 28, 1758

N N

iPr

iPriPr

Ru

H CO

CH2H

N N

iPriPr

BArF4

Page 82: Royal Society of Chemistry Coordination and Organometallic

ACYCLIC TWO-COORDINATE CATIONIC GERMYLENES – METAL ELEMENT BOND FORMATION

Arnab Rit, Simon Aldridge*

Chemistry Research Laboratory, University of Oxford, Mansfield Road, OX1 3TA,

Oxford

[email protected]

The chemistry of thermally stable divalent neutral germanium compounds has attracted

much recent attention.[1]

In contrast to neutral compounds, :GeX2, cationic species of

the type [:Ge(L)X]+

are relatively rare.[2]

Given recent interest in comparisons of

fundamental properties between carbon and its heavier congeners, understanding the

structure and reactivity of cationic germylenes might open up many interesting areas as

has been seen for the corresponding carbenium ions. In addition, the presence of a

cationic charge may lead to the isolation of unsaturated monomeric species (e.g.

containing Ge=E multiple bonds) by discouraging the sorts of oligomerization processes

seen for neutral analogues.

Herein we present the synthesis of rare acyclic two-coordinate cationic germylenes by

chloride abstraction from NHC-stabilized chlorogermylenes. Steric and electronic

tuning of the substituents at germanium allows control of aggregation. Thus, smaller

NHCs result in dimerization to give novel dicationic digermenes [X(L)Ge=GeX(L)]2+

.

Bulkier substituents yield monocationic germylenes which react cleanly with azido or

diazo compounds under N2 release to give monomeric imido or alkylidene complexes.

[1] (a) Kühl, O. Coord. Chem. Rev. 2004, 248, 411; (b) Z. Rappoport, The Chemistry of

Organic Germanium, Tin and Lead compounds; Wiley: Chichester, 2002; Vol. 2, p.

284.

[2] (a) Dias, H. V. R.; Wang, Z. J. Am. Chem. Soc. 1997, 119, 4650; (b) Stender, M.;

Phillips, A. D.; Power, P. P. Inorg. Chem. 2001, 40, 5314.

Page 83: Royal Society of Chemistry Coordination and Organometallic

SMALL MOLECULE ACTIVATION AT A GEOMETRY-

CONSTRAINED PHOSPHORUS CENTRE

T. P. Robinson, D. M. De Rosa, S. Aldridge and J. M. Goicoechea

Department of Chemistry, Chemistry Research Laboratory, University of Oxford, 12 Mansfield Rd, Oxford, UK.

[email protected]

The activation of small molecules and subsequent catalytic transformations has long

been considered the domain of precious transition metals, for which low crustal

abundance, high cost and toxicity are discouraging factors. Consequently there has been

much interest in the development of main-group species that are able to “mimic” this behavior, with a number of systems shown to activate highly challenging small

molecules including dihydrogen and ammonia.[1]

Recent studies have demonstrated the

application of geometry constraining chelating ligands to synthesise P(III) compounds

that facilitate the oxidative addition of polarised E-H bonds (E = O and N).[2-6]

We have

developed this chemistry by investigating the reactivity of a N,N-bis-(2-

phenoxide)amide supported phosphorus system (1) (Scheme 1). The geometry

constrained product is extremely reactive towards a range of small molecule substrates,

resulting in oxidative additions over the phosphorus centre. The rapid and quantitative

reactions with both H2O and NH3 are of particular importance on account of the

difficulties associated with analogous transition-metal mediated activations.

Scheme 1

References:

[1] Power, P. P. Nature 2010, 463, 171–177.

[2] Arduengo III, A. J.; Stewart, C. A.; Davidson, F.; Dixon, D. A.; Becker, J. Y.;

Culley, S. A.; Mizen, M. B. J. Am. Chem. Soc. 1987, 109, 627–647.

[3] McCarthy, S. M.; Lin, Y. –C.; Devarajan, D.; Chang, J. W.; Yennawar, H. P.;

Rioux, R. M.; Ess, D. H.; Radosevich, T. J. Am. Chem. Soc. 2014, 136, 4640–4650.

[4] Dunn, N. L., Ha, M.; Radosevich, A. T. J. Am. Chem. Soc. 2012, 134, 11330–11333.

[5] Zhao, W.; McCarthy, S. M.; Lai, T. Y.; Yennawar, H. P.; Radosevich, A. T. J. Am. Chem. Soc. 2014, 136, 17634−17644.

[6] Cui, J.; Li, Y.; Ganguly, R.; Inthirarajah, A.; Hirao, H.; Kinjo, R. J. Am. Chem. Soc. 2014, 136, 16764−16767.

Page 84: Royal Society of Chemistry Coordination and Organometallic

TETHERED CYCLOPENTADIENYL-STANNYLENE LIGANDS

FOR C-H ACTIVATION

M. Roselló-Merino and S. M. Mansell*

Institute of Chemical Sciences, School of Engineering and Physical Sciences,

Heriot-Watt University, Edinburgh, EH14 4AS, UK.

[email protected], [email protected]

Tethered ligands are interesting candidates for designing the next generation of

homogeneous catalyst as they combine two different ligand classes into one flexible

chelating ligand.[1]

N-heterocyclic stannylenes (NHSns) are an interesting alternative to

the ubiquitous N-heterocyclic carbene (NHC) ligands which have been recently

described,[2]

but have yet to be explored as ligands in catalysis. We are interested in

combining these concepts by synthesising tethered cyclopentadienyl-stannylene ligands

which are expected to open up new modes of reactivity including: 1. Hemilability of the

stannylene moiety, facilitating bond activation at the TM centre, 2. Bridging binding

modes (typically observed in NHSns) promoting unusual bimetallic bond activation, 3.

Transfer of ligands between Sn and the TM, 4. Increased complex stability and

improved product selectivity facilitated by the presence of a tether.

