Co-reporter:Jerico R. Piper;Lewis Cletheroe;Christopher G. P. Taylor;Alexander J. Metherell;Julia A. Weinstein;Igor V. Sazanovich;Michael D. Ward
Chemical Communications 2017 vol. 53(Issue 2) pp:408-411
Publication Date(Web):2016/12/22
DOI:10.1039/C6CC09298E
In a coordination cage which contains an array of twelve naphthyl chromophores surrounding a central cavity, photoinduced energy or electron-transfer can occur from the chromophore array to the bound guest in supramolecular host/guest complexes.
Co-reporter:Zainab N. Zubaidi, Alexander J. Metherell, Elizabeth Baggaley, Michael D. Ward
Polyhedron 2017 Volume 133(Volume 133) pp:
Publication Date(Web):5 September 2017
DOI:10.1016/j.poly.2017.05.017
A new bridging ligand has been prepared in which two chelating bidentate pyrazolyl-pyridine termini are connected to a central naphthalene-2,7-diyl core via methylene spacer units. This ligand has been used to prepared mononuclear and dinuclear Ir(III) complexes in which {Ir(F2ppy)2} [F2ppy = cyclometallating anion of 2-(3,5-difluorophenyl)-pyridine] complex fragments are coordinated to one or both of the pyrazolyl-pyridine termini; in addition a heterodinuclear complex has been prepared containing one {Ir(F2ppy)2} unit and one {Re(CO)3Cl} unit in the two binding sites. X-ray crystallographic studies show that the bridging naphthyl group lies stacked with a coordinated F2ppy ligand from a terminal {Ir(F2ppy)2} unit in every case. Luminescence measurements show that the usual strong Ir(III)-centred blue luminescence is substantially quenched by the presence of a low-lying triplet state on the naphthyl group; in the Ir(III)/Re(I) dyad we observe both weak Ir(III)-based emission as well as lower-energy Re(I)-based emission which overlap; Ir(III) → Re(I) energy-transfer occurs on a timescale of <1 ns as no rise-time for sensitised Re(I)-based emission could be detected, in contrast to other Ir(III)/Re(I) dyads in which the Ir(III) → Re(I) energy-transfer is slower (10–100 ns timescale). We ascribe this to the spatial and energetic intermediacy of the naphthyl group whose triplet energy lies between that of the Ir(III) and Re(I) termini, providing an effective conduit for energy-transfer to occur.Mono- and di-nuclear Ir(III) complexes, and a heterodinuclear Ir(III)/Re(I) complex, of a new bridging ligand containing N,N-bidentate chelating termini are reported. The presence of a naphthyl unit in the bridging ligand, whose triplet energy lies between those of the Ir(III) and Re(I) units, has a substantial effect on the structural and photophysical properties of the complexes.Download high-res image (68KB)Download full-size image
Co-reporter:James S. Wright;Alexander J. Metherell;William M. Cullen;Jerico R. Piper;Robert Dawson;Michael D. Ward
Chemical Communications 2017 vol. 53(Issue 31) pp:4398-4401
Publication Date(Web):2017/04/13
DOI:10.1039/C7CC01959A
Two M8L12 cubic coordination cages, as desolvated crystalline powders, preferentially adsorb CO2 over N2 with ideal selectivity CO2/N2 constants of 49 and 30 at 298 K. A binding site for CO2 is suggested by crystallographic location of CS2 within the cage cavity at an electropositive hydrogen-bond donor site, potentially explaining the high CO2/N2 selectivity compared to other materials with this level of porosity.
Co-reporter:Alexander J. Metherell and Michael D. Ward
Chemical Science 2016 vol. 7(Issue 2) pp:910-915
Publication Date(Web):14 Oct 2015
DOI:10.1039/C5SC03526K
Retrosynthetic analysis of a [M16L24]32+ coordination cage shows how it can be assembled rationally, in a stepwise manner, using a combination of kinetically inert and kinetically labile components. Combination of the components of fac-[Ru(Lph)3](PF6)2, Cd(BF4)2 and Lnaph in the necessary 4:12:12 stoichiometry afforded crystals of [Ru4Cd12(Lph)12(Lnaph)12]X32 (X = a mono-anion) in which the location of the two types of metal ion [Ru(II) or Cd(II)] at specific vertices in the metal-ion array, and the two types of bridging ligand (Lph and Lnaph) along specific edges, is completely controlled by the synthetic strategy. The incorporation of four different types of component at pre-determined positions in a coordination cage superstructure represents a substantial advance in imposing control on the self-assembly of complex metallosupramolecular entities.
Co-reporter:Christopher G. P. Taylor, Jerico R. Piper and Michael D. Ward
Chemical Communications 2016 vol. 52(Issue 37) pp:6225-6228
Publication Date(Web):22 Mar 2016
DOI:10.1039/C6CC02021F
Cubic coordination cages act as competent hosts for several alkyl phosphonates used as chemical warfare agent simulants; a range of cage/guest structures have been determined, contributions to guest binding analysed, and a fluorescent response to guest binding demonstrated.
Co-reporter:Atanu Jana, Bethany J. Crowston, Jonathan R. Shewring, Luke K. McKenzie, Helen E. Bryant, Stanley W. Botchway, Andrew D. Ward, Angelo J. Amoroso, Elizabeth Baggaley, and Michael D. Ward
Inorganic Chemistry 2016 Volume 55(Issue 11) pp:5623
Publication Date(Web):May 24, 2016
DOI:10.1021/acs.inorgchem.6b00702
Luminescent, mixed metal d–f complexes have the potential to be used for dual (magnetic resonance imaging (MRI) and luminescence) in vivo imaging. Here, we present dinuclear and trinuclear d–f complexes, comprising a rigid framework linking a luminescent Ir center to one (Ir·Ln) or two (Ir·Ln2) lanthanide metal centers (where Ln = Eu(III) and Gd(III), respectively). A range of physical, spectroscopic, and imaging-based properties including relaxivity arising from the Gd(III) units and the occurrence of Ir(III) → Eu(III) photoinduced energy-transfer are presented. The rigidity imposed by the ligand facilitates high relaxivities for the Gd(III) complexes, while the luminescence from the Ir(III) and Eu(III) centers provide luminescence imaging capabilities. Dinuclear (Ir·Ln) complexes performed best in cellular studies, exhibiting good solubility in aqueous solutions, low toxicity after 4 and 18 h, respectively, and punctate lysosomal staining. We also demonstrate the first example of oxygen sensing in fixed cells using the dyad Ir·Gd, via two-photon phosphorescence lifetime imaging (PLIM).
Co-reporter:Suad T. Saad, Alexander J. Metherell, Elizabeth Baggaley and Michael D. Ward
Dalton Transactions 2016 vol. 45(Issue 28) pp:11568-11579
Publication Date(Web):27 Jun 2016
DOI:10.1039/C6DT01614F
A series of dinuclear Ir(III)/Re(I) complexes has been prepared based on a family of symmetrical bridging ligands containing two bidentate N,N′-chelating pyrazolyl–pyridine termini, connected by a central aromatic or aliphatic spacer. The Ir(III) termini are based on {Ir(F2ppy)2}+ units (where F2ppy is the cyclometallating anion of a fluorinated phenylpyridine) and the Re(I) termini are based on {Re(CO)3Cl} units. Both types of terminus are luminescent, with the Ir-based unit showing characteristic strong, structured phosphorescence in the blue region (maximum 452 nm) with a triplet excited state energy of 22200 cm−1 and the Re-based unit showing much weaker and lower-energy phosphorescence (maximum 530 nm) with a triplet excited state energy of 21300 cm−1. The energy gradient between the two excited states allows for partial Ir→Re photoinduced energy-transfer, with substantial (but incomplete) quenching of the higher-energy Ir-based emission component and sensitised emission – evidenced by an obvious grow-in component – on the lower-energy Re-based emission. The Ir→Re energy-transfer rate constants vary over the range 1–8 × 107 s−1 depending on the bridging ligand: there is no simple correlation between bridging ligand structure and energy-transfer rate, possibly because this will depend substantially on the conformation of these flexible molecules in solution. To test the role of ligand conformation further, we investigated a complex in which the bridging chain is a (CH2CH2O)6 unit whose conformation is known to be solvent-polarity dependent, with such chains adopting an open, elongated conformation in water and more compact, folded conformations in organic solvents. There was a clear link between the rate and extent of Ir→Re energy-transfer which reduced in polar solvents as the chain became elongated and the Ir/Re separation was larger; and increased in less polar solvents as the chain adopted a more compact conformation and the Ir/Re separation was reduced.
Co-reporter:Alexander J. Metherell and Michael D. Ward
RSC Advances 2016 vol. 6(Issue 13) pp:10750-10762
Publication Date(Web):14 Jan 2016
DOI:10.1039/C5RA22694E
Two families of heteronuclear coordination complexes have been prepared in a stepwise manner using pre-formed, kinetically inert [RuL3]2+ building blocks, in which L is a bis-bidentate bridging ligand with two pyrazole–pyridyl termini, coordinated at one end to the Ru(II) centre. These pre-formed ‘complex ligands’ – with three pendant binding sites – react with additional labile transition metal dications to complete the stepwise assembly of mixed-metal arrays in which labile [Co(II)/Cd(II)] or inert [Ru(II)] ions strictly alternate around the framework. When L = the thiophene-2,5-diyl spaced ligand Lth, the complex [Ru(Lth)3]2+ is formed in the expected 3:1 mer:fac ratio: reaction with labile Co(II) or Cd(II) ions completes formation of a heteronuclear square [Ru2Co2(Lth)6]8+ or one-dimensional coordination polymer {[CdRu(Lth)3]4+}∞, respectively. In these only the mer isomer of [Ru(Lth)3]2+ is selected by the self-assembly process, whereas the fac isomer is not used. When L = a 1,3-benzene-diyl spaced ligand (Lph), the complex ligand [Ru(Lph)3]2+ formed in the initial step is enriched in mer isomer (80–87% mer, depending on reaction conditions). Two quite different products were isolated from reaction of [Ru(Lph)3]2+ with Co(II) depending on the conditions. These are the rectangular, hexanuclear ‘open-book’ array [Ru3Co3(Lph)9]12+ which contains a 2:1 proportion of fac/mer Ru(II) metal centres; and the octanuclear cubic [Ru4Co4(Lph)12{Na(BF4)4}]13+ cage which is a new structural type containing all mer Ru(II) vertices and all fac Co(II) vertices. The cavity of this cubic cage contains a tetrahedral array of fluoroborate anions which in turn coordinate to a central Na(I) ion – an unusual example of a metal complex [Na(BF4)4]3− acting as the guest inside the cage-like metal complex [Ru4Co4(Lph)12]16+.
Co-reporter:Alexander J. Metherell, Michael D. Ward
Polyhedron 2016 Volume 103(Part B) pp:206-216
Publication Date(Web):8 January 2016
DOI:10.1016/j.poly.2015.11.010
The mononuclear complexes [Ru(LoPh)2](PF6)2 and fac-[Ru(L3-Py)3](PF6)2 have been prepared which contain different types of pendant binding site. In [Ru(LoPh)2](PF6)2, one ligand acts as a tetradentate chelate but the other is only bidentate such that there is one bidentate pyrazolyl-pyridine site pendant from the Ru(II) complex core. In fac-[Ru(L3-Py)3](PF6)2 in contrast there are three monodentate 3-pyridyl units pendant from the same face of the Ru(II) core. Two [Ru(LoPh)2](PF6)2 units assemble around an Ag(I) ion to give trinuclear [{Ru(LoPh)2}2Ag](PF6)5 in which the Ag(I) ion is four coordinate. In the solid state three of these associate into a triangular trimeric array via π-stacking and CH⋯F hydrogen-bonding interactions to give an assembly which contains two cup-shaped cavities either side of a central Ag3 face and pointing in opposite directions; these cavities contain H-bonded hexafluorophosphate anions. In MeCN solution at high concentrations (11 mM), according to 1H NMR DOSY measurements the trinuclear complex [{Ru(LoPh)2}2Ag](PF6)5 remains intact but there is no evidence for further aggregation into the trimer. Two fac-[Ru(L3-Py)3](PF6)2 units combine with two {Ag(NCMe)}+ units and an additional Ag(I) ion to give a complex coordination network [{Ru(L3-Py)3}2{Ag(MeCN)2}2Ag](PF6)7. In this structure two fac-[Ru(L3-Py)3](PF6)2 and two {Ag(NCMe)}+ units form tetranuclear Ru2Ag2 rings via Ag–N(pyridyl) interactions; these rings are crosslinked into a one-dimensional chain by additional Ag–N(pyridyl) interactions. Argentophilic Ag⋯Ag interactions then connect the resulting chains into two-dimensional sheets. Both mixed-metal networks are based on initial formation of kinetically inert Ru(II) complexes which then associate in a separate step with Ag(I) ions, via pendant N-donor binding sites. However [{Ru(LoPh)2}2Ag](PF6)5 exploits π–π stacking and hydrogen-bonding interactions to form in the solid state a trimeric assembly which encapsulates anions; in contrast [{Ru(L3-Py)3}2{Ag(MeCN)2}2Ag](PF6)7 exploits argentophilic interactions to complete formation of the network.Simple mononuclear Ru(II) complexes with pendant N-donor bidentate binding sites assemble around Ag(I) ions to give Ru(II)/Ag(I) supramolecular assemblies in which H-bonding, π-stacking and Ag⋯Ag argentophilic interactions contribute to the unusual network structures.
Co-reporter:William Cullen, Simon Turega, Christopher A. Hunter and Michael D. Ward
Chemical Science 2015 vol. 6(Issue 1) pp:625-631
Publication Date(Web):31 Jul 2014
DOI:10.1039/C4SC02090A
A range of organic molecules with acidic or basic groups exhibit strong pH-dependent binding inside the cavity of a polyhedral coordination cage. Guest binding in aqueous solution is dominated by a hydrophobic contribution which is compensated by stronger solvation when the guests become cationic (by protonation) or anionic (by deprotonation). The Parkinson's drug 1-amino-adamantane (‘amantadine’) binds with an association constant of 104 M−1 in the neutral form (pH greater than 11), but the stability of the complex is reduced by three orders of magnitude when the guest is protonated at lower pH. Monitoring the uptake of the guests into the cage cavity was facilitated by the large upfield shift for the 1H NMR signals of bound guests due to the paramagnetism of the host. Although the association constants are generally lower, guests of biological significance such as aspirin and nicotine show similar behaviour, with a substantial difference between neutral (strongly binding) and charged (weakly binding) forms, irrespective of the sign of the charged species. pH-dependent binding was observed for a range of guests with different functional groups (primary and tertiary amines, pyridine, imidazole and carboxylic acids), so that the pH-swing can be tuned anywhere in the range of 3.5–11. The structure of the adamantane-1-carboxylic acid complex was determined by X-ray crystallography: the oxygen atoms of the guest form CH⋯O hydrogen bonds with one of two equivalent pockets on the internal surface of the host. Reversible uptake and release of guests as a function of pH offers interesting possibilities in any application where controlled release of a molecule following an external stimulus is required.
Co-reporter:William Cullen, Simon Turega, Christopher A. Hunter and Michael D. Ward
Chemical Science 2015 vol. 6(Issue 5) pp:2790-2794
Publication Date(Web):10 Mar 2015
DOI:10.1039/C5SC00534E
The protein/ligand docking software GOLD, which was originally developed for drug discovery, has been used in a virtual screen to identify small molecules that bind with extremely high affinities (K ≈ 107 M−1) in the cavity of a cubic coordination cage in water. A scoring function was developed using known guests as a training set and modified by introducing an additional term to take account of loss of guest flexibility on binding. This scoring function was then used in GOLD to successfully identify 15 new guests and accurately predict the binding constants. This approach provides a powerful predictive tool for virtual screening of large compound libraries to identify new guests for synthetic hosts, thereby greatly simplifying and accelerating the process of identifying guests by removing the reliance on experimental trial-and-error.
Co-reporter:William Cullen, Katie A. Thomas, Christopher A. Hunter and Michael D. Ward
Chemical Science 2015 vol. 6(Issue 7) pp:4025-4028
Publication Date(Web):07 May 2015
DOI:10.1039/C5SC01475A
We demonstrate the use of a simple pH swing to control the selection of one of three different guests from aqueous solution by a coordination cage host. Switching between different guests is based on the fact that neutral organic guests bind strongly in the cage due to the hydrophobic effect, but for acidic or basic guests, the charged (protonated or deprotonated) forms are hydrophilic and do not bind. The guests used are adamantane-1,3-dicarboxylic acid (H2A) which binds at low pH when it is neutral but not when it is deprotonated; 1-amino-adamantane (B) which binds at high pH when it is neutral but not when it is protonated; and cyclononanone (C) whose binding is not pH dependent and is therefore the default guest at neutral pH. Thus an increase in pH can reversibly switch the host between the three different bound states cage·H2A (at low pH), cage·C (at medium pH), and cage·B (at high pH) in succession.
Co-reporter:Atanu Jana, Elizabeth Baggaley, Angelo Amoroso and Michael D. Ward
Chemical Communications 2015 vol. 51(Issue 42) pp:8833-8836
Publication Date(Web):21 Apr 2015
DOI:10.1039/C5CC02130H
A new rigid and conjugated ligand structure connecting phenanthroline and poly(amino-carboxylate) binding sites provides d–f complexes which show high potential for use in dual (luminescence + magnetic resonance) imaging and for optimisation of d → f photoinduced energy-transfer.
Co-reporter:William Cullen; Christopher A. Hunter;Michael D. Ward
Inorganic Chemistry 2015 Volume 54(Issue 6) pp:2626-2637
Publication Date(Web):February 20, 2015
DOI:10.1021/ic502780b
The self-assembly between a water-soluble bis-bidentate ligand L18w and Co(II) salts in water affords three high-spin Co(II) products: a dinuclear meso-helicate [Co2(L18w)3]X4; a tetrahedral cage [Co4(L18w)6]X8; and a dodecanuclear truncated-tetrahedral cage [Co12(L18w)18]X24 (X = BF4 or ClO4). All three products were crystallized under different conditions and structurally characterized. In [Co2(L18w)3]X4 all three bridging ligands span a pair of metal ions; in the two larger products, there is a metal ion at each vertex of the Co4 or Co12 polyhedral cage array with a bridging ligand spanning a pair of metal ions along every edge. All three structural types are known: what is unusual here is the presence of all three from the same reaction. The assemblies Co2, Co4, and Co12 are in slow equilibrium (hours/days) in aqueous solution, and this can be conveniently monitored by 1H NMR spectroscopy because (i) the paramagnetism of Co(II) disperses the signals over a range of ca. 200 ppm and (ii) the different symmetries of the three species give characteristically different numbers of independent 1H NMR signals, which makes identification easy. From temperature- and concentration-dependent 1H NMR studies it is clear that increasing temperature and increasing dilution favors fragmentation to give a larger proportion of the smaller assemblies for entropic reasons. High concentrations and low temperature favor the larger assembly despite the unfavorable entropic and electrostatic factors associated with its formation. We suggest that this arises from the hydrophobic effect: reorganization of several smaller complexes into one larger one results in a smaller proportion of the hydrophobic ligand surface being exposed to water, with a larger proportion of the ligand surface protected in the interior of the assembly. In agreement with this, 1H NMR spectra in a nonaqueous solvent (MeNO2) show formation of only [Co2(L18w)3]X4 because the driving force for reorganization into larger assemblies is now absent. Thus, we can identify the contributions of temperature, concentration, and solvent on the result of the metal/ligand self-assembly process and have determined the speciation behavior of the Co2/Co4/Co12 system in aqueous solution.
Co-reporter:Ashley B. Wragg, Alexander J. Metherell, William Cullen and Michael D. Ward
Dalton Transactions 2015 vol. 44(Issue 41) pp:17939-17949
Publication Date(Web):21 Sep 2015
DOI:10.1039/C5DT02957K
Stepwise preparation of the heterometallic octanuclear coordination cages [(Ma)4(Mb)4L12]16+ is reported, in which Ma = Ru or Os and Mb = Cd or Co (all in their +2 oxidation state). This requires initial preparation of the kinetically inert mononuclear complexes [(Ma)L3]2+ in which L is a ditopic ligand with two bidentate chelating pyrazolyl-pyridine units: in the complexes [(Ma)L3]2+ one terminus of each ligand is bound to the metal ion, such that the complex has three pendant bidentate sites at which cage assembly can propagate by coordination to additional labile ions Mb in a separate step. Thus, combination of four [(Ma)L3]2+ units and four [Mb]2+ ions results in assembly of the complete cages [(Ma)4(Mb)4L12]16+ in which a metal ion lies at each of the eight vertices, and a bridging ligand spans each of the twelve edges, of a cube. The different types of metal ion necessarily alternate around the periphery with each bridging ligand bound to one metal ion of each type. All four cages have been structurally characterised: in the Ru(II)/Cd(II) cage (reported in a recent communication) the Ru(II) and Cd(II) ions are crystallographically distinct; in the other three cages [Ru(II)/Co(II), Os(II)/Cd(II) and Os(II)/Co(II), reported here] the ions are disordered around the periphery such that every metal site refines as a 50:50 mixture of the two metal atom types. The incorporation of Os(II) units into the cages results in both redox activity [a reversible Os(II)/Os(III) couple for all four metal ions simultaneously, at a modest potential] and luminescence [the Os(II) units have luminescent 3MLCT excited states which will be good photo-electron donors] being incorporated into the cage superstructure.
Co-reporter:Alexander J. Metherell, Michael D. Ward
Polyhedron 2015 Volume 89() pp:260-270
Publication Date(Web):29 March 2015
DOI:10.1016/j.poly.2015.01.027
The mononuclear complexes [Ru(LoPh)2](PF6)2 and fac-[Ru(L3-Py)3](PF6)2 have been prepared which contain different types of pendant binding site. In [Ru(LoPh)2](PF6)2, one ligand acts as a tetradentate chelate but the other is only bidentate such that there is one bidentate pyrazolyl-pyridine site pendant from the Ru(II) complex core. In fac-[Ru(L3-Py)3](PF6)2 in contrast there are three monodentate 3-pyridyl units pendant from the same face of the Ru(II) core. Two [Ru(LoPh)2](PF6)2 units assemble around an Ag(I) ion to give trinuclear [{Ru(LoPh)2}2Ag](PF6)5 in which the Ag(I) ion is four coordinate. In the solid state three of these associate into a triangular trimeric array via π-stacking and CH⋯F hydrogen-bonding interactions to give an assembly which contains two cup-shaped cavities either side of a central Ag3 face and pointing in opposite directions; these cavities contain H-bonded hexafluorophosphate anions. In MeCN solution at high concentrations (11 mM), according to 1H NMR DOSY measurements the trinuclear complex [{Ru(LoPh)2}2Ag](PF6)5 remains intact but there is no evidence for further aggregation into the trimer. Two fac-[Ru(L3-Py)3](PF6)2 units combine with two {Ag(NCMe)}+ units and an additional Ag(I) ion to give a complex coordination network [{Ru(L3-Py)3}2{Ag(MeCN)2}2Ag](PF6)7. In this structure two fac-[Ru(L3-Py)3](PF6)2 and two {Ag(NCMe)}+ units form tetranuclear Ru2Ag2 rings via Ag–N(pyridyl) interactions; these rings are crosslinked into a one-dimensional chain by additional Ag–N(pyridyl) interactions. Argentophilic Ag⋯Ag interactions then connect the resulting chains into two-dimensional sheets. Both mixed-metal networks are based on initial formation of kinetically inert Ru(II) complexes which then associate in a separate step with Ag(I) ions, via pendant N-donor binding sites. However [{Ru(LoPh)2}2Ag](PF6)5 exploits π–π stacking and hydrogen-bonding interactions to form in the solid state a trimeric assembly which encapsulates anions; in contrast [{Ru(L3-Py)3}2{Ag(MeCN)2}2Ag](PF6)7 exploits argentophilic interactions to complete formation of the network.Simple mononuclear Ru(II) complexes with pendant N-donor bidentate binding sites assemble around Ag(I) ions to give Ru(II)/Ag(I) supramolecular assemblies in which H-bonding, π-stacking and Ag⋯Ag argentophilic interactions contribute to the unusual network structures.
