[go: up one dir, main page]

Academia.eduAcademia.edu
Hybrid Organic-Inorganic Materials: Synthesis, solid state characterisation and solution studies of organoarsonate and -phosphonate functionalised coordination clusters and networks Camelia Ioana Onet A thesis submitted to the University of Dublin for the degree of Doctor of Philosophy School of Chemistry University of Dublin 2013 Declaration I declare that this thesis has not been submitted as an exercise for a degree at this or any other university and it is entirely my own work. I agree to deposit this thesis in the University’s open access institutional repository or allow the Library to do so on my behalf, subject to Irish Copyright Legislation and Trinity College Library conditions of use and acknowledgement. ____________________________ Camelia Ioana Onet I Acknowledgments First and foremost, I would like to thank my supervisor Prof. Dr. Wolfgang Schmitt for offering me the possibility of joining his research group, for his guidance, and encouragement. I am really grateful for his understanding and support throughout my time here in Trinity College Dublin. I am especially grateful to Dr. Tom McCabe, Dr. Lei Zhang and Dr. Nianyong Zhu for their invaluable assistance with the X-ray diffraction measurements and crystal structure determinations. I also would like to express my gratitude to Prof. Rodolphe Clérac and Mathieu Rouzieres for the tireless effort they have put into the magnetic studies. I would sincerely like to thank Dr. J. Bernard Jean-Denis and Dr. Martin Feeney for their help with the ESI-MS measurements and to Dr. John O’Brien and Dr. Manuel Rüther for NMR measurements. Many thanks to all the academic, technical and administrative staff within the School of Chemistry for their help over the years. Special thanks must go to my laboratory colleagues Lei, Ian, John, David, Giuseppe, Bartosz, Gerard, Adam, Anne Marie, Lukas, Theresa, Nianyong, Mariyatra, Anil and Jian Di for all their support on both a professional and personal level. Thanks as well to all the project students who have joined us over the years, especially to Raphaele Clement and Pierre Heijboer. Also, I would like to thank all the members in Professor Draper`s group. I am very grateful to Mrs. Corinne Harrison for all her support during my years of study in Dublin and for treating me like her daughter while living at her house. I would also like to thank SFI for their financial support and Trinity College Dublin for making my training here possible. Finally, I would like to thank my family and friends for their support and encouragement. II Summary The research presented in this thesis focuses on the development of novel hybrid organic-inorganic materials with potential applications in the areas of catalysis, gas sorption and separation, magnetism. The project involves the synthesis and structural characterisation of coordination complexes and networks stabilised by organophosphonate and -arsonate ligands. Solution studies were performed using electrospray ionisation mass spectrometry (ESI-MS). ESI-MS proved to be a very powerful analytical tool that allowed us to monitor the synthetic approach to hybrid organic-inorganic oxo-clusters and investigate their stability in solution. The physicochemical properties of the synthesised compounds have also been studied. The first chapter of this work introduces the reader to the field of research, highlights main achievements of the field and puts the work into context. The second chapter, “Hybrid Organic-Inorganic Polyoxomolybdates”, details the synthesis and characterisation of a series of hybrid polyoxomolybdates stabilised by organoarsonate and –phosphonate ligands: (NH4)2H2[MoV4O8(O3AsC6H5)4]·5H2O (1), (NH4)2H2[MoV4O8(O3AsC6H4NH2)4]·DMF·4H2O (O3AsC6H4OH)5]·9H2O (3), (6), (NH4)5[MoVI2MoV3O11 (NH4)4[MoVI4O10(O3AsC6H3NO2OH)4]·2H2O (NH4)4H4{Mn[MoV6O12(OH)3(O3PC6H5)4]2} 8H2O (O3PC6H5)4]2} 8H2O (2), (5), (NH4)4H4{Fe[MoV6O12(OH)3 (NH4)4H4{Co[MoV6O12(OH)3(O3PC6H5)4]2} 8H2O (NH4)4H4{Ni[MoV6O12(OH)3(O3PC6H5)4]2} 8H2O (8) and (4), (7), (NH4)4H4{Mg[MoV6O12(OH)3 (O3PC6H5)4]2} 8H2O (9). It has been demonstrated that slight perturbations of the ligand functionalities can be exploited to stabilise unprecedented core structures. Additionally, the electrospray ionisation mass spectrometry (ESI-MS) technique has been exploited to investigate the self-assembly process of hybrid organic-inorganic polyoxomolybdates that form upon partial reduction of (NH4)6Mo7O24·4H2O in the presence of aromatic organoarsonates. Functionalised POMs were also obtained by incorporating d-block hetero-atoms (Mn, Fe, Co, Ni) and a main group element (Mg) into defined POM structures. We intended to explore how different heteroatoms would influence the structure and magnetic properties of the compounds. The third chapter, “Polynuclear Manganese Coordination Complexes”, presents the synthesis and characterisation of a series of manganese cluster compounds stabilised by III organophosphonate ligands: ·8H2O 13( 4- (10), (11), [ n [ n 13( 4- [ n 13( 4- [ n 15( 2-H2 )2(C )2( 3- )4( 2- H)2( 2-C )2( 3- )4( 2- H)2( 2-C )2( 3- )4( 2- H)2( 2-C 3 3 )4(C6 )4(C6 5C )4(C6 3 5C 2P )16(C6 3 2P 5P 5P 3)10(C5 3)10(C5 3)10(C6 3)20]Cl5·22CH3OH 5-C3 5 )5Cl]·3H2O 5 )6]Cl·5H2O (12), 6-C5 4 )6]Cl·5H2O K(H2O)4[MnIII12( 3-O)6(CH3OH)6(C6H5CH2PO3H)7(C6H5CH2PO3)8]Cl2·3CH3OH (13), ·4H2O (14), K(H2O)4[MnIII12( 3-O)6(CH3OH)6(C6H5CH2PO3H)7(C6H5CH2PO3)8]Br2 ·2CH3OH·2H2O (15) and (H3O)4[MnIII2MnII4( 4-O)2(H2O)2(CH3CN)2{(C6H5)3CPO3}6]Cl2 ·2CH3CN·4H2O (16). It has been shown that small changes to the reaction conditions result in drastic changes of the cluster nuclearity and topology. The ESI-MS technique was again used to screen the reaction mixtures of 10-16 in order to identify new polynuclear manganese species that form in solution. Magnetic studies performed for these compounds revealed that compounds 11-13 exhibit single molecule magnet (SMM) behaviour. Chapter 4, describes the synthesis and characterisation of two coordination networks employing mononuclear (H3O){ n(H2 )(C 3 )2[C6H3(C6 (H3O){Cu(H2 )2(C 3 )[C6H3(C6 4P 4P SBUs 3 3 and triphosphonate )3]}·xCH3OH·yH2O (17) linkers: and )3]}·xCH3OH·yH2O (18). The compounds exhibit a 2D layered architecture stabilised in the crystal structure by week hydrogen bonds and - interactions. The materials did not show any permanent porosity. Chapter 5 provides a description of the applied materials and methods, and provides the experimental details for the synthesis of the compounds. IV Table of contents Acknowledgements……………………………………………………………………….II Summary…………………………………………………………………………………III List of compounds……………………………………………………………………..VIII Abbreviations………………………………………………………………………..…..IX 1. Introduction……………………………………………………………………………………………………………………....1 1.1 Polyoxometalates……………………………………………………………………………………………………………..2 1.1.1 Historical background of POMs………………………………………………………………….…………………..3 1.1.2 Condensation reactions as synthetic procedure.......................................................................................5 1.1.3 Polyoxomolybdates……………………………………………………………………………………………………….6 1.1.4 Functionalisation of POMs……………………………………………………………………………………………..8 1.1.4.1 Transition metal substituted POMs (TMSPs)…………………………………………………………….8 1.1.4.2 Incorporation of main group elements into POMs……………………………………………………..9 1.1.4.3 Organic and organometallic derivatives of POMs…………………………………………………….10 1.1.4.4 Organophosphonate and –arsonate functionalised polyoxomolybdates……………………11 1.2 Metal-organic frameworks..................................................................................................................................14 1.3 Mass spectrometry in coordination chemistry…………………………………………………………………18 1.4 Molecular magnetism……………………………………………………………………………………………………..24 1.5 Aims and objectives………………………………………………………………………………………………………..29 References…………………………………………………………………………………………………………………………………32 2. Hybrid Organic-Inorganic Polyoxomolybdates…………………………………………………………………37 2.1 Introduction and motivation…………………………………………………………………………………………...38 2.2 Functionalisation of polyoxomolybdates using organo-arsonates and –phosphonates……….40 2.2.1 Synthesis and characterisation of organoarsonate functionalised polyoxomolybdate clusters………………………………………………………………………………………………………………………………….40 2.2.1.1 (NH4)2H2[MoV4O8(O3AsC6H5)4]·5H2O (1)………………………………………………………………….40 2.2.1.2 (NH4)2H2[MoV4O8(O3AsC6H4NH2)4]·DMF·4H2O (2)…………………………………………………...50 2.2.1.3 (NH4)5[MoVI2MoV3O11(O3AsC6H4OH)5]·9H2O (3)……………………………………………………….59 2.2.1.4 (NH4)4[MoVI4O10(O3AsC6H3NO2OH)4]·2H2O (4)………………………………………………………...71 2.2.2 Synthesis and characterisation of organophosphonate functionalised heteropolyoxomolybdate clusters…………………………………………………………………………………………..78 2.2.2.1 (NH4)4H4{Mn[MoV6O12(OH)3(O3PC6H5)4]2}·8H2O (5)………………………………………………...78 2.3 ESI-MS studies of complex reaction mixtures to investigate the formation of hybrid organic- inorganic polyoxomolybdates…………………………………………………………………………………………………….88 2.3.1 Investigation of the reaction system that led to the formation of the cubane structures in 1 and 2……………………………………………………………………………………………………………………………………88 V 2.3.2 Investigation of the reaction system that led to the formation of the {Mo5} complex in 3...94 2.3.3 Investigation of the reaction system that led to the formation of the {Mo4} complex in 4...97 2.4 Conclusion and future work……………….…………………………………………………………………………...99 References……………………………………………………………………………………………………………………………….102 3. Polynuclear Manganese Coordination Complexes…………………………………………………………..105 3.1 Introduction and motivation…………………………………………………………………………………………106 3.2 Phosphonate ligands……………………………………………………………………………………………………108 3.3 Polynuclear manganese complexes stabilised by organophosphonates…………………………...109 3.3.1 Synthesis and characterisation of a pentadecanuclear manganese complex………………….109 3.3.1.1 [ΜnΙΙΙ15(μ2-H2Ο)2(CΗ3ΟΗ)16(C6Η5PΟ3)20]Cl5·22CH3OH·8H2O (10)……………………………109 3.3.2 Synthesis and characterisation of tridecanuclear manganese complexes………………………123 3.3.2.1 [ΜnΙΙΙ13(μ4-Ο)2(μ3-Ο)4(μ2-ΟH)2(μ2-CΗ3Ο)4(C6Η5PΟ3)10(C5Η5Ν)5Cl]·3H2O (11) ............... 123 3.3.2.2 [ΜnΙΙΙ13(μ4-Ο)2(μ3-Ο)4(μ2-ΟH)2(μ2-CΗ3Ο)4(C6Η5CΗ2PΟ3)10(C5Η5Ν)6]Cl·5H2O (12) ....... 143 3.3.2.3 [ΜnΙΙΙ13(μ4-Ο)2(μ3-Ο)4(μ2-ΟH)2(μ2-CΗ3Ο)4(C6Η5CΗ2PΟ3)10(C6Η5-C3Η6-C5Η4Ν)6]Cl·5H2O (13) ................................................................................................................................................................................... 159 3.3.3 Synthesis and characterisation of dodecanuclear manganese complexes ................................ 172 3.3.3.1 K(H2O)4[MnIII12(μ3-O)6(CH3OH)6(C6H5CH2PO3H)7(C6H5CH2PO3)8]Cl2·3CH3OH·4H2O (14) ................................................................................................................................................................................... 172 3.3.3.2 K(H2O)4[MnIII12(μ3-O)6(CH3OH)6(C6H5CH2PO3H)7(C6H5CH2PO3)8]Br2·2CH3OH·2H2O (15) ................................................................................................................................................................................... 187 3.3.4 Synthesis and characterisation of a hexanuclear manganese complex ...................................... 196 3.3.4.1 (H3O)4[MnIII2MnII4(μ4-O)2(H2O)2(CH3CN)2{(C6H5)3CPO3}6]Cl2·2CH3CN·4H2O (16) ....... 196 3.4 ESI-MS studies of complex reaction mixtures to investigate the formation of polynuclear manganese coordination complexes………………………………………………………………………………………….206 3.4.1 Investigation of the reaction system that led to the formation of the {Mn15} complex in 10 ... ........................................................................................................................................................................................ 206 3.4.2 Investigation of the reaction system that led to the formation of {Mn13} complexes observed in 11-13 ........................................................................................................................................................... 211 3.4.3 Investigation of the reaction system that led to the formation of {Mn12} complexes observed in 14-15 ........................................................................................................................................................... 215 3.4.4 Investigation of the reaction system that led to the formation of the {Mn6} complex in 16..... ........................................................................................................................................................................................ 220 3.5 Conclusion and future work…………………….…………………………………………………………………....222 References……………………………………………………………………………………………………………………………….226 4. Supramolecular Coordination Networks Employing Triphosphonate Linkers……………….229 4.1 Introduction…………………………………………………………………………………………………………………230 4.2 Extended triphosphonate ligands………………………………………………………………………………….231 4.2.1 Synthesis of 1,3,5-Tris(4-phosphonophenyl)benzene (P-TPB)……………………………………..233 4.2.2 Synthesis of 1,3,5-Tris(4`-phosphonobiphenyl-4-yl)benzene (P-TBB)………………………….235 VI 4.3 Synthesis and characterisation of coordination networks employing mononuclear inorganic SBUs and triphosphonate linkers……………………………………………………………………………………………...240 4.3.1 (H3O){ΜnII(H2Ο)(CΗ3ΟΗ)2[C6H3(C6Η4PΟ3Η)3]}·xCH3OH·yH2O (17)………………………………240 4.3.2 (H3O){Cu(H2Ο)2(CΗ3ΟΗ)[C6H3(C6Η4PΟ3Η)3]}·xCH3OH·yH2O (18)…………………………………247 4.4 Conclusion and future work…………………..………………………………………………………………………253 References……………………………………………………………………………………………………………………………….254 5. Experimental…………………………………………………………………………………………………………………..256 5.1 Materials and methods………………………………………………………………………………………………….257 5.2 Ligand synthesis…………………………………………………………………………………………………………...260 5.3 Synthesis of the Metal Complexes………………………………………………………………………………….265 References……………………………………………………………………………………………………………………………….272 Appendix…………………………………………………………………………..…….273 VII List of compounds 1 (NH4)2H2[MoV4O8(O3AsC6H5)4]∙5H2O 2 (NH4)2H2[MoV4O8(O3AsC6H4NH2)4]∙DMF∙4H2O 3 (NH4)5∙[MoVI2MoV3O11(O3AsC6H4OH)5]∙9H2O 4 (NH4)4∙[MoVI4O10(O3AsC6H3NO2OH)4]∙2H2O 5 (NH4)4∙H4{Mn[MoV6O12(OH)3(O3PC6H5)4]2}∙8H2O 6 (NH4)4∙H4{Fe[MoV6O12(OH)3(O3PC6H5)4]2}∙8H2O 7 (NH4)4∙H4{Co[MoV6O12(OH)3(O3PC6H5)4]2}∙8H2O 8 (NH4)4∙H4{Ni[MoV6O12(OH)3(O3PC6H5)4]2}∙8H2O 9 (NH4)4∙H4{Mg[MoV6O12(OH)3(O3PC6H5)4]2}∙8H2O 10 [ΜnΙΙΙ15(μ2-H2Ο)2(CΗ3ΟΗ)16(C6Η5PΟ3)20]Cl5·22CH3OH·8H2O 11 [ΜnΙΙΙ13(μ4-Ο)2(μ3-Ο)4(μ2-ΟH)2(μ2-CΗ3Ο)4(C6Η5PΟ3)10(C5Η5Ν)5Cl]·3H2O 12 [ΜnΙΙΙ13(μ4-Ο)2(μ3-Ο)4(μ2-ΟH)2(μ2-CΗ3Ο)4(C6Η5CΗ2PΟ3)10(C5Η5Ν)6]Cl·5H2O 13 [ΜnΙΙΙ13(μ4-Ο)2(μ3-Ο)4(μ2-ΟH)2(μ2-CΗ3Ο)4(C6Η5CΗ2PΟ3)10(C6Η5-C3Η6-C5Η4Ν)6]Cl·5H2O 14 K(H2O)4[MnIII12(μ3-O)6(CH3OH)6(C6H5CH2PO3H)7(C6H5CH2PO3)8]Cl2·3CH3OH·4H2O 15 K(H2O)4[MnIII12(μ3-O)6(CH3OH)6(C6H5CH2PO3H)7(C6H5CH2PO3)8]Br2·2CH3OH·2H2O 16 (H3O)4[MnIII2MnII4(μ4-O)2(H2O)2(CH3CN)2{(C6H5)3CPO3}6]Cl2·2CH3CN·4H2O 17 (H3O){Μn(H2Ο)(CΗ3ΟΗ)2[C6H3(C6Η4PΟ3Η)3]}·xCH3OH·yH2O 18 (H3O){Cu(H2Ο)2(CΗ3ΟΗ)[C6H3(C6Η4PΟ3Η)3]}·xCH3OH·yH2O VIII Abbreviations 3D three dimensional ac alternating current a.m.u. atomic mass units ATR attenuated total reflectance BVS bond valence sum analysis C Currie constant CHN carbon hydrogen nitrogen elemental analysis D zero-field splitting dc direct current DMF N,N`-dimethylformamide DMSO dimethyl sulfoxide EDX energy dispersive X-ray analysis ESI-MS electrospray ionization mass spectrometry FTIR Fourier-transform infrared g Landé g-factor IR infrared J exchange coupling constant M magnetisation MALDI matrix-assisted laser desorption ionisation min minutes MOF metal-organic framework m/z mass to charge ratio NMR nuclear magnetic resonance Oe Oersted POM polyoxometalate P-MOF polyoxometalate metal organic framework S ground spin state SBU secondary building unit SMM single molecule magnet SQUID superconducting quantum interference device T temperature TGA thermogravimetric analysis IX TOF-MS Time-of-flight mass spectrometer U spin reversal barrier UV-vis ultraviolet visible XRD X-ray diffraction chemical shift magnetic susceptibility wavelength QTM quantum tunnelling of the magnetisation X Chapter 1 – Introduction 1. INTRODUCTION 1 Chapter 1 – Introduction 1.1 POLYOXOMETALATES Polyoxometalates (POMs) are anionic metal–oxide clusters of the early transition metals (V, Nb, Ta, Mo, W) in their highest oxidation state, and represent a class of compounds whose preparation is a very active area in chemical and materials research.1-5 The interest in these compounds arises not only from their extraordinary structural characteristics but also from their unique intrinsic electronic, optical and chemical attributes (including redox and photochemical activity, charge distribution and band structures) which promote applications in many diverse disciplines such as catalysis, the development of sensors, photocatalysts, electrochromics, magnetic materials, energy storage and conversion devices.1, 4-16 The field of polyoxometalate chemistry has evolved very rapidly in the last two decades and research developments are governed by newly developed synthetic methodologies that led to the synthesis and characterisation of ultra-large, molecular clusters that contain up to 368 metal atoms.17 Much of the interest in these molecules has arisen because such clusters represent molecular entities whose structures can be rationalised by consecutive condensation reactions of well