We have synthesised new indenyl- and fluorenyl-tethered diamines in order to target

new functionalised NHSn ligands. The formation of transition metal complexes via two

routes has then been explored using reactions of the ligand precursor with

Sn{N(SiMe3)2}2] followed by addition of the transition metal precursor, or introduction

of the transition metal first before reaction with a tin source.

[1] a) a) Butenschön, H. Chem. Rev. 2000, 100, 1527; b) Siemeling, U. Chem. Rev.

2000, 100, 1495.

[2] a) Mansell, S. M.; Russell, C. A.; Wass, D. F. Inorg. Chem. 2008, 47, 11367; b)

Mansell, S. M.; Herber, R. H.; Nowik, I.; Ross, D. H.; Russell, C. A.; Wass, D. F. Inorg.

Chem. 2011, 50, 2252.

Page 85: Royal Society of Chemistry Coordination and Organometallic

NON-INNOCENT ACTIVITIES OF N-HETEROCYCLIC

CARBENES AT COPPER

C. E. Willans and W. Rungtanapirom

School of Chemistry, University of Leeds, Woodhouse Lane, Leeds, LS2 9 JT, UK

[email protected]

N-heterocyclic carbenes (NHCs) have been used as ligands to support copper in a broad

range of catalytic reactions.[1]

However, reactivities in which the NHC ligands become

involved have also been observed, especially reductive elimination of the NHC.[2]

We

have previously observed the reductive elimination of 1-allyl-3-pyridylimidazol-2-

ylidene from copper to form bis-imidazolium salts, which is thought to proceed via

copper(III).[2c]

Furthermore, we discovered that 1-allyl-2-bromo-3-pyridylimidazolium

cuprate salts undergo a cyclisation reaction (Figure 1), similar to those previously

reported with Rh[3]

and Ni.[4]

Further investigation is needed to understand the

mechanism of these cyclisation reactions, with a view to both preventing catalyst

deactivation and using them to our advantage in the synthesis of heterocycles. Further

studies in this work include investigating the deactivation of copper catalysts bearing 1-

alkenyl-3-picolylimidazole-2-ylidene through reductive elimination of the NHC with an

aryl group (Figure 2).

Figure 1. Cyclisation of an allyl-substituted bromoimidazolium salt at copper.

Figure 2. Deactivation of a copper catalyst via reductive elimination of the NHC and aryl group.

[1] J. E. Egbert, C. S. J. Cazin, S. P. Nolan, Catal. Sci. Technol. 2013, 3, 912-926.

[2] aE. L. Kolychev, V. V. Shuntikov, V. N. Khrustalev, A. A. Bush, M. S. Nechaev, Dalton. Trans.

2011, 40, 3074; bB.-L. Lin, P. Kang, T. D. P. Stack, Organometallics 2010, 29, 3683-3685;

cB.

R. M. Lake, A. Ariafard, C. E. Willans, Chem. Eur. 2014, 20, 12729-12733.

[3] K. I. Tan, R. G. Bergman, J. A. Ellman, J. Am. Chem. Soc. 2001, 123, 2685-2686.

[4] A. T. Normand, S. K. Yen, H. V. Huynh, T. S. A. Hor, K. J. Cavell, Organometallics 2008, 27,

3153-3160.

Page 86: Royal Society of Chemistry Coordination and Organometallic

CORRELATIONS OF THE STRUCTURAL PROPERTIES OF A

COMPLETE R2PX SERIES (X = HYDROGEN OR HALOGEN)

Timothy A. Shuttleworth, Jonathan P. Hopewell, Claire L. McMullin, Paul G. Pringle

School of Chemistry, University of Bristol, Bristol, BS8 1TS, UK

[email protected]

The importance of phosphorus(III) ligands in coordination chemistry and homogeneous

catalysis is undeniable. The quest for rational ligand design has engendered numerous

spectroscopic, crystallographic and theoretical studies aimed at quantifying the

stereoelectronic properties of P donors. However, the separation of steric and electronic

effects is a recurrent problem, precluding the presentation of unambiguous empirical

evidence to support even simple concepts. A study of a series of PX3 molecules (X = H

or Hal) would be the most accurate, as variation in the steric properties is minimised.

Unfortunately, the high reactivity of these molecules, even when coordinated to a metal,

makes full characterisation and analysis problematic. However, R2PX molecules offer

greater stability, whilst alteration of only one atom means stereoelectronic deviation is

again minimised. Complete R2PX series are known for R = Me, tBu, Ph and η1

-C5Me5,

although many of these molecules are air sensitive or fluids at room temperature.

CgPX molecules (where R2P = CgP = 6-phospha-2,4,8-trioxa-1,3,5,7-

tetramethyladamant-6-yl cage depicted below; X = H,1 F,

2 Cl,

1 Br

1) have been reported

as air stable, crystalline solids. Building on this, recently we reported the synthesis and,

for the first time, the crystal structures of a complete series of R2PX molecules (R2P =

CgP), along with their corresponding selenium adducts (Scheme 1).3 In this study, we

were able to compare a range of spectroscopic and crystallographic data, allowing

analysis of the stereoelectronic properties of the P donor series.

Scheme 1. Synthesis of CgPX and CgP(Se)X, where X = H, F, Cl, Br and I.

[1] Downing, J. H.; Floure, J.; Heslop, K.; Haddow, M. F.; Hopewell, J.; Lusi, M.;

Phetmung, H.; Orpen, A. G.; Pringle, P. G.; Pugh, R. I.; Zambrano-Williams, D.

Organometallics 2008, 27, 3216–3224.

[2] Fey, N.; Garland, M.; Hopewell, J. P.; McMullin, C. L.; Mastroianni, S.; Orpen, A.

G.; Pringle, P. G. Angew. Chem. Int. Ed. 2012, 124, 122–126.