Co-reporter:Simon Turega ; William Cullen ; Martina Whitehead ; Christopher A. Hunter ;Michael D. Ward
Journal of the American Chemical Society 2014 Volume 136(Issue 23) pp:8475-8483
Publication Date(Web):May 19, 2014
DOI:10.1021/ja504269m
Size and shape criteria for guest binding inside the cavity of an octanuclear cubic coordination cage in water have been established using a new fluorescence displacement assay to quantify guest binding. For aliphatic cyclic ketones of increasing size (from C5 to C11), there is a linear relationship between ΔG for guest binding and the guest’s surface area: the change in ΔG for binding is 0.3 kJ mol–1 Å–2, corresponding to 5 kJ mol–1 for each additional CH2 group in the guest, in good agreement with expectations based on hydrophobic desolvation. The highest association constant is K = 1.2 × 106 M–1 for cycloundecanone, whose volume is approximately 50% of the cavity volume; for larger C12 and C13 cyclic ketones, the association constant progressively decreases as the guests become too large. For a series of C10 aliphatic ketones differing in shape but not size, ΔG for guest binding showed no correlation with surface area. These guests are close to the volume limit of the cavity (cf. Rebek’s 55% rule), so the association constant is sensitive to shape complementarity, with small changes in guest structure resulting in large changes in binding affinity. The most flexible members of this series (linear aliphatic ketones) did not bind, whereas the more preorganized cyclic ketones all have association constants of 104–105 M–1. A crystal structure of the cage·cycloundecanone complex shows that the guest carbonyl oxygen is directed into a binding pocket defined by a convergent set of CH groups, which act as weak hydrogen-bond donors, and also shows close contacts between the exterior surface of the disc-shaped guest and the interior surface of the pseudospherical cage cavity despite the slight mismatch in shape.
Co-reporter:Alexander J. Metherell and Michael D. Ward
Chemical Communications 2014 vol. 50(Issue 75) pp:10979-10982
Publication Date(Web):31 Jul 2014
DOI:10.1039/C4CC05421K
The geometrically pure ‘complex ligand’ fac-[Ru(Lph)3]2+, in which three pendant bidentate binding sites are located on one face of the complex, reacts with Ag(I) ions to form the adamantoid decanuclear cage [{Ru(Lph)3}4Ag6](PF6)14 which contains a 6-coordinate Ru(II) ion at each vertex of a large tetrahedron and a 4-coordinate Ag(I) ion along each edge.
Co-reporter:Alexander J. Metherell, William Cullen, Andrew Stephenson, Christopher A. Hunter and Michael D. Ward
Dalton Transactions 2014 vol. 43(Issue 1) pp:71-84
Publication Date(Web):2013/10/16
DOI:10.1039/C3DT52479E
We have prepared a series of mononuclear fac and mer isomers of Ru(II) complexes containing chelating pyrazolyl-pyridine ligands, to examine their differing ability to act as hydrogen-bond donors in MeCN. This was prompted by our earlier observation that octanuclear cube-like coordination cages that contain these types of metal vertex can bind guests such as isoquinoline-N-oxide (K = 2100 M−1 in MeCN), with a significant contribution to binding being a hydrogen-bonding interaction between the electron-rich atom of the guest and a hydrogen-bond donor site on the internal surface of the cage formed by a convergent set of CH2 protons close to a 2+ metal centre. Starting with [Ru(LH)3]2+ [LH = 3-(2-pyridyl)-1H-pyrazole] the geometric isomers were separated by virtue of the fact that the fac isomer forms a Cu(I) adduct which the mer isomer does not. Alkylation of the pyrazolyl NH group with methyl iodide or benzyl bromide afforded [Ru(LMe)3]2+ and [Ru(Lbz)3]2+ respectively, each as their fac and mer isomers; all were structurally characterised. In the fac isomers the convergent group of pendant –CH2R or –CH3 protons defines a hydrogen-bond donor pocket; in the mer isomer these protons do not converge and any hydrogen-bonding involving these protons is expected to be weaker. For both [Ru(LMe)3]2+ and [Ru(Lbz)3]2+, NMR titrations with isoquinoline-N-oxide in MeCN revealed weak 1:1 binding (K ≈ 1 M−1) between the guest and the fac isomer of the complex that was absent with the mer isomer, confirming a difference in the hydrogen-bond donor capabilities of these complexes associated with their differing geometries. The weak binding compared to the cage however occurs because of competition from the anions, which are free to form ion-pairs with the mononuclear complex cations in a way that does not happen in the cage complexes. We conclude that (i) the presence of fac tris-chelate sites in the cage to act as hydrogen-bond donors, and (ii) exclusion of counter-ions from the central cavity leaving these hydrogen-bonding sites free to interact with guests, are both important design criteria for future coordination cage hosts.
Co-reporter:Daniel Sykes, Ahmet J. Cankut, Noorshida Mohd Ali, Andrew Stephenson, Steven J. P. Spall, Simon C. Parker, Julia A. Weinstein and Michael D. Ward
Dalton Transactions 2014 vol. 43(Issue 17) pp:6414-6428
Publication Date(Web):25 Feb 2014
DOI:10.1039/C4DT00292J
A series of blue-luminescent Ir(III) complexes with a pendant binding site for lanthanide(III) ions has been synthesized and used to prepare Ir(III)/Ln(III) dyads (Ln = Eu, Tb, Gd). Photophysical studies were used to establish mechanisms of Ir→Ln (Ln = Tb, Eu) energy-transfer. In the Ir/Gd dyads, where direct Ir→Gd energy-transfer is not possible, significant quenching of Ir-based luminescence nonetheless occurred; this can be ascribed to photoinduced electron-transfer from the photo-excited Ir unit (*Ir, 3MLCT/3LC excited state) to the pendant pyrazolyl-pyridine site which becomes a good electron-acceptor when coordinated to an electropositive Gd(III) centre. This electron transfer quenches the Ir-based luminescence, leading to formation of a charge-separated {Ir4+}˙—(pyrazolyl-pyridine)˙− state, which is short-lived possibly due to fast back electron-transfer (<20 ns). In the Ir/Tb and Ir/Eu dyads this electron-transfer pathway is again operative and leads to sensitisation of Eu-based and Tb-based emission using the energy liberated from the back electron-transfer process. In addition direct Dexter-type Ir→Ln (Ln = Tb, Eu) energy-transfer occurs on a similar timescale, meaning that there are two parallel mechanisms by which excitation energy can be transferred from *Ir to the Eu/Tb centre. Time-resolved luminescence measurements on the sensitised Eu-based emission showed both fast and slow rise-time components, associated with the PET-based and Dexter-based energy-transfer mechanisms respectively. In the Ir/Tb dyads, the Ir→Tb energy-transfer is only just thermodynamically favourable, leading to rapid Tb→Ir thermally-activated back energy-transfer and non-radiative deactivation to an extent that depends on the precise energy gap between the *Ir and Tb-based 5D4 states. Thus, the sensitised Tb(III)-based emission is weak and unusually short-lived due to back energy transfer, but nonetheless represents rare examples of Tb(III) sensitisation by a energy donor that could be excited using visible light as opposed to the usually required UV excitation.
Co-reporter:Dr. Elizabeth Baggaley; Deng-Ke Cao;Dr. Daniel Sykes; Stanley W. Botchway;Dr. Julia A. Weinstein; Michael D. Ward
Chemistry - A European Journal 2014 Volume 20( Issue 29) pp:8898-8903
Publication Date(Web):
DOI:10.1002/chem.201403618
Abstract
The first example of cell imaging using two independent emission components from a dinuclear d/f complex is reported. A water-stable, cell-permeable IrIII/EuIII dyad undergoes partial IrEu energy transfer following two-photon excitation of the Ir unit at 780 nm. Excitation in the near-IR region generated simultaneously green Ir-based emission and red Eu-based emission from the same probe. The orders-of-magnitude difference in their timescales (Ir ca. μs; Eu ca. 0.5 ms) allowed them to be identified by time-gated detection. Phosphorescence lifetime imaging microscopy (PLIM) allowed the lifetime of the Ir-based emission to be measured in different parts of the cell. At the same time, the cells are simultaneously imaged by using the Eu-based emission component at longer timescales. This new approach to cellular imaging by using dual d/f emitters should therefore enable autofluorescence-free sensing of two different analytes, independently, simultaneously and in the same regions of a cell.
Co-reporter:Michael D. Ward and Paul R. Raithby
Chemical Society Reviews 2013 vol. 42(Issue 4) pp:1619-1636
Publication Date(Web):13 Jul 2012
DOI:10.1039/C2CS35123D
Many naturally occurring systems show us how multi-component supramolecular assemblies can generate useful functional behaviour. In this article the problems and limitations associated with achieving such behaviour in artificial multi-component assemblies is discussed, together with two examples of functions in artificial supramolecular assemblies based on (i) host–guest chemistry in cavities of cages, and (ii) light-harvesting in multi-chromophore arrays. Important challenges for the future are summarised.
Co-reporter:Martina Whitehead, Simon Turega, Andrew Stephenson, Christopher A. Hunter and Michael D. Ward
Chemical Science 2013 vol. 4(Issue 7) pp:2744-2751
Publication Date(Web):10 Apr 2013
DOI:10.1039/C3SC50546D
A water-soluble cubic coordination cage (Hw) has been prepared, which is isostructural with a previously reported organic-soluble cage (H) apart from the hydroxy groups on the external surface which render it water-soluble. These two cages act as hosts for small organic molecules which bind via a combination of (i) hydrogen-bonding interactions with specific sites on the internal surface of the cages; (ii) non-polar interactions such as aromatic and van der Waals interactions between aromatic rings in the guest and the cage internal surface; and (iii) solvophobic interactions. By comparing ΔG° values for guest binding in water (using Hw) and MeCN (using H), and using pairs of related guests that differ in the presence or absence of an aromatic ring substituent, it is possible to construct thermodynamic cycles that allow quantification of the solvophobic contribution to binding. Specifically, this is the difference between the solvophobic contributions to ΔG° in water and MeCN associated with desolvation of both guest and the internal surface of the cage when complexation occurs. A highly consistent value of ca. −10 kJ mol−1 is determined for this solvophobic contribution to ΔG° associated with the aromatic ring in water compared to MeCN, which correlates very well with what would be expected based on the free energy changes associated with transfer of toluene from MeCN to water. Thus, all three contributions to guest binding listed above can be separately quantified. The ability to prepare related pairs of guests with the presence or absence of a wide range of substituents provides a potentially general way to quantify the solvophobic contributions to guest binding of these substituents.
Co-reporter:Simon Turega, Martina Whitehead, Benjamin R. Hall, Anthony J. H. M. Meijer, Christopher A. Hunter, and Michael D. Ward
Inorganic Chemistry 2013 Volume 52(Issue 2) pp:1122-1132
Publication Date(Web):January 9, 2013
DOI:10.1021/ic302498t
The host–guest chemistry of the octanuclear cubic coordination cage [Co8L12]16+ (where L is a bridging ligand containing two chelating pyrazolyl-pyridine units connected to a central naphthalene-1,5-diyl spacer via methylene “hinges”) has been investigated in detail by 1H NMR spectroscopy. The cage encloses a cavity of volume of ca. 400 Å3, which is accessible through 4 Å diameter portals in the centers of the cube faces. The paramagnetism of the cage eliminates overlap of NMR signals by dispersing them over a range of ca. 200 ppm, making changes of specific signals easy to observe, and also results in large complexation-induced shifts of bound guests. The cage, in CD3CN solution, acts as a remarkably size- and shape-selective host for small organic guests such as coumarin (K = 78 M–1) and other bicyclic molecules of comparable size and shape such as isoquinoline-N-oxide (K = 2100 M–1). Binding arises from two independent recognition elements, which have been separately quantified. These are (i) a polar component arising from interaction of the H-bond accepting O atom of the guest with a convergent group of CH protons inside the cavity that lie close to a fac tris-chelate metal center and are therefore in a region of high electrostatic potential; and (ii) an additional component arising from the second aromatic ring (aromatic/van der Waals interactions with the interior surface of the cage and/or solvophobic interactions). The strength of the first component varies linearly with the H-bond-accepting ability of the guest; the second component is fixed at approximately 10 kJ mol–1. We have also used 1H–1H exchange spectroscopy (EXSY) experiments to analyze semiquantitatively two distinct dynamic processes, viz. movement of the guest into and out of the cavity and tumbling of the guest inside the host cavity. Depending on the size of the guest and the position of substituents, the rates of these processes can vary substantially, and the rates of processes that afford observable cross-peaks in EXSY spectra (e.g., between free and bound guest in some cases; between different conformers of a specific host·guest complex in others) can be narrowed down to a specific time window. Overall, the paramagnetism of the host cage has allowed an exceptionally detailed analysis of the kinetics and thermodynamics of its host–guest behavior.
Co-reporter:Daniel Sykes, Simon C. Parker, Igor V. Sazanovich, Andrew Stephenson, Julia A. Weinstein, and Michael D. Ward
Inorganic Chemistry 2013 Volume 52(Issue 18) pp:10500-10511
Publication Date(Web):September 5, 2013
DOI:10.1021/ic401410g
A series of luminescent complexes based on {Ir(phpy)2} (phpy = cyclometallating anion of 2-phenylpyridine) or {Ir(F2phpy)2} [F2phpy = cyclometallating anion of 2-(2′,4′-difluorophenyl)pyridine] units, with an additional 3-(2-pyridyl)-pyrazole (pypz) ligand, have been prepared; fluorination of the phenylpyridine ligands results in a blue-shift of the usual 3MLCT/3LC luminescence of the Ir unit from 477 to 455 nm. These complexes have pendant from the coordinated pyrazolyl ring an additional chelating 3-(2-pyridyl)-pyrazole unit, separated via a flexible chain containing a naphthalene-1,4-diyl or naphthalene-1,5-diyl spacer. Crystal structures show that the flexibility of the pendant chain allows the naphthyl group to lie close to the Ir core and participate in a π-stacking interaction with a coordinated phpy or F2phpy ligand. Luminescence spectra show that, whereas the {Ir(phpy)2(pypz)} complexes show typical Ir-based emission—albeit with lengthened lifetimes because of interaction with the stacked naphthyl group—the {Ir(F2phpy)2(pypz)} complexes are nearly quenched. This is because the higher energy of the Ir-based 3MLCT/3LC excited state can now be quenched by the adjacent naphthyl group to form a long-lived naphthyl-centered triplet (3nap) state which is detectable by transient absorption. Coordination of an {Eu(hfac)3} unit (hfac = 1,1,1,5,5,5-hexafluoro-pentane-2,4-dionate) to the pendant pypz binding site affords Ir–naphthyl–Eu triads. For the triads containing a {Ir(phpy)2} core, the unavailability of the 3nap state (not populated by the Ir-based excited state which is too low in energy) means that direct Ir→Eu energy-transfer occurs in the same way as in other flexible Ir/Eu complexes. However for the triads based on the{Ir(F2phpy)2} core, the initial Ir→3nap energy-transfer step is followed by a second, slower, 3nap→Eu energy-transfer step: transient absorption measurements clearly show the 3nap state being sensitized by the Ir center (synchronous Ir-based decay and 3nap rise-time) and then transferring its energy to the Eu center (synchronous 3nap decay and Eu-based emission rise time). Thus the 3nap state, which is energetically intermediate in the {Ir(F2phpy)2}–naphthyl–Eu systems, can act as a “stepping stone” for two-step d→f energy-transfer.
Co-reporter:Andrew Stephenson, Daniel Sykes and Michael D. Ward
Dalton Transactions 2013 vol. 42(Issue 19) pp:6756-6767
Publication Date(Web):06 Mar 2013
DOI:10.1039/C3DT50161B
The bridging ligand L14Nap, which contains two chelating pyrazolyl-pyridine units separated by a naphthalene-1,4-diyl spacer, has been used in self-assembly of polyhedral coordination cages. The largest such cage is [Cd16(L14Nap)24](BF4)32 which has a tetra-capped truncated tetrahedral Cd16 core with a bridging ligand spanning every edge. The complex is indefinitely stable in dilute solution, which makes it quite different from the previously-reported isostructural cage [Cd16(L14Ph)24](BF4)32 (based on a 1,4-phenyl bridge) that forms on crystallisation but slowly rearranges to smaller cages in solution. The additional inter-ligand π-stacking between ligand fragments associated with replacement of a phenyl group by a naphthyl group allows the complex to be stable in solution, providing conclusive proof of the importance of inter-ligand π-stacking in the assembly of these cages. With Cu(II) in place of Cd(II) a smaller cage [Cu12(L14Nap)15](ClO4)24 was formed which contains a mixture of tris-chelated (six-coordinate) and bis-chelated (four-coordinate, or five-coordinate if an additional monodentate ligand is present) Cu(II) ions; the difference between the two structures arises in part from the different stereoelectronic preferences of the two metal ions. Despite this difference both the Cd16 and Cu12 cages contain {M3(L14Nap)3}6+ triangular helical units as subcomponents which form the triangular faces of the polyhedra. By using a 1:1 ligand:metal ratio in the synthesis examples of these can be isolated and characterised; the structures of the trinuclear cyclic helicates [Cd3(L14Nap)3(BF4)4(EtOAc)2](BF4)2 and [Cu3(L14Nap)3(BF4)(MeCN)2](BF4)5 have also been determined.
Co-reporter:A. J. Metherell and M. D. Ward
RSC Advances 2013 vol. 3(Issue 34) pp:14281-14285
Publication Date(Web):08 Jul 2013
DOI:10.1039/C3RA42598C
A bis-bidentate bridging ligand H2L with inequivalent hard and soft binding sites (catecholate and pyrazolyl-pyridine, respectively) reacts with a mixture of Ti(IV) and Zn(II) ions to afford an octanuclear heterometallic Ti4Zn4 cyclic helicate formed from four dinuclear {TiZn(μ-L)2} units connected in a ring via methoxide ions.
Co-reporter:Victor F. Plyusnin, Arkady S. Kupryakov, Vyacheslav P. Grivin, Alexander H. Shelton, Igor V. Sazanovich, Anthony J. H. M. Meijer, Julia A. Weinstein and Michael D. Ward
Photochemical & Photobiological Sciences 2013 vol. 12(Issue 9) pp:1666-1679
Publication Date(Web):31 May 2013
DOI:10.1039/C3PP50109D
Transient absorption and time resolved luminescence spectroscopy were used to study photophysical processes in the macrocycle-appended 1,8-naphthalimide compound H3L, and its Eu(III) and Gd(III) complexes Eu·L and Gd·L, in particular the naphthalimide–Eu(III) energy-transfer process. In all cases aggregation of the naphthalimide chromophores results in a low-energy emission feature in the 470–500 nm region in addition to the naphthalimide fluorescence; this lower-energy emission has a lifetime longer by an order of magnitude than the monomer naphthalimide fluorescence. Transient absorption spectroscopy was used to measure the decay of the naphthalimide triplet excited state, which occurs in the range 30–50 μs. In Eu·L, partial energy-transfer from the naphthalimide chromophore results in sensitized Eu(III)-based emission in addition to the naphthalimide-based fluorescence features. Time-resolved measurements on the sensitized Eu(III)-based emission reveal both fast (∼109 s−1) and slow (∼104 s−1) energy-transfer processes from the naphthalimide energy-donor, which we ascribe to energy-transfer occurring from the singlet and triplet excited state of naphthalimide respectively. This is an unusual case of observation of sensitization of Eu(III)-based emission from the singlet state of an aromatic chromophore.
Co-reporter:Robert M. Edkins, Daniel Sykes, Andrew Beeby and Michael D. Ward
Chemical Communications 2012 vol. 48(Issue 80) pp:9977-9979
Publication Date(Web):18 May 2012
DOI:10.1039/C2CC33005A
In a pair of Ir/Eu and Ir/Tb dyads, two-photon excitation of the Ir-phenylpyridine chromophore at 780 nm is followed by partial d → f energy-transfer to give a combination of short-lived Ir-based (blue) and long-lived lanthanide-based (red or green) emission; these components can be selected separately by time-gated detection.
Co-reporter:Alexander H. Shelton, Igor V. Sazanovich, Julia A. Weinstein and Michael D. Ward
Chemical Communications 2012 vol. 48(Issue 22) pp:2749-2751
Publication Date(Web):04 Jan 2012
DOI:10.1039/C2CC17182A
A macrocycle-appended naphthalimide derivative and its Eu(III) complex show triple luminescence from isolated naphthalimide (blue), aggregated naphthalimide excimers (green) and Eu centres (red) with the balance being sensitive to the degree of aggregation, allowing white light emission to be obtained from a single molecule.
Co-reporter:James D. Ingram, Paulo J. Costa, Harry Adams, Michael D. Ward, Vítor Félix, and Jim A. Thomas
Inorganic Chemistry 2012 Volume 51(Issue 20) pp:10483-10494
Publication Date(Web):September 25, 2012
DOI:10.1021/ic200814k
The synthesis and characterization of a series of nine new complexes incorporating [RuIICl([n]aneS3)] (n = 12,14, 16) metal centers coordinated to redox active catechol ligands is reported. The solid-state structure of one of these complexes has been determined by X-ray crystallography. The redox properties of these complexes have been probed experimentally through absorption spectroscopy, cyclic voltammetry, and spectroelectrochemistry, as well as computationally through density functional theory calculations. These studies reveal that, whereas the tetrachlorocatechol-based complexes are isolated with the dioxolene unit in the catechol form, the rest of the complexes are isolated in the semiquinone oxidation state. It was also found that the RuIII/II-based couple for the complexes is dependent on the nature of the thiacrown ligand coordinated to the metal center. A combination of optical and theoretical studies revealed that the absorption spectra of the complexes contain contributions from a variety of charge transfer processes; in the case of the tetrachlorocatechol complexes these transitions include catechol-to-thiacrown ligand-to-ligand charge transfer.
Co-reporter:Ashley B. Wragg, Sofia Derossi, Timothy L. Easun, Michael W. George, Xue-Zhong Sun, František Hartl, Alexander H. Shelton, Anthony J. H. M. Meijer and Michael D. Ward
Dalton Transactions 2012 vol. 41(Issue 34) pp:10354-10371
Publication Date(Web):21 Jun 2012
DOI:10.1039/C2DT31001E
The dinuclear complex [{Ru(CN)4}2(μ-bppz)]4− shows a strongly solvent-dependent metal–metal electronic interaction which allows the mixed-valence state to be switched from class 2 to class 3 by changing solvent from water to CH2Cl2. In CH2Cl2 the separation between the successive Ru(II)/Ru(III) redox couples is 350 mV and the IVCT band (from the UV/Vis/NIR spectroelectrochemistry) is characteristic of a borderline class II/III or class III mixed valence state. In water, the redox separation is only 110 mV and the much broader IVCT transition is characteristic of a class II mixed-valence state. This is consistent with the observation that raising and lowering the energy of the d(π) orbitals in CH2Cl2 or water, respectively, will decrease or increase the energy gap to the LUMO of the bppz bridging ligand, which provides the delocalisation pathway via electron-transfer. IR spectroelectrochemistry could only be carried out successfully in CH2Cl2 and revealed class III mixed-valence behaviour on the fast IR timescale. In contrast to this, time-resolved IR spectroscopy showed that the MLCT excited state, which is formulated as RuIII(bppz˙−)RuII and can therefore be considered as a mixed-valence Ru(II)/Ru(III) complex with an intermediate bridging radical anion ligand, is localised on the IR timescale with spectroscopically distinct Ru(II) and Ru(III) termini. This is because the necessary electron-transfer via the bppz ligand is more difficult because of the additional electron on bppz˙− which raises the orbital through which electron exchange occurs in energy. DFT calculations reproduce the electronic spectra of the complex in all three Ru(II)/Ru(II), Ru(II)/Ru(III) and Ru(III)/Ru(III) calculations in both water and CH2Cl2 well as long as an explicit allowance is made for the presence of water molecules hydrogen-bonded to the cyanides in the model used. They also reproduce the excited-state IR spectra of both [Ru(CN)4(μ-bppz)]2– and [{Ru(CN)4}2(μ-bppz)]4− very well in both solvents. The reorganization of the water solvent shell indicates a possible dynamical reason for the longer life time of the triplet state in water compared to CH2Cl2.
Co-reporter:Noorshida Mohd Ali, Voirrey L. MacLeod, Petter Jennison, Igor V. Sazanovich, Christopher A. Hunter, Julia A. Weinstein and Michael D. Ward
Dalton Transactions 2012 vol. 41(Issue 8) pp:2408-2419
Publication Date(Web):03 Jan 2012
DOI:10.1039/C1DT11328C
[Ir(ppy)2(CN)2]− (ppy = anion of 2-phenylpyridine) and some substituted derivatives have been investigated for their ability to interact with additional metal cations, both in solution and the solid state, via the externally-directed cyanide lone pairs, and to act as energy-donors in the resulting assemblies. [Ir(ppy)2(CN)2]− is slightly solvatochromic, showing a blue-shift of the lowest energy absorption manifold in water compared to organic solvents, and the solubilised tBu-substituted analogue [Ir(tBuppy)2(CN)2]− [tBuppy = anion of 2-(4-tBu-phenyl)pyridine] is also metallochromic with coordination of the cyanide lone pairs to two M(II) cations in MeCN (M = Ba, Zn) resulting in blue-shifts of the lowest-energy absorption and emission maxima. These effects are however modest because of (i) the presence of only two cyanide groups, and (ii) the fact that the lowest-energy excited state has a substantial 3LC component and is therefore not purely charge-transfer in nature. Crystallisation of [Ir(ppy)2(CN)2]− as its (PPN)+ salt in the presence of excess of lanthanide(III) salts leads to formation of assemblies based on Ir–CN–Ln bonds, which generate in the solid state either Ir2Ln2(μ-CN)4 square assemblies or linear trinuclear species with Ir–CN–Ln–NC–Ir cores. In the Ir2Eu2(μ-CN)4 and Ir2Nd2(μ-CN)4 complexes the Ir-based emission is substantially quenched due to energy-transfer to lower-lying f–f states of these lanthanide ions. In addition reaction of [Ir(F2ppy)2(CN)2]− [F2ppy = cyclometallating anion of 2-(2,4-difluorophenyl)pyridine] with [Re(phen)(CO)3(MeCN)][PF6] in solution affords dinuclear IrRe and trinuclear IrRe2 species in which {Re(phen)(CO)3} units are attached to the N-donor termini of one or both of the cyanide groups; these complexes have been structurally characterised and display quantitative Ir→Re energy-transfer, showing luminescence only from the Re(I) terminus on excitation of the Ir(III) unit.