defined sub-units. Variations of the reaction conditions can initiate aggregation reactions of specific molecular species producing oligomeric complexes and large nanoscale clusters.1, 17 The versatile nature of this class of compound in terms of structure, size, redox chemistry, photochemistry, and charge distribution, means that the chemistry of polyoxometalates is arguably one of the many areas in inorganic chemistry that is developing very rapidly.1 While many of these systems can be easily prepared, the ability to functionalise polyoxometalates is an ongoing challenge.3, 18-20 The early transition metal ions are able to polarise terminal M-O bonds stabilising large molecular cage structures and other aggregates. These bonds that usually point radial to the outside of the 0-D cluster species are characterised by d- contributions and double-bond character impeding functionalisation. Consequently many advanced developments of polyoxometalatecontaining materials are hindered by the limited availability of functionalised POMs.21 Currently many research groups are developing synthetic approaches to functionalised POMs, whereby the preparation of hybrid POMs appears to be more challenging than ligand–stabilised compounds containing late transition metal ions. 2 Chapter 1 – Introduction 1.1.1 Historical background of POMs The history of polyoxometalate chemistry dates to the discovery of ammonium 12molybdophosphate [PMo12O40]3- by Berzelius in 1826. Many other heteropoly compounds were subsequently reported and analysed throughout the following decades, but the structure of most polyoxometalates remained not fully understood.22 The development of X-ray crystallographic techniques provided a breakthrough for the subject area allowing the detailed elucidation of POM structures and providing access to structure-property relationships. In 1933 employing a powder X-ray diffraction study, J. F. Keggin solved the structure of H3[PW12O40]·5H2O, a nowadays well-known 12:1 type heteropolyanion.23 Named after its discoverer, the Keggin structure (Figure 1.1, a), contains 12 {WO6} octahedra linked by edge and corner sharing motifs, with the heteroatom occupying a tetrahedral coordination site in the centre of the oxo cluster. A further 6:1 type, the Anderson`s heteropolyanion [TeVIMo6O24]6-, was studied by single-crystal X-ray diffraction by Evans in 1948. This structure is now commonly called the Anderson-Evans structure24 (Figure 1.1, b). It consists of six edge sharing {MoO6} octahedra adopting a hexagonal arrangement around a central octahedrally coordinated tellurium atom. Closely related to the Keggin structure, a 18:2 heteropolyanion [P2W18O62]6- was reported in 1953 by Dawson25, referred to as the Wells-Dawson`s structure (Figure 1.1, c). This heteropolyanion consists of two identical subunits in which a central {PO4} tetrahedron is surrounded by 9 {WO6} octahedra linked by edge and corner sharing connectivity motifs. a) b) c) Figure 1.1 – Polyhedral representation of: (a) Keggin structure,23 (WO6 unit – blue polyhedra; P – yellow sphere); (b) Anderson-Evans structure,24 (MoO6 unit – blue polyhedra; Te – red polyhedron); (c) Wells-Dawson structure,25 (WO6 unit – blue polyhedra; P – purple spheres). 3 Chapter 1 – Introduction Another important POM cluster is the isopolyanion [M6O19]n-, (M = Mo, W, Nb, Ta, and V) discovered in 1952 by Lindqvist and known as the Lindqvist structure26 (Figure 1.2). a) b) Figure 1.2 – The Lindqvist Structure: (a) polyhedral representation; (b) ball-and-stick representation.26 Colour code: M blue, O red. Lindqvist POMs have an overall octahedral geometry and consist of one central oxygen atom that is octahedrally surrounded by six metal ions. Each of these six metal atoms is 2-bridged to other metal atoms by four different oxygen ligands and has one terminal oxygen atom.26 One particular Lindqvist POM containing Mo metal ions, hexamolybdate [Mo6O19]2-, is of importance, because it can be directly functionalised to incorporate organoimido ligands. This unique feature has made [Mo6O19]2- a very popular POM cluster that has been studied extensively in recent years.27, 28 Recent exciting new developments in the field of polyoxometalates include: the discovery of large, highly symmetric polyoxomolybdates such as the wheel-shaped [(MoO3)176(H2O)80H32] spherical clusters cluster29 (Figure 1.4, c), giant [{(Mo)Mo5O21(H2O)6}12{MoV2O4(CH3COOH)}30]42- (Figure 1.4, a)30 and Keplerates Na48[H16Mo368O1032(H2O)240(SO4)48] ca. 1000H2O (Figure 1.4, d)17; hybrid organic-inorganic materials that contain POM cores;31, 32 new potential applications based on unusual magnetic7 and optical33 properties of some POMs, and potential medical applications such as anti-tumor and anti-viral uses.34 4 Chapter 1 – Introduction 1.1.2 Condensation reactions as synthetic procedure The main synthetic approaches used to produce oligomeric complexes or large nanoscale clusters of the early transition metal ions are simple and involve usually the acidification of aqueous solutions containing the relevant metal–oxide anions (molybdates, tungstates and vanadates). However, the specific reaction variables and parameters that manipulate the synthesis of POMs are manifold and depend on the type of metal–oxide anion, the concentration, the pH and type of acid, the heteroatom concentration, the ligand types, the reducing agent, temperature and the solvent type. The synthesis of POM clusters is based on self assembly processes involving covalent linking of transferable building blocks, under "one-pot" conditions (Figure 1.3). These structural building blocks can assemble into pre-defined architectures, and consist of aggregates of metal–oxygen units that adopt a defined polyhedral arrangement with the metal ions in the centre and the oxygen ligands as vertices.1, 35-37 a) b) Figure 1.3 – (a) Fundamental units and building blocks;38 (b) Self-assembly of polyoxometalate clusters from {Mo12} to the protein-sized {Mo368}. The Mo centres are shown as green polyhedra.35 5 Chapter 1 – Introduction The overall structure of the POM cluster can be visualised as a combination of a set of polyhedra that adopt corner and/or edge sharing modes. The formation reactions that often occur through a set of consecutive condensation reactions are mechanistically not yet fully understood. Therefore, the concept that regards metal-centered polyhedra as structural building blocks became extremely useful to rationalise the observed POM structures and their formation processes.35, 36, 38 1.1.3 Polyoxomolybdates Polyoxomolybdates represent a very important subclass of POM clusters. Research developments and synthetic approaches in this fascinating area of POM chemistry take advantage of a unique library of molybdenum-oxide based building blocks to generate a large variety of molecular architectures.36 The fundamental units and building blocks of molybdenum-based POMs include {Mo1}, {Mo2}, {(Mo)Mo5}, {Mo8} and {Mo17} = [{Mo8}2{Mo1}] units, as seen in Figure 1.3, a.38 A huge diversity of structures, from small to large nanoscale clusters, can be produced upon combination of these building units. The broad range of polyoxomolybdate structures presently known include {Mo6},24 {Mo7},39 {Mo8},40 {Mo10},41 {Mo12},42 {Mo18},43 {Mo36},44 {Mo37},45 {Mo132},30 {Mo154},46 {Mo176},29 {Mo248}47 and {Mo368}17 oxo-clusters. One of the most exciting developments in POM chemistry is represented by the discovery of a high-nuclearity cluster commonly known as the “big-wheel”, the [Mo154O462H14(H2O)70]14- ≡ {Mo154} cluster, reported by Müller et al. in 1995.46, 48 The {Mo154} cluster anion has a ring topology and consists of 14 {Mo8} building units linked by 14 {Mo2} and 14 {Mo1} units, respectively (Figure 1.4, b). A structurally related wheel-shaped metal oxide cluster, the [(MoO3)176(H2O)80H32] ≡ {Mo176} cluster,29 can be obtained by linking 16 instead of 14 sets of the above mentioned {Mo8}, {Mo2} and {Mo1} building units, respectively (Figure 1.4, c). The {Mo154} cluster has an external diameter of 3.4 nm, while that of the {Mo176} cluster is 4.1 nm. The {Mo8} building block is itself built-up by a densely packed pentagonal unit {(Mo)Mo5} and two more weakly bonded {MoO6} octahedra. The pentagonal unit {(Mo)Mo5} consists of a central bipyramidal {MoO7} unit sharing edges with five {MoO6} octahedra. These pentagonal units are also key for the formation of giant spherical clusters or Keplerates, like the 6 Chapter 1 – Introduction [{(Mo)Mo5O21(H2O)6}12{MoV2O4(CH3COOH)}30]42- {Mo132}30 ≡ and the [H16Mo368O1032(H2O)240(SO4)48]48- ≡ {Mo368}17 clusters. The {Mo368} cluster anion is the largest cluster known to date. It has an approximate D4 symmetry and consists of 368 molybdenum atoms. It can be regarded as composed of 40 pentagonal {(Mo)Mo5} units, 32 dinuclear {Mo2} units and 64 mononuclear {Mo1} units (Figure 1.4, d).1, 48 a) b) c) d) Figure 1.4 – Giant sphere- and wheel-shaped polyoxomolybdates: (a) The spherical {Mo132} cluster; (b,c) The wheel-shaped {Mo154} and {Mo176} clusters; (d) The {Mo368} cluster, highlighting the related pentagonal {(Mo)Mo5} (cyan and blue), dinuclear {Mo2} (red) and mononuclear {Mo1} (yellow) building units.48 7 Chapter 1 – Introduction Despite the fact that the high nuclearity molybdenum clusters, {Mo132}, {Mo154},{Mo176}, {Mo248} and {Mo368}, have complex nanoscale structures, they are quite easy to prepare and handle.1, 48 The versatile nature of polyoxomolybdate chemistry is due to the flexibility of the Mo-O-Mo bridges that allow “split and link” type processes,48 the number of variable redox states of the molybdenum centeres, different coordination numbers, strong hydration stabilisation and the presence of terminal Mo=O groups that ‘protect’ the entities to the outside stabilising discrete molecular clusters.48 1.1.4 Functionalisation of POMs The functionalisation of POMs may provide a tool to fine tune the electronic properties of the parent anions. This can be achieved by replacing the metal-oxo functionality of a POM by an alternate one, formal replacement of some oxo ligands, or grafting of a functional group at the surface of the polyanion.49 The ability to functionalise polyoxometalate clusters is of great importance, as it aims to extend their utility. 1.1.4.1 Transition metal substituted POMs (TMSPs) A simple approach to achieve POM functionalisation is to incorporate transition metal ions into the cluster. This allows us to influence their properties, such as magnetic behaviour50, 51 or catalytic activity4; examples are illustrated in Figure 1.5. All first row and most of the second row transition metals have been incorporated into various POM clusters. Lacunary POM fragments are ideal for the incorporation of transition metals. The most stable lacunary POMs derive from Keggin and Wells-Dawson polyoxotungstates, which were intensively used for the preparation of TMSPs. On the other hand, TMSPs incorporating paramagnetic 3d metals Mn, Fe, Co, Ni and Cu are of particular interest due to their electronic and magnetic properties.51 For instance, the two Fe substituted polyoxotungstate [(FeIII4W9O34(H2O))2(FeIIIW6O26)]19- clusters, [FeIII4(H2O)2(FeIIIW9O34)2]10- and (Figure 1.5, a, b) exhibit fascinating single molecule magnet (SMM) behaviour.52 8 Chapter 1 – Introduction a) b) c) Figure 1.5 – Structural representations of: (a) [Fe4(H2O)2(FeW9O34)2]10- and (b) [(Fe4W9O34(H2O))2 (FeW6O26)2]19- POMs52 (displaying SMM behaviour) and (c) [WZn3(H2O)2(ZnW9O34)2]12- oxidation catalyst.4 Colour code: W red polyhedra, Fe blue spheres, O red spheres, Zn black. 1.1.4.2 Incorporation of main group elements into POMs Polyoxometalates can also incorporate main group elements within their cluster structure. Examples of such main group derivatives are the halogenated POMs, which are useful precursors for further functionalisation.20 This type of functionalised POMs have the halide ions occupying a terminal position, e.g. [Mo4O10(OMe)4Cl2]2-,53 [W6O14Cl10]2-,54 and [PW9O28Br6]3-.55 However, there are also numerous examples of functionalised POMs in which the halide ions are incorporated within the cluster anion, e.g. [H2NaW18O56F6]7-, or where the halide ion acts as a template for the self assembly of a larger cluster shell, e.g. [V18O42Cl]13-.20 Some examples are illustrated in Figure 1.6. Other examples of main group derivatives include Si, Sn or P containing POMs: [SiW11O39{O(SiR)2}]4-, (R = C2H5, C6H5, C3H5), [( -A-XW9O34)(SiR2)3](n-6)- (R = CH3, C6H5; X = Si, n = 10; X = P or As, n = 9), [SiW9O34(SiEt)3(O3SiEt)]4-, [SiM11O39(SnR)]5- (M = W, R = CH3, C2H5, C6H5, C3H5, C4H9; M = Mo, R = C2H5), [PW10O38(SnR)2]5- (R = CH3, C6H5), [PW11O39(SnCl)]3-, [P2W16O60(SnPh)2]8-, [XW11O39Sn]n- (X= P, Si, Ge, B, Ga), a) b) 2-[P2W17O61Sn] 8- 20, 56, 57 . c) Figure 1.6 – Structural representations of: (a) [W6O14Cl10]2-; (b) [H2NaW18O56F6]7- and (c) [V18O42Cl]13cluster anions.20 9 Chapter 1 – Introduction 1.1.4.3 Organic and organometallic derivatives of POMs Numerous experimental studies have been devoted to the development of strategies to functionalise POM species using organic or organometallic precursors as functional groups. Main-group element, organic, and organometallic derivatives of polyoxometalates have been reviewed by A. Proust et al.,3, 20 highlighting several synthetic approaches to obtain such systems. One synthetic approach intensively exploited was the replacement of some terminal oxo ligands with various nitrogen-containing molecules such as nitrido, organoimido, diazenido ligands. The first nitrido POM derivative was reported by Zubieta and was a Lindqvist type POM [Mo6O18N]3-. Later on, Maatta and Proust showed that Reand Os-nitrido species can be incorporated in both Keggin and Dawson type POMs, [PW11O39(MN)]x- and [P2W17O61(MN)]x- (M = Re, Os) (Figure 1.7, a).3, 58, 59 Another exciting feature of the Lindqvist POM is the tendency to undergo multiple functionalisation steps upon reaction with organoimido ligands. Maatta et al. demonstrated that it is possible to replace successively all six terminal oxo groups of the [Mo6O19]2anion by 2,6-(diisopropyl)phenylimido moieties (Figure 1.7, b, c).3, 60, 61 a) b) c) Figure 1.7 – Structural representations of: (a) the nitrido-Keggin ion [PW11O39(MN)]x- (W blue polyhedra, M = Re, Os cian, P purple, N blue spheres, O red); (b) [Mo6O18(NC6H3-2,6-Me2)]2- and (c) [HMo6O13(NC6H3-2,6-Me2)6]- Lindqvist cluster anions (Mo blue, O red, N green, C black).20 Most recent studies exploit the direct grafting of organic substrates and organic ligands at the nucleophilic oxygen atoms of the POM core. Examples of aromatic phosphonate and arsonate stabilised POMs are illustrated in Figure 1.8. 18, 21, 62-66 In these compounds the aromatic ligands stabilise rather symmetrical cluster cores. The ligands 10 Chapter 1 – Introduction point into defined directions of space suggesting that these or related structures could be suitable building units for metal-organic frameworks.61 a) c) b) d) Figure 1.8 – Aromatic phosphonate/arsonate stabilised POMs with regular topology as potential SBUs for MOFs. (a) [Cl2⊂V14O22(OH)4(H2O)2(C6H5PO3)8]6-;62 (b) {Na4(H2O)10}[V12O12(OH)4(H2O)2 (O3AsC6H4NH2)10];21 (c) [Mo12O34(O3AsC6H4NH2)4]4-;63 (d) [Mo12O30(BPO4)2(O3PC6H5)6]5-.64 Colour code: V green, Mo cyan, O red, P purple, N blue, Cl light green, Na yellow, C black, H white. 