[3] Hopewell, J. P.; McMullin, C. L.; Pringle, P. G.; Shuttleworth, T. A.; Woodall, C.

H. Eur. J. Inorg. Chem. 2014, 1843–1849.

Page 87: Royal Society of Chemistry Coordination and Organometallic

REDOX PROMOTED CYCLOMETALATION REACTIONS OF PLATINUM(0) COMPLEXES

Thibault Troadec and Adrian B. Chaplin*

Department of Chemistry, University of Warwick, Coventry CV4 7AL, UK

[email protected]

Due to the diverse potential applications in organic synthesis and petroleum research,

transition-metal-mediated C-H bond activation reactions are of significant contemporary

interest in organometallic chemistry and catalysis. Cyclometalation reactions based on

intramolecular C-H bond oxidative addition are useful model reactions for the more

coveted intermolecular variants, whilst also being of interest in their own right as

pathways in catalytic organic transformations. Building upon previous work by Goel

and Conejero, involving bis-phosphine and bis-(N-Heterocyclic carbene) platinum(II)

complexes (Figure 1),1,2

we present a novel redox-based method for promoting

intramolecular C-H reactions in platinum(0) systems.

Figure 1: Intramolecular cyclometalation reactions in platinum(II) complexes

References

1. Clark, H. C.; Goel, A. B.; Goel, R. G.; Goel, S.; Ogini, W. O. Inorg. Chim. Acta

1978, 31, L441–L442; Clark, H. C.; Goel, A. B.; Goel, R. G.; Goel, S. Inorg. Chem.

1980, 19, 3220–3225; Goel, R. G.; Ogini, W. O.; Srivastava, R. C. Organometallics,

1982, 1, 819-824.

2. Rivada-Wheelaghan, O.; Donnadieu, B.; Maya, C.; Conejero, S. Chem. Eur. J.

2010, 16, 10323 – 10326; Rivada-Wheelaghan, O.; Ortuno, M. A.; Diez, J.; Lledos,

A.; Conejero, S. Angew. Chem. Int. Ed. 2012, 51, 3936 –3939; Rosello-Merino, M.;

Lopez-Serrano, J.; Conejero, S. J. Am. Chem. Soc. 2013, 135, 10910-10913; Rivada-

Wheelaghan, O.; Rosello-Merino, M.; Ortuno, M. A.; Vidossich, P.; Gutierrez-

Puebla, E.; Lledos, A.; Conejero, S. Inorg. Chem. 2014, 53, 4257–4268. Rivada-

Wheelaghan, O.; Rosello-Merino, M.; Diez, J.; Maya, C.; Lopez-Serrano, J.;

Conejero, S. Organometallics 2014, 33, 5944-5947.

Page 88: Royal Society of Chemistry Coordination and Organometallic

DEHYDROCOUPLING OF DIMETHYLAMINE-BORANE BY

CYCLOPENTADIENYL BASED IRON PRECATALYSTS

J. R. Turner, T. Jurca and Ian Manners

School of Chemistry, University of Bristol, Bristol, BS8 1TS, UK

[email protected]

Utilising transition metals to catalyse the dehydrocoupling of main group substrates has

been an expanding field since the 1980s, with the dehydrocoupling of amine-boranes

being a more recent area of interest. This has been extended towards iron, where in

2014, two complexes, CpFe(CO)2I and [CpFe(CO)2]2, were shown to dehydrocouple

dimethylamine-borane.1 Interestingly, these two systems proceed via vastly different

mechanisms, the former via a homogeneous process and the latter heterogeneous.

However, irradiation with UV/vis light has hindered mechanistic investigations, and the

search for an Fe catalyst that does not require photoactivation has been ongoing (Figure

1). The synthesis and characterisation of a number of Cp based Fe complexes and their

room temperature reactivity towards dimethylamine-borane will also be presented.2

Figure 1. Room temperature dehydrocoupling of dimethylamine-borane

[1]Vance, J. R.; Schäfer, A.; Robertson, A. P. M.; Lee, K.; Turner, J.; Whittell, G. R.;

Manners, I., J. Am. Chem. Soc. 2014, 136 (8), 3048-3064.

[2] Turner, J. R.: Titel, J.; Manners, I. Unpublished Results.

Requires hν

Page 89: Royal Society of Chemistry Coordination and Organometallic

GROUP 1 AND 2 CYCLIC (ALKYL)(AMINO)CARBENE

COMPLEXES

Z. R. Turner and J.-C. Buffet

Chemistry Research Laboratory, 12 Mansfield Road, OX1 3TA, Oxford, UK

[email protected]

Cyclic (alkyl)(amino)carbenes (CAACs) are a more recently developed and less

explored class of the ubiquitous N-heterocyclic carbenes (NHC).[1]

They have unique

stereoelectronic properties, being both stronger σ-donors and π-acceptors with respect to

classical NHCs, and demonstrating redox activity.[2]

The alkaline earth metals (Ae; magnesium – barium) are earth abundant and benign

in nature, making them attractive targets for homogeneous catalysis.[3]

There remain

relatively few examples of electropositive metal NHC complexes,[4]

and we recently

described the first examples of CAACs bound to both group 1 and 2 metal ions (Figure

1).[5]

We report here the use of these complexes as efficient catalysts for polar monomer

polymerisation and examples of reactivity unique to these systems.

Figure 1. Solid state molecular structures of potassium and barium CAAC complexes.

[1] Hopkinson, M. N.; Richter, C.; Schedler, M.; Glorius, F. Nature 2014, 510, 485-496.

[2] Soleilhavoup, M.; Bertrand, G. Acc. Chem. Res. 2015, 48, 256-266. [3] Harder, S.;

Editor Alkaline-Earth Metal Compounds: Oddities and Applications. [In: Top.