Co-reporter:Adel M. Najar, Ian S. Tidmarsh and Michael D. Ward
RSC Advances 2012 vol. 2(Issue 4) pp:1326-1328
Publication Date(Web):03 Jan 2012
DOI:10.1039/C2RA01177H
Reaction of Cu(II) salts with a combination of two different types of ligand affords an unusual ‘molecular wheel’, based on a cyclic array of four pyrazolate-bridged dinuclear Cu(II) units which are interconnected by bis-bidentate bridging ligands, in a hierarchical self-assembly process.
Co-reporter:Andrew Stephenson and Michael D. Ward
RSC Advances 2012 vol. 2(Issue 29) pp:10844-10853
Publication Date(Web):11 Sep 2012
DOI:10.1039/C2RA21757K
A series of Ag(I) complexes has been prepared containing bridging ligands based on two pyrazolyl–pyridine ligands connected by a flexible spacer. Crystallographic investigations reveal a remarkable range of structural types from simple mononuclear complexes to dinuclear double helicates, a distorted ‘bow-tie’ metallamacrocycle, and one-dimensional chains including a triple helical chain based on double helical molecular units linked by Ag⋯Ag contacts. The different structures arise as a result of the different interactions dominating in each complex, including aromatic π-stacking, Ag⋯Ag interactions, exocyclic lone pair interactions and inter-ligand π-stacking.
Co-reporter:Adel M. Najar, Ceren Avci, Michael D. Ward
Inorganic Chemistry Communications 2012 Volume 15() pp:126-129
Publication Date(Web):January 2012
DOI:10.1016/j.inoche.2011.10.007
The new ligand bis-bidentate ligand L, containing two pyrazolyl-pyridine chelating units connected to a 1,8-anthracene-diyl core via methylene spacers, reacts with Zn(II), Cd(II) and Cu(II) salts to form trigonal prismatic coordination cages [M6(μ–L)9]12+ in which a metal ion occupies each vertex and a bridging ligand spans each edge; the structure is stabilised by anions which occupy the central cavity and the gaps in the centres of the triangular faces, and also by extensive inter-ligand aromatic stacking between anthracenyl and pyrazolyl-pyridine groups.A new series of trigonal prismatic coordination cages, with a metal ion at each vertex and a bridging ligand along each edge, is described.Highlights► Self assembly of new coordination cages with a trigonal prismatic core structure. ► Cyclic triple helical M3L3 faces connected by bridging ligands. ► Fluorescent ligands containing anthracenyl bridging units.
Co-reporter:Daniel Sykes and Michael D. Ward
Chemical Communications 2011 vol. 47(Issue 8) pp:2279-2281
Publication Date(Web):15 Dec 2010
DOI:10.1039/C0CC04562D
In Ir(III)/Tb(III) dyads in which the excited state energy of the Ir(III) unit lies above 22000 cm−1, visible-light excitation of the Ir(III) chromophore results in sensitised emission from Tb(III) following Ir → Tb energy-transfer.
Co-reporter:Daniel Sykes ; Ian S. Tidmarsh ; Andrea Barbieri ; Igor V. Sazanovich ; Julia A. Weinstein ;Michael D. Ward
Inorganic Chemistry 2011 Volume 50(Issue 22) pp:11323-11339
Publication Date(Web):October 5, 2011
DOI:10.1021/ic2007759
An extensive series of blue-luminescent iridium(III) complexes has been prepared containing two phenylpyridine-type ligands and one ligand containing two pyrazolylpyridine units, of which one is bound to IrIII and the second is pendant. Attachment of {Ln(hfac)3} (Ln = Eu, Gd; hfac = anion of 1,1,1,5,5,5,-hexafluoropentanedione) to the second coordination site affords IrIII/LnIII dyads. Crystallographic analysis of several mononuclear iridium(III) complexes and one IrIII/EuIII dyad reveals that in most cases the complexes can adopt a folded conformation involving aromatic π stacking between a phenylpyridine ligand and the bis(pyrazolylpyridine) ligand, but in one series, based on CF3-substituted phenylpyridine ligands coordinated to IrIII, the steric bulk of the CF3 group prevents this and a quite different and more open conformation arises. Quantum mechanical calculations well reproduce these two types of “folded” and “open” conformations. In the IrIII/EuIII dyads, Ir → Eu energy transfer occurs with varying degrees of efficiency, resulting in partial quenching of the IrIII-based blue emission and the appearance of a sensitized red emission from EuIII. Calculations based on consideration of spectroscopic overlap integrals rule out any significant contribution from Förster (dipole–dipole) energy transfer over the distances involved but indicate that Dexter-type (exchange) energy transfer is possible if there is a small electronic coupling that would arise, in part, through π stacking between components. In some cases, an initial photoinduced electron-transfer step could also contribute to Ir → Eu energy transfer, as shown by studies on isostructural iridium/gadolinium model complexes. A balance between the blue (Ir-based) and red (Eu-based) emission components can generate white light.
Co-reporter:Benjamin R. Hall, Lauren E. Manck, Ian S. Tidmarsh, Andrew Stephenson, Brian F. Taylor, Emma J. Blaikie, Douglas A. Vander Griend and Michael D. Ward
Dalton Transactions 2011 vol. 40(Issue 45) pp:12132-12145
Publication Date(Web):10 Aug 2011
DOI:10.1039/C1DT10781J
The ligand Lbip, containing two bidentate pyrazolyl–pyridine termini separated by a 3,3′-biphenyl spacer, has been used to prepare tetrahedral cage complexes of the form [M4(Lbip)6]X8, in which a bridging ligand spans each of the six edges of the M4 tetrahedron. Several new examples have been structurally characterized with a variety of metal cation and different anions in order to examine interactions between the cationic cage and various anions. Small anions such as BF4− and NO3− can occupy the central cavity where they are anchored by an array of CH⋯F or CH⋯O hydrogen-bonding interactions with the interior surface of the cage, but larger anions such as naphthyl-1-sulfonate or tetraphenylborate lie outside the cavity and interact with the external surface of the cage via CH⋯π interactions or CH⋯O hydrogen bonds. The cages with M = Co and M = Cd have been examined in detail by NMR spectroscopy. For [Co4(Lbip)6](BF4)8 the 1H NMR spectrum is paramagnetically shifted over the range −85 to +110 ppm, but the spectrum has been completely assigned by correlation of measured T1 relaxation times of each peak with Co⋯H distances. 19F DOSY measurements on the anions show that at low temperature a [BF4] − anion diffuses at a similar rate to the cage superstructure surrounding it, indicating that it is trapped inside the central cage cavity. Furthermore, the equilibrium step-by-step self-assembly of the cage superstructure has been elucidated by detailed modeling of spectroscopic titrations at multiple temperatures of an acetonitrile solution of Lbip into an acetonitrile solution of Co(BF4)2. Six species have been identified: [Co2Lbip]4+, [Co2(Lbip)2]4+, [Co4(Lbip)6]8+, [Co4(Lbip)8]8+, [Co2(Lbip)5]4+, and [Co(Lbip)3]2+. Overall the assembly of the cage is entropy, and not enthalpy, driven. Once assembled, the cages show remarkable kinetic inertness due to their mechanically entangled nature: scrambling of metal cations between the sites of pure Co4 and Cd4 cages to give a statistical mixture of Co4, Co3Cd, Co2Cd2, CoCd3 and Cd4 cages takes months in solution at room temperature.
Co-reporter:Andrew Stephenson and Michael D. Ward
Dalton Transactions 2011 vol. 40(Issue 31) pp:7824-7826
Publication Date(Web):14 Feb 2011
DOI:10.1039/C0DT01767A
The octanuclear coordination cage [Ni8(L14Naph)12](BF4)16 has the core structure of a ‘cuneane’ - a toplogical isomer of a cube - with a metal ion at each of the eight vertices and bridging ligand spanning each of the twelve edges; this is the only possible 8-vertex polyhedron other than a cube that will form a cage in which each metal is connected to three others.
Co-reporter:Andrew Stephenson and Michael D. Ward
Dalton Transactions 2011 vol. 40(Issue 40) pp:10360-10369
Publication Date(Web):26 Apr 2011
DOI:10.1039/C1DT10263J
The two new ligands Lfur and Lth consist of two chelating pyrazolyl-pyridine termini connected to furan-2,5-diyl or thiophene-2,5-diyl spacers viamethylene groups. Reaction of these with a range of transition metal dications that prefer octahedral coordination affords a series of unusual structures which are all based on a 2M:3L ratio. [M8(Lfur)12]X16 (M = Co, Cu, X = BF4; and M = Zn, X = ClO4) are octanuclear cubes with approximate D4 symmetry in which two cyclic tetranuclear helicate M4L4 units are connected by four additional ‘pillar’ ligands. In contrast [Ni4(Lfur)6](BF4)8 is a centrosymmetric molecular square consisting of two dinuclear Ni2L2 units of opposite chirality that are connected by a pair of additional Lfur ligands such that the four edges of the Ni4 square are spanned by alternately two and one bridging ligands. [M4(Lth)6](BF4)8 (M = Co, Ni, Cu) are likewise molecular squares with similar structures to [Ni4(Lfur)6](BF4)8 with the significant difference that the two crosslinked double helicate M2L2 units are now homochiral. The Cd(II) complexes both behave quite differently to the first-row metal complexes, with [Cd(Lfur)(BF4)](BF4) being a simple mononuclear complex with a single ligand in which the furan oxygen atom is weakly interacting with the Cd(II) centre. In contrast, in {[Cd2(Lth)3](BF4)4}∞, where this quasi-pentadentate coordination mode of the ligand is not possible because thiophene is too poor an electron donor, the ligand reverts to bis-bidentate bridging coordination to afford a one-dimensional chain consisting of an infinite sequence of crosslinked, homochiral, Cd2(Lth)2 double helicate units.
Co-reporter:Hazel Fenton, Ian S. Tidmarsh and Michael D. Ward
CrystEngComm 2011 vol. 13(Issue 5) pp:1432-1440
Publication Date(Web):16 Nov 2010
DOI:10.1039/C0CE00690D
A series of ligands containing two 3,5-dimethylpyrazole units connected to a central fluorescent naphthyl spacer have been used to prepare Ag(I) networks which have been structurally characterised. The ability of Ag(I) ions to tolerate a wide range of coordination environments has resulted in the formation of three different structural types: linear one-dimensional chains, two-dimensional sheets and three-dimensional networks. Upon complexation the ligands arrange so as to maximise inter-ligand aromatic π-stacking interactions, with the consequence that many of the resulting compounds display in the solid-state low-energy exciplex emission features extending well beyond the normal naphthalene emission range.
Co-reporter:Michael D. Ward
Coordination Chemistry Reviews 2010 Volume 254(21–22) pp:2634-2642
Publication Date(Web):November 2010
DOI:10.1016/j.ccr.2009.12.001
Four sets of dyads are discussed, in all of which near-infrared emitting lanthanide(III) ions such as Nd(III), Er(III) or Yb(III) are energy-acceptors which provide sensitized luminescence following energy-transfer from an antenna group. In three sets of dyads the antenna (energy-donor) group is a luminescent transition metal fragment; in the fourth the antenna is an anthracene group. A combination of photophysical studies and calculations has been used to understand the mechanisms by which energy-transfer to the lanthanide(III) ion occurs. Although definitive answers are not possible in every case due to the presence of several possible energy-transfer pathways, the relative contributions of Förster-type, Dexter-type and redox-mediated energy-transfer pathways have been analysed. Interesting results include (i) the demonstration of pure Dexter energy-transfer over 20 Å in a Ru(II)/Nd(III) dyad, and (ii) the demonstration of a redox-based mechanism for energy-transfer in anthracene/Ln(III) dyads in which the first step is photoinduced electron-transfer from the excited anthracene chromophore to a diimine ligand on the lanthanide(III) to generate a charge-separated state.
Co-reporter:Jonathan Best ; Igor V. Sazanovich ; Harry Adams ; Robert D. Bennett ; E. Stephen Davies ; Anthony J. H. M. Meijer ; Michael Towrie ; Sergei A. Tikhomirov ; Oleg V. Bouganov ; Michael D. Ward ;Julia A. Weinstein
Inorganic Chemistry 2010 Volume 49(Issue 21) pp:10041-10056
Publication Date(Web):September 28, 2010
DOI:10.1021/ic101344t
A series of mononuclear complexes of the type [Pt(Bu2cat)(4,4′-R2-bipy)] [where Bu2cat is the dianion of 3,5-tBu2-catechol and R = H, tBu, or C(O)NEt2] and analogous dinuclear complexes based on the “back-to-back” bis-catechol ligand 3,3′,4,4′-tetrahydroxybiphenyl have been studied in detail in both their ground and excited states by a range of physical methods including electrochemistry, UV/vis/near-IR, IR, and electron paramagnetic resonance spectroelectrochemistry, and time-resolved IR (TRIR) and transient absorption (TA) spectroscopy. Density functional theory calculations have been performed to support these studies, which provide a detailed picture of the ground- and excited-state electronic structures, and excited-state dynamics, of these complexes. Notable observations include the following: (i) for the first time, the lowest-energy catecholate → bipyridine (bpy) ligand-to-ligand charge-transfer (LL′CT) excited states of these chromophores have been studied by TRIR spectroscopy, showing a range of transient bands associated with the bpy radical anion and semiquinone species, and back-electron-transfer occurring in hundreds of picoseconds; (ii) strong electronic coupling between the two catecholate units in the bridging ligand of the dinuclear complexes results in a delocalized, planar (class 3) “mixed-valence” catecholate2−/semiquinone•− state formed by one-electron oxidation of the bridging ligand; (iii) in the LL′CT excited state of the dinuclear complexes, the bridging ligand is symmetrical and delocalized, whereas the bpy radical anion is localized at one terminus of the complex. This study is the first example of an investigation of excited-state behavior in platinum(II) catecholate complexes, performed with the use of picosecond TRIR and femtosecond TA spectroscopy.
Co-reporter:Christopher S. Grange, Anthony J. H. M. Meijer and Michael D. Ward
Dalton Transactions 2010 vol. 39(Issue 1) pp:200-211
Publication Date(Web):20 Nov 2009
DOI:10.1039/B918086A
The trinuclear complexes [{(R2bipy)2Ru}3(µ3-HHTP)](PF6)3 [1(PF6)3, R = H; 2(PF6)3, R = 4-tBu] contain three {Ru(R2bipy)2}2+ fragments connected to the triangular tris-chelating ligand hexahydroxytriphenylene (H6HHTP). This bridging ligand contains three dioxolene-type binding sites, each of which can reversibly convert between dianionic catecholate (cat), monoanionic semiquinone (sq) or neutral quinone (q) redox states. The bridging ligand as a whole can therefore exist in seven different redox states from fully reduced [cat,cat,cat]6− through to fully oxidised, neutral [q,q,q]. Cyclic voltammetry of 1(PF6)3 in MeCN reveals six redox processes of which the three at more positive potentials (the sq/q couples) are reversible but the three at more negative potentials (the sq/cat couples) are irreversible with distorted wave shapes due to the insolubility of the reduced forms of the complex. In contrast, the more soluble complex 2(PF6)3 displays six reversible one-electron redox processes making all components of a seven-membered redox chain accessible. UV/Vis/NIR spectro-electrochemical studies reveal rich spectroscopic behaviour, with—in particular—very intense transitions in the near-IR region in many of the oxidation states associated with Ru(II)→(dioxolene) MLCT and bridging ligand centred π–π* transitions. TDDFT calculations were used to analyse the electronic spectra in all seven oxidation states; the calculated spectra generally show very good agreement with experiment, which has allowed a fairly complete assignment of the low-energy transitions. The strong electrochromism of the complexes in the near-IR region has formed the basis of an ‘optical window’ in which a thin film of 1(PF6)3 or 2(PF6)3 on a conductive glass surface can be reversibly and rapidly switched between redox states that alternate between strongly absorbing or near-transparent at 1100 nm, with—for 2(PF6)3—the switching being stable and reversible in water over thousands of cycles.
Co-reporter:Hazel Fenton, Ian S. Tidmarsh and Michael D. Ward
Dalton Transactions 2010 vol. 39(Issue 16) pp:3805-3815
Publication Date(Web):05 Mar 2010
DOI:10.1039/B926264D
Two ligands L1 and L2 have been prepared which contain a chelating pyrazolyl-pyridine group with a pendant aromatic nitrile (in L1, a benzonitrile; in L2, a naphthonitrile). These ligands react with Ag(I) salts to give a range of infinite coordination networks or dimeric ‘boxes’ in which the pyrazolyl-pyridine chelates and the aromatic nitrile groups both participate in coordination to Ag(I) ions. In contrast, L1 and L2 form simple mononuclear tris-chelates [ML3]2+ with first-row transition metal dications (M = Co, Ni, Zn) in which the aromatic nitrile groups are pendant such that the complexes can be used as ‘complex ligands’. The crystal structures of [M(L2)3](BF4)2 are based on solely the mer tris-chelate geometry although in solution 1H NMR spectroscopy reveals a mixture of both fac and mer isomers of the tris-chelates. Reaction of these with Ag(I) ions allows the interaction of the pendant nitrile groups with Ag(I) ions to generate coordination networks based on [ML3]2+ cations being crosslinked by Ag(I) ions. In these networks the [ML3]2+ cations have solely the fac geometry and lie on threefold rotation axes with all three pendant nitrile groups coordinated to Ag(I) ions which are three-coordinate. {[AgM(L2)3][BF4]3}∞ (M = Co, Ni) consist of two interpenetrated (10,3)a nets which have opposite chirality at the [M(L2)3]2+ centres but are not strictly enantiomorphic as the two nets are not crystallographically equivalent. {[AgNi(L1)3](BF4)3}∞ in contrast contains two-dimensional sheets which have a (6,3) net structure of hexagonal rings of alternative Ni(II) and Ag(I) centres; although not interpenetrating, two such adjacent (and enantiomorphic) sheets interact with each other via numerous CH⋯π interactions between aromatic ligands. Formation of these structures shows that the differential reactivity of the two binding sites in L1 and L2 (pyrazolyl-pyridine, and nitrile) can be used to generate mixed-metal coordination networks in a hierarchical, stepwise manner.
Co-reporter:Adel M. Najar, Ian S. Tidmarsh and Michael D. Ward
CrystEngComm 2010 vol. 12(Issue 11) pp:3642-3650
Publication Date(Web):06 Jul 2010
DOI:10.1039/C0CE00176G
A series of six compartmental ligands, containing two or three bidentate chelating pyridyl-triazole or pyridyl-pyrazole units connected to a central aromatic spacer, has been used to prepare Pb(II) complexes which have been structurally characterised. Five of the complexes form infinite coordination networks including a one-dimensional chain and four two-dimensional sheets comprising three examples of (6,3) nets, and one example of a (4,4) net. One of the complexes is a discrete dinuclear dimeric complex in which two bridging ligands span two metal ions in a ‘mesocate’ box-like arrangement. The Pb(II) centres show a range of coordination numbers and geometries, from distorted six-coordinate with a stereochemically active lone pair of electrons, to fairly regular eight-coordinate with a square antiprismatic geometry.
Co-reporter:Michael D. Ward
Dalton Transactions 2010 vol. 39(Issue 38) pp:8851-8867
Publication Date(Web):08 Jul 2010
DOI:10.1039/C0DT00312C
Luminescent complexes of the [M(diimine)(CN)4]2− family (M = Ru, Os), and their polynuclear analogues, are structurally versatile components for preparation of supramolecular assemblies based on interaction of the cyanide groups with other metal ions or metal complexes via direct coordination, hydrogen bonding, or halogen bonding. In addition their environment-dependent photophysical properties (solvatochromism and metallochromism), and the ability of the CN groups to act as reporters for excited state behaviour via time-resolved IR spectroscopy, make these fragments spectroscopically as well as structurally versatile. This Perspective article summarises work from the author's group over the last decade on the structures and photophysical properties of these fascinating complexes and their supramolecular assemblies.
Co-reporter:Voirrey L. Robinson, Christopher A. Hunter, Michael D. Ward
Inorganica Chimica Acta 2010 Volume 363(Issue 12) pp:2938-2944
Publication Date(Web):15 October 2010
DOI:10.1016/j.ica.2010.03.021
The previously reported complex [Ru(ttpy)(CN)3]− [ttpy = 4′(p-tolyl)-2,2′:6′,2″-terpyridine] is conveniently synthesised by reaction of ttpy with Ru(dmso)4Cl2 to give [Ru(ttpy)(dmso)Cl2], which reacts in turn with KCN in aqueous ethanol to afford [Ru(ttpy)(CN)3]− which was isolated and crystallographically characterised as both its (PPN)+ and K+ salts. The K+ salt contains clusters containing three complex anions and three K+ cations connected by end-on and side-on cyanide ligation to the K+ ions. The solution speciation behaviour of [Ru(ttpy)(CN)3]− was investigated with both Zn2+ and K+ salts in MeCN, a solvent sufficiently non-competitive to allow the added metal cations to associate with the complex anion via the externally-directed cyanide lone pairs. UV–Vis spectroscopic titration of (PPN)[Ru(ttpy)(CN)3] with Zn(ClO4)2 showed a blue shift of 2900 cm−1 in the 1MLCT absorption manifold due to the ‘metallochromism’ effect; a series of distinct binding events could be discerned corresponding to formation of 4:1, 1:1 and then 1:3 anion:cation adducts, all with high formation constants, as the titration proceeded. In contrast titration of (PPN)[Ru(ttpy)(CN)3] with the more weakly Lewis-acidic KPF6 resulted in a much smaller blue-shift of the 1MLCT absorptions, and the titration data corresponded to formation of 1:1 and then 2:1 cation:anion adducts with weaker stepwise association constants of the order of 104 and then 103 M−1. Although association of [Ru(ttpy)(CN)3]− resulted in a blue-shift of the 1MLCT absorptions, the luminescence was steadily quenched, as raising the 3MLCT level makes radiationless decay via a low-lying 3MC state possible.The complex anion [Ru(ttpy)(CN)3]−, for which crystal structures of its K+ and (PPN)+ salts are presented, aggregates with Zn2+ and K+ ions in MeCN via the externally-directed lone pairs on the cyanide groups to form adducts with a range of stiochiometries; this association results in a blue-shift of the 1MLCT absorption maxima but also causes quenching of the luminescence.
Co-reporter:Sofia Derossi, Michael D. Ward
Inorganic Chemistry Communications 2010 Volume 13(Issue 6) pp:741-744
Publication Date(Web):June 2010
DOI:10.1016/j.inoche.2010.03.036
Reaction of [Ru(bpym)3]2+ (bpym = 2,2′-bipyridmidine) with hexacyanoruthenate under forcing conditions affords a mixture of the trinuclear species [(bpym)Ru{(µ-bpym)Ru(CN)4}2]2−, [1]2−, and the tetranuclear species [Ru{(µ-bpym)Ru(CN)4}3]4−, [2]4−, in which two or three (respectively) of the peripheral vacant bpym binding sites of [Ru(bpym)3]2+ are occupied by {Ru(CN)4}2− fragments. Thus, [1]2− and [2]4− have eight and twelve externally-directed cyanide groups respectively for use in forming high connectivity coordination networks. The crystal structure of HK[1]·2MeOH·6.5H2O reveals a one-dimensional ladder structure in which [1]2− anions are connected by (i) cyanide/K+ and (ii) bpym/K+ coordination interactions.Reaction of [Ru(bpym)3]2+ (bpym = 2,2′-bipyrimidine) with K4Ru(CN)6 results in the attachment of {Ru(CN)4}2− fragments at two or three of the peripheral vacant diimine sites, giving trinuclear [1]2− (with eight externally-directed cyanide ligands) and tetranuclear [2]4− (with 12 externally-directed cyanides in a D3 symmetric array). These are interesting building blocks for crystal engineering applications providing potentially very high connectivity at a single node in cyanide-bridged networks.