1.1.4.4 Organophosphonate and –arsonate functionalised polyoxomolybdates Metal organophosphonates and -arsonates have attracted considerable research interest due to their interesting coordination chemistry and potential applications as catalysts and ion exchange materials.67 The structural diversity of these hybrid organicinorganic materials spans over a range of structures from molecular clusters to threedimensional (3D) frameworks.68 A very important subclass of these materials is represented by oganophosphonate and -arsonate functionalised polyoxometalates. The 11 Chapter 1 – Introduction tetrahedral functional group of the organophosphonate and -arsonate ligands displays geometrical and electronic similarities to metal-oxygen units in polyoxometalates suggesting that these ligands could be good candidates for POM functionalisation. The V/O/RPO32- system proved to be quite fruitful in yielding a range of structurally unique clusters and extended compounds, while the corresponding Mo/O/RPO32- system remain relatively undeveloped, being mostly limited to some molecular species. The most common structural type of the Mo/O/RPO32- system is represented by the cyclic pentanuclear core structure, with the general formula [Mo5O15(RPO3)2]4-.3, 61 These pentamolybdobisphosphonates were isolated using different phosphonate ligands (R = CH3, C2H5, C6H5, CH2C6H5, C2H4NH3+, p-CH2C6H4NH3+). The {Mo5} clusters have a ring topology with two phosphonate ligands capping both sides of the ring (Figure 1.9, a).3, 20, 69 Tetra-, hexa- and dodecanuclear architectures can also be isolated. Examples [Mo4O10(PhPO3)4]4-, include [Mo6O18(ButPO3)2]4-, [{(C6H5P)Mo6O21(H2O)3}2]4-, [RPMo6O21(O2CCH2NH3)3]2- (R = OH, CH3, C2H5), [(RPO3)4Mo12O34]4- (R = CH3, C2H5), [Na{Mo6O12(OH)3(PhPO3)4}2]9-, [Mo12O30(BPO4)2(O3PC6H5)6]5-.20, 61, 64, 69-71 The tetranuclear complex [Mo4O10(PhPO3)4]4- consists of two {Mo2O10} dinuclear units connected by two phosphonate ligands to form a six membered {Mo4P2} ring. The remaining two phosphonate ligands cap both sides of the central cavity of the ring (Figure 1.9, b).20 The core structure of the hexanuclear complexes, [{(C6H5P)Mo6O21(H2O)3}2]4-, [RPMo6O21(O2CCH2NH3)3]2-, adopts a similar ring motif capped by only one organophosphonate ligand, while the [Mo6O18(ButPO3)2]4- species adopts a non-planar conformation as seen in Figure 1.9, c.69, 70 The dodecanuclear complex [(RPO3)4Mo12O34]4is stabilised by methyl- or ethyl- phosphonates and adopts a so-called inverted-Keggin structure (Figure 1.9, d).71 The sandwich type dodecanuclear compounds [Na{Mo6O12(OH)3(PhPO3)4}2]9- consist of two hexanuclear MoV moieties linked by a Na ion (Figure 1.9, e), while the dodecanuclear polyanion, [Mo12O30(BPO4)2(O3PC6H5)6]5(Figure 1.8, d), is part of a class of POMs that resemble Dawson anions.20, 64, 72, 73 The incorporation of diphosphonate ligands into POM structures was also investigated. Methylenediphosphonate (O3PCH2PO3)4- was the first to be explored for this purpose. Dolbecq and Mialane reported a series of cyclic compounds with the general formula {(MoV2O4)(O3PCH2PO3)}n] (n = 3, 4, 10) (Figure 1.9, f).74, 75 The compounds 12 Chapter 1 – Introduction consist of alternating {MoV2O4} units and diphosphonate ligands, and their geometry depends on the nature of the co-ligand used and the nature of the counterions (NH4+, Na+, Li+).61 The [(Mo2O4)10(O3PCH2PO3)10(CH3COO)8(H2O)4]28- anion differentiates itself from conventional cyclic, adopting a double wheel architecture built from interconnected octa- and dodecanuclear wheels that are linked via two sodium ions (Figure 1.9, g).75 Other diphosphonate functionalised POMs include {Mo5O15(O3P(CH2)4PO3)}, [MoV6MoVIO16(O3PCH2PO3)]8- and [MoV7MoVIO16(O3PPhPO3H)]3- (Figure 1.9, h).65, 76, 77 a) b) c) d) e) f) g) h) Figure 1.9 – (a) [Mo5O15(CH3PO3)2]4-;20 (b) [Mo4O10(REO3)4]4- R = Ph, E = P, As;69 (c) 72, 73 [Mo6O18(ButPO3)2]4-;69 (d) [(EtPO3)4Mo12O34]4-;71 (e) [Na{Mo6O12(OH)3(PhPO3)4}2]9-; (f) [(Mo2O4)4(O3PCH2PO3)4(CO3)2]12-,74 (g) [(Mo2O4)10(O3PCH2PO3)10(CH3COO)8(H2O)4]28-,75 ({MoO6} orange octahedra, {PO3C} green tetrahedra, Na blue/gray spheres, C black spheres); (h) [MoV7MoVIO16(O3PPhPO3H)]3-.65 13 Chapter 1 – Introduction Even though organoarsonates display reactivities and structural characteristics close to that of organophosphonates, organoarsonate-stabilised POM derivatives have far less been isolated.61 Examples of organoarsonate-functionalised polyoxomolybdates include [Mo4O10(RAsO3)4]4-, [Mo5O15(RAsO3)2]4-, [Mo6O18(RAsO3)2]4-, [Mo12O34(RAsO3)4]4- (R = CH3, C3H7, C6H5, p-C6H4NH3+, C2H4OH, C6H4COOH).3, 20, 78, 79 The [Mo12O34(RAsO3)4]4- (Figure 1.8, c) species are analogous to the organophosphonate derivatives that display an inverted-Keggin core with a tetrahedral arrangement of the organic ligands. These type of clusters may be suitable to be used as building units for metal-organic frameworks.61 1.2 METAL-ORGANIC FRAMEWORKS Metal-organic frameworks (MOFs) represent a class of porous materials consisting of clusters or metal ions linked through rigid organic ligands. MOFs can be prepared using a fundamental construction principle, based on building block approaches where molecules, metal ions or clusters are considered as nodes, and coordination bonds represent node connectors. Choosing the desired combination of nodes and connectors, and using suitable building blocks, microporous networks with unprecedented surface areas can be obtained. The extraordinary low densities (1.00 to 0.20 gcm-3) and high surface areas (500 to 7000 m2 g-1) make MOFs ideal candidates for the storage and separation of gases (N2, Ar, CO2, CH4, and H2).80-82 The conceptual approach used to synthesise materials designed to have predetermined structures, compositions and properties is known as ‘the reticular synthetic approach’. This concept was described by O. M. Yaghi as “the process of assembling judiciously designed rigid molecular building blocks into predetermined ordered structures (networks), which are held together by strong bonding.”83-85 This approach requires the use of secondary building units (SBU). The term ‘secondary building unit’ refers to the geometry of the units (molecules, metal ions or clusters) defined by the points of extension (ligand coordination modes and metal coordination environments). These SBUs have intrinsic geometric properties that are used to direct the assembly of ordered frameworks. New materials with pre-desired structure, composition and properties can be obtainable by 14 Chapter 1 – Introduction linking these SBUs through coordination bonds. A few examples of inorganic and organic SBUs with varying geometries are presented in Figure 1.10. 81, 83-87 Inorganic units SBUs Organic units SBUs Figure 1.10 – Examples of SBUs from carboxylate MOFs. Colour code: O red, N green, C black. In inorganic units metal-oxygen polyhedra are blue.83 The design of rigid frameworks based on such SBUs has been highly successful as demonstrated by Yaghi and co-workers. Their strategy was to use {Zn4O(CO2)6} octahedral units and polyaromatic carboxylate ligands (Figure 1.10) to assemble extended 15 Chapter 1 – Introduction network structures with high surface areas and permanent porosity (Figure 1.11).83 For instance, by combining octahedral {Zn4O(CO2)6} units and large triangular tricarboxylate ligands, MOFs exhibiting exceptional porosity were generated. Examples include [Zn4O(BTB)2] (MOF-177, BTB3- = 4,4 ,4 -benzene-1,3,5-triyl-tribenzoate) (Figure 1.11), [Zn4O(BBC)2] (MOF-200, BBC3- = 4,4 ,4 -[benzene-1,3,5-triyl-tris(benzene-4,1- diyl)]tribenzoate) and [Zn4O(BTE)(BPDC)] (MOF-210, BTE3- = 4,4 ,4 -[benzene-1,3,5triyl-tris(ethyne-2,1-diyl)]tribenzoate and BPDC2- = biphenyl-4,4 -dicarboxylate).85 MOF177 shows a surface area of 4500 m2 g-1 and has a hydrogen gas uptake capacity of 7.5 wt % at 77 K.85 COO- COO- MOF-5 COO- + MOF-10 COO- COO- -OOC COO- MOF-177 Figure 1.11 – Representation of the reticular synthesis concept which can be applied to prepare metalorganic frameworks.83-85 16 Chapter 1 – Introduction Current developments have led to the formation of metal-organic frameworks with ultrahigh surface areas. For instance, [Cu3(L6-(109))(H2O)3]n (NU-109, LH6(109) = 1,3,5tris[(1,3-carboxylic acid-5-(4-(ethynyl)phenyl))butadiynyl]-benzene, Figure 1.12) and [Cu3(L6-(110))(H2O)3]n (NU-110, LH6(110) = 1,3,5-tris[((1,3-carboxylic acid-5-(4- (ethynyl)phenyl))ethynyl)phenyl]-benzene) display the highest surface areas reported to date (∼ 7000 m2 g-1).82 Figure 1.12 – NU-109 showing the presence of different types of cages within the structure and the packing arrangement viewed in the direction of the crystallographic a-axis.82 The striking increase of the surface areas reported for MOFs is primarily due to the development of new SBUs with extended structures and multiple binding sites. However, it is to a certain extend surprising that the vast number of existing POM clusters have not yet been systematically exploited as building blocks for MOFs. Only a very limited number of polyoxometalate metal–organic frameworks (P-MOFs) have been reported.80, 88, 89 P-MOFs are hybrid organic-inorganic compounds that might be obtainable by functionalisation of POMs with rigid organic ligands. They may be regarded as a new generation of solid-state materials with promising attributes for advanced applications in the areas of gas storage materials and catalysis.80 In contrast to traditional POMs, P-MOFs might provide additional functionalities arising from their organic components. Highly ordered and amendable cavities could give rise to shape-selectivity and may impact directly on important catalytic acid-base and redox processes. 17 Chapter 1 – Introduction 1.3 MASS SPECTROMETRY IN COORDINATION CHEMISTRY Mass spectrometry (MS) is a very powerful analytical technique used for both quantitative and qualitative investigations of a wide range of compounds ranging from small inorganic and organic molecules to biological macromolecules. MS technologies have faced huge progress in the past 20 years, leading to the development of highly sensitive mass spectrometers.90-92 The ability of MS to be used to analyse complex systems and labile solution species is due to the advances in the development of soft ionisation techniques such as electrospray ionisation (ESI) and matrix-assisted laser desorption ionisation (MALDI) processes. Furthermore, the sensitivity of a mass spectrometer is correlated to the mass analyser used, which separates the ions according to their mass-to-charge ratio (m/z). The most commonly used analysers are quadrupole and time of flight mass analysers. Different ionisation techniques and mass analysers have different advantages and disadvantages that one has to be aware of when performing MS analyses. For a specific class of compounds or for particular applications it is a requirement to choose the optimal ionisation technique and mass analyser.90, 91 For instance, electrospray ionisation mass spectrometry (ESI-MS) has been successfully applied in the study of coordination compounds. The growing interest in the use of ESI-MS to characterise metal complexes arises from the simplicity of the technique which is capable of transferring non-volatile solution phase ions to the gas phase.91-94 Furthermore, the soft character of ESI permits the transfer of the ions in the gas phase with minimum fragmentation, reflecting the actual species present in the solution. ESI-MS is a sensitive and direct method of detection that enables the analysis of highly diluted solutions and the examination of several species in solution simultaneously.91, 94, 95 Another attractive feature of ESI-MS is the flexibility with regard to the sample medium, as a wide range of solvents can be used and a wide pH range can be tolerated.93 Mass spectrometry investigations of complex coordination clusters is furthermore facilitated by the large number of natural isotopes that some of the elements possess. As a result, the mass spectrum of metal complexes displays characteristic wide distributions of isotopic envelopes that can be used for the interpretation of the spectra. Thus, the isotopic distribution gives the possibility to identify the presence and the number of polyisotopic 18 Chapter 1 – Introduction metals by comparison of experimental spectra with theoretically calculated isotopic patterns.93, 95, 96 The shape of the isotopic envelope is determined by the type and number of polyisotopic elements present. A high number of polyisotopic elements, e.g. molybdenum (92Mo, 14.8%; 94 Mo, 9.3%; 95 Mo, 15.9%; 96 Mo, 16.7%; 97 Mo, 9.6%; 98 Mo, 24.1%; 100Mo, 9.6%) and tungsten isotopes (180W, 0.1%; 182W, 26.5%; 183W, 14.3%; 184W, 186 W, 28.4%) even can give rise to a Gaussian type distribution of the resulting 30.6%; isotopic envelope (see Figure 1.13). Some other non-metallic elements have distinct isotopic distributions that can be used as “fingerprints” for their identification, i.e. chlorine (35Cl, 75.8%; 37Cl, 24.2%) and bromine atoms (79Br, 50.7%; 81Br, 49.3%).95, 97 An early example that demonstrated that a polynuclear cluster can be examined by ESI-MS was conducted by Colton and co-workers in 1992 using the heteropolyoxomolybdate (NEt4)4[S2Mo18O62] dissolved in acetonitrile solution. The study demonstrated that the intact [S2Mo18O62]4- cluster ion can be observed in the mass spectrum and its isotopic pattern was compared with the calculated spectrum (Figure 1.13).97 Figure 1.13 – Comparison of calculated and experimental negative-ion ESI mass spectral isotope pattern for [S2Mo18O62]4-.97 In addition to the here discussed main advantages of ESI-MS, a number of potential limitations must also be considered for data interpretation. Limitations when investigating 19 Chapter 1 – Introduction solution processes may include artefacts or signals in the mass spectrum due to ionisation processes or gas-phase reactions such as adduction, fragmentation or polymerisation, as well as redox processes.92, 93 Until very recently, mass spectrometry in coordination chemistry was mainly used as a complementary tool for the characterisation of novel compounds. Such studies involved structural characterisations of metal complexes dissolved in various solvent systems.92, 93, 95 In these studies, MS was assisted by other techniques such as NMR, X-ray crystallography, electrochemistry, etc.92, 96 On the other hand, recent directions of MS demonstrate that this technique is also capable of monitoring reaction mixtures in order to investigate the formation processes and reactivity of complex coordination clusters in solution.98-100 Only a few studies of this kind were carried out to date. An early example of real-time monitoring of reacting polyoxometalate species in solution by ESI-MS is represented by the work performed by Howarth et al. in 1997.101 Aqueous solutions of two separate tungsten and molybdenum complexes were mixed and the ESI-MS spectra of the reaction mixture were recorded over a time period of 45 min (Figure 1.14). The experimental results showed that the two anions [HPW12O40]2- and [HPMo12O40]2interchange metal ions with one another. The most intense signal in the mass spectrum recorded after 2 min after the solutions were mixed is due to the [HPW9Mo3O40]2- ion. During a time period of 45 min the relative distribution of the metal species for the anions reaches an equilibrium and the most intense signal in the mass spectrum corresponds to the [HPW6Mo6O40]2- ion.101 Figure 1.14 – The ESI mass spectra of mixtures of phosphododecatungstate and phosphododecamolybdate acquired (a) 2, (b) 17, (c) 32 and (d) 47 min after the two solutions were initially mixed. Labels a-m correspond, in order, to the formulae [HPWnMo12-nO40]2- with n = 0 to 12.101 20 Chapter 1 – Introduction These experiments indeed demonstrated that it is possible to monitor formation and possible rearrangement processes of polyoxometalate species that occur in solution. The power of MS technique in the polyoxometalate field has recently been highlighted by Cronin et al., who used it to study the self-assembly processes of polyoxometalate systems that emerge in solution. For example, cryospray ionisation mass spectrometry (CSI-MS) was used to analyse the reaction solution of the compound ((nC4H9)4N)2n)[Ag2Mo8O26]n.98 a) b) c) Figure 1.15 – (a) CSI-MS data collected of the reaction solution of ((n-C4H9)4N)2n)[Ag2Mo8O26]n; (b) Representation of the [AgMo2O7]- and [AgMo4O13]- species as building blocks of the (Ag{Mo8}Ag) synthon showing the isotopic envelopes for their corresponding mass peaks at 410.7 m/z and 700.5 m/z, respectively; (c) Structural representation of the higher mass fragments identified within the CSI-MS analysis of the reaction solution of ((n-C4H9)4N)2n)[Ag2Mo8O26]n.98 21 Chapter 1 – Introduction CSI-MS allowed access to direct observation of the rearrangement of Lindqvist anions into the (Ag{Mo8}Ag) synthon units and subsequent addition of the organic cations. The mass spectrum of the reaction mixture is presented in Figure 1.15 and reveals the presence of six monoanionic series: [MomO3m]- (m = 2, 3 or 5); [HMomO3m+1]- (m = 2 to 6); [H7MomO3m+2]- (m = 2 to 6); [H7MomO3m+3]- (m = 2 to 5); [H9MomO3m+4]- (m = 2 to 6); [AgMomO3m+1]- (m = 2 to 4). The detection of the [AgMo2O7]- (signal at m/z = 410.7 a.m.u.) and [AgMo4O13]- (signal at m/z = 700.5 a.m.u.) fragments of the (Ag{Mo8}Ag) synthon units are particularly important in order to understand the formation of the compound. Also the detection of the species [(AgMo8O26)TBA2]- (signal at m/z = 1776.6 a.m.u.), [(Ag2Mo8O26)(Mo4O13)TBA3]- (signal at m/z = 2718.3 a.m.u.), and [(Ag2Mo8O26)(Mo8O26)TBA5]- (signal at m/z = 3796.5 a.m.u.) (Figure 1.15), containing an increasing number of organic cations is of great importance. The analysis demonstrates that the nuclearity of the chain-type compound increases with the number of organic cations present highlighting that the formation process involves “monomeric” units that assemble into larger fragments which eventually lead to crystals of ((n- C4H9)4N)2n)[Ag2Mo8O26]n. The experimental results also demonstrate how organic cations can impose structure-directing effects on POM structures that assemble in solution.98 A related study, reported in 2011, employs electrospray ionisation mass spectrometry (ESI-MS) to monitor in real-time the self-assembly of an organic-inorganic POM cluster. The reaction system involves the rearrangement of [ -Mo8O26]4- into ((nC4H9)4N)3[MnMo6O18((OCH2)3CNH2)2], a Mn-Anderson type cluster (Figure 1.16).99 ESIMS spectra of the reaction mixture containing ((n-C4H9)4N)3[MnMo6O18((OCH2)3CNH2)2] were recorded at different time intervals over the course of the reaction (∼ 30 h), in order to identify different fragment ions present in the reaction mixture. A very important observation of the study is that the intensity of the signal attributed to the reactant ions decreases exponentially, whilst the intensity of the product signal increases at a lower rate over the course of the reaction. This suggests that the mechanism of formation of [MnIIIMo6O18((OCH2)3CNH2)2TBA2]- proceeds through further intermediates. ESI-MS studies allowed the identification of intermediate fragment ions involved in the rearrangement of [ -Mo8O26]4- into ((n-C4H9)4N)3[MnMo6O18((OCH2)3CNH2)2]. It was proposed that the rearrangement initially results in the formation of the [Mo4O13]2- cluster species (i.e. [Mo4O13Na]-, m/z = 614.6 a.m.u. and [Mo4O13TBA]-, m/z = 833.8 a.m.u.) representing half-fragments of the {Mo8} clusters. It is believed that further decomposition to smaller fragment ions (i.e. [Mo2O7H]-, m/z = 304.8 a.m.u. and [Mo3O10TBA]-, m/z = 22 Chapter 1 – Introduction 690.0 a.m.u.) which subsequently bind to tris(hydroxymethyl)aminomethane ligands (TRIS) ([Mo2O5((OCH2)3CNH2)]- m/z = 389.8 a.m.u.), manganese ions ([MnIIIMo3O8((OCH2)3CNH2)2- m/z = 706.7 a.m.u.), and anionic molybdate units, leads to the formation of the final Mn-Anderson-TRIS cluster, [MnIIIMo6O18((OCH2)3CNH2)2TBA2]- (m/z = 1640.0 a.m.u.) (Figure 1.16).99 a) b) Figure 1.16 – (a) ESI mass spectrum recorded of the reaction mixture of ((nC4H9)4N)3[MnMo6O18((OCH2)3CNH2)2]; spectrum was recorded after it has been heated at reflux (80°C) for approximately 30 h; (b) Illustration to visualize the prominent, intermediate fragment ions identified in this study (labelled b-f), which are involved in the rearrangement of the [ -Mo8O26]4- anion (labelled a), into the symmetrical Mn-Anderson anion [MnMo6O18((OCH2)3CNH2)2]3- (labelled g). Structures (bf) represent the following fragment ions identified in the ESI-MS investigations: b) [Mo4O13TBA]-, c) [Mo2O7H]-, d) [Mo3O10TBA]-, e) [Mo2O5((OCH2)3CNH2)]-, f) [MnIIIMo3O8((OCH2)3CNH2)2-. Colour scheme: Mo green polyhedra, Mn orange polyhedra, O red, N blue, C gray spheres.99 Mass spectrometry proves to be an indispensable tool in coordination chemistry which when used in combination with other techniques (i.e. X-ray crystallography, NMR 23 Chapter 1 – Introduction spectroscopy, electrochemistry, etc.) provides important information about the analyte (e.g. structure, purity and composition). Most remarkably the analytical technique provides the possibility of exploring real-time growth reactions of complex coordination clusters.91, 92, 100 1.4 MOLECULAR MAGNETISM Magnetic materials are considered to be indispensable in our daily lives finding applications in electronics, acoustic and telecommunication devices, data storage and readout devices, diagnostic equipment and medical therapy, etc.102 Molecular magnetism is an interdisciplinary area that deals with the design, synthesis, and characterisation of molecular-based magnetic materials.103 Single molecule magnets (SMMs) are isolated molecules with a finite number of interacting spin centres. SMMs represent a class of compounds that combine a defined spin ground state (S) and a large and negative magnetic anisotropy (D). This type of compound shows slow relaxation of the magnetisation at low temperatures, due to the large spin reversal barrier (U).102-104 SMMs are fascinating materials as they show magnetic hysteresis effects as classical magnets do, but at the same time their nanoscopic nature gives rise to quantum effects. The combination of classical and quantum mechanical phenomena makes SMMs highly attractive for theoretical studies.102, 105 Furthermore, the well defined spin or quantum states of SMMs may lead to future prospective applications in the IT sector whereby information could be stored at high density using a single molecule. This conceptional approach could satisfy the need for the miniaturisation of electronic devices and increase the data storage capacity.102, 105, 106 The most representative class of SMMs is based on manganese clusters.105, 107-109 Other large molecular clusters such as iron-oxo clusters110-112, polyoxovanadates113-116 or lanthanide based compounds117-119 have also shown to give rise to SMM behaviour (Table 1.1). 24 Chapter 1 – Introduction Table 1.1 Examples of single molecule magnets.120 Molecule S D (cm-1) Ref. 51/2 - 0.02 [109] 33/2 - 0.04 [111] 3 - 1.5 [113] 11 - 0.4 [119] [Mn25O18(OH)2(N3)12(pdm)6(pdmH)6](Cl)2 (pdmH2 = pyridine-2,6-dimethanol) {Mn25} [Fe19O6(OH)14(metheidi)10(H2O)12]+ (metheidi = N-(1-hydroxymethylethyl)iminodiacetic acid) {Fe19} [V4O2(EtCO2)7(bpy)2]+ (bpy = 2,2`-bipyridine) {V4} [Mn21DyO20(OH)2(ButCO2)20(HCO2)4(NO3)3(H2O)7] {Mn21Dy} 25 Chapter 1 – Introduction The first reported example of a single-molecule magnet is a dodecanuclear manganese cluster, [MnIV4MnIII8O12(CH3COO)16(H2O)4]⋅2CH3COOH⋅4H2O (Figure 1.17), synthesised by Lis in 1980121 and intensively studied in the 1990s.106, 122, 123 In 1993 Novak et al. reported that the magnetisation of the {Mn12} cluster is highly anisotropic and that the relaxation time of the magnetisation becomes very long below 4K, giving rise to pronounced hysteresis effects (Figure 1.17).123 The compound possesses a large spin ground state which arises from antiferromagnetic interactions between the spins of the four MnIV ions (S = 3/2) and the spins of the eight MnIII ions (S = 2) to give a total spin of S = [(4 x 3/2) – (8 x 2)] = 10. The magnetisation relaxation half-life of the {Mn12} cluster is more than 2 months at 2 K. Unfortunately the relaxation time of the magnetisation decreases dramatically with increasing temperature and the hysteresis effects are no longer observed above 4 K.103, 104 The hysteresis loop observed for a single crystal at low temperature (Figure 1.17) shows step-like features that occur at values of the applied field where the energies of different spin states coincide. At these particular values of the field the relaxation from one spin state to another is enhanced, due to quantum tunnelling of the magnetisation (QTM) through the energy barrier U.104, 106 Since the discovery of the first SMM, significant research work has been devoted to the development of molecule-based magnets that display spin reversal barriers (U) large enough to be able to show magnetic hysteresis effects at higher temperature.102, 104 [Mn12O12(CH3COO)16(H2O)4]⋅2CH3COOH⋅4H2O and its magnetisation hysteresis loops measured at low temperature between magnetic fields of 5 Figure 1.17 – The crystal structure of 106, 121 26 Chapter 1 – Introduction Even though an increasing number of metal complexes have been demonstrated to behave as SMMs, {Mn12} still remains the SMM with the most striking features. The largest SMM yet reported is a {Mn84} cluster, [Mn84O72(O2CMe)78(OMe)24(MeOH)12(H2O)42(OH)6], synthesised by Christou et al. in 2004 (Figure 1.18).105 The cluster adopts a circular torus motif with a diameter of ∼ 4.2 nm, a height of ∼ 1.2 nm and an inner free space with a diameter of ∼ 1.9 nm. It has a ground state spin of S = 6 and exhibits both magnetic hysteresis and QTM.103, 105 Figure 1.18 – The crystal structure of the {Mn84} torus. The rectangle shows the repeating {Mn14} unit that represents the contents of the asymmetric unit. Magnetisation hysteresis loops measured on single crystals of {Mn84} at low temperature showing SMM behaviour.105 Another interesting SMM is a {Mn6} complex, [MnIII6O2(Et- sao)6(O2CPh(Me)2)2(EtOH)6] (Et-saoH2 = 2-hydroxyphenylpropane oxime), that exhibits the highest energy barrier for the reversal of the magnetisation Ueff = 86.4 K and a blocking temperature of ∼ 4.5 K. This complex breaks the record held by the {Mn12} family which was known to show the highest blocking temperatures (∼ 3.5 K). {Mn6} has a ground state spin of S = 12 and exhibits both magnetic hysteresis and QTM (Figure 1.19).124 The greatest spin ground state of S = 83/2 was achieved in a {Mn19} complex, [MnIII12MnII7(µ4-O)8(µ3,η1-N3)8(HL)12-(MeCN)6]Cl2·10MeOH·MeCN (H3L = 2,6- bis(hydroxymethyl)-4-methylphenol), which comprises twelve MnIII (S = 2) and seven MnII (S = 5/2) ions. Despite this large spin, {Mn19} does not exhibit SMM behaviour due to a very small positive D.125 Anisotropy can be introduced in the {Mn19} complex by 27 Chapter 1 – Introduction replacing the central MnII with a DyIII ion to generate a {Mn18Dy} complex, [MnIII12MnII6DyIII(µ4-O)8(µ3-Cl)6.5(µ3-N3)1.5(HL)12(MeOH)6]Cl3·25MeOH, with the same core topology but showing SMM behaviour (Figure 1.20).117 Figure 1.19 – The crystal structure of the {Mn6} complex and its hysteresis loops measured at a constant field sweep rate of 1 mT s-1 at low temperature.124 and its hysteresis loops measured at a field scan rate of 2 mT s in the temperature range 0.5-0.04 K.117 Figure 1.20 – The crystal structure of the {Mn18Dy} complex -1 28 Chapter 1 – Introduction 1.5 AIMS AND OBJECTIVES The main aim of this research project is to synthesise hybrid organic-inorganic materials using a series of mono- and trifunctional organophosphonate and -arsonate ligands. These products could possess advantageous properties which could lend themselves to applications in the areas of catalysis, gas sorption and separation, magnetism. The research developed in the course of this project is focused on the solution behaviour, synthesis and structure elucidation of novel hybrid organic-inorganic materials. In the first part of the work we set out to explore the formation of mixed-valent molybdates in the presence of organoarsonates and -phosphonates. As already discussed in the previous sections of the introduction, early transition metal ions in their higher oxidation states have the ability to polarise terminal O2– ligands efficiently, often resulting in stable purely ‘inorganic’ oxo-clusters.21 The terminal metal-oxygen bonds limit the functionalisation of these clusters and advanced applications, for instance in catalysis are hampered by the limited availability of hybrid polyoxometalates that contain organic molecules. Organophosphonates and -arsonates appear to be good candidates for POM functionalisation as they have tetrahedral binding sites which display geometrical and electronic similarities to metal-oxygen units in polyoxometalates. Functionalisation of spherical clusters with rigid polyphosphonate and –arsonate ligands that point in defined directions of space might link the clusters into polyoxometalate metal-organic frameworks (P-MOFs) that are highly desired for catalysis, gas sorption and separation studies. The second part of this project is focused on the preparation of novel polynuclear manganese clusters stabilised by organophosphonates. The resulting coordination clusters may display interesting magnetic properties. The chemistry of manganese coordination clusters has drawn much attention due to their potential to exhibit properties of single molecule magnets. Carboxylate ligands have previously been intensively exploited to stabilise high-nuclearity molecular compounds of manganese, whereas organophosphonates have rarely been employed.126 However, organophosphonates appear to be good candidates to stabilise such species as they are versatile ligands that show high binding affinities towards metal ions.67, 127-129 A third aspect of the project explores the use of rigid extended triphosphonate ligands for the formation of supramolecular coordination networks. Porous hybrid organic29 Chapter 1 – Introduction inorganic materials are of particular interest for potential applications in catalysis, gas storage and separation. Several classes of MOFs have been explored, of which the carboxylate-based compounds are the most studied. This class of MOFs may possess high surface areas and uniform pore size distribution, but unfortunately the structural stability of many MOFs can be quite limited. This prompted us to explore the preparation of network structures using organic triphosphonate ligands. Organophosphonates are known to form stronger bonds with metal ions than carboxylates and therefore have a greater potential to form robust porous solids.67 Furthermore, the multidentate binding mode of the triphosphonate ligands makes them ideal for the preparation of hybrid materials with unprecedented topologies or enhanced porosity. Another aspect of the work aims to investigate the stability, formation processes and the reactivity of complex polynuclear clusters in solution. For this study we decided to take advantage of electrospray ionisation mass spectrometry (ESI-MS) which is a very powerful analytical tool. ESI-MS proved to be very efficient in monitoring complex reaction mixtures in order to gain a better understanding of how polynuclear clusters form and how we can influence the formation processes towards desired materials. Solution studies of polyoxometalates that involve ESI-MS screening of reaction mixtures is still at its infancy, while ESI-MS studies that investigate complex Mn coordination cluster systems have not been reported in the literature. This analytical technique was also applied to investigate the stability of the synthesised compounds in solution, an essential requirement for potential applications, e.g. as homogeneous catalysts. The organophosphonate and -arsonate ligands used for the preparation of these materials are presented in Figure 1.21. In order to accomplish the main aims of the research project, we initially set out to attain the following specific objectives: To synthesise and characterise rigid polyphosphonate and -arsonate ligands; To stabilise supramolecular coordination complexes and networks using organophosphonates and -arsonates; To monitor the formation of these compounds via mass spectrometry and to investigate the stability of these compounds in solution; To structurally characterise the compounds using single crystal and powder X-ray diffraction; 30 Chapter 1 – Introduction To investigate the magnetic properties of all relevant compounds; To determine other supplemental physicochemical properties (optical properties, thermal stability, surface areas etc.) of the compounds. AsO3H2 AsO3H2 AsO3H2 AsO3H2 NO2 NH2 PAA OH APAA HPAA OH HNPAA PO3H2 PO3H2 PO3H2 PPA BPA P-TPM PO3H2 PO3H2 H2O3P P-TPB PO3H2 H2O3P P-TBB PO3H2 Figure 1.21 – Phosphonate and arsonate ligands used to stabilise supramolecular coordination complexes and networks: PAA – phenylarsonic acid; APAA – (4-aminophenyl)arsonic acid (p-arsanilic acid); HPAA – (4-hydroxyphenyl)arsonic acid; HNPAA – (4-hydroxy-3-nitrophenyl)arsonic acid; PPA - phenylphosphonic acid; BPA – benzylphosphonic acid; P-TPM – triphenylmethylphosphonic acid; PTPB – 1,3,5-tris(4-phosphonophenyl)benzene; P-TBB – 1,3,5-tris(4`-phosphonobiphenyl-4-yl)benzene. 31 Chapter 1 – Introduction REFERENCES 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. D.-L. Long, E. Burkholder and L. Cronin, Chemical Society Reviews, 2007, 36, 105-121. B. Hasenknopf, K. Micoine, E. Lacôte, S. Thorimbert, M. Malacria and R. Thouvenot, European Journal of Inorganic Chemistry, 2008, 2008, 5001-5013. A. Proust, R. Thouvenot and P. Gouzerh, Chemical Communications, 2008, 18371852. C. L. Hill and C. M. Prosser-McCartha, Coordination Chemistry Reviews, 1995, 143, 407-455. T. Yamase, Chemical Reviews, 1998, 98, 307-326. I. Kozhevnikov and S. M. Roberts, Chichester: Wiley, 2002, 2. A. Müller, M. Luban, C. Schröder, R. Modler, P. Kögerler, M. Axenovich, J. Schnack, P. Canfield, S. Bud'ko and N. Harrison, A European Journal of Chemical Physics and Physical Chemistry, 2001, 2, 517-521. P. Wang, X. Wang and G. Zhu, Electrochimica Acta, 2000, 46, 637-641. M.-S. Wang, G. Xu, Z.-J. Zhang and G.-C. Guo, Chemical Communications, 2010, 46, 361-376. W. Qi, Y. Wang, W. Li and L. Wu, Chemistry – A European Journal, 2010, 16, 1068-1078. K. C. Dey and V. Sharma, ChemInform, 2010, 41, 368-375. Z. Wang, R. Zhang, Y. Ma, L. Zheng, A. Peng, H. Fu and J. Yao, Journal of Materials Chemistry, 2010, 20, 1107-1111. Z. Wang, R. Zhang, Y. Ma, A. Peng, H. Fu and J. Yao, Journal of Materials Chemistry, 2010, 20, 271-277. U. Kortz, A. Müller, J. van Slageren, J. Schnack, N. S. Dalal and M. Dressel, Coordination Chemistry Reviews, 2009, 253, 2315-2327. D. E. Katsoulis, Chemical Reviews, 1998, 98, 359-388. N. Mizuno and M. Misono, Chemical Reviews, 1998, 98, 199-218. A. Müller, E. Beckmann, H. Bögge, M. Schmidtmann and A. Dress, Angewandte Chemie International Edition, 2002, 41, 1162-1167. P. Mialane, A. Dolbecq and F. Secheresse, Chemical Communications, 2006, 34773485. D.-L. Long and L. Cronin, Chemistry – A European Journal, 2006, 12, 3698-3706. P. Gouzerh and A. Proust, Chemical Reviews, 1998, 98, 77-112. J. M. Breen and W. Schmitt, Angewandte Chemie International Edition, 2008, 47, 6904-6908. L. C. W. Baker and D. C. Glick, Chemical Reviews, 1998, 98, 3-50. J. W. Illingworth and J. F. Keggin, Journal of the Chemical Society, 1935, 575-580. H. T. Evans, Journal of the American Chemical Society, 1948, 70, 1291-1292. B. Dawson, Acta Crystallographica, 1953, 6, 113-126. W. A. LaRue, A. T. Liu and J. San Filippo, Inorganic Chemistry, 1980, 19, 315320. Y. Du, A. Rheingold and E. A. Maatta, Journal of the American Chemical Society, 1992, 114, 345-346. J. Strong, R. Ostrander, A. Rheingold and E. A. Maatta, Journal of the American Chemical Society, 1994, 116, 3601-3602. A. Müller, E. Krickemeyer, H. Bögge, M. Schmidtmann, C. Beugholt, P. Kögerler and C. Lu, Angewandte Chemie International Edition, 1998, 37, 1220-1223. 32 Chapter 1 – Introduction 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. A. Müller, P. Kögerler and A. W. M. Dress, Coordination Chemistry Reviews, 2001, 222, 193-218. Y.-F. Song, D.-L. Long and L. Cronin, Angewandte Chemie International Edition, 2007, 46, 3900-3904. H.-X. Guo and S.-X. Liu, Inorganic Chemistry Communications, 2004, 7, 12171220. J. Schnack, M. Brüger, M. Luban, P. Kögerler, E. Morosan, R. Fuchs, R. Modler, H. Nojiri, R. C. Rai, J. Cao, J. L. Musfeldt and X. Wei, Physical Review B, 2006, 73, 094401. R. Prudent, V. Moucadel, B. Laudet, C. Barette, L. Lafanechère, B. Hasenknopf, J. Li, S. Bareyt, E. Lacôte, S. Thorimbert, M. Malacria, P. Gouzerh and C. Cochet, Chemistry and Biology, 2008, 15, 683-692. H. N. Miras, E. F. Wilson and L. Cronin, Chemical Communications, 2009, 12971311. L. Cronin, Comprehensive Coordination Chemistry II, 2003, 7, 1-56. D.