Organomet. Chem., 2013; 45]; Springer GmbH, 2013. [4] a) Bellemin-Laponnaz, S.;

Dagorne, S. Chem. Rev. 2014, 114, 8747-8774; b) Arnold, P. L.; Casely, I. J. Chem.

Rev. 2009, 109, 3599-3611. [5] Turner, Z. R.; Buffet, J.-C. Dalton Trans. 2015, 44,

12985-12989.

Page 90: Royal Society of Chemistry Coordination and Organometallic

PROTON COUPLED MIXED VALENCY IN HYDROGEN BONDED

DIMERS

K.B. Vincent, L.A. Wilkinson, L. Brown and N.J. Patmore

University of Huddersfield, Department of Chemical Sciences, Queensgate,

Huddersfield, HD1 3DH, UK

[email protected]

The study of organometallic mixed valence systems has achieved much attention over

recent decades with potential applications in photochromic systemsI and as components

in future molecular electronics technology.II Such systems are of the archetypal donor-

bridge-acceptor (D-B-A) architecture bearing redox active metal units linked through

covalently bonded bridging units, often organic in nature. To this extent there have been

many reports of systems comprising two multiply bonded di-metallic moieties linked

through covalent bridges where the redox chemistry of the metal fragments is exploited

to give rise to D-B-A systems.III

In contrast to this we present here a ‘new’ class of systems where metal-bridge-metal systems are prepared by exploiting hydrogen

bonding functionality of the ligand to form dimeric species, Figure 1a.IV

Electrochemical and spectroelectrochemical analyses of these hydrogen bonded dimers

show that in weakly / non- coordinating solvents the thermodynamic stability of the

cation is observed giving rise to voltammograms showing two consecutive reversible

one-electron oxidations. The requirement for the hydrogen bond is then observed by the

addition of DMSO which serves to break the dimeric bond as evidenced by the

breakdown two a single redox process, Figure 1b. Further evidence to support the

requirement for the hydrogen bond is observed in UV-vis NIR spectroelectrochemistry

of the mixed valence state where no IVCT band is observed, as would be expected in

classic mixed valence systems.

Figure 1: a) Metal dimers prepared by hydrogen bonding ligands. c) CV of bimetallic species showing

the dimeric species (green) and monomer (red).

I Sakamoto, R. et al. Chem. Eur. J. 2008, 14, 6978.

II Low, P. J. Dalton Trans. 2005 2821.

III Casas, J. M; Cayton, R. H; Chisolm, M. H. Inorg. Chem., 1991, 30, 360.

IV Patmore, N. J. et al, J. Am. Chem. Soc., 2013, 135, 1723. Wilkinson, L. A; McNeill, L;

Scattergood, P; Patmore, N. J. Inorg. Chem., 2013, 52, 9683

Page 91: Royal Society of Chemistry Coordination and Organometallic

SYNTHESIS AND REACTIVITY OF DITOPIC CARBANIONIC N-

HETEROCYCLIC CARBENE COMPLEXES

Lajoy S. Tucker, Jose M. Goicoechea and Jordan B. Waters

Department of Chemistry, Chemistry Research Laboratory, University of Oxford, Mansfield Road, OX1 3TA

[email protected]

Since the discovery of the first isolable N-heterocyclic carbene (NHC) 24 years ago,1

these compounds have gone from chemical curiosities to ubiquitous ligands in modern

organometallic chemistry. NHCs can bond to metals through the “conventional” (C2) or “abnormal” (C4) positions.2 Recently, Robinson reported the synthesis of the first N-

heterocyclic “dicarbene” (NHDC), by deprotonation of an NHC resulting in a species

capable of binding through the C2 and C4 positions simultaneously.3 Accordingly, we

have recently reviewed the known chemistry of ditopic carbanionic NHCs.4

We have been investigating the reactivity of the lithium salt of the deprotonated IPr

carbene towards E[N(SiMe3)2]2 complexes (E = Ge, Sn, Pb). Such reactions in an

equimolar ratio result in the formation of novel germylene, stannylene or plumbylene

species. We have also synthesised the analogous potassium salt (KIPr) and studied the

reactivity towards E[N(SiMe3)2]2 (E = Ge, Sn, Pb and Zn).5 The products are all C4

bonded adducts E[:CCH{[N(2,6-iPr2C6H3)]2C:}][N(SiMe3)2]2]

− which in the case for E

= Ge and Pb, will slowly rearrange to afford the bis-carbene adduct along with the

homoleptic metal tris-amide.

We have also synthesised the neutral

organomercury species Hg(aIPr:)2 by

the reaction of two equivalents of

KIPr with HgCl2 (see figure). This

complex consists of two abnormally

bonded NHCs; the carbenic carbons

can then be protonated resulting in

the formation of the dicationic salt

[Hg(aIPr)]2+

[BAr4F]2.

6

References

1. Arduengo, A. J.; Harlow, R. L.; Kline, M. J. Am. Chem. Soc., 1991, 113, 361.

2. Aldeco-Perez, E.; Rosenthal, A. J.; Donnadieu, B.; Parameswaran, P.; Frenking, G.

Bertrand, G. Science, 2009, 326, 556.

3. Wang, Y.; Xie, Y.; Abraham, M. Y.; Wei, P.; Schaefer, H. F.; Schleyer, P. v. R.;

Robinson, G. H. J. Am. Chem. Soc., 2010, 132, 14370.

4. Waters, J. B.; Goicoechea, J. M. Coord. Chem. Rev., 2014, 293-294, 80.

5. Waters, J. B.; Goicoechea, J. M. Dalton Trans., 2014, 43, 14239. 6. Waters, J. B.; Goicoechea, J. M. manuscript in preparation.