Co-reporter:Nicholas M. Tart, Daniel Sykes, Igor Sazanovich, Ian S. Tidmarsh and Michael D. Ward
Photochemical & Photobiological Sciences 2010 vol. 9(Issue 7) pp:886-889
Publication Date(Web):29 Mar 2010
DOI:10.1039/C0PP00011F
Luminescent iridium(III) complex units bearing pendant 2,2′-bipyridyl-type binding sites can be used to generate Ir/Ln dyads in which the Ir(III) luminophore acts as an energy donor to the lanthanide by the Dexter mechanism, generating sensitised emission in the visible (from Eu) or near-infrared (Nd, Yb) regions.
Co-reporter:Michael D. Ward
Chemical Communications 2009 (Issue 30) pp:4487-4499
Publication Date(Web):10 Jun 2009
DOI:10.1039/B906726B
Reaction of simple bis-bidentate ligands, containing two chelating pyrazolyl-pyridine units connected to a central aromatic spacer, with six-coordinate transition metal dications results in self-assembly of an extensive series of polyhedral cage complexes. These include M4L6 tetrahedra, M8L12 cubes, M12L18 truncated tetrahedra and M16L24 tetra-capped truncated tetrahedra. In all cases the metal : ligand ratio is 2 : 3, reflecting the combination of six-coordinate metal ions with tetradentate ligands. The resulting structures are based on those polyhedra which have a 2 : 3 ratio of vertices to faces, with a metal ion at each vertex and bridging ligand spanning each edge. The cages display a range of interesting properties such as an anion-based template effect in the smaller examples; host–guest chemistry associated with the central cavity; aromatic stacking around the periphery between electron-poor and electron-rich ligand fragments which appears to contribute substantially to their stability; and modified fluorescence properties arising from the aromatic stacking of fluorophores such as naphthyl and anthracenyl groups built into the ligand backbone. Even more complex structural types are available using a mixture of face-capping (tris-bidentate) and edge-bridging (bis-bidentate) ligands, such as examples of M12 cuboctahedra which select a combination of two types of ligand during the self-assembly process.
Co-reporter:Sofia Derossi ; Lee Brammer ; Christopher A. Hunter ;Michael D. Ward
Inorganic Chemistry 2009 Volume 48(Issue 4) pp:1666-1677
Publication Date(Web):January 9, 2009
DOI:10.1021/ic8021529
The interactions between the [Ru(bipy)(CN)4]2− anion and N-methyl-halopyridinium cations have been examined in both the solid state and in solution. In the solid state, crystal structures of [Ru(bipy)(CN)4]2− salts containing iodinated cations (N-methyl-3-iodopyridinium and N-methyl-3,5-diiodopyridinium) show clear C−I···NC(Ru) halogen bonds between the externally directed cyanide lone pairs of the anion and the iodine atoms of the cation which dominates the structures. In contrast the analogous brominated cations (N-methyl-3-bromopyridinium and N-methyl-3,5-dibromopyridinium) do not exhibit C−Br···NC(Ru) interactions in the solid state, with the cyanide groups instead involved in hydrogen bonding, principally to lattice water molecules. The charge-assisted C−I···NC(Ru) interactions are therefore clearly of value as synthons in crystal engineering applications. In CH2Cl2 solution, spectroscopic titrations between [Ru(4,4′-tBu2-bipy)(CN)4]2− and both N-methyl-3-iodopyridinium and N-methyl-3-bromopyridinium cations show clear evidence for formation of distinct 1:1, 3:2, and then 2:1 cation/anion adducts with high association constants (>107 M−1 for the first 1:1 association constant). However the presence of identical results using the non-halogenated cation N-methyl-pyridinium indicates that this strong cation/anion association in CH2Cl2 is dominated by electrostatic effects: either C−H···NC(Ru) hydrogen bonds or C−X···NC(Ru) halogen bonds could be involved in the ion pairs but it is the charge-assistance that makes the association strong. This is confirmed by a titration between [Ru(4,4′-tBu2-bipy)(CN)4]2− and the neutral halogen-bond acceptor C6F5I for which the first association constant is very low (ca. 6 M−1). The formation of adducts between [Ru(4,4′-tBu2-bipy)(CN)4]2− and the various N-methyl-pyridinium cations in solution results in a clear blue-shift of the 1MLCT absorption maxima associated with the Ru(II) unit, a characteristic consequence of interaction of the cyanide lone pairs with a Lewis-acidic site on the cation. The 3MLCT luminescence from the [Ru(4,4′-tBu2-bipy)(CN)4]2− center, however, does not show the usual associated increase in intensity associated with this blue shift in the 1MLCT absorptions, most likely because of electron-transfer quenching by the N-methyl-pyridinium cations in the assemblies.
Co-reporter:Timothy L. Easun, Wassim Z. Alsindi, Nina Deppermann, Michael Towrie, Kate L. Ronayne, Xue-Zhong Sun, Michael D. Ward and Michael W. George
Inorganic Chemistry 2009 Volume 48(Issue 18) pp:8759-8770
Publication Date(Web):August 17, 2009
DOI:10.1021/ic900924w
The exploitation of the dramatic negative solvatochromism of the [Ru(bipy)(CN)4] moiety (bipy = 2,2′-bipyridine) allows a change in solvent to reverse the direction of photoinduced energy transfer (PEnT) in two related dinuclear complexes. Both dyads consist of a [Ru(bpyam)2Ln]2+ (Ru-bpyam) unit (bpyam = 4,4′-diethylamido-2,2′-bipyridine; Ln = bis-bipyridyl-based bridging ligand) and a [Ru(Ln)(CN)4]2− (Ru-CN) unit. Both termini have IR-active spectroscopic handles (amide carbonyl or cyanide, respectively) allowing the excited-state dynamics to be studied by time-resolved IR (TRIR) spectroscopy. One dyad (1) contains a relatively rigid exoditopic macrocyclic bis-bipyridyl bridging ligand (L1) and the other (2) contains a more flexible bis-bipyridyl bridging ligand with only one covalent linkage between the two bipyridyl binding sites (L2). The conformational effects on PEnT rates in these dyads are probed using a combination of luminescence and TRIR studies. In both 1 and 2 in D2O it is demonstrated that Ru-CN → Ru-bpyam PEnT occurs (PEnT time scales were in the range 10 ps−3 ns) because the 3MLCT energy of the Ru-CN terminus is higher than that of the Ru-bpyam terminus. Changing the solvent from D2O to CH3CN results in lowering the 3MLCT energy of the Ru-CN unit below that of the Ru-bpyam unit such that in both dyads a reversal in the direction of PEnT to Ru-bpyam → Ru-CN (time scales of 10 ps−2 ns) occurs. Complex kinetic behavior results from the presence of a dark 3MLCT excited state formulated as {(bpyam)2Ru3+(Ln●−)} and by the presence of multiple conformers in solution which have different Ru···Ru separations giving rise to different energy transfer rates.
Co-reporter:Adel M. Najar, Ian S. Tidmarsh, Harry Adams and Michael D. Ward
Inorganic Chemistry 2009 Volume 48(Issue 24) pp:11871-11881
Publication Date(Web):November 19, 2009
DOI:10.1021/ic901892y
Reaction of two structurally related bridging ligands L26Py and L13Ph, in which two bidentate chelating pyrazolyl-pyridine units are connected to either a 2,6-pyridine-diyl or 1,3-benzene-diyl central group via methylene spacers, with first-row transition metal dications, results in a surprising variety of structures. The commonest is that of an octanuclear coordination cage [M8L12]X16 [M = Co(II) or Zn(II); X = perchlorate or tetrafluoroborate] in which a metal ion is located at each of the eight vertices of an approximate cube, and one bis-bidentate bridging ligand spans each edge. The arrangement of fac and mer tris-chelate metal centers around the inversion center results in approximate (non-crystallographic) S6 symmetry. Another structural type observed in the solid state is a hexanuclear complex [Co6(L13Ph)9](ClO4)12 in which the six metal ions are in a rectangular array (two rows of three), folded about the central Co−Co vector like a partially open book, with each metal−metal edge containing one bridging ligand apart from the two outermost metal−metal edges which are spanned by a pair of bridging ligands in a double helical array. The final structural type we observed is a tetranuclear square [Ni4(L26Py)6](BF4)8, with the four Ni−Ni edges spanned alternately by one and two bridging ligand such that it effectively consists of two dinuclear double helicates cross-linked by additional bridging ligands. A balance between the “cube” and “book” forms, which varied from compound to compound, was observed in solution in many cases by 1H NMR and ES mass spectrometry studies.
Co-reporter:Sofia Derossi, Svetlana G. Baca, Thomas A. Miller, Harry Adams, John C. Jeffery, Michael D. Ward
Inorganica Chimica Acta 2009 Volume 362(Issue 4) pp:1282-1288
Publication Date(Web):2 March 2009
DOI:10.1016/j.ica.2008.06.017
Crystallisation of simple cyanoruthenate complex anions [Ru(NN)(CN)4]2− (NN = 2,2′-bipyridine or 1,10-phenanthroline) in the presence of Lewis-acidic cations such as Ln(III) or guanidinium cations results, in addition to the expected [Ru(NN)(CN)4]2− salts, in the formation of small amounts of salts of the dinuclear species [Ru2(NN)2(CN)7]3−. These cyanide-bridged anions have arisen from the combination of two monomer units [Ru(NN)(CN)4]2− following the loss of one cyanide, presumably as HCN. The crystal structures of [Nd(H2O)5.5][Ru2(bipy)2(CN)7] · 11H2O and [Pr(H2O)6][Ru2(phen)2(CN)7] · 9H2O show that the cyanoruthenate anions form Ru–CN–Ln bridges to the Ln(III) cations, resulting in infinite coordination polymers consisting of fused Ru2Ln2(μ-CN)4 squares and Ru4Ln2(μ-CN)6 hexagons, which alternate to form a one-dimensional chain. In [CH6N3]3[Ru2(bipy)2(CN)7] · 2H2O in contrast the discrete complex anions are involved in an extensive network of hydrogen-bonding involving terminal cyanide ligands, water molecules, and guanidinium cations. In the [Ru2(NN)2(CN)7]3− anions themselves the two NN ligands are approximately eclipsed, lying on the same side of the central Ru–CN–Ru axis, such that their peripheries are in close contact. Consequently, when NN = 4,4′-tBu2-2,2′-bipyridine the steric bulk of the t-butyl groups prevents the formation of the dinuclear anions, and the only product is the simple salt of the monomer, [CH6N3]2[Ru(tBu2bipy)(CN)4] · 2H2O. We demonstrated by electrospray mass spectrometry that the dinuclear by-product [Ru2(phen)2(CN)7]3− could be formed in significant amounts during the synthesis of monomeric [Ru(phen)(CN)4]2− if the reaction time was too long or the medium too acidic. In the solid state the luminescence properties of [Ru2(bipy)2(CN)7]3− (as its guanidinium salt) are comparable to those of monomeric [Ru(bipy)(CN)4]2−, with a 3MLCT emission at 581 nm.During crystallisation of [Ru(NN)(CN)4]2− (NN = 2,2′-bipyridine or 1,10-phenanthroline) with Lewis-acidic cations [Ln(III), guanidinium] the cyanide-bridged dimers [Ru2(NN)2(CN)7]3− form, which generate one-dimensional coordination polymers with Ln(III) cations or a hydrogen-bonded network with guanidinium cations.
Co-reporter:Nawal K. Al-Rasbi, Sofia Derossi, Daniel Sykes, Stephen Faulkner, Michael D. Ward
Polyhedron 2009 28(2) pp: 227-232
Publication Date(Web):
DOI:10.1016/j.poly.2008.10.046
Co-reporter:Juan Manuel Herrera, Enrique Colacio, Corine Mathonière, Duane Choquesillo-Lazarte and Michael D. Ward
Chemical Communications 2008 (Issue 37) pp:4460-4462
Publication Date(Web):25 Jul 2008
DOI:10.1039/B807364C
Reaction of the anionic cyanometallate chromophore [{Ru(CN)4}3(μ3-HAT)]6− with [MII(tren)]2+ complexes (M = ZnII, CuII) provides discrete tetradecanuclear clusters of formula [{MII(tren)(μ-CN)}11{Ru3(HAT)(CN)}]16+; the weak luminescence of the Ru3 chromophore is substantially enhanced in the presence of ZnII ions, whereas it is completely quenched when CuII centers are present.
Co-reporter:Svetlana G. Baca ; Simon J. A. Pope ; Harry Adams ;Michael D. Ward
Inorganic Chemistry 2008 Volume 47(Issue 9) pp:3736-3747
Publication Date(Web):March 29, 2008
DOI:10.1021/ic702353c
Slow evaporation of aqueous solutions containing mixtures of Na2[Os(phen)(CN)4], Ln(III) salts (Ln = Pr, Nd, Gd, Er, Yb), and (in some cases) an additional ligand such as 1,10-phenanthroline (phen) or 2,2′-bipyrimidine (bpym) afforded crystalline coordination networks in which the [Os(phen)(CN)4]2− anions are coordinated to Ln(III) cations via Os−CN−Ln cyanide bridges. The additional diimine ligands, if present, also coordinate to the Ln(III) centers. Several types of structure have been identified by X-ray crystallographic studies. Photophysical studies showed that the characteristic emission of the [Os(phen)(CN)4]2− chromophore, which occurs at ∼680 nm in this type of coordination environment with a triplet metal-to-ligand charge transfer (3MLCT) energy content of ∼16 000 cm−1, is quenched by energy transfer to those Ln(III) centers (Pr, Nd, Er, Yb) that have low-lying f−f states capable of accepting energy from the Os(II)-based 3MLCT state. Time-resolved studies on the residual (partially quenched) Os(II)-based luminescence allowed the rates of Os(II) → Ln(III) energy transfer to be evaluated. The measured rates varied substantially, having values of >5 × 108, ∼1 × 108, and 2.5 × 107 s−1 for Ln = Nd, Er or Yb, and Pr, respectively. These differing rates of Os(II) → Ln(III) energy transfer can be rationalized on the basis of the availability of f−f states of the different Ln(III) centers that are capable of acting as energy acceptors. In general, the rates of Os(II) → Ln(III) energy transfer are an order of magnitude faster than the rates of Ru(II) → Ln(III) energy transfer in a previously described series of [Ru(bipy)(CN)4]2−/Ln(III) networks. This is ascribed principally to the lower energy of the Os(II)-based 3MLCT state, which provides better spectroscopic overlap with the low-lying f−f states of the Ln(III) ions.
Co-reporter:Timothy L. Easun, Wassim Z. Alsindi, Michael Towrie, Kate L. Ronayne, Xue-Zhong Sun, Michael D. Ward and Michael W. George
Inorganic Chemistry 2008 Volume 47(Issue 12) pp:5071-5078
Publication Date(Web):May 14, 2008
DOI:10.1021/ic702005w
The dyad RuLRe contains {Re(bpy)(CO)3Cl} and {Ru(bpy)(bpyam)2}2+ termini (bpy = 2,2′-bipyridine; bpyam = 4,4′-diethylamido-2,2′-bipyridine) separated by a flexible ethylene spacer. Luminescence studies reveal the expected Re → Ru photoinduced energy transfer, with partial quenching of ReI-based triplet metal-to-ligand charge-transfer (3MLCT) luminescence and consequent sensitization of the RuII-based 3MLCT luminescence, which has a component with a grow-in lifetime of 0.76 (±0.2) ns. The presence of IR-active spectroscopic handles on both termini [CO ligands directly attached to ReI and amide carbonyl substituents on the bpy ligands coordinated to RuII] allowed the excited-state dynamics to be studied by time-resolved IR (TRIR) spectroscopy in much more detail than allowed by luminescence methods. A combination of picosecond- and nanosecond-time-scale TRIR studies revealed the presence of at least three distinct Re → Ru energy-transfer processes, with lifetimes of ca. 20 ps and 1 and 13 ns. This complex behavior occurs because of a combination of two different Ru-based 3MLCT states (Ru → L and Ru → bpyam), which are sensitized by energy transfer from the ReI donor at different rates; and the presence of at least two conformers of the flexible molecule RuLRe, which have different Re···Ru separations.
Co-reporter:Theodore Lazarides, Harry Adams, Daniel Sykes, Stephen Faulkner, Giuseppe Calogero and Michael D. Ward
Dalton Transactions 2008 (Issue 5) pp:691-698
Publication Date(Web):23 Nov 2007
DOI:10.1039/B714640J
The complexes [Ru(tBu2bipy)(bpym)X2] (X = Cl, NCS) and [M(tBu2bipy)2(bpym)][PF6]2 (M = Ru, Os) all have a low-energy LUMO arising from the presence of a 2,2′-bipyrimidine ligand, and consequently have lower-energy 1MLCT and 3MLCT states than analogous complexes of bipyridine. The vacant site of the bpym ligand provides a site at which {Ln(diketonate)3} units can bind to afford bipyrimidine-bridged dinuclear Ru–Ln and Os–Ln dyads; four such complexes have been structurally characterised. UV/Vis and luminescence spectroscopic studies show that binding of the Ln(III) fragment at the second site of the bpym ligand reduces the 3MLCT energy of the Ru or Os fragment still further. The result is that in the dyads [Ru(tBu2bipy)X2(µ-bpym)Ln(diketonate)3] (X = Cl, NCS) and [Os(tBu2bipy)2(µ-bpym)Ln(diketonate)3][PF6]2 the 3MLCT is too low to sensitise the luminescent f-f states of Nd(III) and Yb(III), but in [Ru(tBu2bipy)2(µ-bpym)Ln(diketonate)3][PF6]2 the 3MLCT energy of 13500 cm−1 permits energy transfer to Yb(III) and Nd(III) resulting in sensitised near-infrared luminescence on the microsecond timescale.
Co-reporter:Qiao-Hua Wei, Stephen P. Argent, Harry Adams and Michael D. Ward
New Journal of Chemistry 2008 vol. 32(Issue 1) pp:73-82
Publication Date(Web):10 Sep 2007
DOI:10.1039/B708572A
The simple mononuclear complex [Ru(H2bpp)2][PF6]2 [H2bpp = 2,6-bis(pyrazol-3-yl)pyridine] contains four coordinated pyrazolyl ligands which each have a reactive NH site at the position adjacent to the coordinated N atom. Alkylation of these with either 2-[1-{4-(bromomethyl)benzyl}-1H-pyrazol-3-yl]pyridine or 4′-[(4-bromomethyl)phenyl]terpyridine allows attachment of four additional chelating groups, either bidentate pyrazolyl–pyridine and terdentate terpyridyl units, respectively, which are pendant from the central kinetically inert RuIIN6 complex core. These functionalised mononuclear complexes [Ru(L1)2][PF6]2 (with four pendant pyrazolyl–pyridine bidentate sites) and [Ru(L2)2][PF6]2 (with four pendant terpyridyl sites) can be used as the starting point for polynuclear assemblies by attachment of additional labile metal ions as the secondary sites. As examples of this we prepared and structurally characterised the trinuclear complex [RuAg2(L1)2][ClO4]4, an unusual example of a polynuclear helicate containing a kinetically inert metal centre, and the pentanuclear complex [RuCu4(MeCN)5(H2O)1.5(L2)2](SbF6)6(BF4)4 in which each of the pendant terpyridyl sites of the [Ru(L2)2]2+ core is coordinated to a Cu(II) ion.
Co-reporter:Theodore Lazarides Dr.;Daniel Sykes Dr.;Stephen Faulkner ;Andrea Barbieri Dr.;MichaelD. Ward
Chemistry - A European Journal 2008 Volume 14( Issue 30) pp:9389-9399
Publication Date(Web):
DOI:10.1002/chem.200800600
Abstract
We have used time-resolved luminescence methods to study rates of photoinduced energy transfer (PEnT) from [M(bipy)3]2+ (M=Ru, Os) chromophores to LnIII ions with low-energy f–f states (Ln=Yb, Nd, Er) in d–f dyads in which the metal fragments are separated by a saturated CH2CH2 spacer, a p-C6H4 spacer, or a p-(C6H4)2 spacer. The finding that df PEnT is much faster across a conjugated p-C6H4 spacer than it is across a shorter CH2CH2 spacer points unequivocally to a Dexter-type energy transfer, involving electronic coupling mediated by the bridging ligand orbitals (superexchange) as the dominant mechanism. Comparison of the distance dependence of the RuNd energy-transfer rate across different conjugated spacers [p-C6H4 or p-(C6H4)2 groups] is also consistent with this mechanism. Observation of RuNd PEnT (as demonstrated by partial quenching of the RuII-based 3 MLCT emission (MLCT=metal-to-ligand charge transfer), and the growth of sensitised NdIII-based emission at 1050 nm) over approximately 20 Å by an exchange mechanism is a departure from the normal situation with lanthanides, in which long-range energy transfer often involves through-space Coulombic mechanisms.
Co-reporter:Richard Frantz, Christopher S. Grange, Nawal K. Al-Rasbi, Michael D. Ward and Jérôme Lacour
Chemical Communications 2007 (Issue 14) pp:1459-1461
Publication Date(Web):30 Jan 2007
DOI:10.1039/B618092B
The addition of enantiopure TRISPHAT anions to chiral cationic cages of type [Co4(L)6(BF4)]7+ leads to the enantiodifferentiation of the ligands of the racemic salts and, even more effectively, of the achiral tetrafluoroborate anion trapped inside.
Co-reporter:Frazer Kennedy, Nail M. Shavaleev, Thelma Koullourou, Zoë R. Bell, John C. Jeffery, Stephen Faulkner and Michael D. Ward
Dalton Transactions 2007 (Issue 15) pp:1492-1499
Publication Date(Web):28 Feb 2007
DOI:10.1039/B616423D
The luminescent transition metal complexes [Re(CO)3Cl(bppz)] and [Pt(CC–C6H4CF3)2(bppz)] [bppz = 2,3-bis(2-pyridyl)pyrazine], in which one of the diimine binding sites of the potentially bridging ligand bppz is vacant, have been used as ‘complex ligands’ to make heterodinuclear d–f complexes by attachment of a {Ln(dik)3} fragment (dik = a 1,3-diketonate) at the vacant site. When Ln = Pr, Nd, Er or Yb the lanthanide centre has low-energy f–f excited states capable of accepting energy from the 3MLCT excited state of the Pt(II) or Re(I) centre, quenching the 3MLCT luminescence and affording sensitised lanthanide(III)-based luminescence in the near-IR region. UV/Vis and luminescence spectroscopic titrations allowed measurement of (i) the association constants for binding of the {Ln(dik)3} fragment at the vacant diimine site of [Re(CO)3Cl(bppz)] or [Pt(CC–C6H4CF3)2(bppz)], and (ii) the degree of quenching of the 3MLCT luminescence according to the nature of the Ln(III) centre. In all cases Nd(III) was found to be the most effective of the series at quenching the 3MLCT luminescence of the d-block component because the high density of f–f excited states of the appropriate energy make it a particularly effective energy-acceptor.
Co-reporter:Theodore Lazarides, Mohammed A. H. Alamiry, Harry Adams, Simon J. A. Pope, Stephen Faulkner, Julia A. Weinstein and Michael D. Ward
Dalton Transactions 2007 (Issue 15) pp:1484-1491
Publication Date(Web):20 Feb 2007
DOI:10.1039/B700714K
The ligand L1, which contains a chelating 2-(2-pyridyl)benzimidazole (PB) unit with a pendant anthacenyl group An connected via a methylene spacer, (L1 = PB-An), was used to prepare the 8-coordinate lanthanide(III) complexes [Ln(hfac)3(L1)] (Ln = Nd, Gd, Er, Yb) which have been structurally characterised and all have a square antiprismatic N2O6 coordination geometry. Whereas free L1 displays typical anthracene-based fluorescence, this fluorescence is completely quenched in its complexes. The An group in L1 acts as an antenna unit: in the complexes [Ln(hfac)3(L1)] (Ln = Nd, Er, Yb) selective excitation of the anthracene results in sensitised near-infrared luminescence from the lanthanide centres with concomitant quenching of An fluorescence. Surprisingly, the anthracene fluorescence is also quenched even in the Gd(III) complex and in its Zn(II) adduct in which quenching via energy transfer to the metal centre is not possible. It is proposed that the quenching of anthracene fluorescence in coordinated L1 arises due to intra-ligand photoinduced electron-transfer from the excited anthracene chromophore 1An* to the coordinated PB unit generating a short-lived charge-separated state [An˙+–PB˙−] which collapses by back electron-transfer to give 3An*. This electron-transfer step is only possible upon coordination of L1 to the metal centre, which strongly increases the electron acceptor capability of the PB unit, such that 1An* → PB PET is endoergonic in free L1 but exergonic in its complexes. Thus, rather than a conventional set of steps (1An* →
3An* → Ln), the sensitization mechanism now includes 1An* → PB photoinduced electron transfer to generate charge-separated [An˙+–PB˙−], then back electron-transfer to generate 3An* which finally sensitises the Ln(III) centre via energy transfer. The presence of 3An* in L1 and its complexes is confirmed by nanosecond transient absorption studies, which have also shown that the 3An* lifetime in the Nd(III) complex matches the rise time of Nd-centred near-infrared emission, confirming that the final step of the sequence is 3An* → Ln(III) energy-transfer.