-L. Long, R. Tsunashima and L. Cronin, Angewandte Chemie International Edition, 2010, 49, 1736-1758. A. Müller, P. Kögerler and C. Kuhlmann, Chemical Communications, 1999, 13471358. H. T. Evans, B. M. Gatehouse and P. Leverett, Journal of the Chemical Society, Dalton Transactions, 1975, 505-514. A. J. Wilson, V. McKee, B. R. Penfold and C. J. Wilkins, Acta Crystallographica Section C, 1984, 40, 2027-2030. J. L. Garin and J. A. Costamagna, Acta Crystallographica, 1988, 44, 779-782. J. Fielden, K. Quasdorf, L. Cronin and P. Kogerler, Dalton Transactions, 2012, 41, 9876-9878. H. Yu, X. Zhang, L. Kong and J. Xu, Acta Crystallographica, 2009, 65, 1698-1699. Z. Han, W. Pei, J. Xie, Y. Zou and X.-M. Ren, Inorganic Chemistry Communications, 2012, 16, 61-64. A. Müller, J. Meyer, E. Krickemeyer, C. Beugholt, H. Bögge, F. Peters, M. Schmidtmann, P. Kögerler and M. J. Koop, Chemistry – A European Journal, 1998, 4, 1000-1006. A. Müller, E. Krickemeyer, J. Meyer, H. Bögge, F. Peters, W. Plass, E. Diemann, S. Dillinger, F. Nonnenbruch, M. Randerath and C. Menke, Angewandte Chemie International Edition in English, 1995, 34, 2122-2124. A. Muller, S. Q. N. Shah, H. Bogge and M. Schmidtmann, Nature, 1999, 397, 4850. A. Muller and P. Gouzerh, Chemical Society Reviews, 2012. J. J. Borrás-Almenar, E. Coronado, A. Müller and M. Pope, Polyoxometalate Molecular Science, NATO Science Series, Kluwer Academic Publishers, 2003, 98. E. Coronado and C. J. Gómez-García, Chemical Reviews, 1998, 98, 273-296. S.-T. Zheng and G.-Y. Yang, Chemical Society Reviews, 2012. J.-D. Compain, P. Mialane, A. Dolbecq, I. M. Mbomekallé, J. Marrot, F. Sécheresse, E. Rivière, G. Rogez and W. Wernsdorfer, Angewandte Chemie International Edition, 2009, 48, 3077-3081. H. Kang, S. Lui, S. Shaikh, T. Nicholson and J. Zubieta, Inorganic Chemistry, 1989, 28, 920-933. W. Clegg, R. Errington, D. Hockless and C. Radshaw, Polyhedron, 1989, 8, 17881789. R. Errington, R. Wingad, W. Clegg and M. Elsegood, Angewandte Chemie International Edition in English, 2000, 39, 3889. 33 Chapter 1 – Introduction 56. 57. 58. 59. 60. 61. 62. 63. 64. 65. 66. 67. 68. 69. 70. 71. 72. 73. 74. 75. 76. 77. 78. 79. 80. 81. 82. A. Mazeaud, N. Ammari, F. Robert and R. Thouvenot, Angewandte Chemie International Edition in English, 1996, 35, 1961-1964. F. Xin, M. T. Pope, G. J. Long and U. Russo, Inorganic Chemistry, 1996, 35, 12071213. H. Kwen, S. Tomlinson, E. A. Maatta, C. Dablemont, R. Thouvenot, A. Proust and P. Gouzerh, Chemical Communications, 2002, 2970-2971. C. Dablemont, C. G. Hamaker, R. Thouvenot, Z. Sojka, M. Che, E. A. Maatta and A. Proust, Chemistry – A European Journal, 2006, 12, 9150-9160. J. B. Strong, B. S. Haggerty, A. L. Rheingold and E. A. Maatta, Chemical Communications, 1997. A. Dolbecq, E. Dumas, C. d. R. Mayer and P. Mialane, Chemical Reviews, 2010, 110, 6009-6048. A. Müller, K. Hovemeier, E. Krickemeyer and H. Bögge, Angewandte Chemie International Edition in English, 1995, 34, 779-781. B. J. S. Johnson, S. A. Geers, W. W. Brennessel, J. V. G. Young and A. Stein, Dalton Transactions, 2003, 4678-4681. C. Sassoye, K. Norton and S. C. Sevov, Inorganic Chemistry, 2003, 42, 1652-1655. N. Calin and S. C. Sevov, Inorganic Chemistry, 2003, 42, 7304-7308. A. Müller, K. Hovemeier and R. Rohlfing, Angewandte Chemie, 1992, 104, 12141216. A. Clearfield and K. Demadis, Royal Society of Chemistry, 2011. P. DeBurgomaster, A. Aldous, H. Liu, C. J. O’Connor and J. Zubieta, Crystal Growth & Design, 2010, 10, 2209-2218. Y.-D. Chang and J. Zubieta, Inorganica Chimica Acta, 1996, 245, 177-198. D.-G. Lyxell, D. Bostrom, M. Hashimoto and L. Pettersson, Acta Crystallographica Section B, 1998, 54, 424-430. T. Ueda, T. Yonemura, M. Shiro, M. Fukudome and M. Hojo, Inorganic Chemistry Communications, 2007, 10, 1301-1303. A. Gammersbach, P. Angenault, W. F. Hölderich, W. Clegg and M. R. J. Elsegood, Polyhedron, 1998, 17, 2535-2541. M. I. Khan, Q. Chen and J. Zubieta, Inorganica Chimica Acta, 1995, 235, 135-145. A. Dolbecq, L. Lisnard, P. Mialane, J. Marrot, M. Bénard, M.-M. Rohmer and F. Sécheresse, Inorganic Chemistry, 2006, 45, 5898-5910. C. d. Peloux, A. Dolbecq, P. Mialane, J. Marrot and F. Secheresse, Dalton Transactions, 2004, 1259-1263. R. C. Finn, E. Burkholder and J. Zubieta, Chemical Communications, 2001, 18521853. E. Dumas, C. Sassoye, K. D. Smith and S. C. Sevov, Inorganic Chemistry, 2002, 41, 4029-4032. W. Kwak, L. M. Rajkovic, J. K. Stalick, M. T. Pope and C. O. Quicksall, Inorganic Chemistry, 1976, 15, 2778-2783. K. M. Barkigia, L. M. Rajkovic-Blazer, M. T. Pope and C. O. Quicksall, Inorganic Chemistry, 1981, 20, 3318-3323. Y.-Q. Lan, S.-L. Li, X.-L. Wang, K.-Z. Shao, D.-Y. Du, H.-Y. Zang and Z.-M. Su, Inorganic Chemistry, 2008, 47, 8179-8187. J. L. C. Rowsell, E. C. Spencer, J. Eckert, J. A. K. Howard and O. M. Yaghi, Science, 2005, 309, 1350-1354. O. K. Farha, I. Eryazici, N. C. Jeong, B. G. Hauser, C. E. Wilmer, A. A. Sarjeant, R. Q. Snurr, S. T. Nguyen, A. Ö. Yazaydın and J. T. Hupp, Journal of the American Chemical Society, 2012, 134, 15016-15021. 34 Chapter 1 – Introduction 83. 84. 85. 86. 87. 88. 89. 90. 91. 92. 93. 94. 95. 96. 97. 98. 99. 100. 101. 102. 103. 104. 105. 106. 107. 108. 109. O. M. Yaghi, M. O'Keeffe, N. W. Ockwig, H. K. Chae, M. Eddaoudi and J. Kim, Nature, 2003, 423, 705. M. Eddaoudi, J. Kim, N. Rosi, D. Vodak, J. Wachter, M. O'Keeffe and O. M. Yaghi, Science, 2002, 295, 469-472. H. Furukawa, N. Ko, Y. B. Go, N. Aratani, S. B. Choi, E. Choi, A. Ö. Yazaydin, R. Q. Snurr, M. O’Keeffe, J. Kim and O. M. Yaghi, Science, 2010, 329, 424-428. D. J. Tranchemontagne, J. L. Mendoza-Cortes, M. O'Keeffe and O. M. Yaghi, Chemical Society Reviews, 2009, 38, 1257-1283. A. R. Millward and O. M. Yaghi, Journal of the American Chemical Society, 2005, 127, 17998-17999. L. Yuan, C. Qin, X. Wang, Y. Li and E. Wang, Dalton Transactions, 2009, 41694175. Xiang, Wu, ZhangZhang, Fu, Hu and ZhangZhang, Journal of the American Chemical Society, 2005, 127, 16352-16353. A. El-Aneed, A. Cohen and J. Banoub, Applied Spectroscopy Reviews, 2009, 44, 210-230. R. Colton, A. D'Agostino and J. C. Traeger, Mass Spectrometry Reviews, 1995, 14, 79-106. D. Schaumlöffel and A. Tholey, Analytical and Bioanalytical Chemistry, 2011, 400, 1645-1652. M. J. Keith-Roach, Analytica Chimica Acta, 2010, 678, 140-148. I. I. Stewart, Spectrochimica Acta Part B: Atomic Spectroscopy, 1999, 54, 16491695. R. Jirásko and M. Hol apek, Mass Spectrometry Reviews, 2011, 30, 1013-1036. B. F. G. Johnson J and S. McIndoe, Coordination Chemistry Reviews, 2000, 200– 202, 901-932. R. Colton and J. C. Traeger, Inorganica Chimica Acta, 1992, 201, 153-155. E. F. Wilson, H. Abbas, B. J. Duncombe, C. Streb, D.-L. Long and L. Cronin, Journal of the American Chemical Society, 2008, 130, 13876-13884. E. F. Wilson, H. N. Miras, M. H. Rosnes and L. Cronin, Angewandte Chemie International Edition, 2011, 50, 3720-3724. H. N. Miras, M. Sorus, J. Hawkett, D. O. Sells, E. J. L. McInnes and L. Cronin, Journal of the American Chemical Society, 2012, 134, 6980-6983. M. J. Deery, O. W. Howarth and K. R. Jennings, Journal of the Chemical Society, Dalton Transactions, 1997, 4783-4788. I.-R. Jeon and R. Clerac, Dalton Transactions, 2012, 41, 9569-9586. T. Glaser, Chemical Communications, 2011, 47, 116-130. J. R. Long, in The Chemistry Of Nanostructured Materials, World Scientific Publishing: Hong Kong, Editon edn., 2003, pp. 291-315. A. J. Tasiopoulos, A. Vinslava, W. Wernsdorfer, K. A. Abboud and G. Christou, Angewandte Chemie International Edition, 2004, 43, 2117-2121. L. Thomas, F. Lionti, R. Ballou, D. Gatteschi, R. Sessoli and B. Barbara, Nature, 1996, 383, 145-147. J. T. Brockman, J. C. Huffman and G. Christou, Angewandte Chemie International Edition, 2002, 41, 2506-2508. M. A. Bolcar, S. M. J. Aubin, K. Folting, D. N. Hendrickson and G. Christou, Chemical Communications, 1997, 1485-1486. M. Murugesu, M. Habrych, W. Wernsdorfer, K. A. Abboud and G. Christou, Journal of the American Chemical Society, 2004, 126, 4766-4767. 35 Chapter 1 – Introduction 110. 111. 112. 113. 114. 115. 116. 117. 118. 119. 120. 121. 122. 123. 124. 125. 126. 127. 128. 129. A. K. Powell, S. L. Heath, D. Gatteschi, L. Pardi, R. Sessoli, G. Spina, F. Del Giallo and F. Pieralli, Journal of the American Chemical Society, 1995, 117, 24912502. J. C. Goodwin, R. Sessoli, D. Gatteschi, W. Wernsdorfer, A. K. Powell and S. L. Heath, Journal of the Chemical Society, Dalton Transactions, 2000, 1835-1840. C. Sangregorio, T. Ohm, C. Paulsen, R. Sessoli and D. Gatteschi, Physical Review Letters, 1997, 78, 4645-4648. Z. Sun, D. N. Hendrickson, Z. Sun, C. M. Grant, S. L. Castro and G. Christou, Chemical Communications, 1998, 721-722. A. Barra, D. Gatteschi, L. Pardi, A. Müller and J. Döring, Journal of the American Chemical Society, 1992, 114, 8509-8514. A. Müller, R. Sessoli, E. Krickemeyer, H. Bögge, J. Meyer, D. Gatteschi, L. Pardi, J. Westphal, K. Hovemeier, R. Rohlfing, J. Döring, F. Hellweg, C. Beugholt and M. Schmidtmann, Inorganic Chemistry, 1997, 36, 5239-5250. D. Gatteschi, L. Pardi, A. L. Barra, A. Muller and J. Doring, Nature, 1991, 354, 463-465. A. M. Ako, V. Mereacre, R. Clerac, W. Wernsdorfer, I. J. Hewitt, C. E. Anson and A. K. Powell, Chemical Communications, 2009, 544-546. R. Sessoli and A. K. Powell, Coordination Chemistry Reviews, 2009, 253, 23282341. C. Papatriantafyllopoulou, W. Wernsdorfer, K. A. Abboud and G. Christou, Inorganic Chemistry, 2011, 50, 421-423. J. R. Long, Chemistry of Nanostructured Materials, 2003, 291-315. T. Lis, Acta Crystallographica Section B, 1980, 36, 2042-2046. J. R. Friedman, M. P. Sarachik, J. Tejada and R. Ziolo, Physical Review Letters, 1996, 76, 3830-3833. R. Sessoli, D. Gatteschi, A. Caneschi and M. A. Novak, Nature, 1993, 365, 141143. C. J. Milios, A. Vinslava, W. Wernsdorfer, S. Moggach, S. Parsons, S. P. Perlepes, G. Christou and E. K. Brechin, Journal of the American Chemical Society, 2007, 129, 2754-2755. A. M. Ako, I. J. Hewitt, V. Mereacre, R. Clérac, W. Wernsdorfer, C. E. Anson and A. K. Powell, Angewandte Chemie International Edition, 2006, 45, 4926-4929. G. E. Kostakis, A. M. Ako and A. K. Powell, Chemical Society Reviews, 2010, 39, 2238-2271. Y.-S. Ma, Y. Song, Y.-Z. Li and L.-M. Zheng, Inorganic Chemistry, 2007, 46, 5459-5461. M. Shanmugam, M. Shanmugam, G. Chastanet, R. Sessoli, T. Mallah, W. Wernsdorfer and R. E. P. Winpenny, Journal of Materials Chemistry, 2006, 16, 2576-2578. M. Wang, C. Ma, H. Wen and C. Chen, Dalton Transactions, 2009, 994-1003. 36 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates 2. HYBRID ORGANIC-INORGANIC POLYOXOMOLYBDATES 37 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates 2.1 INTRODUCTION AND MOTIVATION The chemistry of the polyoxometalates (POMs) is one of the most active and rapidly advancing areas of inorganic chemistry.1-4 Functionalisation approaches to generate hybrid polyoxometalate structures or advanced inorganic materials remain an ongoing challenge and new synthetic and analytical approaches are pivotal to the progress of the field.2-5 We are interested in the chemistry of hybrid organic-inorganic materials, and we decided to explore the formation of mixed-valent molybdates in the presence of organoarsonates and phosphonates. Hybrid organic-inorganic polyoxomolybdates would combine the characteristics of both the inorganic oxo clusters and the organic moieties resulting in new products with potentially new or even enhanced properties. Organophosphonate and -arsonate stabilised oxo clusters that contain redox-active metal centres are of interest for applications in catalysis.2, 6, 7 A limited number of organoarsonate and -phosphonate stabilised molybdates have previously been reported in the literature.8-19 Despite an ever increasing interest in POM complexes, it is surprising that their solution behaviour, i.e. formation reactions that often prevail through a set of consecutive condensation reactions, are often fairly poorly understood. However, speciation within POM solutions is not only important to devise rational synthetic approaches to desired products but it is moreover a pre-requisite to improve catalytic processes that draw on the active sites or electronic characteristics of POMs. Recent accounts demonstrate that electrospray ionisation mass spectrometry (ESI-MS) provides a very powerful analytical tool to characterise the formation and the reactivity of complex polyoxometalate clusters in solution.20-33 We decided to use electrospray ionisation mass spectrometry (ESI-MS) to investigate the self-assembly process of hybrid polyoxomolybdates that form upon partial reduction of (NH4)6Mo7O24·4H2O in the presence of aromatic organoarsonates. We were interested in exploring how perturbations of the ligand functionality influence the formation of unprecedented species, and we used ESI-MS to screen the reaction mixtures prior to crystallisation attempts. Although, several arsonate stabilised molybdates have previously been reported,9, 11, 34-39 this mass spectrometry-guided approach allowed us to 38 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates selectively identify species with unprecedented core structures. The structure and composition of these species were later confirmed by single crystal XRD studies. The first part of this chapter presents the synthesis and characterisation of four unprecedented hybrid cluster compounds stabilised by organoarsonate ligands: (NH4)2H2[MoV4O8(O3AsC6H5)4]·5H2O (1), (NH4)2H2[MoV4O8(O3AsC6H4NH2)4]·DMF·4H2O (2), (NH4)5[MoVI2MoV3O11(O3AsC6H4OH)5]·9H2O (3), (NH4)4[MoVI4O10(O3AsC6H3NO2OH)4]·2H2O (4). Stimulated by the interest in transition metal functionalised POMs and in so-called “sandwiched atoms” involving POMs,40 we additionally decided to explore how d-block hetero-atoms can be incorporated into defined POM structures. A known molybdenum phosphonate is (NH4)5Na4{Na[Mo6O12(OH)3(O3PC6H5)4)]2} 6H2O, in which two {Mo6} fragments are linked through a Na ion.13 Considering this, we sought to investigate if transition metal ions could be used to link the hexanuclear subunits to produce sandwichtype compounds. We intended to explore how different heteroatoms would influence the structure and magnetic properties of the compounds. This part of the chapter presents the synthetic approach, the structure characterisation and properties of five isostructural compounds (4 transition metal derivatives and 1 main group compound) that self-assemble under closely related reaction conditions: (NH4)4H4{Mn[MoV6O12(OH)3(O3PC6H5)4]2} 8H2O (5), (NH4)4H4{Fe[MoV6O12(OH)3(O3PC6H5)4]2} 8H2O (6), (NH4)4H4{Co[MoV6O12(OH)3(O3PC6H5)4]2} 8H2O (7), (NH4)4H4{Ni[MoV6O12(OH)3(O3PC6H5)4]2} 8H2O (8), (NH4)4H4{Mg[MoV6O12(OH)3(O3PC6H5)4]2} 8H2O (9). 39 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates 2.2 FUNCTIONALISATION OF POLYOXOMOLYBDATES USING ORGANOARSONATES AND -PHOSPHONATES 2.2.1 Synthesis and characterisation of organoarsonate functionalised polyoxomolybdate clusters 2.2.1.1 The Phenylarsonic Acid – MoVI/MoV Reaction System (NH4)2H2[MoV4O8(O3AsC6H5)4]·5H2O (1) Partial reduction of (NH4)6Mo7O24 4H2O in the presence of phenylarsonic acid and acetic acid using N2H4 H2SO4 results in a deep blue solution from which dark red crystals of (NH4)2H2[MoV4O8(O3AsC6H5)4]·5H2O (1) separate within a time period of three weeks. Single crystal X-ray analysis reveals that 1 crystallises in the triclinic crystal system in the space group P and contains a tetranuclear MoV complex in which the Mo atoms and bridging O donors adopt a typical [Mo4(µ 3-O)4]12+ cubane structure. The structure is shown in Figure 2.1. In this structure, the four Mo atoms are situated diagonally across from each other, occupying four corners of a cube. Four O2- oxo ligands occupy the remaining four corners of the cube, and each one binds to three Mo atoms. The structure is further stabilised by four fully deprotonated phenylarsonic acid ligands that each bind in a η1:η1:μ2 bridging mode to two Mo ions, resulting in an overall octanuclear compound (the As atoms are also considered to be core atoms). The slightly distorted octahedral coordination environment of each of the four Mo ions in 1 comprises of three µ 3-O oxo ligands, two O donors originating from two different arsonate ligands and is completed by a terminal Mo=O bond. The distorted nature of the octahedra arises from the geometrical restrictions of the binding ligands and can be observed from the bond angles and bond lengths of the MoV metal centres. The bond lengths of the Mo=O bonds range between 1.660(4) Å – 1.679(4) Å and are, as expected, significantly shorter than the Mo–(µ 3-O) distances of 1.969(4) Å – 2.424(4) Å and the Mo-Oarsonate bond distances of 2.043(4) Å – 2.090(4) Å, (Table 2.1). The following bond angles: O(12)-Mo(1)-O(6), O(7)-Mo(1)-O(10) and O(8)Mo(1)-O(11) of 170.50(17)°, 152.42(14)° and 151.57(14)°, respectively, deviate from the ideal octahedral angle of 180°, while the bond angles O(12)-Mo(1)-O(7), O(12)-Mo(1)O(11), O(8)-Mo(1)-O(7), O(7)-Mo(1)-O(11), O(8)-Mo(1)-O(10), O(7)-Mo(1)-O(6) and 40 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates O(8)-Mo(1)-O(6) of 110.67(16)°, 97.79(19)°, 88.68(17)°, 86.42(17)°, 85.21(16)°, 76.37(14)° and 75.74(15)°, respectively, deviate from the ideal angle of 90°. Similar values can be found for the bond angles of the Mo(2), Mo(3) and Mo(4) ions, which are summarized in Table 2.2. a) b) Figure 2.1 – Crystal structure of the tetranuclear Mo complex in 1: (a) ball-and-stick representation, (b) polyhedral presentation. Colour code: MoV blue, As orange, O red, C grey, H white (hydrogen atoms have been omitted for clarity in b). The [Mo4(µ 3-O)4]12+ core structure deviates slightly from the geometry of an ideal cube. The Mo–O–Mo and O–Mo–O angles range between 83.82(3)º – 104.04(1)º and 76.18(1)º – 89.72(3)º, respectively, differing from the ideal angle of 90º. Bond valence sum analysis41 confirms that all four Mo atoms in 1 adopt the oxidation state +V. Short Mo(1)Mo(2) and Mo(3)-Mo(4) contacts of 2.647(1) Å and 2.642(1) Å, respectively, are in agreement with the assigned +V oxidation states. The cluster is further stabilised by weak intramolecular - interactions42 with an interplanar separation distance that range between 3.604(21) Å – 4.091(19) Å. The packing arrangement of the [MoV4O8(O3AsC6H5)4]4- clusters in the crystal structure is stabilised by weak hydrogen bonds between the cluster anions and constitution water molecules (O-O distances: 2.769(48) Å – 2.915(57) Å) and also between the cluster anions and NH4+ counterions (N-O distances: 2.698(36) Å – 3.045(30) Å). The solvent molecules are situated in small channels that run in the direction of the crystallographic aaxis (Figure 2.3). Very weak intermolecular - interactions (shortest contact 3.975(28) Å) of the aromatic rings of the arsonate ligands also contribute to the stability of 1 in the solid state. 