Hg1 C1

C2

C3

C28

C29

C30

Page 92: Royal Society of Chemistry Coordination and Organometallic

SWAPPING THE ROLE OF DONOR/ACCEPTOR AND BRIDGE IN

A DYCLIC DIMER OF DIMERS

L. A. Wilkinson1, and N. J. Patmore

2

1University of Sheffield, Western Bank, Sheffield, S10 2TN 2University of Huddersfield, Queensgate, Huddersfield, HD13DH

[email protected]

The fundamental study of electron transfer is paramount for understanding many

complicated processes in nature and for applications in molecular electronics1 and

mixed valence systems have been extensively employed to study this process.2 In many

systems, the donor/acceptor units are metal-based and the bridge is an organic linker

(such as the Creutz-Taube ion)3, however there are also multiple reports of the

inverse.4,5

To our knowledge, there has never been a report of a single molecule wherein

electron transfer could occur between metal centres via an organic bridge or between

organic redox sites via a metal-based bridge. Our laboratory has developed a cyclic

dimer of dimers of the form [Mo2(TiPB)2]2(μ-ADC)2 (where TiPB = 2,4,6-

triisopropylbenzoic acid and H2ADC = anthracene-1,8-dicarboxylic acid). DFT

calculations show that the HOMO is based on the Mo2 units whereas the LUMO is

based on the ADC fragments. This allows the possibility of forming two mixed valence

states: a metal based radical cation and a ligand based radical anion. This concept is

investigated by cyclic voltammetry and EPR spectroscopy.

References.

[1] Ward, M. D. Chem. Soc. Rev. 1995, 24, 121–134.

[2] Winter, R. F. Organometallics 2014.

[3] Creutz, C.; Taube, H. J. Am. Chem. Soc. 1969, 91, 3988–3989.

[4] Jones, S. C.; Coropceanu, V.; Barlow, S.; Kinnibrugh, T.; Timofeeva, T.; Bredas,

J.; Marder, S. R. J. Am. Chem. Soc. 2004, 126, 11782–11783.

[5] Lu, C. C.; Bill, E.; Weyhermüller, T.; Bothe, E.; Wieghardt, K. J. Am. Chem. Soc. 2008, 130, 3181–3197.

Figure 1: Cyclic dimer of dimers and the possible redox processes. Ancillary ligands omitted for clarity.

Page 93: Royal Society of Chemistry Coordination and Organometallic

WELL DEFINED TUNGSTEN CATALYSTS FOR SELECTIVE

ETHYLENE OLIGOMERISATION

C. M. R. Wright, J.-C. Buffet, D. O’Hare

Chemistry Research laboratory, University of Oxford, 12 Mansfield Road, Oxford, OX1 3TA.

[email protected] Feedstocks of ethane which can be cracked on a huge scale to produce ethylene have

increased due to shale gas exploration.1 Processes that upgrade such low molecular

weight α-olefins to higher homologues are therefore of increasing value. Olefin

oligomerisation is one such method employed by industry, although selectivity for such

transformations is still an issue.2,3

Ill-defined tungsten systems formed from WCl6,

aniline and triethylamine have shown promise in this area, displaying good selectivity in

ethylene oligomerisation when a Lewis acid cocatalyst is employed.4 Here we report a

well-defined tungsten imido complex which is active for ethylene dimerisation in the

presence of methylaluminoxane. The cocatalyst ratio has a marked effect on the

stability, activity and selectivity of the complex.

Left: Scheme showing the dimerisation of ethylene to 1-butene utilising a well-defined tungsten imido complex

with MAO as a cocatalyst. Right: Graph showing conversion of ethylene to 1-butene with varying cocatalyst

ratios.

Utilising the technique of surface organometallic chemistry (SOMC),5 tungsten imido

complexes have been grafted onto solid supports. The activity and selectivity for the

dimerisation can be maintained and in some cases enhanced compared to the

homogeneous reaction.

[1] Chem. World 2015, 12, 18–19.

[2] Grasset, F.; Magna, L. EP2388069 (A1), 2011.

[3] Mol, J. J. Mol. Catal. A Chem 2004, 213, 39–45.

[4] Hanton, M. J.; Daubney, L.; Lebl, T.; Polas, S.; Smith, D. M.; Willemse, A. Dalt. Trans. 2010, 39, 7025–7037.

[5] Copéret, C. New J. Chem. 2004, 28, 1–10.

Page 94: Royal Society of Chemistry Coordination and Organometallic

USING IN SITU X-RAY DIFFRACTION TO OBSERVE SOLVENT

EXCHANGE DURING MOF SYNTHESIS

Y. Wu,a M. I. Breeze,

b G. J. Clarkson,

b F. Millange, D. O’Hare,a and R. I. Walton*

b

a Department of Chemistry University of Oxford, Oxford, OX1 3TA, U.K.

b Department of Chemistry, University of Warwick, Coventry, CV4 7AL, U.K.

c Département de Chimie, Université de Versailles-St-Quentin-en-Yvelines, France

[email protected]

In most cases of metal-organic framework (MOF) synthesis using solvothermal

methods, it is unclear whether the framework is initially formed with coordinated

solvent that is then exchanged with another ligand to reach the final product, or if the

final product is formed from the start as the sole species. Using time-resolved

monochromatic high energy X-ray diffraction with a custom IR furnace cell (‘ODISC’, the Oxford-Diamond In-Situ Cell, Fig. 1, left),

1 we have been able to understand this

behaviour at an unprecedented level of detail.

Figure 1. ODISC furnace (left); Observation of in situ guest exchange (right)

We present an in situ study of the solvothermal crystallisation of a new ytterbium MOF

from a water/DMF mixture under solvothermal conditions. Analysis of high resolution

powder patterns using Rietveld refinement reveals an evolution of lattice parameters

and electron density during the crystallisation process. Quenching studies confirm that

this is due to a gradual topochemical replacement of coordinated solvent molecules: the

water initially coordinated to Yb3+

is replaced by DMF as the reaction progresses (Fig.