Co-reporter:Tanya K. Ronson, Harry Adams, Lindsay P. Harding, Simon J. A. Pope, Daniel Sykes, Stephen Faulkner and Michael D. Ward
Dalton Transactions 2007 (Issue 10) pp:1006-1022
Publication Date(Web):01 Feb 2007
DOI:10.1039/B618258E
A set of three potentially bridging ligands containing two tridentate chelating N,N′,O-donor (pyrazole–pyridine–amide) donors separated by an o, m, or p-phenylene spacer has been prepared and their coordination chemistry with lanthanide(III) ions investigated. Ligand L1 (p-phenylene spacer) forms complexes with a 2 : 3 M : L ratio according to the proportions used in the reaction mixture; the Ln2(L1)3 complexes contain two 9-coordinate Ln(III) centres with all three bridging ligands spanning both metal ions, and have a cylindrical (non-helical) ‘mesocate’ architecture. The 1 : 1 complexes display a range of structural types depending on the conditions used, including a cyclic Ln4(L1)4 tetranuclear helicate, a Ln2(L1)2 dinuclear mesocate, and an infinite one-dimensional coordination polymer in which metal ions and bridging ligands alternate along the sequence. ESMS studies indicate that the 1 : 1 complexes form a mixture of oligonuclear species {Ln(L1)}n in solution (n up to 5) which are likely to be cyclic helicates. In contrast, ligands L2 and L3 (with o- and m-phenylene spacers, respectively) generally form dinuclear Ln2L2 Ln(III) complexes in which the two ligands may be arranged in a helical or non-helical architecture about the two metal ions. These complexes also contain an additional exogenous bidentate bridging ligand, either acetate or formate, which has arisen from hydrolysis of solvent molecules promoted by the Lewis-acidity of the Ln(III) ions. Luminescence studies on some of the Nd(III) complexes showed that excitation into ligand-centred π–π* transitions result in the characteristic near-infrared luminescence from Nd(III) at 1060 nm.
Co-reporter:Sofia Derossi, Harry Adams and Michael D. Ward
Dalton Transactions 2007 (Issue 1) pp:33-36
Publication Date(Web):17 Nov 2006
DOI:10.1039/B614346F
The complex cations [RuL2(H2biim)]2+ (L = bipy, 4,4′-tBu2-bipy) interact with cyanometallate anions via a chelating hydrogen-bonding interaction between the two N–H donors of the complex cation and the N lone pair of one cyanide ligand in the complex anion; the anion hexacyanoferrate(III) quenches the Ru(II)-based luminescence in CH2Cl2 solution by photoinduced electron-transfer within the H-bonded assembly, whereas hexacyanocobaltate(III) enhances the Ru(II)-based luminescence.
Co-reporter:Svetlana G. Baca, Harry Adams, Daniel Sykes, Stephen Faulkner and Michael D. Ward
Dalton Transactions 2007 (Issue 23) pp:2419-2430
Publication Date(Web):05 Apr 2007
DOI:10.1039/B702235B
A series of cyanide-bridged coordination networks has been prepared which contain [Ru(phen)(CN)4]2− anions, Ln(III) cations, and additional oligopyridine ligands (1,10-phenanthroline, 2,2′:6′,2‴-terpyridine or 2,2′-bipyrimidine) which coordinate to the Ln(III) centres. Five structural types have been identified and examples of each type of structure are described: these are hexanuclear Ru4Ln2 clusters; two-dimensional Ru–Ln sheets with a honeycomb pattern of edge-linked Ru6Ln6 hexagons; one-dimensional chains consisting of two parallel cross-linked strands in a ladder-like arrangement; simple single-stranded chains of alternating Ru/Ln components; and a one-dimensional ‘chain of squares’ in which Ru2Ln2 squares are linked by bipyrimidine bridging ligands which connect to the Ln(III) corners of adjacent squares in the sequence. The 3MLCT luminescence characteristic of the [Ru(phen)(CN)4]2− units is quenched in those networks containing Ln(III) which have low-lying near-infrared luminescent f–f states [Pr(III), Nd(III), Er(III), Yb(III)], with sensitised Ln(III)-based near-IR luminescence generated by d → f energy-transfer. The rate of d → f energy-transfer, and hence the degree of quenching of the 3MLCT luminescence from the [Ru(phen)(CN)4]2− units, depends on the availability of f–f levels of an appropriate energy on the Ln(III) centre, with Nd(III) (with a high density of low-lying f–f states) being the most effective energy-acceptor and Yb(III) (with a single low-lying f–f state) being the least effective. Rates of d → f energy-transfer to different Ln(III) centres could be determined from both the residual (partially quenched) lifetimes of the 3MLCT luminescence, and—in the case of the Yb(III) networks—by a rise-time for the sensitised near-IR luminescence. The presence of the ‘blocking’ polypyridyl ligands, which reduced the number of cyanide and water ligands that would otherwise coordinate to the Ln(III) centres, resulted in increases in the Ln(III)-based emission lifetimes compared to networks where these blocking ligands were not used.
Co-reporter:Nawal K. Al-Rasbi;Harry Adams;Lindsay P. Harding;Michael D. Ward
European Journal of Inorganic Chemistry 2007 Volume 2007(Issue 30) pp:
Publication Date(Web):29 AUG 2007
DOI:10.1002/ejic.200700715
Three new ligands have been prepared in which two terdentate chelating pyrazolyl-bipyridine units are connected by a central aromatic spacer via methylene “hinges”: the spacers are o-phenylene (LPh), 2,6-pyridine-diyl (LPy) and 2,3-naphthalenediyl (Lnaph). The ligands act as potentially hexadentate bridging ligands, with the central pyridyl N atom of LPy not involved in coordination. The following complexes were prepared and structurally characterised: [M2(LPh)2][ClO4]4 (M = Ni, Cu), which are dinuclear double helicates; [Ag2(LPh)(MeCN)2][BF4]2, a dinuclear complex with an Ag···Ag bond in which the ligand adopts a helical twist around the pair of metal ions; [Ni2(LPy)2][BF4]4, an achiral “mesocate” with a box-like structure and a face-to-face arrangement of ligands; [Ag3(Lnaph)2](BF4)3, which contains a linear trinuclear array of AgI ions with the two ligands arranged in a shallow helical twist, each ligand spanning one terminal and the central metal ion; and [Cd6(Lnaph)6](ClO4)12, a cyclic hexanuclear helicate with a perchlorate anion in the central cavity. Both [Cu2(LPh)2][ClO4]4 and [Cd6(Lnaph)6](ClO4)12, which have architecturally similar bridging ligands, show evidence by electrospray mass spectrometry for formation of a range of cyclic oligomers in solution up to 11-mers for the CdII/Lnaph system.(© Wiley-VCH Verlag GmbH & Co. KGaA, 69451 Weinheim, Germany, 2007)
Co-reporter:Theodore Lazarides, Andrea Barbieri, Cristiana Sabatini, Francesco Barigelletti, Harry Adams, Michael D. Ward
Inorganica Chimica Acta 2007 Volume 360(Issue 3) pp:814-824
Publication Date(Web):15 February 2007
DOI:10.1016/j.ica.2006.04.029
A new ligand L1 has been prepared in which two 1,10-phenanthroline fragments are separated by an 18-crown-6 macrocyclic spacer. This was used to prepare the heterodinuclear complex [(bipy)2Ru(μ-L1)Re(CO)3Cl][PF6]2 [Ru(L1)Re] in which the {Ru(bipy)2(phen)}2+ and {Re(CO)Cl(phen)} chromophores are separated by a saturated and fairly flexible crown-ether fragment. On the basis of photophysical studies on Ru(L1)Re and associated mononuclear Ru(II) and Re(I) complexes, Re → Ru photoinduced energy-transfer occurs with a rate constant of 1.9 × 108 s−1 in solution room temperature leading to near-complete quenching of the Re(I)-based luminescence. At 77 K the Re(I)-based luminescence component is completely quenched. Calculations on the efficiency of both Förster and Dexter energy-transfer as a function of Re⋯Ru distance in this system suggest that a folded conformation of the complex, in which the Re⋯Ru separation is much shorter than that implied by the extended conformation detected crystallographically, is responsible for the energy-transfer, since neither Förster nor Dexter Re → Ru energy-transfer should be possible with the complex in an extended conformation. Addition of K+ or Ba2+ salts to solutions of Ru(L1)Re had no effect on the photophysical properties, probably because the association constants are too low to give significant metal-ion binding in the macrocycle at the low concentrations employed.The rate of photoinduced energy-transfer from the Re(I) to the Ru(II) terminus of the complex shown, based on a new bis-phenanthroline bridging ligand containing an 18-crown-6 spacer between the termini, has been evaluated and related to the molecular conformation.
Co-reporter:Theodore Lazarides, Graham M. Davies, Harry Adams, Cristiana Sabatini, Francesco Barigelletti, Andrea Barbieri, Simon J. A. Pope, Stephen Faulkner and Michael D. Ward
Photochemical & Photobiological Sciences 2007 vol. 6(Issue 11) pp:1152-1157
Publication Date(Web):26 Sep 2007
DOI:10.1039/B708683K
Crystallisation of [Co(CN)6]3− or [Cr(CN)6]3− with Ln(III) salts (Ln = Nd, Gd, Yb) from aqueous dmf afforded the cyanide-bridged d/f systems [Ln(dmf)4(H2O)3(μ-CN)Co(CN)5] (Co–Ln, discrete dinuclear species) and {[Cr(CN)4(μ-CN)2Ln(H2O)2(dmf)4]}∞ (Cr–Ln, infinite cyanide-bridged chains with alternating Cr and Ln centres). With Ln = Gd the characteristic long-lived phosphorescence from d–d excited states of the [M(CN)6]3− units was apparent in the red region of the spectrum, with lifetimes of the order of 1 μs, since the heavy atom effect of the Gd(III) promotes inter-system crossing at the [M(CN)6]3− units to generate the phosphorescent spin-forbidden excited states. With Ln = Yb or Nd however, the d-block luminescence was completely quenched due to fast (>108 s−1) energy-transfer to the Ln(III) centre, resulting in the characteristic sensitised emission from Yb(III) and Nd(III) in the near-IR region. For both Co–Nd and Co–Yb, calculations based on spectroscopic overlap between emission of the donor (Co) and absorption of the acceptor (Ln) suggest that the Dexter energy-transfer mechanism is responsible for the complete quenching that we observe.
Co-reporter:Juan-Manuel Herrera, Michael D. Ward, Harry Adams, Simon J. A. Pope and Stephen Faulkner
Chemical Communications 2006 (Issue 17) pp:1851-1853
Publication Date(Web):21 Mar 2006
DOI:10.1039/B601876A
The complexes [Ru(CN)4(HAT)]2−, [{Ru(CN)4}2(μ2-HAT)]4− and [{Ru(CN)4}3(μ3-HAT)]6− (HAT = hexaaza-triphenylene) contain four, eight and twelve externally-directed cyanide ligands, respectively; they show strongly solvatochromic and intense MLCT absorptions, and [3]6− forms a high-dimensionality cyanide-bridged coordination network with Nd(III), in which Ru → Nd energy transfer results in sensitised near-IR luminescence.
Co-reporter:Nawal K. Al-Rasbi, Cristiana Sabatini, Francesco Barigelletti and Michael D. Ward
Dalton Transactions 2006 (Issue 40) pp:4769-4772
Publication Date(Web):15 Sep 2006
DOI:10.1039/B609809F
Incorporation of ligands containing substituted naphthalene cores into coordination cages results in extensive aromatic π-stacking between ligands; this results in a red-shifted ‘excimer-like’ luminescence component from the naphthyl groups compared to the free ligands, which is diagnostic of, and can be used to monitor, cage assembly.
Co-reporter:Stephen P. Argent, Harry Adams, Lindsay P. Harding and Michael D. Ward
Dalton Transactions 2006 (Issue 4) pp:542-544
Publication Date(Web):28 Nov 2005
DOI:10.1039/B515296H
Reaction of the bis-bidentate bridging ligand L1 with Co(ClO4)2 or Zn(BF4)2 affords a mixture of complexes [M8(L1)12]X16 and [M6(L1)9]X12 having the same metal : ligand ratio: the former is a molecular cube with a metal ion at each vertex and a bridging ligand spanning each edge, whereas the latter has a metal framework like that of an ‘open book’ containing cross-linked double helical metal–ligand subunits.
Co-reporter:Harry Adams, Wassim Z. Alsindi, Graham M. Davies, Martin B. Duriska, Timothy L. Easun, Hazel E. Fenton, Juan-Manuel Herrera, Michael W. George, Kate L. Ronayne, Xue-Zhong Sun, Michael Towrie and Michael D. Ward
Dalton Transactions 2006 (Issue 1) pp:39-50
Publication Date(Web):14 Sep 2005
DOI:10.1039/B509042C
A series of complexes of the type K2[Ru(NN)(CN)4] has been prepared, in which NN is a diimine ligand, and were investigated for both their structural and photophysical properties. The ligands used (and the abbreviations for the resulting complexes) are 3-(2-pyridyl)pyrazole (Ru-pypz), 2,2′-bipyrimidine (Ru-bpym), 5,5′-dimethyl-2,2′-bipyridine (Ru-dmb), 1-ethyl-2-(2-pyridyl)benzimidazole (Ru-pbe), bidentate 2,2′:6′,2‴-terpyridine (Ru-tpy). The known complexes with NN
= 2,2′-bipyridine (Ru-bpy) and 1,10-phenathroline (Ru-phen) were also included in this work. A series of crystallographic studies showed that the [Ru(NN)(CN)4]2− complex anions form a range of elaborate coordination networks when crystallised with either K+ or Ln3+ cations. The K+ salts are characterised by a combination of near-linear Ru–CN–K bridges, with the cyanides coordinating to K+ in the usual ‘end-on’ mode, and unusual side-on π-type coordination of cyanide ligands to K+ ions. With Ln3+ cations in contrast only Ru–CN–Ln near-linear bridges occurred, affording 1-dimensional helical or diamondoid chains, and 2-dimensional sheets constituted from linked metallamacrocyclic rings. All of the K2[Ru(NN)(CN)4] complexes show a reversible Ru(II)/Ru(III) couple (ca.
+0.9 V vs. Ag/AgCl in water), the exception being Ru-tpy whose oxidation is completely irreversible. Luminescence studies in water showed the presence of 3MLCT-based emission in all cases apart from Ru-bpym with lifetimes of tens/hundreds of nanoseconds. Time-resolved infrared studies showed that in the 3MLCT excited state the principal C–N stretching vibration shifts to positive energy by ca. 50 cm−1 as a consequence of the transient oxidation of the metal centre to Ru(III) and the reduction in back-bonding to the cyanide ligands; measurement of transient decay rates allowed measurements of 3MLCT lifetimes for those complexes which could not be characterised by luminescence spectroscopy. A few complexes were also examined in different solvents (MeCN, dmf) and showed much weaker emission and shorter excited-state lifetimes in these solvents compared to water.
Co-reporter:Stephen P. Argent, Harry Adams, Thomas Riis-Johannessen, John C. Jeffery, Lindsay P. Harding, William Clegg, Ross W. Harrington and Michael D. Ward
Dalton Transactions 2006 (Issue 42) pp:4996-5013
Publication Date(Web):07 Sep 2006
DOI:10.1039/B607541J
The coordination chemistry of a series of di- and tri-nucleating ligands with Ag(I), Hg(I) and Hg(II) has been investigated. Most of the ligands contain two or three N,N′-bidentate chelating pyrazolyl-pyridine units pendant from a central aromatic spacer; one contains three binding sites (2 + 3 + 2-dentate) in a linear sequence. A series of thirteen complexes has been structurally characterised displaying a wide range of structural types. Bis-bidentate bridging ligands react with Ag(I) to give complexes in which Ag(I) is four-coordinate from two bidentate donors, but the complexes can take the form of one-dimensional coordination polymers, or dinuclear complexes (mesocate or helicate). A tris-bidentate triangular ligand forms a complicated two-dimensional coordination network with Ag(I) in which Ag⋯Ag contacts, as well as metal–ligand coordination bonds, play a significant role. Three dinuclear Hg(I) complexes were isolated which contain an {Hg2}2+ metal–metal bonded core bound to a single bis-bidentate ligand which can span both metal ions. Also characterised were a series of Hg(II) complexes comprising a simple mononuclear four-coordinate Hg(II) complex, a tetrahedral HgII4 cage which incorporates a counter-ion in its central cavity, a trinuclear double helicate, and a trinuclear catenated structure in which two long ligands have spontaneously formed interlocked metallomacrocyclic rings thanks to cyclometallation of two of the Hg(II) centres.
Co-reporter:Tanya K. Ronson, Harry Adams, T. Riis-Johannessen, John C. Jeffery and Michael D. Ward
New Journal of Chemistry 2006 vol. 30(Issue 1) pp:26-28
Publication Date(Web):08 Dec 2005
DOI:10.1039/B515154F
A 1 : 1 mixture of the homoleptic double helicates [MII(L1)2] and [MII(L2)2]
[M = Cu, Zn; (L1)2− and (L2)2− are bis-bidentate ligands containing two pyrazolyl-phenolate termini but with different spacers separating them] affords the mixed ligand complexes [M(L1)(L2)] in high yield, indicative of a favourable inter-ligand interaction in the mixed-ligand complexes. Whereas [Cu(L1)(L2)] is a double helicate, [Zn(L1)(L2)] is a mesocate with a ‘face to face’ arrangement of the two ligands.
Co-reporter:Tanya K. Ronson;Theodore Lazarides Dr.;Harry Adams;Simon J. A. Pope Dr.;Daniel Sykes;Stephen Faulkner ;Simon J. Coles Dr.;Michael B. Hursthouse ;William Clegg ;Ross W. Harrington Dr.;Michael D. Ward
Chemistry - A European Journal 2006 Volume 12(Issue 36) pp:
Publication Date(Web):22 SEP 2006
DOI:10.1002/chem.200600698
The complexes [Pt(bipy){CC-(4-pyridyl)}2] (1) and [Pt(tBu2bipy){CC-(4-pyridyl)}2] (2) and [Pt(tBu2-bipy)(CC-phen)2] (3) all contain a Pt(bipy)(diacetylide) core with pendant 4-pyridyl (1 and 2) or phenanthroline (3) units which can be coordinated to {Ln(diketonate)3} fragments (Ln = a lanthanide) to make covalently-linked PtII/LnIII polynuclear assemblies in which the PtII chromophore, absorbing in the visible region, can be used to sensitise near-infrared luminescence from the LnIII centres. For 1 and 2 one-dimensional coordination polymers [1⋅Ln(tta)3]∞ and [2⋅Ln(hfac)3]∞ are formed, whereas 3 forms trinuclear adducts [3⋅{Ln(hfac)3}2] (tta=anion of thenoyl-trifluoroacetone; hfac=anion of hexafluoroacetylacetone). Complexes 1–3 show typical PtII-based 3MLCT luminescence in solution at ≈510 nm, but in the coordination polymers [1⋅Ln(tta)3]∞ and [2⋅Ln(hfac)3]∞ the presence of stacked pairs of PtII units with short Pt⋅⋅⋅Pt distances means that the chromophores have 3MMLCT character and emit at lower energy (≈630 nm). Photophysical studies in solution and in the solid state show that the 3MMLCT luminescence in [1⋅Ln(tta)3]∞ and [2⋅Ln(hfac)3]∞ in the solid state, and the 3MLCT emission of [3⋅{Ln(hfac)3}2] in solution and the solid state, is quenched by PtLn energy transfer when the lanthanide has low-energy f–f excited states which can act as energy acceptors (Ln=Yb, Nd, Er, Pr). This results in sensitised near-infrared luminescence from the LnIII units. The extent of quenching of the PtII-based emission, and the PtLn energy-transfer rates, can vary over a wide range according to how effective each LnIII ion is at acting as an energy acceptor, with YbIII usually providing the least quenching (slowest PtLn energy transfer) and either NdIII or ErIII providing the most (fastest PtLn energy transfer) according to which one has the best overlap of its f–f absorption manifold with the PtII-based luminescence.
Co-reporter:Stephen P. Argent, Thomas Riis-Johannessen, John C. Jeffery, Lindsay P. Harding and Michael D. Ward
Chemical Communications 2005 (Issue 37) pp:4647-4649
Publication Date(Web):17 Aug 2005
DOI:10.1039/B509239F
A chiral bridging ligand affords a single diastereoisomer of tetrahedral M4L6 cage complex in which the optical rotation of each ligand is increased by a factor of 5 on coordination.
Co-reporter:Theodore Lazarides, Thomas A. Miller, John C. Jeffery, Tanya K. Ronson, Harry Adams and Michael D. Ward
Dalton Transactions 2005 (Issue 3) pp:528-536
Publication Date(Web):05 Jan 2005
DOI:10.1039/B416293E
A range of ligands in which a macrocyclic unit is fused to a 1,10-phenanthroline unit has been prepared starting from 5,6-dihydroxyphenanthroline. The ligands are L1 in which the pendant ligand is 18-crown-6; L2, in which the pendant ligand is benzo-24-crown-8; and L3, in which the macrocycle contains two carboxamide units. Ligands L1 and L2 can bind Group 1 and 2 metal cations in their crown-ether cavities; L3 contains two H-bond (amide) donors and is suitable for anion-binding. Luminescent complexes of the form [Ru(bipy)2L]2+, [ReL(CO)3Cl] and [RuL(CN)4]2− were prepared and some were structurally characterised; their interactions with various guest species were investigated by luminescence and NMR spectroscopy. For complexes with the crown ethers (L1 and L2), binding of K+ was rather weak, but the electrostatic effect due to the charge on the host complex was clear with [RuL1(CN)4]2− binding K+ more strongly than [Ru(bipy)2L1]2+. Binding to the pendant crown ethers was much stronger with Ba2+, and both [ReL1(CO)3Cl] and [ReL2(CO)3Cl] showed substantial luminescence quenching in MeCN on addition of Ba2+ ions, with binding constants of 4.5 × 104 M−1 for [ReL1(CO)3Cl]/Ba2+ and 1.3 × 105 M−1 for [ReL2(CO)3Cl]/Ba2+. Complexes [Ru(bipy)2L3]2+ and [ReL3(CO)3Cl], due to their H-bond donor sites, showed binding of dihydrogenphosphate to the macrocycle. Whereas [ReL3(CO)3Cl] showed 1 : 1 binding with (H2PO4)− in dmso with a binding constant of 65 M−1, [Ru(bipy)2L3]2+ showed 1 : 2 binding, with microscopic association constants of ca. 1 × 106 and 1.6 × 106 M−1 in MeCN. The fact that K2 > K1 suggests a cooperative interaction whereby binding of the first anion makes binding of the second one easier to an extent which overcomes electrostatic effects, and a model for this is proposed which also accounts for the substantial increase in luminescence from [Ru(bipy)2L3]2+
(5-fold enhancement) when the second (H2PO4)− anion binds. Both [Ru(bipy)2L3]2+ and [ReL3(CO)3Cl] undergo complete luminescence quenching and a change in colour to near-black in the presence of (anhydrous) fluoride in MeCN, probably due to deprotonation of the carboxamide group. These changes are however irreversible on a long timescale and lead to slow decomposition.
Co-reporter:Harry Adams, Stuart R. Batten, Graham M. Davies, Martin B. Duriska, John C. Jeffery, Paul Jensen, Jinzhen Lu, Graham R. Motson, Simon J. Coles, Michael B. Hursthouse and Michael D. Ward
Dalton Transactions 2005 (Issue 11) pp:1910-1923
Publication Date(Web):28 Apr 2005
DOI:10.1039/B502892B
The new ligands dihydrobis[3-(4-pyridyl)pyrazol-1-yl]borate [Bp4py]−, hydrotris[3-(4-pyridyl)pyrazol-1-yl]borate [Tp4py]−, tetrakis[3-(4-pyridyl)pyrazol-1-yl]borate [Tkp4py]−, dihydrobis[3-(3-pyridyl)pyrazol-1-yl]borate [Bp3py]−, hydrotris[3-(3-pyridyl)pyrazol-1-yl]borate [Tp3py]− and tetrakis[3-(3-pyridyl)pyrazol-1-yl]borate [Tkp4py]− are derivatives of the well known bis-, tris- and tetrakis-(pyrazolyl)borate cores, bearing 4-pyridyl or 3-pyridyl substituents attached to the pyrazolyl C3 positions. These pyridyl groups cannot chelate to the metal ions in the poly(pyrazolyl) cavity but are externally directed. Structural studies on a range of metal complexes show how, in many cases, coordination of these pendant pyridyl groups to the M(pyrazolyl)n core of an adjacent metal complex fragment results in formation of coordination oligomers or polymeric networks. [Tl(Bp3py)], [Tl(Bp4py)] and [Tl(Tp4py)] form one-dimensional polymeric chains via coordination of one of their pendant pyridyl units to the Tl(I) centre of an adjacent complex fragment; in contrast, in [Tl(Tp3py)] coordination of all three pendant pyridyl units to separate Tl(I) neighbours results in formation of a two-dimensional polymeric sheet. In [Tl(Tkp3py)] and [Tl(Tkp4py)] the Tl(I) is coordinated by two or three of the four pyrazolyl arms, respectively; bridging interactions of pendant 4-pyridyl groups with adjacent Tl(I) centres result in a two-dimensional sheet forming in each case. In Ag(Tkp4py) each Ag(I) ion is coordinated by two pyrazolyl rings, and two bridging pyridyl ligands from other complex units, resulting in a one-dimensional chain consisting of pairs of cross-linked zigzag chains. In contrast to these polymeric coordination networks, the structures of [Cu(Tp4py)] and [(Tp3py)Cd(CH3CO2)] are dimers, with a pendant pyridyl residue from the first metal centre attaching to a vacant coordination site on the second, and vice versa; these dimers are stabilised by π-stacking interactions between sections of the two ligands. [Ni(Tp3py)2] is monomeric, with an octahedral coordination geometry arising from two tris(pyrazolyl)borate chelates; the array of pendant 3-pyridyl groups is involved only in intramolecular hydrogen-bonding. [(Tp4py)Re(CO)3] is also monomeric, with a facial arrangement of three pyrazolyl ligands and three carbonyls, with the pendant 4-pyridyl groups not further coordinated. [(Tp2py)Re(CO)3], based on the related ligand hydrotris[3-(2-pyridyl)pyrazol-1-yl]borate, has a similar fac-(CO)3(pyrazolyl)3 coordination geometry.