41 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates Figure 2.2 – The core structure of the tetranuclear Mo complex in 1. Colour code: MoV blue, As orange, O red, C grey. a) b) Figure 2.3 – Packing arrangement of 1 in the crystal structure viewed in the direction of the crystallographic: (a) a-axis and (b) b-axis. Colour code: MoV blue, As orange, O red, C grey (crystallization water molecules, NH4+ counterions and hydrogen atoms have been omitted for clarity). 42 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates Table 2.1 − Selected bond lengths [Å] and bond valence sum (BVS) for compound 1. Atom Bond Bond distances (Å) BVS Assigned oxidation state Mo(1) Mo(1)-O(12) Mo(1)-O(7) Mo(1)-O(8) Mo(1)-O(11) Mo(1)-O(10) Mo(1)-O(6) 1.660(4) 1.989(4) 1.983(4) 2.055(4) 2.065(4) 2.405(4) 5.159 +5 Mo(2) Mo(2)-O(19) Mo(2)-O(7) Mo(2)-O(8) Mo(2)-O(15) Mo(2)-O(16) Mo(2)-O(5) 1.673(4) 1.974(4) 1.986(4) 2.043(4) 2.046(4) 2.424(4) 5.161 +5 Mo(3) Mo(3)-O(2) Mo(3)-O(5) Mo(3)-O(6) Mo(3)-O(14) Mo(3)-O(9) Mo(3)-O(8) 1.667(4) 1.970(4) 1.983(4) 2.047(4) 2.090(4) 2.364(4) 5.165 +5 Mo(4) Mo(4)-O(1) Mo(4)-O(5) Mo(4)-O(6) Mo(4)-O(3) Mo(4)-O(4) Mo(4)-O(7) 1.679(4) 1.969(4) 1.972(4) 2.066(4) 2.087(4) 2.386(4) 5.085 +5 Mo(1) ··· Mo(2) Mo(3) ··· Mo(4) 2.647(1) 2.642(1) Structurally related cubane structures were previously isolated using squaric acid ((ColMe)4[Mo4O8(C4O4)4]·2MeOH·2Col, (ColH)4(PyEt)[Mo4O8(C4O4)4]Br; Col = 2,4,6collidine), diphenylphosphinic acid (Mo4(µ 3-O)4(µ-O2PPh2)4O4) and dimethylphosphinic acid (Mo4(µ 3-O)4(µ-O2PMe2)4O4) as ligands.43-45 However, it is noteworthy that the cubane arrangement of MoV atoms is far less frequently observed than the structurally related rhombic planar arrangement.43 43 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates Table 2.2 − Selected bond angles [º] for compound 1. Bond Angle (º) Bond Angle (º) O(12)-Mo(1)-O(6) O(7)-Mo(1)-O(10) O(8)-Mo(1)-O(11) O(12)-Mo(1)-O(7) O(12)-Mo(1)-O(11) O(8)-Mo(1)-O(7) O(7)-Mo(1)-O(11) O(8)-Mo(1)-O(10) O(7)-Mo(1)-O(6) O(8)-Mo(1)-O(6) 170.50(17) 152.42(14) 151.57(14) 110.67(16) 97.79(19) 88.68(17) 86.42(17) 85.21(16) 76.37(14) 75.74(15) O(19)-Mo(2)-O(5) O(7)-Mo(2)-O(16) O(8)-Mo(2)-O(15) O(19)-Mo(2)-O(7) O(19)-Mo(2)-O(15) O(19)-Mo(2)-O(16) O(7)-Mo(2)-O(8) O(8)-Mo(2)-O(16) O(16)-Mo(2)-O(5) O(8)-Mo(2)-O(5) 171.66(17) 154.38(15) 152.70(14) 109.20(17) 98.17(17) 96.04(18) 89.02(16) 86.95(16) 77.54(15) 76.61(14) O(2)-Mo(3)-O(8) O(5)-Mo(3)-O(9) O(6)-Mo(3)-O(14) O(2)-Mo(3)-O(6) O(2)-Mo(3)-O(9) O(5)-Mo(3)-O(6) O(6)-Mo(3)-O(9) O(14)-Mo(3)-O(9) O(5)-Mo(3)-O(8) O(14)-Mo(3)-O(8) 170.03(19) 154.78(15) 154.28(16) 109.62(19) 95.60(18) 89.45(16) 88.05(17) 84.41(17) 78.38(15) 77.62(14) O(1)-Mo(4)-O(7) O(5)-Mo(4)-O(4) O(6)-Mo(4)-O(3) O(1)-Mo(4)-O(6) O(1)-Mo(4)-O(3) O(5)-Mo(4)-O(6) O(5)-Mo(4)-O(3) O(3)-Mo(4)-O(4) O(5)-Mo(4)-O(7) O(3)-Mo(4)-O(7) 170.26(17) 154.73(14) 152.73(14) 109.66(17) 96.97(18) 89.81(18) 87.12(16) 82.92(16) 77.92(15) 75.73(15) 44 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates - PHYSICOCHEMICAL CHARACTERISATION - FT-IR spectroscopy Infrared spectroscopy was used as a basic tool for preliminary characterisation of compound 1. The IR spectrum of 1 (Figure 2.4) reveals some characteristic stretches of the organoarsonate ligands. These include signals at 1093 cm-1 attributed to As–C vibrations and 1439 cm-1 associated with C–C skeletal vibrations of the phenyl rings. Additionally observed bands arising from the molybdenum-oxygen and arsenic-oxygen stretching vibrations appear in the 1000 – 650 cm-1 region.8, 15, 37, 46-48 Figure 2.4 – Infrared spectrum of 1. - Thermogravimetric analysis The thermal stability of compound 1 was investigated by thermogravimetric analysis (TGA) using a freshly prepared crystalline sample. The analysis was carried out in the temperature range between 30 and 900 °C, in an N2 atmosphere. The TGA of 1 reveals a weight loss of 6.4 % in the temperature range between 30 – 90 °C. This 45 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates thermogravimetric step can be attributed to the loss of five water molecules (calcd: 6.2 %). A further increase in the temperature up to 500 °C produces a weight loss that occurs in a two step process. This weight loss can be associated with the decomposition of the organic ligands of the cluster. The weight loss centered at ca. 770 °C is most likely being associated with the formation of oxides. 100 Weight % (%) 90 80 70 60 50 40 30 0 200 400 600 800 1000 Temperature (°C) Figure 2.5 – Thermogravimetric analysis of 1. - Mass spectrometry The stability of compound 1 in solution was investigated by electrospray ionization mass spectrometry (ESI-MS). The negative mode ESI-MS spectrum of compound 1 dissolved in DMSO is presented in Figure 2.6. The mass spectrum displays only one major isotopic envelope in the high molecular mass region, centered at m/z = 1315.3 a.m.u. The signal was assigned to the {H3[MoV4O8(O3AsC6H5)4]}- species. The expected signal was modelled (Figure 2.6, Inset) and good fits between the experimental and simulated isotopic envelopes confirm the assignment. 46 Relative ionic abundance Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates Figure 2.6 – Negative-mode ESI-MS spectra for crystals of 1 dissolved in DMSO. Inset: Comparison of the experimental isotopic envelopes (black spectrum) with simulated patterns (red spectrum) for {H3[MoV4O8(O3AsC6H5)4]}- centered at m/z = 1315.3 a.m.u. (cone voltage: 30 V). - NMR spectroscopy The 1H NMR spectrum of 1 recorded in deuterated DMSO further confirms the stability of the compound in DMSO environment. The spectrum of 1 exhibits resonance shifts at 7.97 ppm (d, 8H) and 7.67 ppm (m, 12H) corresponding to aromatic H-atoms of the arsonate ligands. A section of the 1H NMR spectrum of 1 is presented in Figure 2.7. 1 2 Figure 2.7 – A section of the 1H NMR (400MHz, DMSO) spectrum showing the assignment for 1. 47 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates - UV-Vis spectroscopy The UV-Vis absorption spectrum of 1 was recorded in DMF (Figure 2.8). The spectrum displays an absorption band at ca. 457 nm ( = 1000 L mol-1 cm-1) due to d – d transitions involving the MoV metal centeres. These d – d transitions can be assigned to a Eg 2 T2g transition for the octahedral MoV ions within 1.38 0.12 Absorbance (a.u.) 2 0.10 0.08 0.06 0.04 0.02 400 450 500 550 600 Wavelength (nm) Figure 2.8 –A section of the UV-Vis spectrum of a 10-4 M solution of 1 in DMF. 48 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates Table 2.3 − Crystal data and structural refinement parameters for 1. Compound 1 Empirical formula Molecular mass/g mol C24H40As4Mo4N2O25 -1 Crystal colour/shape 3 1440.02 Red / triangular plate Crystal size/mm 0.20×0.20×0.05 Crystal system Triclinic Space group P a/ Å 10.009(4) b/ Å 15.767(5) c/ Å 17.287(5) /º 117.09(2) /º 96.03(2) /º 98.454(18) 3 V/ Å 2356.1(14) Z 2 Temperature (K) 118(2) -3 Density/Mg m 2.001 -1 Absorp. coef./mm 3.906 F(000) 1356 2 50 max/º Reflections collected 36989 Independent reflections 8303 [R(int) = 0.1174] Data / restraints / parameters 8303 / 2128 / 577 2 S on F 1.036 R1, wR2 [I>2 (I)] 0.0985, 0.2684 R1, wR2 (all data) 0.1171, 0.2869 Largest diff. peak and hole/e.Å-3 2.217 and -2.503 49 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates 2.2.1.2 The p- Aminophenylarsonic Acid – MoVI/MoV Reaction System (NH4)2H2[MoV4O8(O3AsC6H4NH2)4]·DMF·4H2O (2) Encouraged by these results we decided to alter the ligand functionality and introduced an amino group in the para position to the arsonate functionality. When this ligand is used to control the self-assembly process upon reduction of (NH4)6Mo7O24 4H2O, a blue solution is obtained, as in the previous case. During a time period of one week, rodshaped red crystals of (NH4)2H2[MoV4O8(O3AsC6H4NH2)4] ·DMF·4H2O (2) were obtained. 2 was characterised by single crystal X-ray diffraction measurements. 2 crystallises in the triclinic crystal system in the space group P . The core structure within 2 (Figure 2.10) is almost isostructural to the cluster core found in 1, having only slightly different structural and geometrical parameters. The bond valence sums and underlying M−O and Mo−Mo distances agree very well with those observed for 1. Selected interatomic distances and angles for compound 2 are given in Table 2.4 and Table 2.5, respectively. Similar to 1 the core structure within 2, [Mo4(µ 3-O)4]12+, deviates slightly from the geometry of an ideal cube. The Mo–O–Mo and O–Mo–O angles range between 83.19(29)º – 104.78(30)º and 75.16(27)º – 90.33(30)º, respectively, differing from the ideal angle of 90º. a) b) Figure 2.9 – Crystal structure of the tetranuclear Mo complex in 2: (a) ball-and-stick representation, (b) polyhedral presentation. Colour code: MoV blue, As orange, O red, N light blue, C grey, H white (hydrogen atoms have been omitted for clarity in b). 50 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates Figure 2.10 – The core structure of the tetranuclear Mo complex in 2. Colour code: MoV blue, As orange, O red, C grey. a) b) Figure 2.11 – Packing arrangement of 2 in the crystal structure viewed in the direction of the crystallographic: (a) a-axis and (b) b-axis. Colour code: MoV blue, As orange, O red, N light blue, C grey (crystallisation solvent molecules, NH4+ counterions and hydrogen atoms have been omitted for clarity). The amine functionalities of the organic ligands only impose an influence on the packing arrangement of the clusters in the crystal structure, resulting in a grid-like packing arrangement, with small intercluster cavities (filled with solvent molecules and NH4+ counterions) running in the direction of the crystallographic a-axis (Figure 2.11). The 51 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates packing arrangement of the clusters in the crystal structure is stabilised by weak hydrogen bonds between the cluster anions and constitution solvent molecules (O-O distances: 2.847(21) Å – 3.060(24) Å) and also between the cluster anions and NH4+ counterions (NO distances: 2.861(19) Å – 3.083(23) Å). Very weak offset π-π interactions42 between two adjacent clusters with interplanar separation distances that range between 3.475(29) Å – 4.145(32) Å also contribute to the stability of 2 in the solid state. The protonation of the tetranuclear Mo clusters in 1 and 2, respectively, are in agreement with the elemental analysis and the constutional representation in the form of Hx[clustercore] is commonly applied in this field of research.9 Table 2.4 − Selected bond lengths [Å] and bond valence sum (BVS) for compound 2. Atom Bond Bond distances (Å) BVS Assigned oxidation state Mo(1) Mo(1)-O(17) Mo(1)-O(18) Mo(1)-O(13) Mo(1)-O(7) Mo(1)-O(6) Mo(1)-O(14) 1.679(9) 1.978(8) 1.976(7) 2.024(8) 2.079(8) 2.420(8) 5.122 +5 Mo(2) Mo(2)-O(19) Mo(2)-O(15) Mo(2)-O(14) Mo(2)-O(2) Mo(2)-O(1) Mo(2)-O(13) 1.663(9) 1.966(7) 1.967(8) 2.060(8) 2.084(8) 2.353(8) 5.227 +5 Mo(3) Mo(3)-O(20) Mo(3)-O(14) Mo(3)-O(15) Mo(3)-O(8) Mo(3)-O(12) Mo(3)-O(18) 1.678(8) 1.978(8) 1.985(8) 2.036(8) 2.067(8) 2.411(8) 5.112 +5 Mo(4) Mo(4)-O(16) Mo(4)-O(13) Mo(4)-O(18) Mo(4)-O(11) Mo(4)-O(3) Mo(4)-O(15) 1.674(9) 1.984(7) 1.983(8) 2.044(7) 2.074(8) 2.393(8) 5.110 +5 Mo(1) ··· Mo(4) Mo(2) ··· Mo(3) 2.630(1) 2.615(1) 52 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates Table 2.5 − Selected bond angles [º] for compound 2. Bond Angle (º) Bond Angle (º) O(17)-Mo(1)-O(14) O(18)-Mo(1)-O(6) O(13)-Mo(1)-O(7) O(17)-Mo(1)-O(13) O(17)-Mo(1)-O(7) O(17)-Mo(1)-O(6) O(13)-Mo(1)-O(18) O(7)-Mo(1)-O(6) O(13)-Mo(1)-O(6) O(6)-Mo(1)-O(14) 171.1(3) 155.5(3) 152.0(3) 109.6(4) 97.9(3) 95.0(3) 89.5(3) 87.4(3) 85.4(3) 78.0(3) O(19)-Mo(2)-O(13) O(14)-Mo(2)-O(2) O(15)-Mo(2)-O(1) O(19)-Mo(2)-O(14) O(19)-Mo(2)-O(1) O(15)-Mo(2)-O(14) O(2)-Mo(2)-O(1) O(15)-Mo(2)-O(2) O(2)-Mo(2)-O(13) O(15)-Mo(2)-O(13) 170.0(3) 155.9(3) 154.7(3) 109.4(4) 95.7(4) 90.6(3) 88.1(3) 85.7(3) 78.7(3) 77.5(3) O(20)-Mo(3)-O(18) O(15)-Mo(3)-O(8) O(14)-Mo(3)-O(12) O(20)-Mo(3)-O(15) O(20)-Mo(3)-O(12) O(20)-Mo(3)-O(8) O(14)-Mo(3)-O(15) O(15)-Mo(3)-O(12) O(14)-Mo(3)-O(8) O(8)-Mo(3)-O(18) 171.4(3) 155.2(3) 153.9(3) 108.9(4) 97.0(4) 95.6(4) 89.7(3) 86.8(3) 85.8(3) 78.8(3) O(16)-Mo(4)-O(15) O(18)-Mo(4)-O(3) O(13)-Mo(4)-O(11) O(16)-Mo(4)-O(13) O(16)-Mo(4)-O(11) O(16)-Mo(4)-O(3) O(18)-Mo(4)-O(13) O(3)-Mo(4)-O(15) O(18)-Mo(4)-O(15) O(11)-Mo(4)-O(15) 171.5(3) 155.9(3) 152.0(3) 109.3(4) 98.2(3) 94.7(4) 89.1(3) 79.0(3) 76.9(3) 75.8(3) 53 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates - PHYSICOCHEMICAL CHARACTERISATION - FT-IR spectroscopy The IR spectrum of 2 is presented in Figure 2.12 and reveals some characteristic stretches of the organic ligands. The set of bands observed between 1600 – 1400 cm-1 can be attributed to C–C skeletal vibrations of the phenyl rings while the signal at 1095 cm-1 can be associated with As–C vibrations. Other characteristic stretches observed in the IR spectrum of 2 appear as strong bands in the 1000 – 650 cm-1 region. Some of these bands can be attributed to the molybdenum-oxygen and arsenic-oxygen stretching vibrations. The O–H stretching vibrations and H–O–H bending vibrations of the crystallisation water molecules engaged in H-bonds appear as broad bands centered at ca. 3200 cm-1 and ca. 1630 cm-1, respectively. Stretches relating to NH2 group of the organic ligands are expected to appear in the same region as the O–H stretches of the crystallisation water molecules.8, 15, 37, 46-48 Figure 2.12 – Infrared spectrum of 2. 54 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates - Thermogravimetric analysis The thermal stability of 2 was investigated by thermogravimetric analysis using a freshly prepared crystalline sample (Figure 2.13). Compound 2 reveals a weight loss of about 9.7 % below 150 °C corresponding to the loss of four water and one DMF molecules (calcd: 9.3 %). From 150 to 600 °C a gradual weight loss is associated with the decomposition of organic moieties finally producing oxide materials. 100 Weight % (%) 80 60 40 20 0 0 100 200 300 400 500 600 700 Temperature (°C) Figure 2.13 – Thermogravimetric analysis of 2. - Mass spectrometry ESI-MS studies on pristine crystals of 2 dissolved in DMSO were performed in order to investigate the stability of the molybdenum cluster in solution. Similar to compound 1, the mass spectrum of 2 (Figure 2.14) reveals only one major isotopic envelope in the high molecular mass region of the spectrum. The signal is centered at m/z = 1374.4 a.m.u. and corresponds to the {H3[MoV4O8(O3AsC6H4NH2)4]}- species. Again, the signal of the species was modelled and a comparison of the experimental isotopic envelopes with the simulated envelope is presented in Figure 2.14, Inset. It is noteworthy that both compounds 1 and 2 are insoluble in H2O. 55 Relative ionic abundance Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates Figure 2.14 – Negative-mode ESI-MS spectra for crystals of 2 dissolved in DMSO. Inset: Comparison of the experimental isotopic envelopes (black spectrum) with simulated patterns (red spectrum) for {H3[MoV4O8(O3AsC6H4NH2)4]}- centered at m/z = 1374.4 a.m.u. (cone voltage: 30 V). - NMR spectroscopy Similar to 1, the cluster of 2 appears to be stable in a DMSO environment. The 1H NMR spectrum of 2 (Figure 2.15) exhibits signals at 7.56 ppm (d, 8H), 6.73 ppm (d, 8H) and 5.80 ppm (s, 8H). The two most downfield signals were assigned to the aromatic Hatoms of the (4-aminophenyl)arsonate ligands, while the signal at 5.80 ppm was assigned to the amine H-atoms of the organic ligands. 1 2 3 Figure 2.15 – A section of the 1H NMR (400MHz, DMSO) spectrum showing the assignment for 2. 