1, right).

(1) Moorhouse, S. J.; Vranješ, N.; Jupe, A.; Drakopoulos, M.; O’Hare, D. Rev. Sci.

Instrum. 2012, 83 (8), 084101.

Page 95: Royal Society of Chemistry Coordination and Organometallic

SYNTHESIS AND REACTIVITY OF ZIRCONIUM AND HAFNIUM

BORYLIMIDO COMPOUNDS

Bowen Xie, Benjamin A. Clough, Philip Mountford*

Chemistry Research Laboratory, University of Oxford, Mansfield Road, Oxford, OX1 3TA, U.K.

[email protected]

Complexes containing transition metal-nitrogen multiple bonds have been applied in a range of important organic and inorganic transformations.1 The synthesis and small molecule reactivity of Group 4 imido and hydrazido complexes, (L)M=NR and (L)M=NNR2, have been extensively studied within this research group and by others.2

Of relevance to this contribution, imidozirconocene complexes Cp2Zr=NR(THF) have shown a wild range of reactivity with different small molecules.3 Only five examples of the related borylimides, (L)M=NBR2, have been reported across all of the transition metals to date,4 and we here report the synthesis the first zirconium and halnium borylimido complex analogue by using the borylamine H2NB(NAr'CH)2 (Ar' = 2,6-C6H3

iPr2) (see Figure 1). Furthermore, the zirconium complex was found to give some unusual E-H (E = H, C) activation reactions, which we also report here.

Figure 1: Examples of new zirconium and hafnium borylimido complexes.

References. [1] For example: Acc. Chem. Res., 2005, 38, 955; Acc. Chem. Res., 2005, 38, 839; J.

Am. Chem. Soc., 2001, 123, 2923; John Wily and Sons, 1988 [2] For example: Organometallics, 2011, 30, 1182; J. Am. Chem. Soc., 2004, 126, 1794; Organometallics, 2009, 28, 4747. [3] Chem Rec, 2002, 2, 431 [4] Polyhedron, 1993, 12, 1061; Z. Anorg. Allg. Chem., 2003, 629, 744; Angew. Chem.

Int. Ed. Engl., 2002, 41, 3709-3712; J. Am. Chem. Soc., 2014, 136, 8197.

Page 96: Royal Society of Chemistry Coordination and Organometallic

TRIS(PYRAZOLYL)METHANIDE MAGNESIUM COMPLEXES AS

CATALYSTS FOR THE RING-OPENING POLYMERISATION OF

ε-CL AND rac-LA

Xinning Yin, Insun Yu, Junjuda Unruangsri, and Philip Mountford*

Chemistry Research Laboratory, University of Oxford, Mansfield Road, Oxford, OX1 3TA

[email protected]

Among many face-capping, six-electron N-donor ligands reported in the literature,

tris(pyrazolyl)hydroborate is one of the most widely used. The isoelectronic

tris(pyrazolyl)methane (Tpm) was later introduced as a supporting ligand because C-N

bonds are stronger than B-N bonds, therefore providing robustness in catalyst

activities.1,2

Herein, we report the synthesis and transformation of a series of

tris(pyrazolyl)methanide (Tpmd) magnesium complexes, Scheme 1. (Tpm

d)MgBH4 and

(Tpmd)MgOBn were efficient catalysts for the ring-opening polymerisation (ROP) of ε-

CL and rac-LA. In addition, borate ester derivatives were applied as chain transfer

agents and some ROP results were shown in Figure 1.

Scheme 1. Synthesis of a series of tris(pyrazolyl)methanide magnesium compounds.

0

5000

10000

15000

20000

25000

30000

0 20 40 60 80 100

Mn

(GP

C)/

gm

ol-1

% Conversion

Mn (GPC) 8PhCHOMn (GPC) B(OBn)3Mn (GPC) none线性 (Calcd for 1 chain)线性 (Calcd for 4 chains)

References [1]. Cushion, M.; Meyer, J.; Heath, A. and Mountford, P. Organometallics, 2010, 29, 1174.

[2]. Bigmore, H.; Meyer, J.; Krummenacher, I. and Mountford, P. Chem. Eur. J., 2008, 14, 5918.

Figure 1. Mn (GPC) vs % conversion

plot of ε-CL ROP using TpmdMgBH4,

TpmdMgOBn and various additives in

THF.

Page 97: Royal Society of Chemistry Coordination and Organometallic

OLEFIN POLYMERISATION WITH CARBON-BASED TITANIUM

COMPLEXES

George J.P. Britovsek1, Serge Bettonville

2, Maria Gragert

1, Craig T. Young

1*

1Department of Chemistry, Imperial College London, U.K.

2INEOS Technologies, Brussels, Belgium

*[email protected]

Long chain branching (LCB) in polyethylene (PE) results in behaviour considerably

different to that of purely linear polymer. In particular, LCB levels influence (a) swell

during blow moulding, (b) bubble stability, orientation, melt fracture, and extensional

viscosity in film blowing, (c) melt strength in geomembrane resins, (d) sag during large

diameter pipe extrusion and (e) various mechanical properties in finished articles.[1]

Certain Group 4 single-site catalysts (SSCs) are capable of producing long chain

branching, but the mechanism of formation is debatable.[2-4]

The majority of these SSCs

are metallocene based and/or contain Group 15-16 donor atoms.[3-5]

In comparison,

Group 4 SSCs with η1-bound carbon donor ligands have thus far received only little

attention.

Our aim was to synthesise Group 4 metal SSCs with aromatic carbon donor ligands,

which are η1-bound to the metal centre. We present herein the synthesis of new titanium

complexes with biphenyl (A), diphenylpyridine (B) and terphenyl (C, D) ligands. The

dimethylamido complexes could easily be prepared as opposed to the corresponding

metal halide complexes, which suffered significant decomposition. Preliminary ethylene

polymerisation experiments have shown that these complexes are active polymerisation

catalysts. The syntheses of ligands and metal complexes will be discussed.