Co-reporter:Tanya K. Ronson;Harry Adams;Michael D. Ward
European Journal of Inorganic Chemistry 2005 Volume 2005(Issue 22) pp:
Publication Date(Web):4 OCT 2005
DOI:10.1002/ejic.200500704
We have prepared a series of five ligands with potentially N,S-bidentate chelating arms derived from 3-[2-(methylsulfanyl)phenyl]pyrazole linked to central aromatic spacers by methylene units. Complexes with a variety of architectures have been obtained, including simple mononuclear complexes and polynuclear chain complexes. The p-xylyl-spaced ligand L1 forms one-dimensional helical coordination polymers with copper(I) and silver(I) ions. These polymers display interligand aromatic stacking interactions within each helical chain. The m-xylyl-spaced ligand L2 forms a coordination polymer with copper(I) but a mononuclear complex with the larger silver(I) ion in which the central phenyl ring is involved in an η1 π-type Ag···C interaction with the AgI. The 3,3′-biphenyl-spaced ligand L3 also forms one-dimensional polymers with silver(I) and copper(I) ions, but in this case the sequence of bridging ligands between one metal centre and the next follows a zig-zag path rather than being helical. The 1,8-naphthyl-spaced ligand L4 only forms mononuclear complexes with copper(I) and silver(I) ions showing that this spacer is not large enough to enforce a bridging coordination mode. The three-armed ligand L5, prepared from 2,4,6-tris(bromomethyl)mesitylene, also forms a mononuclear complex with silver(I) ions, where one of the three arms is pendant. However, when excess silver(I) ions are present two of these mononuclear complexes can be assembled into the trinuclear complex [Ag3(L5)2](ClO4)3. (© Wiley-VCH Verlag GmbH & Co. KGaA, 69451 Weinheim, Germany, 2005)
Co-reporter:Stephen P. Argent, Harry Adams, Lindsay P. Harding, T. Riis-Johannessen, John C. Jeffery and Michael D. Ward
New Journal of Chemistry 2005 vol. 29(Issue 7) pp:904-911
Publication Date(Web):18 May 2005
DOI:10.1039/B502423D
An octadentate ligand L has been prepared which contains a sequence of bidentate (pyrazolyl-pyridine), terdentate [bis(pyrazolyl)pyridine] and bidentate (pyrazolyl-pyridine) binding sites separated by p-xylyl spacers. This forms a range of double helical complexes in which the two ligands define 4-, 6-, and 4-coordinate binding sites, and there is substantial π-stacking between overlapping parallel areas of the ligands. In [Cu3L2][PF6]4 the sequence of oxidation states for the copper ions is +1, +2, +1 with the Cu(I) ions being four-coordinate at the terminal sites and Cu(II) being in the central six-coordinate site. In [Cu3(OAc)2L2][PF6]4 all copper centres are in oxidation state +2, with the terminal ions having an additional monodentate acetate ligand giving them a five-coordinate geometry. The 4 + 6 + 4 arrangement of coordination numbers means that reaction of L with a mixture of Fe(II) and Ag(I) results in high yield formation of [Ag2FeL2][BF4]4 in which Ag(I) ions occupy the terminal 4-coordinate sites and Fe(II) occupies the central pseudo-octahedral site. Reaction of L with Ag(I) produced a mixture of [Ag3L2][BF4]3
(major product) and [Ag4L2][BF4]4
(minor product). In [Ag3L2][BF4]3 the central Ag(I) ion is, unusually, in a pseudo-octahedral coordination environment from the two meridional, terdentate bis(pyrazolyl)pyridine donors. In [Ag4L2][BF4]4 in contrast the central 6-coordinate cavity is occupied by two Ag(I) ions separated by 2.85 Å. The terdentate chelating bis(pyrazolyl)pyridine units at the centre of the helicate are now substantially twisted such that each donates a bidentate pyrazolyl-pyridine to one Ag(I) centre and a monodentate pyrazole unit to the other. In solution, 1H NMR and mass spectroscopic evidence indicates that the fourth Ag(I) ion is lost and [Ag3L2][BF4]3 forms, unless a large excess of Ag(I) is present in which case traces of [Ag4L2][BF4]4 can be detected by mass spectrometry.
Co-reporter:Tanya K. Ronson, Harry Adams, Michael D. Ward
Inorganica Chimica Acta 2005 Volume 358(Issue 13) pp:4104
Publication Date(Web):1 September 2005
DOI:10.1016/j.ica.2005.07.007
Co-reporter:Graham M. Davies, Harry Adams, Simon J. A. Pope, Stephen Faulkner and Michael D. Ward
Photochemical & Photobiological Sciences 2005 vol. 4(Issue 10) pp:829-834
Publication Date(Web):12 Aug 2005
DOI:10.1039/B508382F
The complexes [M(L1)2(NO3)] and [M(L2)(NO3)2]
(M = Pr, Er; L1
= the tetradentate ligand dihydrobis-[3-(2-pyridyl)pyrazolyl]borate; L2
= the hexadentate ligand hydrotris-[3-(2-pyridyl)pyrazolyl]borate) were prepared and their structural and photophysical properties studied. All complexes are 10-coordinate. Crystallographic analysis of [M(L1)2(NO3)]
(M = Pr, Er) showed that for the smaller Er(III) ions steric congestion at the metal centre results in two of the Er–N(pyridyl) distances being particularly long, which does not occur with the larger Pr(III) ion that is better able to accommodate 10-fold coordination. On UV irradiation, both Pr(III) complexes show, in the visible region of their luminescence spectra, transitions originating from both the 3P0 level (at ca. 21000 cm−1) and the 1D2 level (at ca. 17000 cm−1), a consequence of the fact that the lowest triplet state of the coordinated pyrazolylborate ligands lies at ca. 24000 cm−1 in each case so is high enough in energy to populate both levels. This contrasts with Pr(III) complexes based on diketonate ligands in which the lower triplet energies of the ligands result in emission from the 1D2 level only. At longer wavelengths, near-infrared luminescence arising from the 1D2 emissive level is observed with lifetimes (in both the solid state and solution) being in the range 50–110 ns. For both Er(III) complexes, luminescence at 1530 nm occurs following UV excitation of ligand-centred transitions. In CH2Cl2 both complexes gave dual-exponential luminescence, with the major component having a lifetime characteristic of an intact Er(III) complex (≈ 1.5 µs) and the minor component being much shorter lived (0.2–0.5 µs), suggestive of a species in which a ligand is partially detached and the metal is solvated, with the two forms interconverting slowly. This behaviour is consistent with the steric congestion and long M–N(pyridyl) bonds that were observed in [Er(L1)2(NO3)]. In the solid state both Er(III) complexes gave very weak luminescence, which could be fitted to a single exponential decay with a lifetime similar to the longer-lived of the solution components.
Co-reporter:Rowena L. Paul, Stephen P. Argent, John C. Jeffery, Lindsay P. Harding, Jason M. Lynam and Michael D. Ward
Dalton Transactions 2004 (Issue 21) pp:3453-3458
Publication Date(Web):16 Sep 2004
DOI:10.1039/B409809A
Reaction of the bis-bidentate bridging ligand L3, in which two bidentate chelating 3(2-pyridyl)pyrazole units are separated by a 3,3′-biphenyl spacer, with Co(II) salts affords tetranuclear cage complexes of composition [Co4(L3)6]X8
(X =
[BF4]−, [ClO4]−, [PF6]− or I−) in which four 6-coordinate Co(II) ions in an approximately tetrahedral array are connected by six bis-bidentate bridging ligands, one spanning each of the six edges of the Co4 tetrahedron. In every case, X-ray crystallography reveals that the ‘apical’ Co(II) ion has a fac tris-chelate geometry, whereas the other three Co(II) ions have mer tris-chelate geometries, resulting in (non-crystallographic)
C3 symmetry for the cages; that this structure is retained in solution is confirmed by 1H NMR spectroscopy of the paramagnetic cages. In every case one of the anions is located inside the central cavity of the cage, with the remaining seven outside. We found no clear evidence for an anion-based templating effect. The cage superstructure is sufficiently large to leave gaps in the centres of the faces through which the internal and external anions can exchange. Variable-temperature 19F NMR spectroscopy was used to investigate the dynamic behaviour of the cages with X =
[BF4]− and [PF6]− in MeCN solution: in both cases two separate signals, corresponding to external and internal anions, are clear at 233 K which have coalesced to a single signal at room temperature. Analysis of the linewidth of the minor signal (for the internal anion) at various temperatures below coalescence gave an activation energy for anion exchange of ca. 50 kJ mol−1 in each case, a figure which suggests that anion exchange can occur via a conformational rearrangement of the cage superstructure in solution rather than opening of the cavity by cleavage of metal–ligand bonds.
Co-reporter:Graham M. Davies, Rebecca J. Aarons, Graham R. Motson, John C. Jeffery, Harry Adams, Stephen Faulkner and Michael D. Ward
Dalton Transactions 2004 (Issue 8) pp:1136-1144
Publication Date(Web):17 Mar 2004
DOI:10.1039/B400992D
The ligands tris[3-(2-pyridyl)pyrazol-1-yl]hydroborate (L1, potentially hexadentate) and bis[3-(2-pyridyl)pyrazol-1-yl]dihydroborate (L2, potentially tetradentate) have been used to prepare ternary lanthanide complexes in which the remaining ligands are dibenzoylmethane anions (dbm). [Eu(L1)(dbm)2] is eight-coordinate, with L1 acting only as a tetradentate chelate (with one potentially bidentate arm pendant) and two bidentate dbm ligands. [Nd(L1)(dbm)2] was also prepared but on recrystallisation some of it rearranged to [Nd(L1)2][Nd(dbm)4], which contains a twelve-coordinate [Nd(L1)2]+ cation (two interleaved hexadentate podand ligands) and the eight-coordinate anion [Nd(dbm)4]− which, uniquely amongst eight-coordinate complexes having four diketonate ligands, has a square prismatic structure with near-perfect O8 cubic coordination. Formation of this sterically unfavourable geometry is assumed to arise from favourable packing with the pseudo-spherical cation. The isostructural series of complexes [Ln(L2)(dbm)2]
(Ln = Pr, Nd, Eu, Gd, Tb, Er, Yb) was also prepared and all members structurally characterised; again the metal ions are eight-coordinate, from one tetradentate ligand L2 and two bidentate dbm ligands. Photophysical studies on the complexes with Ln = Pr, Nd, Er, and Yb were carried out; all show the near-IR luminescence characteristic of these metal ions, with longer lifetimes in CD3OD than in CH3OH. For [Yb(L2)(dbm)2], two species with different luminescence lifetimes were observed in CH3OH solution, corresponding to species with zero or one coordinated solvent molecules, in slow exchange on the luminescence timescale. For [Nd(L2)(dbm)2] a single average solvation number of 0.7 was observed in MeOH. For [Pr(L2)(dbm)2] a range of emission lines in the visible and NIR regions was detected; time-resolved measurements show a particularly high susceptibility to quenching by solvent CH and OH oscillators.
Co-reporter:Thomas A. Miller, John C. Jeffery, Michael D. Ward, Harry Adams, Simon J. A. Pope and Stephen Faulkner
Dalton Transactions 2004 (Issue 10) pp:1524-1526
Publication Date(Web):21 Apr 2004
DOI:10.1039/B404820B
Co-crystallisation of the anionic cyanometallate chromophore [Ru(bipy)(CN)4]2− with Yb(III) provides coordination polymers or oligomers containing Ru–CN–Yb bridges; in [K(H2O)4][Yb(H2O)6][Ru(bipy)(CN)4]2·5H2O Ru → Yb energy-transfer (k > 5 × 106 s−1) results in partial quenching of the Ru-based luminescence and sensitised near-IR luminescence from the Yb(III) unit.
Co-reporter:Nail M. Shavaleev, Zöe R. Bell, Timothy L. Easun, Ramune Rutkaite, Linda Swanson and Michael D. Ward
Dalton Transactions 2004 (Issue 21) pp:3678-3688
Publication Date(Web):06 Oct 2004
DOI:10.1039/B411341A
N,N′-Chelating ligands based on the 2-(2-pyridyl)benzimidazole (PB) core have been prepared with a range of substituents (phenyl, pentafluorophenyl, naphthyl, anthracenyl, pyrenyl) connected to the periphery via alkylation of the benzimidazolyl unit at one of the N atoms. These PB ligands have been used to prepare a series of complexes of the type [Re(PB)(CO)3Cl], [Pt(PB)(CCR)2]
(where –CCR is an acetylide ligand) and [Ru(bpy)2(PB)][PF6]2
(bpy = 2,2′-bipyridine). Six of the complexes have been structurally characterised. Electrochemical and luminescence studies show that all three series of complexes behave in a similar manner to the analogous complexes with 2,2′-bipyridine in place of PB. In particular, all three series of complexes show luminescence in the range 553–605 nm (Pt series), 620–640 nm (Re series) and 626–645 nm (Ru series) arising from the 3MLCT state, with members of the Pt(II) series being the most strongly emissive with lifetimes of up to 500 ns and quantum yields of up to 6% in air-saturated CH2Cl2 at room temperature. In the Re and Ru series there was clear evidence for inter-component energy-transfer processes in both directions between the 3MLCT state of the metal centre and the singlet and triplet states of the pendant organic luminophores (naphthalene, pyrene, anthracene). For example the pyrene singlet is almost completely quenched by energy transfer to a Re-based MLCT excited state, which in turn is completely quenched by energy transfer to the lower-lying pyrene triplet state. For the analogous Ru(II) complexes the inter-component energy transfer is less effective, with 1anthracene → Ru(3MLCT) energy transfer being absent, and Ru(3MLCT)
→
3anthracene energy transfer being incomplete. This is rationalised on the basis of a greater effective distance for energy transfer in the Ru(II) series, because the MLCT excited states are localised on the bpy ligands which are remote from the pendant aromatic group; in the Re series in contrast, the MLCT excited states involve the PB ligand to which the pendant aromatic group is directly attached, giving more efficient energy transfer.
Co-reporter:Zöe R. Bell, Lindsay P. Harding and Michael D. Ward
Chemical Communications 2003 (Issue 19) pp:2432-2433
Publication Date(Web):04 Sep 2003
DOI:10.1039/B307172N
Reaction of the bis-bidentate ligand L1, having two bidentate pyrazolyl-pyridine termini, with Co(II) or Zn(II) results in formation of the complexes [M8(L1)12]X16
(X = perchlorate or tetrafluoroborate); [Zn8(L1)12](ClO4)16 has been structurally characterised and is a cube with a metal ion at each corner, a bridging ligand along each edge, and an anion in the central cavity.
Co-reporter:Thomas A. Miller, John C. Jeffery and Michael D. Ward
CrystEngComm 2003 vol. 5(Issue 88) pp:495-497
Publication Date(Web):04 Dec 2003
DOI:10.1039/B313122J
This communication describes the crystal structure of the mixed-metal coordination polymer {[Sm(H2O)5][Ru(CN)3(bpy)-(μ-CN)-Ru(CN)3(bpy)]·11H2O}∞, which was isolated in low yield from co-crystallisation of K2[Ru(bpy)(CN)4] with Sm(NO3)3·6H2O. This polymeric chain extends along the a-axis and is constructed from cyanide ligands bridging Ru and Sm metal vertices; it contains both hydrophobic and hydrophilic channels, the latter filled with an H-bonded network of water molecules.
Co-reporter:Nail M. Shavaleev Dr.;Lucy P. Moorcraft;Simon J. A. Pope Dr.;Zöe R. Bell Dr.;Stephen Faulkner Dr.;Michael D. Ward
Chemistry - A European Journal 2003 Volume 9(Issue 21) pp:
Publication Date(Web):3 NOV 2003
DOI:10.1002/chem.200305132
A series of dinuclear platinum(II)–lanthanide(iii) complexes has been prepared in which a square-planar PtII unit, either [(PPh3)2Pt(pdo)] (H2pdo=5,6-dihydroxyphenanthroline) or [Cl2Pt(dppz)] [dppz=2,3-bis(2-pyridyl)pyrazine], is connected to a {Ln(dik)3} unit (“dik”=a 1,3-diketonate ligand). The mononuclear complexes [(PPh3)2Pt(pdo)] and [Cl2Pt(dppz)] both have external, vacant N,N-donor diimine-type binding sites that react with various [Ln(dik)3(H2O)2] units to give complexes [(PPh3)2Pt(μ-pdo)Ln(tta)3] (series A; Htta=thenoyltrifluoroacetone), [Cl2Pt(μ-dppz)Ln(tta)3] (series B); and [Cl2Pt(μ-dppz)Ln(btfa)3] (series C; Hbtfa=benzoyltrifluoroacetone); in all of these the lanthanide centres are eight-coordinate. The lanthanides used exhibit near-infrared luminescence (Nd, Yb, Er). Crystal structures of members of each series are described. In all complexes, excitation into the Pt-centred absorption band (at 520 nm for series A complexes; 440 nm for series B and C complexes) results in characteristic near-IR luminescence from the Nd, Yb or Er centres in both the solid state and in CH2Cl2, following energy-transfer from the Pt antenna chromophore. This work demonstrates how d-block-derived chromophores, with their intense and tunable electronic transitions, can be used as sensitisers to achieve near-infrared luminescence from lanthanides in suitably designed heterodinuclear complexes based on simple bridging ligands.
Co-reporter:Atanu Jana, Simon J.A. Pope, Michael D. Ward
Polyhedron (8 May 2017) Volume 127() pp:
Publication Date(Web):8 May 2017
DOI:10.1016/j.poly.2017.02.016
A series of d/f dinuclear complexes Ir•Ln [where Ln = Eu(III), Gd(III), Yb(III) and Nd(III)] are reported. The core structure consists of a rigid skeleton containing two different types of receptor site in a single molecular motif designed to combine a transition metal ion [Ir(III)] and a lanthanide ion [Ln(III)] in different binding sites at either end of fully conjugated bridge to facilitate d → f energy-transfer following photoexcitation of the Ir(III)-based antenna unit. Steady state and time-resolved photophysical experiments on these compounds revealed that the energy-transfer is feasible only in case of Ir•Yb and Ir•Nd systems, affording sensitized emission from the Yb(III) or Nd(III) centres. Such EnT is not possible in the Ir•Eu dyad as the excited state energy of the Ir(III) unit is insufficient to sensitise the excited state of the Eu(III) centre.In a series of Ir(III)/Ln(III) dinuclear complexes, photoinduced energy-transfer from the emissive Ir(III) unit to the lanthanide unit occurs when Ln = Yb or Nd, generating sensitised near-infrared luminescence, but not when Ln = Eu.Download high-res image (79KB)Download full-size image
Co-reporter:Andrew Stephenson ; Stephen P. Argent ; Thomas Riis-Johannessen ; Ian S. Tidmarsh ;Michael D. Ward
Journal of the American Chemical Society () pp:
Publication Date(Web):December 22, 2010
DOI:10.1021/ja107403p
The bis-bidentate bridging ligand L {α,α′-bis[3-(2-pyridyl)pyrazol-1-yl]-1,4-dimethylbenzene}, which contains two chelating pyrazolyl-pyridine units connected to a 1,4-phenylene spacer via flexible methylene units, reacts with transition metal dications to form a range of polyhedral coordination cages based on a 2M:3 L ratio in which a metal ion occupies each vertex of a polyhedron, a bridging ligand lies along every edge, and all metal ions are octahedrally coordinated. Whereas the Ni(II) complex [Ni8L12](BF4)12(SiF6)2 is an octanuclear cubic cage of a type we have seen before, the Cu(II), Zn(II), and Cd(II) complexes form new structural types. [Cu6L9](BF4)12 is an unusual example of a trigonal prismatic cage, and both Zn(II) and Cd(II) form unprecedented hexadecanuclear cages [M16L24]X32(X = ClO4 or BF4) whose core is a skewed tetracapped truncated tetrahedron. Both Cu6L9 and M16L24 cages are based on a cyclic helical M3L3 subunit that can be considered as a triangular “panel”, with the cages being constructed by interconnection of these (homochiral) panels with additional bridging ligands in different ways. Whereas [Cu6L9](BF4)12 is stable in solution (by electrospray mass spectrometry, ES-MS) and is rapidly formed by combination of Cu(BF4)2 and L in the correct proportions in solution, the hexadecanuclear cage [Cd16L24](BF4)32 formed on crystallization slowly rearranges in solution over a period of several weeks to the trigonal prism [Cd6L9](BF4)12, which was unequivocally identified on the basis of its 1H NMR spectrum. Similarly, combination of Cd(BF4)2 and L in a 2:3 ratio generates a mixture whose main component is the trigonal prism [Cd6L9](BF4)12. Thus the hexanuclear trigonal prism is the thermodynamic product arising from combination of Cd(II) and L in a 2:3 ratio in solution, and arises from both assembly of metal and ligand (minutes) and rearrangement of the Cd16 cage (weeks); the large cage [Cd16L24](BF4)32 is present as a minor component of a mixture of species in solution but crystallizes preferentially.
Co-reporter:Theodore Lazarides, Mohammed A. H. Alamiry, Harry Adams, Simon J. A. Pope, Stephen Faulkner, Julia A. Weinstein and Michael D. Ward
Dalton Transactions 2007(Issue 15) pp:NaN1491-1491
Publication Date(Web):2007/02/20
DOI:10.1039/B700714K
The ligand L1, which contains a chelating 2-(2-pyridyl)benzimidazole (PB) unit with a pendant anthacenyl group An connected via a methylene spacer, (L1 = PB-An), was used to prepare the 8-coordinate lanthanide(III) complexes [Ln(hfac)3(L1)] (Ln = Nd, Gd, Er, Yb) which have been structurally characterised and all have a square antiprismatic N2O6 coordination geometry. Whereas free L1 displays typical anthracene-based fluorescence, this fluorescence is completely quenched in its complexes. The An group in L1 acts as an antenna unit: in the complexes [Ln(hfac)3(L1)] (Ln = Nd, Er, Yb) selective excitation of the anthracene results in sensitised near-infrared luminescence from the lanthanide centres with concomitant quenching of An fluorescence. Surprisingly, the anthracene fluorescence is also quenched even in the Gd(III) complex and in its Zn(II) adduct in which quenching via energy transfer to the metal centre is not possible. It is proposed that the quenching of anthracene fluorescence in coordinated L1 arises due to intra-ligand photoinduced electron-transfer from the excited anthracene chromophore 1An* to the coordinated PB unit generating a short-lived charge-separated state [An˙+–PB˙−] which collapses by back electron-transfer to give 3An*. This electron-transfer step is only possible upon coordination of L1 to the metal centre, which strongly increases the electron acceptor capability of the PB unit, such that 1An* → PB PET is endoergonic in free L1 but exergonic in its complexes. Thus, rather than a conventional set of steps (1An* →
3An* → Ln), the sensitization mechanism now includes 1An* → PB photoinduced electron transfer to generate charge-separated [An˙+–PB˙−], then back electron-transfer to generate 3An* which finally sensitises the Ln(III) centre via energy transfer. The presence of 3An* in L1 and its complexes is confirmed by nanosecond transient absorption studies, which have also shown that the 3An* lifetime in the Nd(III) complex matches the rise time of Nd-centred near-infrared emission, confirming that the final step of the sequence is 3An* → Ln(III) energy-transfer.