56 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates - UV-Vis spectroscopy The UV-Vis absorption spectrum of 2 recorded in DMF is similar to that of 1 displaying an absorption band at ca. 460 nm ( = 1000 L mol-1 cm-1). This singnal most likely originates from d – d transitions that can be assigned to 2Eg 2 T2g transitions involving the octahedrally coordinated MoV ions within 2.38 Absorbance (a.u.) 0.10 0.08 0.06 0.04 0.02 0.00 400 450 500 550 600 Wavelength (nm) Figure 2.16 –A section of the UV-Vis spectrum of a 10-4 M solution of 2 in DMF. 57 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates Table 2.6 − Crystal data and structural refinement parameters for 2. Compound 2 Empirical formula Molecular mass/g mol C27H49As4Mo4N7O25 -1 Crystal colour/shape 3 1555.16 Red / rod-shaped Crystal size/mm 0.20×0.15×0.10 Crystal system Triclinic Space group P a/ Å 10.207(4) b/ Å 13.740(5) c/ Å 19.700(8) /º 87.426(13) /º 82.265(10) /º 71.005(11) 3 V/ Å 2588.6(17) Z 2 Temperature (K) 118(2) -3 Density/Mg m 1.978 -1 Absorp. coef./mm 3.567 F(000) 1494 2 50 max/º Reflections collected 39817 Independent reflections 9114 [R(int) = 0.1044] Data / restraints / parameters 9114 / 13 / 604 2 S on F 1.006 R1, wR2 [I>2 (I)] 0.0836, 0.1961 R1, wR2 (all data) 0.0892, 0.2006 Largest diff. peak and hole/e.Å-3 2.238 and -1.076 58 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates 2.2.1.3 The p-Hydroxyphenylarsonic Acid – MoVI/MoV Reaction System (NH4)5[MoVI2MoV3O11(O3AsC6H4OH)5]·9H2O (3) Surprisingly when p-hydroxyphenylarsonic acid is used as a stabilising ligand under the same reaction conditions as described above, we obtain a reaction mixture from which blue crystals of (NH4)5[MoVI2MoV3O11(O3AsC6H4OH)5]·9H2O (3), separate within a time period of one week. The structure of 3 is significantly different to that of 1 and 2. Single crystal X-ray diffraction measurements reveal that compound 3 crystallises in the triclinic crystal system in the space group P , and contains a pentanuclear, mixedvalent cluster stabilised by five p-hydroxyphenyl arsonate ligands. The structure of the cluster anion is shown in Figure 2.17. All Mo ions in the structure are distorted octahedrally surrounded by O donors. The coordination spheres of Mo(1) and Mo(2), each consist of three bridging 2- 2- oxo ligands, two O donors that derive from two deprotonated organoarsonate ligands and are completed by a terminal Mo=O bonds. The bond distances between these Mo ions and the bridging 2- 2- oxo ligands range between 1.930(5) Å – 2.060(5) Å. The bond distances between the Mo ions and the O donors originating from the organic arsonate ligands range between 2.083(5) Å – 2.299(5) Å, while the Mo=O bond distances between Mo(1)-O(9) and Mo(2)-O(10) are 1.701(5) Å and 1.685(5) Å, respectively (Table 2.7). The remaining Mo ions in 3, Mo(3), Mo(4) and Mo(5), also show distorted octahedral geometries each consisting of two bridging 2- 2- oxo ligands, three Oarsonate donors and a terminal Mo=O bond. The Mo – ( 2- ) bond distances range between 1.817(5) Å – 1.998(5) Å; the Mo – Oarsonate bond distances range between 2.013(5) Å – 2.312(5) Å, while the Mo=O bond distances between Mo(3)-O(11), Mo(4)-O(12) and Mo(5)-O(13) are 1.701(5) Å, 1.671(5) Å and 1.682(5) Å, respectively. The distorted nature of the octahedral geometries of the coordinated metal centres can also be observed from their bond angles that deviate from the ideal octahedral geometry. The bond angle O(2)-Mo(1)-O(7) of 155.3(2)° shows the greatest deviation from the ideal octahedral angle of 180°, while the bond angle O(9)-Mo(1)-O(2) of 106.5(2)° shows the greatest deviation from the ideal octahedral angle of 90°. Selected bond lengths and bond angles for the Mo ions in 3 are listed in Table 2.7 and Table 2.8. 59 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates a) b) Figure 2.17 – Crystal structure of the pentanuclear Mo complex in 3: (a) ball-and-stick representation, (b) polyhedral presentation. Colour code: MoVI lavender blue, MoV blue, As orange, O red, C grey (hydrogen atoms have been omitted for clarity). Figure 2.18 – The core structure of the pentanuclear Mo complex in 3. Colour code: MoVI lavender blue, MoV blue, As orange, O red, C grey. 60 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates The coordination polyhedra of Mo(1), Mo(2), Mo(3) and Mo(5) share common edges. The polyhedron of Mo(4), shares common vertices with the polyhedra of Mo(3) and Mo(5) and closes the resulting circular entity in 3. Four organic ligands are situated on the outer side of the ring, and each binds via its deprotonated arsonate functionalities in a η1:η1:μ2 bridging mode to two Mo ions. The organic ligands bridge between Mo(2) and Mo(3), Mo(3) and Mo(4), Mo(4) and Mo(5) and Mo(5) and Mo(1), respectively. The fifth ligand is situated above the centre of the {Mo5} ring and binds with its three O atoms to all Mo ions (O(26) binds to Mo(4), and O(3) and O(8) bridge between Mo(2) and Mo(3), and Mo(5) and Mo(1), respectively). The aromatic ring systems of the ligands all point in one direction and project perpendicularly to the virtual plane of the {Mo5} ring. The phenolic OH functionalities of the organic ligands remain protonated and are engaged in hydrogen bonding within the crystal structure. a) b) Figure 2.19 – Packing arrangement of 3 in the crystal structure viewed in the direction of the crystallographic: (a) a-axis and (b) c-axis. Colour code: Mo blue, As orange, O red, C grey (crystallisation water molecules, NH4+ counterions and hydrogen atoms have been omitted for clarity). The Mo−O bond distances of Mo(3) and Mo(5) are significantly shorter than the corresponding distances of the other Mo ions in the structure. Further bond valence sum analyses confirm that these two Mo ions adopt the oxidation state +VI, while Mo(1), Mo(2) and Mo(4) are in the oxidation state +V. The assignment of the oxidation states is further substantiated by the observed Mo−Mo distances within 3. The short interatomic distance of 2.574(1) Å between Mo(1) and Mo(2) is indicative for Mo−Mo contacts and in agreement with the +V oxidation states. Comparable MoV−MoVI distances are expectedly 61 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates significantly longer (Mo(1)-Mo(5) 3.359(1) Å, Mo(2)-Mo(3) 3.353(3) Å, Mo(3)-Mo(4) 3.680(2) Å and Mo(4)-Mo(5) 3.671(5) Å). Its mixed valence nature distinguishes 3 from many other organoarsonate-stabilised polyoxomolybdates, which tend to exist, with few exceptions,38, 39 predominantly in their fully oxidised forms. Table 2.7 − Selected bond lengths [Å] and bond valence sum (BVS) for compound 3. Atom Bond Bond distances (Å) BVS Assigned oxidation state Mo(1) Mo(1)-O(9) Mo(1)-O(1) Mo(1)-O(2) Mo(1)-O(7) Mo(1)-O(15) Mo(1)-O(8) 1.701(5) 1.930(5) 1.954(5) 2.060(5) 2.083(5) 2.252(5) 5.251 +5 Mo(2) Mo(2)-O(10) Mo(2)-O(2) Mo(2)-O(1) Mo(2)-O(4) Mo(2)-O(17) Mo(2)-O(3) 1.685(5) 1.938(5) 1.940(5) 2.053(5) 2.101(6) 2.299(5) 5.277 +5 Mo(3) Mo(3)-O(11) Mo(3)-O(5) Mo(3)-O(4) Mo(3)-O(18) Mo(3)-O(20) Mo(3)-O(3) 1.701(5) 1.826(5) 1.828(5) 2.033(6) 2.069(5) 2.312(5) 5.919 +6 Mo(4) Mo(4)-O(12) Mo(4)-O(5) Mo(4)-O(6) Mo(4)-O(21) Mo(4)-O(23) Mo(4)-O(26) 1.671(5) 1.980(5) 1.998(5) 2.013(5) 2.022(5) 2.183(5) 5.462 +5 Mo(5) Mo(5)-O(13) Mo(5)-O(6) Mo(5)-O(7) Mo(5)-O(24) Mo(5)-O(14) Mo(5)-O(8) 1.682(5) 1.817(5) 1.843(5) 2.045(5) 2.085(5) 2.272(5) 5.981 +6 Mo(1) ··· Mo(2) 2.574(1) 62 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates Table 2.8 − Selected bond angles [º] for compound 3. Bond Angle (º) Bond Angle (º) O(9)-Mo(1)-O(8) O(1)-Mo(1)-O(15) O(2)-Mo(1)-O(7) O(9)-Mo(1)-O(2) O(9)-Mo(1)-O(7) O(1)-Mo(1)-O(2) O(7)-Mo(1)-O(15) O(1)-Mo(1)-O(7) O(2)-Mo(1)-O(15) O(15)-Mo(1)-O(8) 165.6(2) 161.5(2) 155.3(2) 106.5(2) 97.6(2) 94.7(2) 87.3(2) 84.5(2) 85.9(2) 77.56(19) O(10)-Mo(2)-O(3) O(1)-Mo(2)-O(17) O(2)-Mo(2)-O(4) O(10)-Mo(2)-O(2) O(10)-Mo(2)-O(4) O(2)-Mo(2)-O(1) O(2)-Mo(2)-O(17) O(1)-Mo(2)-O(3) O(2)-Mo(2)-O(3) O(17)-Mo(2)-O(3) 166.9(2) 161.0(2) 153.6(2) 106.4(2) 99.4(2) 94.9(2) 86.7(2) 85.1(2) 81.8(2) 76.3(2) O(11)-Mo(3)-O(3) O(5)-Mo(3)-O(18) O(4)-Mo(3)-O(20) O(11)-Mo(3)-O(5) O(11)-Mo(3)-O(20) O(11)-Mo(3)-O(18) O(4)-Mo(3)-O(18) O(20)-Mo(3)-O(3) O(18)-Mo(3)-O(3) O(4)-Mo(3)-O(3) 175.1(2) 158.2(2) 156.3(2) 101.1(2) 99.0(2) 96.4(2) 88.3(2) 83.64(18) 80.1(2) 75.4(2) O(12)-Mo(4)-O(26) O(5)-Mo(4)-O(23) O(6)-Mo(4)-O(21) O(12)-Mo(4)-O(21) O(12)-Mo(4)-O(5) O(5)-Mo(4)-O(6) O(5)-Mo(4)-O(21) O(6)-Mo(4)-O(23) O(5)-Mo(4)-O(26) O(21)-Mo(4)-O(26) 178.8(2) 166.1(2) 165.6(2) 97.1(2) 96.8(2) 95.4(2) 88.8(2) 86.6(2) 83.72(19) 81.7(2) O(13)-Mo(5)-O(8) O(6)-Mo(5)-O(14) O(7)-Mo(5)-O(24) O(6)-Mo(5)-O(7) O(13)-Mo(5)-O(7) O(13)-Mo(5)-O(14) O(6)-Mo(5)-O(24) O(6)-Mo(5)-O(8) O(14)-Mo(5)-O(8) O(7)-Mo(5)-O(8) 172.6(2) 157.4(2) 155.9(2) 103.7(2) 99.3(2) 97.2(2) 87.8(2) 84.2(2) 78.76(19) 74.3(2) The pentanuclear polyoxomolybdate cluster in 3 carries an overall charge of -5, which is compensated for by NH4+ counterions. In the crystal structure, the cluster packs to form a layered lamellar structure with alternating inorganic and organic areas. This structure, viewed in the direction of the crystallographic c-axis, is shown in Figure 2.19, b. The counterions and the crystallisation water molecules are engaged in hydrogen bonds (O-O distances: 2.497(67) Å – 3.142(95) Å, N-O distances: 2.589(74) Å – 3.122(66) Å) and prevail in the hydrophilic part of the structure. 63 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates - PHYSICOCHEMICAL CHARACTERISATION - FT-IR spectroscopy The IR spectrum of 3 (Figure 2.20) is similar to that of 1 and 2. Characteristic C–C skeletal vibrations of the phenyl rings can be observed in the range 1600 – 1400 cm-1. The signal at 1093 cm-1 can be associated with As–C vibrations, while some bands in the 1000 – 650 cm-1 region can be attributed to the molybdenum-oxygen and arsenic-oxygen stretching vibrations.8, 15, 37, 46-48 Figure 2.20 – Infrared spectrum of 3. 64 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates - Thermogravimetric analysis The thermal stability of compound 3 was investigated by thermogravimetric analysis in the temperature range between 30 and 900 °C, in a N2 atmosphere (Figure 2.21). Upon thermolysis, 3 decomposes in a similar sequence as 2. A weight loss of about 8.3 % occurs below 150 °C. This weight loss can be attributed to the loss of nine water molecules (calcd: 8.1 %). Decomposition of the organic ligands occurs between 200 – 400 °C and a step at 700 °C indicates the transformation into oxides. 100 Weight % (%) 90 80 70 60 50 40 0 200 400 600 800 1000 Temperature (°C) Figure 2.21 – Thermogravimetric analysis of 3. 65 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates - Magnetism The assignments of the oxidation states of the Mo atoms in 3 are in agreement with the observed magnetic properties. In the {Mo5} units, three S = ½ MoV (light pink) and two diamagnetic MoVI (purple) metal ions are present (Figure 2.22 a). Considering the topology of the complex, two different magnetic exchange pathways could possibly be envisaged for the {Mo5} core. These magnetic pathways are highlighted in Figure 2.22. Moreover, it is worth noting that the interaction denoted with J' will likely be very strong as two oxo bridges mediate the magnetic interactions between MoV centres. a) b) χT /cm3 K mol-1 0.5 0.4 0.3 0.2 0.1 0 1000 Oe 10000 Oe 0 50 100 150 200 250 300 T /K Figure 2.22 – (a) {Mo5} units of compound 3 showing: the S = ½ Mo(V) (light pink), the diamagnetic Mo(VI) (purple) metal ions, and the J and J` magnetic interactions; (b) Temperature dependence of the χT product (with χ defined as the magnetic susceptibility and equal to M/H after diamagnetic and experimental corrections) at 0.1 and 1 T. The temperature dependence of the χT product of 3 is shown in Figure 2.22, b. At room temperature, the χT product is 0.37 cm3 K mol-1 which is in good agreement with the presence of one S = ½ MoV metal ion (expected value: 0.375 cm3 K mol-1 with g = 2). The two other MoV ions appear to be magnetically silent as expected for two spin centres that are strongly antiferromagnetically coupled (|J`| >> 500 K) to give a diamagnetic dinuclear unit. Such strong antiferomagnetic exchange is common in polyoxomolybdates that feature edge- and face-sharing dinuclear molybdenum octahedra.49, 50 When the temperature is lowered, the χT product at 1000 Oe remains constant at 0.37 cm3 K mol-1 down to 1.8 K indicating a Curie type paramagnetism. From the Curie constant (0.37 cm3 K mol-1), the g value is estimated to be very close to 2. 66 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates a) b) 1.2 1.2 1.8 K 3K 5K 8K 1 1 0.8 M / µB M / µB 0.8 0.6 0.6 0.4 0.4 0.2 0.2 0 1.8 K 3K 5K 8K 0 0 20000 40000 H / Oe 60000 b Chisq R 0 y = bri1d(1,2,2) Value Error 2.0503 0.00084494 0.0010978 NA 0.99996 NA 20000 40000 H T-1 / Oe K -1 Figure 2.23 – (a) M vs H and (b) M vs H/T data below 8 K. The solid lines are guides for eyes on the left figure while the blue line on the right figure is the best fit obtained with an S = ½ Brillouin function and g = 2.05. The magnetisation as a function of field at low temperatures has been measured up to 7 T between 1.8 and 8 K (Figure 2.23). It is worth noting that no hysteresis effect were detected even at 1.8 K and that the magnetisation plotted as a function of H/T is perfectly superimposed on a single master curve, as expected for isotropic systems. At 7 T and 1.8 K, the magnetic moment saturates at 1.01 µB in perfect agreement with the presence of an S = ½ spin. The S = ½ spin ground state of the {Mo5} unit is confirmed by the fit of the magnetisation to an S = ½ Brillouin function that reproduces very well the experimental data with g = 2.05. 67 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates - Mass spectrometry Negative mode ESI-MS spectra of compound 3 dissolved in a suitable solvent (H2O and DMSO, respectively), were recorded. ESI-MS studies using the crystalline material reveal that compound 3 undergoes structural transformations when dissolved in distilled water but is stable in a DMSO environment. The mass spectrum of 3 dissolved in DMSO (Figure 2.24) reveals the presence of a signal centered at m/z = 1739.1 a.m.u. which corresponds to the {H3[MoVI3MoV2O11(O3AsC6H4OH)5]}- species. A comparison of the experimental isotopic envelopes with simulated patterns is presented in Figure 2.24 and Relative ionic abundance confirms our assignment. Figure 2.24 – Negative-mode ESI-MS spectra for crystals of 3 dissolved in DMSO. Inset: Comparison of the experimental isotopic envelopes (black spectrum) with simulated envelope (red spectrum) for {H3[MoVI3MoV2O11(O3AsC6H4OH)5]}- centered at m/z = 1739.1 a.m.u. (cone voltage: 30 V). 68 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates - UV-Vis spectroscopy The UV-Vis absorption spectrum of 3 recorded in DMF displays an absorption band at ca. 531 nm ( = 2700 L mol-1 cm-1). Compared to 1 and 2, the absorption band observed for 3 is much broader and shifted to higher wavelengths. Thus, this band is a result of two different types of absorptions: d – d transitions involving the MoV metal centeres and intervalence charge transfer between MoV and MoVI via an oxo bridge. The d – d transitions can be assigned to a 2Eg 2 T2g transition of the octahedrally coordinated MoV ions.38, 51, 52 0.13 Absorbance (a.u.) 0.12 0.11 0.10 0.09 0.08 0.07 0.06 0.05 400 450 500 550 600 650 700 Wavelength (nm) Figure 2.25 – A section of the UV-Vis spectrum of a 10-4 M solution of 3 in DMF. 69 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates Table 2.9 − Crystal data and structural refinement parameters for 3. Compound 3 Empirical formula Molecular mass/g mol C30H63As5Mo5N5O40 -1 Crystal colour/shape 3 1988.13 Blue / rectangular plate Crystal size/mm 0.50×0.40×0.40 Crystal system Triclinic Space group P a/ Å 19.417(4) b/ Å 21.040(4) c/ Å 21.562(4) /º 113.09(3) /º 112.90(3) /º 99.16(3) 3 V/ Å 6941(2) Z 2 Temperature (K) 108(2) -3 Density/Mg m 1.864 -1 Absorp. coef./mm 3.334 F(000) 3732 2 50 max/º Reflections collected 103006 Independent reflections 24429 [R(int) = 0.0648] Data / restraints / parameters 24429 / 15280 / 1587 2 S on F 1.059 R1, wR2 [I>2 (I)] 0.0704, 0.1844 R1, wR2 (all data) 0.0789, 0.1930 Largest diff. peak and hole/e.Å-3 2.729 and -2.598 70 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates 2.2.1.4 The (4-Hydroxy-3-Nitrophenyl)arsonic Acid – MoVI/MoVI Reaction System (NH4)4[MoVI4O10(O3AsC6H3NO2OH)4]·2H2O (4) It was decided to investigate the influence of a disubstituted arsonate ligand on the self-assembly process of hybrid polyoxomolybdates under the same outlined reaction conditions that led to the formation of 1−3. Using a substituted arsonate ligand with a hydroxyl group in the para position and a nitro group in the meta position, a dark-green solution is obtained, which undergoes a colour change to red-orange within approximately ten days. Orange crystals of (NH4)4[MoVI4O10(O3AsC6H3NO2OH)4]·2H2O (4) were obtained from the reaction mixture within a time period of three weeks, and single crystal X-ray studies were performed. 