[1] McDaniel, M. P.; Rohlfing, D. C.; Benham, E. A. Polym. React. Eng. 2003, 11, 101.

[2] Yang, Q.; Jensen, M. D.; McDaniel, M. P. Macromolecules 2010, 43, 8836.

[3] Reinking, M. K.; Orf, G.; Mcfaddin, D. J. Polym. Sci. A Polym. Chem. 1998, 36,

2889.

[4] Budzelaar, P. H. M. WIREs Comput. Mol. Sci. 2012, 2, 221.

[5] Piel, C.; Stadler, F. J.; Kaschta, J.; Rulhoff, S.; Münstedt, H.; Kaminsky, W.

Macromol. Chem. Phys. 2006, 207, 26.

Page 98: Royal Society of Chemistry Coordination and Organometallic

UNIQUE GROUP 1 CATIONS STABILISED BY HOMOLEPTIC PHOSPHINE COORDINATION

Marina Carravetta, Maria Concistre, William Levason, Gillian Reid and Wenjian Zhang

School of Chemistry, University of Southampton, Southampton UK SO17 1BJ.

[email protected]

Phosphine ligands are ubiquitous in transition metal chemistry, owing to their capacity

to tune the electronic and steric properties, and hence the reactivity, of the complexes.

This has led to wide utilisation of phosphine co-ligands in many transition metal

catalysts and reagents. Phosphine complexes containing p-block acceptors have also

been developed substantially in recent years. However, examples of soft phosphine

coordination towards the hard s-block elements, particularly the Group 1 cations, have

remained extremely elusive. To-date there has been only one reported example of

neutral phosphine coordination to an alkali metal cation, although this is in an

organometallic silylamide dimer [Li{N(Ar)CC(R)Si(R)2NAr}(-Me2PCH2CH2PMe2)]2.

In this poster we report the first series of homoleptic phosphine complexes with group 1

cations, in the form of distorted octahedral Li+ and Na

+ cations containing tris-

diphosphine coordination. These complexes can be achieved by using

Li[Al{OC(CF3)3}4] and Na[BArF] ([BAr

F] = [B{3,5-(CF3)2-C6H3}4]

), which bear

large weakly-coordinating anion and ‘naked’ electrophilic Group 1 cations, together with strong -donor diphosphines, Me2P(CH2)2PMe2 and o-C6H4(PMe2)2.

1. G. Wilkinson, R. D. Gillard, J. A. McCleverty (Eds.), Comprehensive

Coordination Chemistry, Pergamon Press, Oxford, 1987; J. A. McCleverty, T. J.

Meyer (Eds.), Comprehensive Coordination Chemistry II, Elsevier, Oxford, 2004;

2. J. Burt, W. Levason, G. Reid, Coord. Chem. Rev., 2014, 260, 65.

3. R, J. Bowen, M. A. Fernandes, P. B. Hitchcock, M. F. Lappert, M. Lay, J. Chem. Soc., Dalton Trans., 2002, 3253.

4. M. Carravetta, M. Concistre, W. Levason, G. Reid, W. Zhang, Chem. Commun., 2015, 51, 9555.

Page 99: Royal Society of Chemistry Coordination and Organometallic

NAME AFFILIATION E-MAIL

Miss Heba Abdelgawad University of Leeds [email protected]

Ms Gemma Adams University of Oxford [email protected]

Miss Hope Aitchison University of Bristol [email protected]

Mrs Zainab Al-Ali University of Hull [email protected]

Prof Simon Aldridge University of Oxford [email protected]

Mr Matthew Allison University of Leeds [email protected]

Ms Rhiann Andrew University of Warwick [email protected]

Mr Phakpoom

Angpanitcharoen University of Oxford [email protected]

Prof Steve Archibald University of Hull [email protected]

Dr Nicola Bell University of Edinburgh [email protected]

Dr Sophie Benjamin Nottingham Trent University [email protected]

Dr Matthew Blake University of Oxford [email protected]

Dr George Britovsek Imperial College London [email protected]

Mr Lee Brown University of Huddersfield [email protected]

Dr Jean-Charles Buffet University of Oxford [email protected]

Miss Jennifer Burt University of Southampton [email protected]

Mr Michael Butler Imperial College London [email protected]

Ms Alexa Caise University of Oxford [email protected]

Dr Jesus Campos University of Oxford [email protected]

Dr Mark Chadwick University of Oxford [email protected]

Prof Neil Champness University of Nottingham [email protected]

Mr Yao-Pang Chang University of Southampton [email protected]

Dr Adrian Chaplin University of Warwick [email protected]

Mr Michael Chapman University of Leeds [email protected]

Mr Benjamin Clough University of Oxford [email protected]

Mr Lee Collins University of Bath [email protected]

Mr Richard Collins University of Oxford [email protected]

Dr Mark Crimmin Imperial College London [email protected]

Mr Ben Crozier University of Sheffield [email protected]

Mr Massimiliano Curcio University of Edinburgh [email protected]

Mr Mateusz Cybulski University of Bath [email protected]

Dr Scott Dalgarno Heriot-Watt University [email protected]

Miss Nichabhat Diteepeng University of Oxford [email protected]

Mr Do Dinh Cao Huan University of Oxford [email protected]

Dr Juozas Domarkas University of Hull [email protected]

Ms Katie Dryden-Holt RSC [email protected]

Dr Philip Dyer University of Durham [email protected]

Dr Timothy Easun University of Cardiff [email protected]

Mr Jack Emerson-King University of Warwick [email protected]

Mrs Vicki Emms University of Leicester [email protected]

Dr Nicola Farrer University of Oxford [email protected]

Prof Joshua Figueroa UC San Diego [email protected]