Co-reporter:Frazer Kennedy, Nail M. Shavaleev, Thelma Koullourou, Zoë R. Bell, John C. Jeffery, Stephen Faulkner and Michael D. Ward
Dalton Transactions 2007(Issue 15) pp:NaN1499-1499
Publication Date(Web):2007/02/28
DOI:10.1039/B616423D
The luminescent transition metal complexes [Re(CO)3Cl(bppz)] and [Pt(CC–C6H4CF3)2(bppz)] [bppz = 2,3-bis(2-pyridyl)pyrazine], in which one of the diimine binding sites of the potentially bridging ligand bppz is vacant, have been used as ‘complex ligands’ to make heterodinuclear d–f complexes by attachment of a {Ln(dik)3} fragment (dik = a 1,3-diketonate) at the vacant site. When Ln = Pr, Nd, Er or Yb the lanthanide centre has low-energy f–f excited states capable of accepting energy from the 3MLCT excited state of the Pt(II) or Re(I) centre, quenching the 3MLCT luminescence and affording sensitised lanthanide(III)-based luminescence in the near-IR region. UV/Vis and luminescence spectroscopic titrations allowed measurement of (i) the association constants for binding of the {Ln(dik)3} fragment at the vacant diimine site of [Re(CO)3Cl(bppz)] or [Pt(CC–C6H4CF3)2(bppz)], and (ii) the degree of quenching of the 3MLCT luminescence according to the nature of the Ln(III) centre. In all cases Nd(III) was found to be the most effective of the series at quenching the 3MLCT luminescence of the d-block component because the high density of f–f excited states of the appropriate energy make it a particularly effective energy-acceptor.
Co-reporter:Juan Manuel Herrera, Enrique Colacio, Corine Mathonière, Duane Choquesillo-Lazarte and Michael D. Ward
Chemical Communications 2008(Issue 37) pp:NaN4462-4462
Publication Date(Web):2008/07/25
DOI:10.1039/B807364C
Reaction of the anionic cyanometallate chromophore [{Ru(CN)4}3(μ3-HAT)]6− with [MII(tren)]2+ complexes (M = ZnII, CuII) provides discrete tetradecanuclear clusters of formula [{MII(tren)(μ-CN)}11{Ru3(HAT)(CN)}]16+; the weak luminescence of the Ru3 chromophore is substantially enhanced in the presence of ZnII ions, whereas it is completely quenched when CuII centers are present.
Co-reporter:Richard Frantz, Christopher S. Grange, Nawal K. Al-Rasbi, Michael D. Ward and Jérôme Lacour
Chemical Communications 2007(Issue 14) pp:NaN1461-1461
Publication Date(Web):2007/01/30
DOI:10.1039/B618092B
The addition of enantiopure TRISPHAT anions to chiral cationic cages of type [Co4(L)6(BF4)]7+ leads to the enantiodifferentiation of the ligands of the racemic salts and, even more effectively, of the achiral tetrafluoroborate anion trapped inside.
Co-reporter:Jerico R. Piper, Lewis Cletheroe, Christopher G. P. Taylor, Alexander J. Metherell, Julia A. Weinstein, Igor V. Sazanovich and Michael D. Ward
Chemical Communications 2017 - vol. 53(Issue 2) pp:NaN411-411
Publication Date(Web):2016/12/06
DOI:10.1039/C6CC09298E
In a coordination cage which contains an array of twelve naphthyl chromophores surrounding a central cavity, photoinduced energy or electron-transfer can occur from the chromophore array to the bound guest in supramolecular host/guest complexes.
Co-reporter:Martina Whitehead, Simon Turega, Andrew Stephenson, Christopher A. Hunter and Michael D. Ward
Chemical Science (2010-Present) 2013 - vol. 4(Issue 7) pp:NaN2751-2751
Publication Date(Web):2013/04/10
DOI:10.1039/C3SC50546D
A water-soluble cubic coordination cage (Hw) has been prepared, which is isostructural with a previously reported organic-soluble cage (H) apart from the hydroxy groups on the external surface which render it water-soluble. These two cages act as hosts for small organic molecules which bind via a combination of (i) hydrogen-bonding interactions with specific sites on the internal surface of the cages; (ii) non-polar interactions such as aromatic and van der Waals interactions between aromatic rings in the guest and the cage internal surface; and (iii) solvophobic interactions. By comparing ΔG° values for guest binding in water (using Hw) and MeCN (using H), and using pairs of related guests that differ in the presence or absence of an aromatic ring substituent, it is possible to construct thermodynamic cycles that allow quantification of the solvophobic contribution to binding. Specifically, this is the difference between the solvophobic contributions to ΔG° in water and MeCN associated with desolvation of both guest and the internal surface of the cage when complexation occurs. A highly consistent value of ca. −10 kJ mol−1 is determined for this solvophobic contribution to ΔG° associated with the aromatic ring in water compared to MeCN, which correlates very well with what would be expected based on the free energy changes associated with transfer of toluene from MeCN to water. Thus, all three contributions to guest binding listed above can be separately quantified. The ability to prepare related pairs of guests with the presence or absence of a wide range of substituents provides a potentially general way to quantify the solvophobic contributions to guest binding of these substituents.
Co-reporter:Suad T. Saad, Alexander J. Metherell, Elizabeth Baggaley and Michael D. Ward
Dalton Transactions 2016 - vol. 45(Issue 28) pp:NaN11579-11579
Publication Date(Web):2016/06/27
DOI:10.1039/C6DT01614F
A series of dinuclear Ir(III)/Re(I) complexes has been prepared based on a family of symmetrical bridging ligands containing two bidentate N,N′-chelating pyrazolyl–pyridine termini, connected by a central aromatic or aliphatic spacer. The Ir(III) termini are based on {Ir(F2ppy)2}+ units (where F2ppy is the cyclometallating anion of a fluorinated phenylpyridine) and the Re(I) termini are based on {Re(CO)3Cl} units. Both types of terminus are luminescent, with the Ir-based unit showing characteristic strong, structured phosphorescence in the blue region (maximum 452 nm) with a triplet excited state energy of 22200 cm−1 and the Re-based unit showing much weaker and lower-energy phosphorescence (maximum 530 nm) with a triplet excited state energy of 21300 cm−1. The energy gradient between the two excited states allows for partial Ir→Re photoinduced energy-transfer, with substantial (but incomplete) quenching of the higher-energy Ir-based emission component and sensitised emission – evidenced by an obvious grow-in component – on the lower-energy Re-based emission. The Ir→Re energy-transfer rate constants vary over the range 1–8 × 107 s−1 depending on the bridging ligand: there is no simple correlation between bridging ligand structure and energy-transfer rate, possibly because this will depend substantially on the conformation of these flexible molecules in solution. To test the role of ligand conformation further, we investigated a complex in which the bridging chain is a (CH2CH2O)6 unit whose conformation is known to be solvent-polarity dependent, with such chains adopting an open, elongated conformation in water and more compact, folded conformations in organic solvents. There was a clear link between the rate and extent of Ir→Re energy-transfer which reduced in polar solvents as the chain became elongated and the Ir/Re separation was larger; and increased in less polar solvents as the chain adopted a more compact conformation and the Ir/Re separation was reduced.
Co-reporter:Christopher G. P. Taylor, Jerico R. Piper and Michael D. Ward
Chemical Communications 2016 - vol. 52(Issue 37) pp:NaN6228-6228
Publication Date(Web):2016/03/22
DOI:10.1039/C6CC02021F
Cubic coordination cages act as competent hosts for several alkyl phosphonates used as chemical warfare agent simulants; a range of cage/guest structures have been determined, contributions to guest binding analysed, and a fluorescent response to guest binding demonstrated.
Co-reporter:Alexander J. Metherell and Michael D. Ward
Chemical Communications 2014 - vol. 50(Issue 75) pp:NaN10982-10982
Publication Date(Web):2014/07/31
DOI:10.1039/C4CC05421K
The geometrically pure ‘complex ligand’ fac-[Ru(Lph)3]2+, in which three pendant bidentate binding sites are located on one face of the complex, reacts with Ag(I) ions to form the adamantoid decanuclear cage [{Ru(Lph)3}4Ag6](PF6)14 which contains a 6-coordinate Ru(II) ion at each vertex of a large tetrahedron and a 4-coordinate Ag(I) ion along each edge.
Co-reporter:Daniel Sykes, Ahmet J. Cankut, Noorshida Mohd Ali, Andrew Stephenson, Steven J. P. Spall, Simon C. Parker, Julia A. Weinstein and Michael D. Ward
Dalton Transactions 2014 - vol. 43(Issue 17) pp:NaN6428-6428
Publication Date(Web):2014/02/25
DOI:10.1039/C4DT00292J
A series of blue-luminescent Ir(III) complexes with a pendant binding site for lanthanide(III) ions has been synthesized and used to prepare Ir(III)/Ln(III) dyads (Ln = Eu, Tb, Gd). Photophysical studies were used to establish mechanisms of Ir→Ln (Ln = Tb, Eu) energy-transfer. In the Ir/Gd dyads, where direct Ir→Gd energy-transfer is not possible, significant quenching of Ir-based luminescence nonetheless occurred; this can be ascribed to photoinduced electron-transfer from the photo-excited Ir unit (*Ir, 3MLCT/3LC excited state) to the pendant pyrazolyl-pyridine site which becomes a good electron-acceptor when coordinated to an electropositive Gd(III) centre. This electron transfer quenches the Ir-based luminescence, leading to formation of a charge-separated {Ir4+}˙—(pyrazolyl-pyridine)˙− state, which is short-lived possibly due to fast back electron-transfer (<20 ns). In the Ir/Tb and Ir/Eu dyads this electron-transfer pathway is again operative and leads to sensitisation of Eu-based and Tb-based emission using the energy liberated from the back electron-transfer process. In addition direct Dexter-type Ir→Ln (Ln = Tb, Eu) energy-transfer occurs on a similar timescale, meaning that there are two parallel mechanisms by which excitation energy can be transferred from *Ir to the Eu/Tb centre. Time-resolved luminescence measurements on the sensitised Eu-based emission showed both fast and slow rise-time components, associated with the PET-based and Dexter-based energy-transfer mechanisms respectively. In the Ir/Tb dyads, the Ir→Tb energy-transfer is only just thermodynamically favourable, leading to rapid Tb→Ir thermally-activated back energy-transfer and non-radiative deactivation to an extent that depends on the precise energy gap between the *Ir and Tb-based 5D4 states. Thus, the sensitised Tb(III)-based emission is weak and unusually short-lived due to back energy transfer, but nonetheless represents rare examples of Tb(III) sensitisation by a energy donor that could be excited using visible light as opposed to the usually required UV excitation.
Co-reporter:Andrew Stephenson, Daniel Sykes and Michael D. Ward
Dalton Transactions 2013 - vol. 42(Issue 19) pp:NaN6767-6767
Publication Date(Web):2013/03/06
DOI:10.1039/C3DT50161B
The bridging ligand L14Nap, which contains two chelating pyrazolyl-pyridine units separated by a naphthalene-1,4-diyl spacer, has been used in self-assembly of polyhedral coordination cages. The largest such cage is [Cd16(L14Nap)24](BF4)32 which has a tetra-capped truncated tetrahedral Cd16 core with a bridging ligand spanning every edge. The complex is indefinitely stable in dilute solution, which makes it quite different from the previously-reported isostructural cage [Cd16(L14Ph)24](BF4)32 (based on a 1,4-phenyl bridge) that forms on crystallisation but slowly rearranges to smaller cages in solution. The additional inter-ligand π-stacking between ligand fragments associated with replacement of a phenyl group by a naphthyl group allows the complex to be stable in solution, providing conclusive proof of the importance of inter-ligand π-stacking in the assembly of these cages. With Cu(II) in place of Cd(II) a smaller cage [Cu12(L14Nap)15](ClO4)24 was formed which contains a mixture of tris-chelated (six-coordinate) and bis-chelated (four-coordinate, or five-coordinate if an additional monodentate ligand is present) Cu(II) ions; the difference between the two structures arises in part from the different stereoelectronic preferences of the two metal ions. Despite this difference both the Cd16 and Cu12 cages contain {M3(L14Nap)3}6+ triangular helical units as subcomponents which form the triangular faces of the polyhedra. By using a 1:1 ligand:metal ratio in the synthesis examples of these can be isolated and characterised; the structures of the trinuclear cyclic helicates [Cd3(L14Nap)3(BF4)4(EtOAc)2](BF4)2 and [Cu3(L14Nap)3(BF4)(MeCN)2](BF4)5 have also been determined.
Co-reporter:Noorshida Mohd Ali, Voirrey L. MacLeod, Petter Jennison, Igor V. Sazanovich, Christopher A. Hunter, Julia A. Weinstein and Michael D. Ward
Dalton Transactions 2012 - vol. 41(Issue 8) pp:NaN2419-2419
Publication Date(Web):2012/01/03
DOI:10.1039/C1DT11328C
[Ir(ppy)2(CN)2]− (ppy = anion of 2-phenylpyridine) and some substituted derivatives have been investigated for their ability to interact with additional metal cations, both in solution and the solid state, via the externally-directed cyanide lone pairs, and to act as energy-donors in the resulting assemblies. [Ir(ppy)2(CN)2]− is slightly solvatochromic, showing a blue-shift of the lowest energy absorption manifold in water compared to organic solvents, and the solubilised tBu-substituted analogue [Ir(tBuppy)2(CN)2]− [tBuppy = anion of 2-(4-tBu-phenyl)pyridine] is also metallochromic with coordination of the cyanide lone pairs to two M(II) cations in MeCN (M = Ba, Zn) resulting in blue-shifts of the lowest-energy absorption and emission maxima. These effects are however modest because of (i) the presence of only two cyanide groups, and (ii) the fact that the lowest-energy excited state has a substantial 3LC component and is therefore not purely charge-transfer in nature. Crystallisation of [Ir(ppy)2(CN)2]− as its (PPN)+ salt in the presence of excess of lanthanide(III) salts leads to formation of assemblies based on Ir–CN–Ln bonds, which generate in the solid state either Ir2Ln2(μ-CN)4 square assemblies or linear trinuclear species with Ir–CN–Ln–NC–Ir cores. In the Ir2Eu2(μ-CN)4 and Ir2Nd2(μ-CN)4 complexes the Ir-based emission is substantially quenched due to energy-transfer to lower-lying f–f states of these lanthanide ions. In addition reaction of [Ir(F2ppy)2(CN)2]− [F2ppy = cyclometallating anion of 2-(2,4-difluorophenyl)pyridine] with [Re(phen)(CO)3(MeCN)][PF6] in solution affords dinuclear IrRe and trinuclear IrRe2 species in which {Re(phen)(CO)3} units are attached to the N-donor termini of one or both of the cyanide groups; these complexes have been structurally characterised and display quantitative Ir→Re energy-transfer, showing luminescence only from the Re(I) terminus on excitation of the Ir(III) unit.
Co-reporter:Andrew Stephenson and Michael D. Ward
Dalton Transactions 2011 - vol. 40(Issue 40) pp:NaN10369-10369
Publication Date(Web):2011/04/26
DOI:10.1039/C1DT10263J
The two new ligands Lfur and Lth consist of two chelating pyrazolyl-pyridine termini connected to furan-2,5-diyl or thiophene-2,5-diyl spacers viamethylene groups. Reaction of these with a range of transition metal dications that prefer octahedral coordination affords a series of unusual structures which are all based on a 2M:3L ratio. [M8(Lfur)12]X16 (M = Co, Cu, X = BF4; and M = Zn, X = ClO4) are octanuclear cubes with approximate D4 symmetry in which two cyclic tetranuclear helicate M4L4 units are connected by four additional ‘pillar’ ligands. In contrast [Ni4(Lfur)6](BF4)8 is a centrosymmetric molecular square consisting of two dinuclear Ni2L2 units of opposite chirality that are connected by a pair of additional Lfur ligands such that the four edges of the Ni4 square are spanned by alternately two and one bridging ligands. [M4(Lth)6](BF4)8 (M = Co, Ni, Cu) are likewise molecular squares with similar structures to [Ni4(Lfur)6](BF4)8 with the significant difference that the two crosslinked double helicate M2L2 units are now homochiral. The Cd(II) complexes both behave quite differently to the first-row metal complexes, with [Cd(Lfur)(BF4)](BF4) being a simple mononuclear complex with a single ligand in which the furan oxygen atom is weakly interacting with the Cd(II) centre. In contrast, in {[Cd2(Lth)3](BF4)4}∞, where this quasi-pentadentate coordination mode of the ligand is not possible because thiophene is too poor an electron donor, the ligand reverts to bis-bidentate bridging coordination to afford a one-dimensional chain consisting of an infinite sequence of crosslinked, homochiral, Cd2(Lth)2 double helicate units.
Co-reporter:Michael D. Ward and Paul R. Raithby
Chemical Society Reviews 2013 - vol. 42(Issue 4) pp:NaN1636-1636
Publication Date(Web):2012/07/13
DOI:10.1039/C2CS35123D
Many naturally occurring systems show us how multi-component supramolecular assemblies can generate useful functional behaviour. In this article the problems and limitations associated with achieving such behaviour in artificial multi-component assemblies is discussed, together with two examples of functions in artificial supramolecular assemblies based on (i) host–guest chemistry in cavities of cages, and (ii) light-harvesting in multi-chromophore arrays. Important challenges for the future are summarised.
Co-reporter:Michael D. Ward
Chemical Communications 2009(Issue 30) pp:
Publication Date(Web):
DOI:10.1039/B906726B
Co-reporter:Alexander J. Metherell and Michael D. Ward
Chemical Science (2010-Present) 2016 - vol. 7(Issue 2) pp:NaN915-915
Publication Date(Web):2015/10/14
DOI:10.1039/C5SC03526K
Retrosynthetic analysis of a [M16L24]32+ coordination cage shows how it can be assembled rationally, in a stepwise manner, using a combination of kinetically inert and kinetically labile components. Combination of the components of fac-[Ru(Lph)3](PF6)2, Cd(BF4)2 and Lnaph in the necessary 4:12:12 stoichiometry afforded crystals of [Ru4Cd12(Lph)12(Lnaph)12]X32 (X = a mono-anion) in which the location of the two types of metal ion [Ru(II) or Cd(II)] at specific vertices in the metal-ion array, and the two types of bridging ligand (Lph and Lnaph) along specific edges, is completely controlled by the synthetic strategy. The incorporation of four different types of component at pre-determined positions in a coordination cage superstructure represents a substantial advance in imposing control on the self-assembly of complex metallosupramolecular entities.
Co-reporter:Ashley B. Wragg, Alexander J. Metherell, William Cullen and Michael D. Ward
Dalton Transactions 2015 - vol. 44(Issue 41) pp:NaN17949-17949
Publication Date(Web):2015/09/21
DOI:10.1039/C5DT02957K
Stepwise preparation of the heterometallic octanuclear coordination cages [(Ma)4(Mb)4L12]16+ is reported, in which Ma = Ru or Os and Mb = Cd or Co (all in their +2 oxidation state). This requires initial preparation of the kinetically inert mononuclear complexes [(Ma)L3]2+ in which L is a ditopic ligand with two bidentate chelating pyrazolyl-pyridine units: in the complexes [(Ma)L3]2+ one terminus of each ligand is bound to the metal ion, such that the complex has three pendant bidentate sites at which cage assembly can propagate by coordination to additional labile ions Mb in a separate step. Thus, combination of four [(Ma)L3]2+ units and four [Mb]2+ ions results in assembly of the complete cages [(Ma)4(Mb)4L12]16+ in which a metal ion lies at each of the eight vertices, and a bridging ligand spans each of the twelve edges, of a cube. The different types of metal ion necessarily alternate around the periphery with each bridging ligand bound to one metal ion of each type. All four cages have been structurally characterised: in the Ru(II)/Cd(II) cage (reported in a recent communication) the Ru(II) and Cd(II) ions are crystallographically distinct; in the other three cages [Ru(II)/Co(II), Os(II)/Cd(II) and Os(II)/Co(II), reported here] the ions are disordered around the periphery such that every metal site refines as a 50:50 mixture of the two metal atom types. The incorporation of Os(II) units into the cages results in both redox activity [a reversible Os(II)/Os(III) couple for all four metal ions simultaneously, at a modest potential] and luminescence [the Os(II) units have luminescent 3MLCT excited states which will be good photo-electron donors] being incorporated into the cage superstructure.
Co-reporter:Alexander J. Metherell, William Cullen, Andrew Stephenson, Christopher A. Hunter and Michael D. Ward
Dalton Transactions 2014 - vol. 43(Issue 1) pp:NaN84-84
Publication Date(Web):2013/10/16
DOI:10.1039/C3DT52479E
We have prepared a series of mononuclear fac and mer isomers of Ru(II) complexes containing chelating pyrazolyl-pyridine ligands, to examine their differing ability to act as hydrogen-bond donors in MeCN. This was prompted by our earlier observation that octanuclear cube-like coordination cages that contain these types of metal vertex can bind guests such as isoquinoline-N-oxide (K = 2100 M−1 in MeCN), with a significant contribution to binding being a hydrogen-bonding interaction between the electron-rich atom of the guest and a hydrogen-bond donor site on the internal surface of the cage formed by a convergent set of CH2 protons close to a 2+ metal centre. Starting with [Ru(LH)3]2+ [LH = 3-(2-pyridyl)-1H-pyrazole] the geometric isomers were separated by virtue of the fact that the fac isomer forms a Cu(I) adduct which the mer isomer does not. Alkylation of the pyrazolyl NH group with methyl iodide or benzyl bromide afforded [Ru(LMe)3]2+ and [Ru(Lbz)3]2+ respectively, each as their fac and mer isomers; all were structurally characterised. In the fac isomers the convergent group of pendant –CH2R or –CH3 protons defines a hydrogen-bond donor pocket; in the mer isomer these protons do not converge and any hydrogen-bonding involving these protons is expected to be weaker. For both [Ru(LMe)3]2+ and [Ru(Lbz)3]2+, NMR titrations with isoquinoline-N-oxide in MeCN revealed weak 1:1 binding (K ≈ 1 M−1) between the guest and the fac isomer of the complex that was absent with the mer isomer, confirming a difference in the hydrogen-bond donor capabilities of these complexes associated with their differing geometries. The weak binding compared to the cage however occurs because of competition from the anions, which are free to form ion-pairs with the mononuclear complex cations in a way that does not happen in the cage complexes. We conclude that (i) the presence of fac tris-chelate sites in the cage to act as hydrogen-bond donors, and (ii) exclusion of counter-ions from the central cavity leaving these hydrogen-bonding sites free to interact with guests, are both important design criteria for future coordination cage hosts.
Co-reporter:Atanu Jana, Elizabeth Baggaley, Angelo Amoroso and Michael D. Ward
Chemical Communications 2015 - vol. 51(Issue 42) pp:NaN8836-8836
Publication Date(Web):2015/04/21
DOI:10.1039/C5CC02130H
A new rigid and conjugated ligand structure connecting phenanthroline and poly(amino-carboxylate) binding sites provides d–f complexes which show high potential for use in dual (luminescence + magnetic resonance) imaging and for optimisation of d → f photoinduced energy-transfer.
Co-reporter:Benjamin R. Hall, Lauren E. Manck, Ian S. Tidmarsh, Andrew Stephenson, Brian F. Taylor, Emma J. Blaikie, Douglas A. Vander Griend and Michael D. Ward
Dalton Transactions 2011 - vol. 40(Issue 45) pp:NaN12145-12145
Publication Date(Web):2011/08/10
DOI:10.1039/C1DT10781J
The ligand Lbip, containing two bidentate pyrazolyl–pyridine termini separated by a 3,3′-biphenyl spacer, has been used to prepare tetrahedral cage complexes of the form [M4(Lbip)6]X8, in which a bridging ligand spans each of the six edges of the M4 tetrahedron. Several new examples have been structurally characterized with a variety of metal cation and different anions in order to examine interactions between the cationic cage and various anions. Small anions such as BF4− and NO3− can occupy the central cavity where they are anchored by an array of CH⋯F or CH⋯O hydrogen-bonding interactions with the interior surface of the cage, but larger anions such as naphthyl-1-sulfonate or tetraphenylborate lie outside the cavity and interact with the external surface of the cage via CH⋯π interactions or CH⋯O hydrogen bonds. The cages with M = Co and M = Cd have been examined in detail by NMR spectroscopy. For [Co4(Lbip)6](BF4)8 the 1H NMR spectrum is paramagnetically shifted over the range −85 to +110 ppm, but the spectrum has been completely assigned by correlation of measured T1 relaxation times of each peak with Co⋯H distances. 19F DOSY measurements on the anions show that at low temperature a [BF4] − anion diffuses at a similar rate to the cage superstructure surrounding it, indicating that it is trapped inside the central cage cavity. Furthermore, the equilibrium step-by-step self-assembly of the cage superstructure has been elucidated by detailed modeling of spectroscopic titrations at multiple temperatures of an acetonitrile solution of Lbip into an acetonitrile solution of Co(BF4)2. Six species have been identified: [Co2Lbip]4+, [Co2(Lbip)2]4+, [Co4(Lbip)6]8+, [Co4(Lbip)8]8+, [Co2(Lbip)5]4+, and [Co(Lbip)3]2+. Overall the assembly of the cage is entropy, and not enthalpy, driven. Once assembled, the cages show remarkable kinetic inertness due to their mechanically entangled nature: scrambling of metal cations between the sites of pure Co4 and Cd4 cages to give a statistical mixture of Co4, Co3Cd, Co2Cd2, CoCd3 and Cd4 cages takes months in solution at room temperature.