Mr Duncan Fraser University of Oxford [email protected]

Miss Kimberley Gallagher University of Bath [email protected]

Miss Karen Gamero-Vega University of Oxford [email protected]

Miss Laura Ghandhi University of Leeds [email protected]

Prof Jose Goicoechea University of Oxford [email protected]

Dr Lucero Gonzalez-

Sebastian University of Warwick [email protected]

Page 100: Royal Society of Chemistry Coordination and Organometallic

NAME AFFILIATION E-MAIL

Mr Ross Green University of Oxford [email protected]

Mr David Griffin University of Sheffield [email protected]

Mr Matthew Gyton University of Warwick [email protected]

Ms Harina Amer Hamzah University of Bath [email protected]

Miss Samantha Hawken University of Southampton [email protected]

Dr Anthony Haynes University of Sheffield [email protected]

Miss Louise Hazeland University of Bristol [email protected]

Miss Alexandra Hicken Imperial College London [email protected]

Dr Graeme Hogarth King's College London [email protected]

Mr Jordan Holmes University of Leeds [email protected]

Dr Thomas Hooper University of Oxford [email protected]

Ms Chloe Johnson University College Dublin [email protected]

Dr Guy Jones RSC [email protected]

Miss Boon-Uma Jowanaridhi University of Hull [email protected]

Mr Andrew Jupp University of Oxford [email protected]

Dr Ann Keep Industrialist [email protected]

Dr Sandy Kilpatrick University of Oxford [email protected]

Mr Andrew King University of Bath [email protected]

Miss Jane Knichal University of Bath [email protected]

Dr Richard Knighton University of Warwick [email protected]

Dr Evgeny Kolychev University of Oxford [email protected]

Dr George Kostakis University of Sussex [email protected]

Ms Lilja Kristinsdottir University of Oxford [email protected]

Mr Amit Kumar University of Oxford [email protected]

Prof Richard Layfield University of Manchester [email protected]

Mr Siu Kwan Lo University of Oxford [email protected]

Dr Jason Love University of Edinburgh [email protected]

Dr Mark Lowe University of Leicester [email protected]

Miss Cecilia Madzivire University of Leeds [email protected]

Dr Stephen Mansell Heriot-Watt University [email protected]

Dr Patrick McGowan University of Leeds [email protected]

Miss Simona Mellino University of Oxford [email protected]

Mr Owen Metters University of Bristol [email protected]

Miss Krishna Mistry University of Bristol [email protected]

Mr Stefan Mitzinger University of Marburg [email protected]

Dr Zhenbo Mo University of Oxford [email protected]

Mr Francesco Matteo

Monzittu University of Southampton [email protected]

Prof Philip Mountford University of Oxford [email protected]

Ms Raihan Musa University of Oxford [email protected]

Miss Rebecca Musgrave University of Bristol [email protected]

Dr Simon Neil RSC [email protected]

Mr Rob Newland Heriot-Watt University [email protected]

Dr Paul Newman University of Cardiff [email protected]

Mr Haoyu Niu University of Oxford [email protected]

Mr James Pankhurst University of Edinburgh [email protected]

Ms Ruth Patchett University of Warwick [email protected]

Dr Nathan Patmore University of Huddersfield [email protected]

Mr Ben Peerless University of Sheffield [email protected]

Mrs Katy Pellow University of Bristol [email protected]

Page 101: Royal Society of Chemistry Coordination and Organometallic

NAME AFFILIATION E-MAIL

Dr Andreas Phanopoulos Imperial College London [email protected]

Dr Simon Pope University of Cardiff [email protected]

Dr Peter Portius University of Sheffield [email protected]

Dr Amparo Prades University of Oxford [email protected]

Dr Andrey Protchenko University of Oxford [email protected]

Dr David Pugh University of Southampton [email protected]

Prof Gill Reid University of Southampton [email protected]

Mr Diego A. Resendiz Lara University of Bristol [email protected]

Mr Jose Adan Reyes-Sanchez University of Oxford [email protected]

Dr Ian Riddlestone University of Bath [email protected]

Dr. Arnab Rit University of Oxford [email protected]

Dr Thomas Robinson University of Oxford [email protected]

Dr Marta Rosello-Merino Heriot-Watt University [email protected]

Mr Wasupol Rungtanapirom University of Leeds [email protected]

Mrs Sara Shahruddin Imperial College London [email protected]

Mr Tim Shuttleworth University of Bristol [email protected]

Ms Anna Spearing-Ewyn University of Oxford [email protected]

Ms Caroline Storey University of Warwick [email protected]

Prof Peter Tasker University of Edinburgh [email protected]

Dr Atanas Tomov Imperial College London [email protected]

Dr Thibault Troadec University of Warwick [email protected]

Mr Josh Turner University of Bristol [email protected]

Dr Zoe Turner University of Oxford [email protected]

Dr Kevin Vincent University of Huddersfield [email protected]

Prof Kylie Vincent University of Oxford [email protected]

Dr Charles Vriamont Imperial College London [email protected]

Dr Richard Walker RSC [email protected]

Dr Benjamin Ward University of Cardiff [email protected]

Mr Jordan Waters University of Oxford [email protected]

Prof Andrew Weller University of Oxford [email protected]

Dr Charlotte Willans University of Leeds [email protected]

Mr Luke Wilkinson University of Sheffield [email protected]

Mr Christopher Wright University of Oxford [email protected]

Dr Yue Wu University of Oxford [email protected]

Mr Bowen Xie University of Oxford [email protected]

Miss Xinning Yin University of Oxford [email protected]

Mr Craig Young Imperial College London [email protected]

Dr Insun Yu University of Oxford [email protected]

Dr Wenjian Zhang University of Southampton [email protected]

Dr Sergey Zlatogorsky

University of Central

Lancashire [email protected]