Co-reporter:Michael D. Ward
Dalton Transactions 2010 - vol. 39(Issue 38) pp:NaN8867-8867
Publication Date(Web):2010/07/08
DOI:10.1039/C0DT00312C
Luminescent complexes of the [M(diimine)(CN)4]2− family (M = Ru, Os), and their polynuclear analogues, are structurally versatile components for preparation of supramolecular assemblies based on interaction of the cyanide groups with other metal ions or metal complexes via direct coordination, hydrogen bonding, or halogen bonding. In addition their environment-dependent photophysical properties (solvatochromism and metallochromism), and the ability of the CN groups to act as reporters for excited state behaviour via time-resolved IR spectroscopy, make these fragments spectroscopically as well as structurally versatile. This Perspective article summarises work from the author's group over the last decade on the structures and photophysical properties of these fascinating complexes and their supramolecular assemblies.
Co-reporter:Christopher S. Grange, Anthony J. H. M. Meijer and Michael D. Ward
Dalton Transactions 2010 - vol. 39(Issue 1) pp:NaN211-211
Publication Date(Web):2009/11/20
DOI:10.1039/B918086A
The trinuclear complexes [{(R2bipy)2Ru}3(µ3-HHTP)](PF6)3 [1(PF6)3, R = H; 2(PF6)3, R = 4-tBu] contain three {Ru(R2bipy)2}2+ fragments connected to the triangular tris-chelating ligand hexahydroxytriphenylene (H6HHTP). This bridging ligand contains three dioxolene-type binding sites, each of which can reversibly convert between dianionic catecholate (cat), monoanionic semiquinone (sq) or neutral quinone (q) redox states. The bridging ligand as a whole can therefore exist in seven different redox states from fully reduced [cat,cat,cat]6− through to fully oxidised, neutral [q,q,q]. Cyclic voltammetry of 1(PF6)3 in MeCN reveals six redox processes of which the three at more positive potentials (the sq/q couples) are reversible but the three at more negative potentials (the sq/cat couples) are irreversible with distorted wave shapes due to the insolubility of the reduced forms of the complex. In contrast, the more soluble complex 2(PF6)3 displays six reversible one-electron redox processes making all components of a seven-membered redox chain accessible. UV/Vis/NIR spectro-electrochemical studies reveal rich spectroscopic behaviour, with—in particular—very intense transitions in the near-IR region in many of the oxidation states associated with Ru(II)→(dioxolene) MLCT and bridging ligand centred π–π* transitions. TDDFT calculations were used to analyse the electronic spectra in all seven oxidation states; the calculated spectra generally show very good agreement with experiment, which has allowed a fairly complete assignment of the low-energy transitions. The strong electrochromism of the complexes in the near-IR region has formed the basis of an ‘optical window’ in which a thin film of 1(PF6)3 or 2(PF6)3 on a conductive glass surface can be reversibly and rapidly switched between redox states that alternate between strongly absorbing or near-transparent at 1100 nm, with—for 2(PF6)3—the switching being stable and reversible in water over thousands of cycles.
Co-reporter:Alexander H. Shelton, Igor V. Sazanovich, Julia A. Weinstein and Michael D. Ward
Chemical Communications 2012 - vol. 48(Issue 22) pp:NaN2751-2751
Publication Date(Web):2012/01/04
DOI:10.1039/C2CC17182A
A macrocycle-appended naphthalimide derivative and its Eu(III) complex show triple luminescence from isolated naphthalimide (blue), aggregated naphthalimide excimers (green) and Eu centres (red) with the balance being sensitive to the degree of aggregation, allowing white light emission to be obtained from a single molecule.
Co-reporter:Robert M. Edkins, Daniel Sykes, Andrew Beeby and Michael D. Ward
Chemical Communications 2012 - vol. 48(Issue 80) pp:NaN9979-9979
Publication Date(Web):2012/05/18
DOI:10.1039/C2CC33005A
In a pair of Ir/Eu and Ir/Tb dyads, two-photon excitation of the Ir-phenylpyridine chromophore at 780 nm is followed by partial d → f energy-transfer to give a combination of short-lived Ir-based (blue) and long-lived lanthanide-based (red or green) emission; these components can be selected separately by time-gated detection.
Co-reporter:William Cullen, Simon Turega, Christopher A. Hunter and Michael D. Ward
Chemical Science (2010-Present) 2015 - vol. 6(Issue 5) pp:NaN2794-2794
Publication Date(Web):2015/03/10
DOI:10.1039/C5SC00534E
The protein/ligand docking software GOLD, which was originally developed for drug discovery, has been used in a virtual screen to identify small molecules that bind with extremely high affinities (K ≈ 107 M−1) in the cavity of a cubic coordination cage in water. A scoring function was developed using known guests as a training set and modified by introducing an additional term to take account of loss of guest flexibility on binding. This scoring function was then used in GOLD to successfully identify 15 new guests and accurately predict the binding constants. This approach provides a powerful predictive tool for virtual screening of large compound libraries to identify new guests for synthetic hosts, thereby greatly simplifying and accelerating the process of identifying guests by removing the reliance on experimental trial-and-error.
Co-reporter:William Cullen, Simon Turega, Christopher A. Hunter and Michael D. Ward
Chemical Science (2010-Present) 2015 - vol. 6(Issue 1) pp:NaN631-631
Publication Date(Web):2014/07/31
DOI:10.1039/C4SC02090A
A range of organic molecules with acidic or basic groups exhibit strong pH-dependent binding inside the cavity of a polyhedral coordination cage. Guest binding in aqueous solution is dominated by a hydrophobic contribution which is compensated by stronger solvation when the guests become cationic (by protonation) or anionic (by deprotonation). The Parkinson's drug 1-amino-adamantane (‘amantadine’) binds with an association constant of 104 M−1 in the neutral form (pH greater than 11), but the stability of the complex is reduced by three orders of magnitude when the guest is protonated at lower pH. Monitoring the uptake of the guests into the cage cavity was facilitated by the large upfield shift for the 1H NMR signals of bound guests due to the paramagnetism of the host. Although the association constants are generally lower, guests of biological significance such as aspirin and nicotine show similar behaviour, with a substantial difference between neutral (strongly binding) and charged (weakly binding) forms, irrespective of the sign of the charged species. pH-dependent binding was observed for a range of guests with different functional groups (primary and tertiary amines, pyridine, imidazole and carboxylic acids), so that the pH-swing can be tuned anywhere in the range of 3.5–11. The structure of the adamantane-1-carboxylic acid complex was determined by X-ray crystallography: the oxygen atoms of the guest form CH⋯O hydrogen bonds with one of two equivalent pockets on the internal surface of the host. Reversible uptake and release of guests as a function of pH offers interesting possibilities in any application where controlled release of a molecule following an external stimulus is required.
Co-reporter:Theodore Lazarides, Harry Adams, Daniel Sykes, Stephen Faulkner, Giuseppe Calogero and Michael D. Ward
Dalton Transactions 2008(Issue 5) pp:NaN698-698
Publication Date(Web):2007/11/23
DOI:10.1039/B714640J
The complexes [Ru(tBu2bipy)(bpym)X2] (X = Cl, NCS) and [M(tBu2bipy)2(bpym)][PF6]2 (M = Ru, Os) all have a low-energy LUMO arising from the presence of a 2,2′-bipyrimidine ligand, and consequently have lower-energy 1MLCT and 3MLCT states than analogous complexes of bipyridine. The vacant site of the bpym ligand provides a site at which {Ln(diketonate)3} units can bind to afford bipyrimidine-bridged dinuclear Ru–Ln and Os–Ln dyads; four such complexes have been structurally characterised. UV/Vis and luminescence spectroscopic studies show that binding of the Ln(III) fragment at the second site of the bpym ligand reduces the 3MLCT energy of the Ru or Os fragment still further. The result is that in the dyads [Ru(tBu2bipy)X2(µ-bpym)Ln(diketonate)3] (X = Cl, NCS) and [Os(tBu2bipy)2(µ-bpym)Ln(diketonate)3][PF6]2 the 3MLCT is too low to sensitise the luminescent f-f states of Nd(III) and Yb(III), but in [Ru(tBu2bipy)2(µ-bpym)Ln(diketonate)3][PF6]2 the 3MLCT energy of 13500 cm−1 permits energy transfer to Yb(III) and Nd(III) resulting in sensitised near-infrared luminescence on the microsecond timescale.
Co-reporter:Andrew Stephenson and Michael D. Ward
Dalton Transactions 2011 - vol. 40(Issue 31) pp:NaN7826-7826
Publication Date(Web):2011/02/14
DOI:10.1039/C0DT01767A
The octanuclear coordination cage [Ni8(L14Naph)12](BF4)16 has the core structure of a ‘cuneane’ - a toplogical isomer of a cube - with a metal ion at each of the eight vertices and bridging ligand spanning each of the twelve edges; this is the only possible 8-vertex polyhedron other than a cube that will form a cage in which each metal is connected to three others.
Co-reporter:Theodore Lazarides, Nicholas M. Tart, Daniel Sykes, Stephen Faulkner, Andrea Barbieri and Michael D. Ward
Dalton Transactions 2009(Issue 20) pp:NaN3979-3979
Publication Date(Web):2009/03/10
DOI:10.1039/B901560D
The complexes RuL and OsL contain [M(bipy)3]2+ chromophores with a pendant aza-18-crown-6 macrocycle for binding of lanthanide(III) ions. The photophysical properties of the adducts RuL·Ln and OsL·Ln, prepared by addition of excess Ln(NO3)3 (Ln = Nd, Yb) to solutions of RuL and OsL in MeCN, were examined using time-resolved and steady-state luminescence methods. Whereas RuL does not act as an energy-donor to Yb(III), it will transfer energy to (and generate sensitised near-infrared luminescence from) Nd(III) with a Ru(II)→Nd(III) energy-transfer rate constant of 6.8 × 106 s−1. In contrast, OsL is quenched by both Yb(III) and Nd(III), but with faster energy-transfer to Yb(III) (2.6 × 107 s−1) than to Nd(III) (1.4 × 107 s−1). Thus d → f energy transfer is in both cases faster for Os(II) than for Ru(II), but the relative ability of Nd(III) and Yb(III) to act as energy-acceptors is inverted from RuL·Ln to OsL·Ln. Reasons for this are discussed with reference to contributions from the Förster and Dexter mechanism for energy-transfer in RuL·Nd and OsL·Nd, using calculated spectroscopic overlap integrals coupled with molecular modelling to estimate inter-chromophore separations. The particular effectiveness of Os(II) → Yb(III) energy-transfer in OsL·Yb is explained in terms of the Horrocks redox mechanism involving an initial *Os(II) → Yb(III) photoinduced electron transfer step generating an Os(III)/Yb(II) state, which is shown to be marginally favourable for OsL·Yb, but not for RuL·Yb in which the [Ru(bipy)3]2+ unit is a poorer excited-state electron-donor by about 0.1 eV.
Co-reporter:Alexander J. Metherell and Michael D. Ward
Dalton Transactions 2016 - vol. 45(Issue 41) pp:NaN16111-16111
Publication Date(Web):2016/09/01
DOI:10.1039/C6DT03041F
This ‘Perspective’ article summarises recent work from the authors’ research group on the exploitation of the simple fac/mer geometric isomerism of octahedral metal tris-chelates as a tool to control the chemistry of coordination cages based on bis(pyrazolyl-pyridine) ligands, in two different respects. Firstly this geometric isomerism plays a major role in controlling the guest binding properties of cages because a fac tris-chelate arrangement of pyrazolyl-pyridine chelates around a metal ion vertex results in formation of a convergent set of inwardly-directed C–H protons in a region of high positive electrostatic potential close to a metal cation. This collection of δ+ protons therefore provides a charge-assisted hydrogen-bond donor site, which interacts with the electron-rich regions of guest molecules that are of the correct size and shape to occupy the cage cavity, and the strength of this hydrogen-bonding interaction plays a major role in guest recognition in non-aqueous solvents. Secondly the ability to prepare mononuclear complexes with either a fac or mer arrangement of ligands provides an entry into the controlled, stepwise assembly of heterometallic cages based on a combination of kinetically inert and kinetically labile metal ions at different sites. This has allowed introduction of useful physical properties such as redox activity or luminescence, commonly associated with inert metal ions which are not amenable to participation in thermodynamic self-assembly processes, to be incorporated in a predictable way into the superstructures of coordination cages at specific sites.
Co-reporter:James S. Wright, Alexander J. Metherell, William M. Cullen, Jerico R. Piper, Robert Dawson and Michael D. Ward
Chemical Communications 2017 - vol. 53(Issue 31) pp:NaN4401-4401
Publication Date(Web):2017/03/31
DOI:10.1039/C7CC01959A
Two M8L12 cubic coordination cages, as desolvated crystalline powders, preferentially adsorb CO2 over N2 with ideal selectivity CO2/N2 constants of 49 and 30 at 298 K. A binding site for CO2 is suggested by crystallographic location of CS2 within the cage cavity at an electropositive hydrogen-bond donor site, potentially explaining the high CO2/N2 selectivity compared to other materials with this level of porosity.
Co-reporter:Tanya K. Ronson, Harry Adams, Lindsay P. Harding, Simon J. A. Pope, Daniel Sykes, Stephen Faulkner and Michael D. Ward
Dalton Transactions 2007(Issue 10) pp:NaN1022-1022
Publication Date(Web):2007/02/01
DOI:10.1039/B618258E
A set of three potentially bridging ligands containing two tridentate chelating N,N′,O-donor (pyrazole–pyridine–amide) donors separated by an o, m, or p-phenylene spacer has been prepared and their coordination chemistry with lanthanide(III) ions investigated. Ligand L1 (p-phenylene spacer) forms complexes with a 2 : 3 M : L ratio according to the proportions used in the reaction mixture; the Ln2(L1)3 complexes contain two 9-coordinate Ln(III) centres with all three bridging ligands spanning both metal ions, and have a cylindrical (non-helical) ‘mesocate’ architecture. The 1 : 1 complexes display a range of structural types depending on the conditions used, including a cyclic Ln4(L1)4 tetranuclear helicate, a Ln2(L1)2 dinuclear mesocate, and an infinite one-dimensional coordination polymer in which metal ions and bridging ligands alternate along the sequence. ESMS studies indicate that the 1 : 1 complexes form a mixture of oligonuclear species {Ln(L1)}n in solution (n up to 5) which are likely to be cyclic helicates. In contrast, ligands L2 and L3 (with o- and m-phenylene spacers, respectively) generally form dinuclear Ln2L2 Ln(III) complexes in which the two ligands may be arranged in a helical or non-helical architecture about the two metal ions. These complexes also contain an additional exogenous bidentate bridging ligand, either acetate or formate, which has arisen from hydrolysis of solvent molecules promoted by the Lewis-acidity of the Ln(III) ions. Luminescence studies on some of the Nd(III) complexes showed that excitation into ligand-centred π–π* transitions result in the characteristic near-infrared luminescence from Nd(III) at 1060 nm.
Co-reporter:Hazel Fenton, Ian S. Tidmarsh and Michael D. Ward
Dalton Transactions 2010 - vol. 39(Issue 16) pp:NaN3815-3815
Publication Date(Web):2010/03/05
DOI:10.1039/B926264D
Two ligands L1 and L2 have been prepared which contain a chelating pyrazolyl-pyridine group with a pendant aromatic nitrile (in L1, a benzonitrile; in L2, a naphthonitrile). These ligands react with Ag(I) salts to give a range of infinite coordination networks or dimeric ‘boxes’ in which the pyrazolyl-pyridine chelates and the aromatic nitrile groups both participate in coordination to Ag(I) ions. In contrast, L1 and L2 form simple mononuclear tris-chelates [ML3]2+ with first-row transition metal dications (M = Co, Ni, Zn) in which the aromatic nitrile groups are pendant such that the complexes can be used as ‘complex ligands’. The crystal structures of [M(L2)3](BF4)2 are based on solely the mer tris-chelate geometry although in solution 1H NMR spectroscopy reveals a mixture of both fac and mer isomers of the tris-chelates. Reaction of these with Ag(I) ions allows the interaction of the pendant nitrile groups with Ag(I) ions to generate coordination networks based on [ML3]2+ cations being crosslinked by Ag(I) ions. In these networks the [ML3]2+ cations have solely the fac geometry and lie on threefold rotation axes with all three pendant nitrile groups coordinated to Ag(I) ions which are three-coordinate. {[AgM(L2)3][BF4]3}∞ (M = Co, Ni) consist of two interpenetrated (10,3)a nets which have opposite chirality at the [M(L2)3]2+ centres but are not strictly enantiomorphic as the two nets are not crystallographically equivalent. {[AgNi(L1)3](BF4)3}∞ in contrast contains two-dimensional sheets which have a (6,3) net structure of hexagonal rings of alternative Ni(II) and Ag(I) centres; although not interpenetrating, two such adjacent (and enantiomorphic) sheets interact with each other via numerous CH⋯π interactions between aromatic ligands. Formation of these structures shows that the differential reactivity of the two binding sites in L1 and L2 (pyrazolyl-pyridine, and nitrile) can be used to generate mixed-metal coordination networks in a hierarchical, stepwise manner.
Co-reporter:William Cullen, Katie A. Thomas, Christopher A. Hunter and Michael D. Ward
Chemical Science (2010-Present) 2015 - vol. 6(Issue 7) pp:NaN4028-4028
Publication Date(Web):2015/05/07
DOI:10.1039/C5SC01475A
We demonstrate the use of a simple pH swing to control the selection of one of three different guests from aqueous solution by a coordination cage host. Switching between different guests is based on the fact that neutral organic guests bind strongly in the cage due to the hydrophobic effect, but for acidic or basic guests, the charged (protonated or deprotonated) forms are hydrophilic and do not bind. The guests used are adamantane-1,3-dicarboxylic acid (H2A) which binds at low pH when it is neutral but not when it is deprotonated; 1-amino-adamantane (B) which binds at high pH when it is neutral but not when it is protonated; and cyclononanone (C) whose binding is not pH dependent and is therefore the default guest at neutral pH. Thus an increase in pH can reversibly switch the host between the three different bound states cage·H2A (at low pH), cage·C (at medium pH), and cage·B (at high pH) in succession.
Co-reporter:Ashley B. Wragg, Sofia Derossi, Timothy L. Easun, Michael W. George, Xue-Zhong Sun, František Hartl, Alexander H. Shelton, Anthony J. H. M. Meijer and Michael D. Ward
Dalton Transactions 2012 - vol. 41(Issue 34) pp:NaN10371-10371
Publication Date(Web):2012/06/21
DOI:10.1039/C2DT31001E
The dinuclear complex [{Ru(CN)4}2(μ-bppz)]4− shows a strongly solvent-dependent metal–metal electronic interaction which allows the mixed-valence state to be switched from class 2 to class 3 by changing solvent from water to CH2Cl2. In CH2Cl2 the separation between the successive Ru(II)/Ru(III) redox couples is 350 mV and the IVCT band (from the UV/Vis/NIR spectroelectrochemistry) is characteristic of a borderline class II/III or class III mixed valence state. In water, the redox separation is only 110 mV and the much broader IVCT transition is characteristic of a class II mixed-valence state. This is consistent with the observation that raising and lowering the energy of the d(π) orbitals in CH2Cl2 or water, respectively, will decrease or increase the energy gap to the LUMO of the bppz bridging ligand, which provides the delocalisation pathway via electron-transfer. IR spectroelectrochemistry could only be carried out successfully in CH2Cl2 and revealed class III mixed-valence behaviour on the fast IR timescale. In contrast to this, time-resolved IR spectroscopy showed that the MLCT excited state, which is formulated as RuIII(bppz˙−)RuII and can therefore be considered as a mixed-valence Ru(II)/Ru(III) complex with an intermediate bridging radical anion ligand, is localised on the IR timescale with spectroscopically distinct Ru(II) and Ru(III) termini. This is because the necessary electron-transfer via the bppz ligand is more difficult because of the additional electron on bppz˙− which raises the orbital through which electron exchange occurs in energy. DFT calculations reproduce the electronic spectra of the complex in all three Ru(II)/Ru(II), Ru(II)/Ru(III) and Ru(III)/Ru(III) calculations in both water and CH2Cl2 well as long as an explicit allowance is made for the presence of water molecules hydrogen-bonded to the cyanides in the model used. They also reproduce the excited-state IR spectra of both [Ru(CN)4(μ-bppz)]2– and [{Ru(CN)4}2(μ-bppz)]4− very well in both solvents. The reorganization of the water solvent shell indicates a possible dynamical reason for the longer life time of the triplet state in water compared to CH2Cl2.
Co-reporter:Sofia Derossi, Harry Adams and Michael D. Ward
Dalton Transactions 2007(Issue 1) pp:NaN36-36
Publication Date(Web):2006/11/17
DOI:10.1039/B614346F
The complex cations [RuL2(H2biim)]2+ (L = bipy, 4,4′-tBu2-bipy) interact with cyanometallate anions via a chelating hydrogen-bonding interaction between the two N–H donors of the complex cation and the N lone pair of one cyanide ligand in the complex anion; the anion hexacyanoferrate(III) quenches the Ru(II)-based luminescence in CH2Cl2 solution by photoinduced electron-transfer within the H-bonded assembly, whereas hexacyanocobaltate(III) enhances the Ru(II)-based luminescence.
Co-reporter:Svetlana G. Baca, Harry Adams, Daniel Sykes, Stephen Faulkner and Michael D. Ward
Dalton Transactions 2007(Issue 23) pp:NaN2430-2430
Publication Date(Web):2007/04/05
DOI:10.1039/B702235B
A series of cyanide-bridged coordination networks has been prepared which contain [Ru(phen)(CN)4]2− anions, Ln(III) cations, and additional oligopyridine ligands (1,10-phenanthroline, 2,2′:6′,2‴-terpyridine or 2,2′-bipyrimidine) which coordinate to the Ln(III) centres. Five structural types have been identified and examples of each type of structure are described: these are hexanuclear Ru4Ln2 clusters; two-dimensional Ru–Ln sheets with a honeycomb pattern of edge-linked Ru6Ln6 hexagons; one-dimensional chains consisting of two parallel cross-linked strands in a ladder-like arrangement; simple single-stranded chains of alternating Ru/Ln components; and a one-dimensional ‘chain of squares’ in which Ru2Ln2 squares are linked by bipyrimidine bridging ligands which connect to the Ln(III) corners of adjacent squares in the sequence. The 3MLCT luminescence characteristic of the [Ru(phen)(CN)4]2− units is quenched in those networks containing Ln(III) which have low-lying near-infrared luminescent f–f states [Pr(III), Nd(III), Er(III), Yb(III)], with sensitised Ln(III)-based near-IR luminescence generated by d → f energy-transfer. The rate of d → f energy-transfer, and hence the degree of quenching of the 3MLCT luminescence from the [Ru(phen)(CN)4]2− units, depends on the availability of f–f levels of an appropriate energy on the Ln(III) centre, with Nd(III) (with a high density of low-lying f–f states) being the most effective energy-acceptor and Yb(III) (with a single low-lying f–f state) being the least effective. Rates of d → f energy-transfer to different Ln(III) centres could be determined from both the residual (partially quenched) lifetimes of the 3MLCT luminescence, and—in the case of the Yb(III) networks—by a rise-time for the sensitised near-IR luminescence. The presence of the ‘blocking’ polypyridyl ligands, which reduced the number of cyanide and water ligands that would otherwise coordinate to the Ln(III) centres, resulted in increases in the Ln(III)-based emission lifetimes compared to networks where these blocking ligands were not used.
Co-reporter:Daniel Sykes and Michael D. Ward
Chemical Communications 2011 - vol. 47(Issue 8) pp:NaN2281-2281
Publication Date(Web):2010/12/15
DOI:10.1039/C0CC04562D
In Ir(III)/Tb(III) dyads in which the excited state energy of the Ir(III) unit lies above 22000 cm−1, visible-light excitation of the Ir(III) chromophore results in sensitised emission from Tb(III) following Ir → Tb energy-transfer.
Co-reporter:Alexander J. Metherell, Christophe Curty, Andreas Zaugg, Suad T. Saad, Genevieve H. Dennison and Michael D. Ward
Journal of Materials Chemistry A 2016 - vol. 4(Issue 41) pp:NaN9668-9668
Publication Date(Web):2016/10/03
DOI:10.1039/C6TC03754B
Interaction of the V-series chemical warfare agent simulant ‘VO’ with Eu(III) results in selective quenching of the red Eu-based emission component from a dual-luminescent (blue/red) Ir(III)/Eu(III) dyad, resulting in a colour change in the luminescence from red to blue in the presence of the simulant.