4 crystallises in the triclinic crystal system in the space group P . The anionic cluster [MoVI4O10(O3AsC6H3NO2OH)4]4– in 4 (Figure 2.26) contains two {Mo2O10}8subunits in which octahedrally coordinated Mo atoms share a common edge (Mo(1) with Mo(2) and Mo(1`) with Mo(2`)). The dimeric units are connected by four deprotonated arsonate ligands. Two of the ligands have their As atoms (As(3) and As(3`)) arranged nearly coplanar to the four Mo atoms. Each of the ligands is bridging between two Mo ions in a µ 2-syn, syn bridging mode to form a six membered {As2Mo4} ring. The remaining two arsonate ligands cap both sides of the central cavity of the ring and each provides one μ2bridging O-donor (O(4) and O(4`)) of the common edges of the {Mo2O10}8- dimer moieties. The other two O atoms of the tetrahedral arsonate functionality of each of the ligands (O(7), O(13) and O(7`), O(13`) respectively) link to the adjacent {Mo2O10}8moiety and interconnect the dimeric subunits. The involved O donors bridge in a µ 2-syn, syn bridging mode and occupy two apical positions of the octahedrally coordinated Mo atoms within the dimer units. All four Mo atoms show distorted octahedral coordination environments. Each Mo ion in 4 is involved in two short Mo=O bonds (range between 1.692(3) - 1.720(3) Å), three elongated Mo–Oarsonate bonds that range between 2.009(3) 2.427(3) Å and one Mo–(µ 2-O) bond with a bond length of 1.915(3) Å for Mo(1)-O(1) and 1.894(3) Å for Mo(2)-O(1) (Table 2.10). The distorted nature of the octahedral coordination environments of the Mo ions can be exemplified by examining the bond lengths and bond angles of the metal centres. The bond angles O(1)-Mo(1)-O(5) and O(1)Mo(1)-O(4) of 156.65(11)° and 75.20(10)°, respectively, show the greatest deviation from the ideal octahedral geometry (Table 2.11). 71 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates a) b) Figure 2.26 – Crystal structure of the tetranuclear Mo complex in 4: (a) ball-and-stick representation, (b) polyhedral presentation. Colour code: MoVI sky blue, As orange, O red, N light blue, C grey (hydrogen atoms have been omitted for clarity). Figure 2.27 – The core structure of the tetranuclear Mo complex in 4. Colour code: MoVI sky blue, As orange, O red, C grey. The core structure of (Et3NH)4[Mo4O10(C6H5PO3)4]·2CH3CN 4 is and related to those observed in (Et3NH)4[Mo4O10(C6H5AsO3)4]·4H2O, 72 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates respectively.11 In these compounds the anionic clusters share a common structural motif but are stabilised by different organic ligands. Bond valence sum analyses confirm that all four Mo atoms in 4 adopt the oxidation state +VI (Table 2.10). Table 2.10 − Selected bond lengths [Å] and bond valence sum (BVS) for compound 4. Atom Bond Bond distances (Å) BVS Assigned oxidation state Mo(1) Mo(1)-O(3) Mo(1)-O(2) Mo(1)-O(1) Mo(1)-O(5) Mo(1)-O(13) Mo(1)-O(4) 1.692(3) 1.720(3) 1.915(3) 2.009(3) 2.223(3) 2.276(3) 5.979 +6 Mo(2) Mo(2)-O(8) Mo(2)-O(9) Mo(2)-O(1) Mo(2)-O(6) Mo(2)-O(7) Mo(2)-O(4) 1.707(3) 1.714(3) 1.894(3) 2.049(3) 2.113(3) 2.427(3) 5.937 +6 Table 2.11 − Selected bond angles [º] for compound 4. Bond Angle (º) O(2)-Mo(1)-O(13) O(3)-Mo(1)-O(4) O(1)-Mo(1)-O(5) O(3)-Mo(1)-O(2) O(3)-Mo(1)-O(5) O(2)-Mo(1)-O(1) O(2)-Mo(1)-O(4) O(3)-Mo(1)-O(13) O(5)-Mo(1)-O(4) O(1)-Mo(1)-O(4) 167.73(12) 162.67(12) 156.65(11) 104.73(14) 98.60(13) 97.13(13) 92.55(12) 86.93(12) 82.40(11) 75.20(10) Bond Angle (º) O(8)-Mo(2)-O(4) O(9)-Mo(2)-O(7) O(1)-Mo(2)-O(6) O(8)-Mo(2)-O(9) O(8)-Mo(2)-O(1) O(8)-Mo(2)-O(6) O(6)-Mo(2)-O(4) O(9)-Mo(2)-O(6) O(6)-Mo(2)-O(7) O(7)-Mo(2)-O(4) 169.07(12) 154.73(13) 158.36(12) 104.19(14) 99.20(13) 96.61(13) 90.51(10) 88.55(13) 77.79(11) 74.87(10) As expected, the di-substituted arsonate ligand imposes an influence on the selfassembly process under the investigated conditions that produce 1 and 2. In fact, the composition of the cluster anion 4 is closely related to that of 1 and 2. However, it seems 73 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates that the -OH and -NO2 functionalities of the organic ligands enforce a rearrangement of the cubane core structure. The rearrangement can formally be explained by an intramolecular condensation reaction under acidic conditions (Figure 2.28). +2H2O +4H++4e- Figure 2.28 – Formal rearrangement of the {Mo4} core structure in 1 or 2 to give 4. The packing arrangement of the [MoVI4O10(O3AsC6H3NO2OH)4]4– clusters in the crystal structure is characterised by small intercluster channels (filled with constitution water molecules and NH4+ counterions) that run in the direction of the crystallographic aaxis (Figure 2.29). The counterions and the crystallisation water molecules are engaged in H bonds whose N-O and O-O contacts range between 2.697(10) Å – 3.112(34) Å. a) b) Figure 2.29 – Packing arrangement of 4 in the crystal structure viewed in the direction of the crystallographic: (a) a-axis and (b) b-axis. Colour code: MoVI sky blue, As orange, O red, N light blue, C grey (crystallisation water molecules, NH4+ counterions and hydrogen atoms have been omitted for clarity). 74 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates - PHYSICOCHEMICAL CHARACTERISATION - FT-IR spectroscopy The IR spectrum of 4 shown in Figure 2.30 is comparable to that of 1 – 3. Notably, the 1620 – 1050 cm-1 region of the spectrum is quite crowded displaying a set of sharp bands of medium intensity. These bands arise from C–C skeletal vibrations of the phenyl rings, As–C vibrations and N–O stretches of the organoarsonate ligands.8, 15, 37, 46-48 Figure 2.30 – Infrared spectrum of 4. 75 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates - Thermogravimetric analysis Compound 4 was also analysed by TGA (Figure 2.31). A gradual weight loss of 2.3 % observed below 200 °C which corresponds to the loss of two water molecules (calcd: 2.1 %). A further weight loss between 200 – 450 °C indicates the decomposition of the organic moieties, while the thermogravimetric step between 450 – 550 °C can be associated with the decomposition into oxides. 100 Weight % (%) 80 60 40 20 0 0 200 400 600 800 1000 Temperature (°C) Figure 2.31 – Thermogravimetric analysis of 4. ESI-MS studies reveal that 4 is unstable in solution. The compound most likely decomposes into insoluble inorganic or polymeric materials when dissolved in H2O or DMSO. 76 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates Table 2.12 − Crystal data and structural refinement parameters for 4. Compound 4 Empirical formula Molecular mass/g mol C24H36As4Mo4N8O36 -1 Crystal colour/shape 3 1696.02 Orange / rectangular plate Crystal size/mm 0.20×0.20×0.20 Crystal system Triclinic Space group P a/ Å 10.7255(19) b/ Å 10.937(3) c/ Å 12.000(3) /º 89.474(13) /º 66.129(10) /º 63.926(7) 3 V/ Å 1131.6(5) Z 1 Temperature (K) 108(2) -3 Density/Mg m 2.465 -1 Absorp. coef./mm 4.109 F(000) 808 2 50.04 max/º Reflections collected 14710 Independent reflections 3987 [R(int) = 0.0313] Data / restraints / parameters 3987 / 3 / 343 2 S on F 1.036 R1, wR2 [I>2 (I)] 0.0282, 0.0817 R1, wR2 (all data) 0.0303, 0.0831 Largest diff. peak and hole/e.Å-3 1.502 and -0.739 77 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates 2.2.2 Synthesis and characterisation of organophosphonate functionalised heteropolyoxomolybdate clusters 2.2.2.1 The Phenylphosphonic Acid – MoVI/MoV Reaction System (NH4)4H4{Mn[MoV6O12(OH)3(O3PC6H5)4]2}·8H2O (5) Partial reduction of molybdic acid using N2H4 H2O in the presence of phenylphosphonic acid and MnCl2 4H2O results in a red-brown solution from which orange octahedral crystals of (NH4)4H4{Mn[MoV6O12(OH)3(O3PC6H5)4]2} 8H2O (5) separate within a time period of three days. Single crystal X-ray analysis reveals that 5 crystallises in the orthorhombic crystal system in the space [MoV6O12(OH)3(O3PC6H5)4]5- group Pbca. The structure consists of two II fragments linked through a Mn ion (Figure 2.32). Within each fragment, six Mo ions form distorted octahedral coordination polyhedra, each sharing two common edges with adjacent polyhedra to form a six membered ring. Three organophosphonate ligands are situated on the periphery of the resulting ring, each bridging two Mo ions in a η1:η1:μ2 bridging mode. The O donors of a fourth organic ligand cap the central cavity of the ring and provide µ2 bridging O donors (O(2), O(4) and O(12)) to three, dimeric edges-sharing subunits (Mo(2) and Mo(3), Mo(4) and Mo(5), and Mo(6) and Mo(1), respectively). The aromatic moieties of the four organic ligands are arranged approximately perpendicular to the mean plane of the six membered ring. The six molybdenum atoms within the ring display distorted octahedral coordination environments. This distortion mainly results from the presence of short Mo=O bonds (range between 1.678(7) - 1.699(6) Å) situated in trans positions to significantly longer Mo–(µ 2-O) bonds (range between 2.302(6) - 2.333(6) Å). The µ 2 bridging O-donors (O(2), O(4) and O(12)) engaged in these elongated bonds originate from the phosphonate functionality of the central ligand. The remaining coordination sites of the Mo ions are occupied by one O donor originating from a peripheral organic ligand, two two Mo ions and a 3- 2- 2- 2- ions that bridge between ligand bridging between two Mo ions and the central Mn ion. The bond distances between the Mo ions and the O donors of these peripheral phosphonate ligands range between 2.066(6) Å – 2.093(7) Å, the Mo-( 2between 1.941(7) Å – 2.135(6) Å, while the Mo-( 3- 2- ) bond distances range 2- ) bond distances range between 1.972(6) Å – 1.989(6) Å. The distorted octahedral coordination environment of the Mo 78 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates ions in 5 can also be exemplified by examining the bond angles that deviate from the ideal octahedral geometry. Selected bond lengths and bond angles for the Mo ions in 5 are listed in Table 2.13, Table 2.14. a) b) Figure 2.32 – Crystal structure of the dodecanuclear Mo complex in 5. Colour code: MoV blue, P purple, O red, Mn cyan, C grey (hydrogen atoms have been omitted for clarity). Figure 2.33 – The {Mn[MoV6O12(OH)3(O3PC6H5)4]}3- fragment in 5. Colour code: MoV blue, P purple, O red, Mn cyan, C grey. 79 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates Table 2.13 − Selected bond lengths [Å] and bond valence sum (BVS) for compound 5. Atom Bond Bond distances (Å) BVS Assigned oxidation state Mo(1) Mo(1)-O(21) Mo(1)-O(25) Mo(1)-O(3) Mo(1)-O(16) Mo(1)-O(11) Mo(1)-O(12) 1.699(6) 1.951(7) 1.987(6) 2.084(7) 2.127(6) 2.326(6) 5.174 +5 Mo(2) Mo(2)-O(20) Mo(2)-O(25) Mo(2)-O(3) Mo(2)-O(15) Mo(2)-O(13) Mo(2)-O(2) 1.689(7) 1.945(7) 1.987(6) 2.093(7) 2.109(6) 2.302(6) 5.047 +5 Mo(3) Mo(3)-O(19) Mo(3)-O(5) Mo(3)-O(1) Mo(3)-O(14) Mo(3)-O(13) Mo(3)-O(2) 1.678(7) 1.943(7) 1.972(6) 2.079(7) 2.118(6) 2.318(6) 5.134 +5 Mo(4) Mo(4)-O(18) Mo(4)-O(5) Mo(4)-O(1) Mo(4)-O(8) Mo(4)-O(6) Mo(4)-O(4) 1.689(7) 1.941(7) 1.989(6) 2.092(7) 2.127(6) 2.333(6) 5.001 +5 Mo(5) Mo(5)-O(23) Mo(5)-O(9) Mo(5)-O(10) Mo(5)-O(7) Mo(5)-O(6) Mo(5)-O(4) 1.692(7) 1.943(6) 1.978(6) 2.068(6) 2.135(6) 2.320(6) 5.044 +5 Mo(6) Mo(6)-O(22) Mo(6)-O(9) Mo(6)-O(10) Mo(6)-O(17) Mo(6)-O(11) Mo(6)-O(12) 1.687(7) 1.945(6) 1.978(6) 2.066(6) 2.117(6) 2.325(6) 5.091 +5 Mo(1) ··· Mo(2) Mo(3) ··· Mo(4) Mo(5) ··· Mo(6) 2.6058(14) 2.6027(14) 2.5934(15) Bond valence sum analysis confirms that all six Mo atoms in 5 adopt the oxidation state +V. In the crystal structure of 5, the molecular entities pack to form a layered lamellar structure with alternating hydrophobic organic and hydrophilic inorganic regions (Figure 2.34). The NH4+ counterions and the crystallisation water molecules are situated in the hydrophilic part of the structure being engaged in hydrogen bonds (O-O and O-N distances: 2.901(26) Å – 3.084(21) Å). Weak offset π-π interactions between two adjacent 80 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates clusters with an interplanar separation distance of about 3.9 Å also contribute to the stability of 5 in the solid state. Table 2.14 − Selected bond angles [º] for compound 5. Bond Angle (º) a) Bond Angle (º) O(21)-Mo(1)-O(12) O(3)-Mo(1)-O(16) O(25)-Mo(1)-O(11) O(21)-Mo(1)-O(25) O(21)-Mo(1)-O(11) O(25)-Mo(1)-O(3) O(25)-Mo(1)-O(16) O(16)-Mo(1)-O(11) O(3)-Mo(1)-O(12) O(11)-Mo(1)-O(12) 169.8(3) 160.2(3) 155.6(3) 106.2(3) 97.3(3) 95.4(3) 86.5(3) 84.1(3) 81.3(2) 73.0(2) O(20)-Mo(2)-O(2) O(3)-Mo(2)-O(15) O(25)-Mo(2)-O(13) O(20)-Mo(2)-O(25) O(20)-Mo(2)-O(15) O(25)-Mo(2)-O(3) O(3)-Mo(2)-O(13) O(15)-Mo(2)-O(13) O(3)-Mo(2)-O(2) O(13)-Mo(2)-O(2) 168.7(3) 159.3(3) 155.9(3) 105.9(3) 97.5(3) 95.6(3) 86.5(3) 84.6(3) 80.9(2) 72.1(2) O(19)-Mo(3)-O(2) O(1)-Mo(3)-O(14) O(5)-Mo(3)-O(13) O(19)-Mo(3)-O(5) O(19)-Mo(3)-O(13) O(5)-Mo(3)-O(1) O(1)-Mo(3)-O(13) O(5)-Mo(3)-O(2) O(1)-Mo(3)-O(2) O(13)-Mo(3)-O(2) 168.1(3) 159.5(3) 155.1(3) 106.7(3) 96.9(3) 95.4(3) 86.6(3) 84.2(2) 80.7(2) 71.7(2) O(1)-Mo(4)-O(8) O(18)-Mo(4)-O(4) O(5)-Mo(4)-O(6) O(18)-Mo(4)-O(5) O(18)-Mo(4)-O(6) O(5)-Mo(4)-O(1) O(5)-Mo(4)-O(8) O(8)-Mo(4)-O(6) O(1)-Mo(4)-O(4) O(6)-Mo(4)-O(4) 160.6(3) 169.8(3) 155.8(3) 106.1(3) 97.3(3) 94.9(3) 87.2(3) 84.0(3) 81.5(2) 73.0(2) O(23)-Mo(5)-O(4) O(10)-Mo(5)-O(7) O(9)-Mo(5)-O(6) O(23)-Mo(5)-O(9) O(23)-Mo(5)-O(7) O(9)-Mo(5)-O(10) O(9)-Mo(5)-O(7) O(7)-Mo(5)-O(6) O(10)-Mo(5)-O(4) O(6)-Mo(5)-O(4) 169.5(3) 158.6(3) 156.9(3) 105.5(3) 97.6(3) 95.5(2) 86.6(3) 84.6(3) 80.5(2) 73.2(2) O(22)-Mo(6)-O(12) O(10)-Mo(6)-O(17) O(9)-Mo(6)-O(11) O(22)-Mo(6)-O(9) O(22)-Mo(6)-O(17) O(9)-Mo(6)-O(10) O(9)-Mo(6)-O(17) O(17)-Mo(6)-O(11) O(10)-Mo(6)-O(12) O(11)-Mo(6)-O(12) 169.0(3) 158.9(3) 156.8(3) 105.9(3) 97.1(3) 95.4(3) 86.4(3) 84.3(3) 80.4(2) 73.2(2) b) Figure 2.34 – Packing arrangement of 5 in the crystal structure viewed in the direction of the crystallographic: (a) a-axis and (b) c-axis. Colour code: MoV blue, P purple, O red, Mn cyan, C grey (crystallisation water molecules, NH4+ counterions and hydrogen atoms have been omitted for clarity). 81 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates Several of these sandwich-type compounds were successfully synthesised using similar synthetic procedure as for 5, but replacing the MnII ion with FeIII, CoII, NiII, and MgII metal ions, respectively, in order to link the two {Mo6} fragments: (NH4)4H4{Fe[MoV6O12(OH)3(O3PC6H5)4]2} 8H2O (6), (NH4)4H4{Co[MoV6O12(OH)3(O3PC6H5)4]2} 8H2O (7), (NH4)4H4{Ni[MoV6O12(OH)3(O3PC6H5)4]2} 8H2O (8), (NH4)4H4{Mg[MoV6O12(OH)3 (O3PC6H5)4]2} 8H2O (9). The resulting compounds are isostructural to 5 and have very similar structural and geometrical parameters. A single crystal X-ray diffraction study was carried out on the FeIII complex, confirming the expected sandwich structure (Table 2.16). XRD powder diffraction analyses were performed on the CoII, NiII, and MgII complexes to certify that the compounds are isostructural. The measured powder pattern of each compound fits to the simulated pattern which is based on the single crystal X-ray diffraction data of 5 (Figure 2.35). 4000 2000 Measured Intensity/a.u 0 Calculated -2000 -4000 -6000 5 7 8 9 -8000 -10000 5 10 15 20 25 2 Theta/° Figure 2.35 – X-ray powder diffraction analysis, comparing the experimental patterns of 5 (black), 7 (pink), 8 (purple) and 9 (blue) complexes with simulated pattern (red) based on the single crystal X-ray diffraction data of 5. EDX analyses further confirm the presence of the transition metal ions within the corresponding compounds. 82 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates - PHYSICOCHEMICAL CHARACTERISATION - FT-IR spectroscopy The IR spectrum of 5 is presented in Figure 2.36 and reveals some characteristic stretches of the organophosphonate ligands. The C–C skeletal vibrations of the phenyl rings can be identified in the range 1600 – 1400 cm-1, while some of the bands in the 800 – 650 cm-1 region arise from C–H out-of-plane bending vibrations of the aromatic rings. Different P–O stretching vibrations of the phosphonate groups can be observed in the range 1200 – 1000 cm-1. The very strong band at ca. 950 cm-1 is characteristic of Mo=O stretching vibrations. Finally, the broad bands at ca. 3200 cm-1 and ca. 1600 cm-1 arise from O–H stretching vibrations and H–O–H bending vibrations of the crystallisation water molecules engaged in H-bonds.13, 48, 53, 54 Figure 2.36 – Infrared spectrum of 5. 83 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates - Thermogravimetric analysis The TGA curve of compound 5 is presented in Figure 2.37. It reveals a weight loss of 4.6 % below 100 °C, which was attributed to the loss of eight water molecules (calcd: 4.5 %). The decomposition of the organic ligands occurs between 300 – 600 °C and further cluster degradation processes result in the formation of metal oxides. The thermal stability of compounds 6, 7, 8 and 9 was also investigated. As expected, all five isostructural compounds reveal similar decomposition behaviour. 105 100 Weight % (%) 95 90 85 80 75 70 65 60 0 200 400 600 800 1000 Temperature (°C) Figure 2.37 – Thermogravimetric analysis of 5. The CHN analysis of the compounds is also in good agreement with the composition of the compounds. 84 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates - UV-Vis spectroscopy The UV-Vis absorption spectrum of 5 recorded in CH3CN is presented in Figure 2.38. The strong absorption band observed at ca. 305 nm ( = 11000 L mol-1 cm-1) is due to charge transfer transitions involving either the phosphonate ligands or the {Mo-O} moieties. A very weak absorption band can also be observed at ca. 443 nm ( = 1050 L mol-1 cm-1) involving a d – d transition that can be assigned to 2Eg 2 T2g transitions of the octahedrally coordinated MoV ions within 5. Compounds 6, 7, 8 and 9 display similar absorption spectra as 5 showing characteristic charge transfer absorptions around 300 nm and very weak absorption bands due to d – d transitions at ca. 450 nm.38, 55 0.18 0.12 0.16 0.10 Absorbance (a.u.) 0.14 0.08 0.12 0.06 0.10 0.04 0.08 400 450 500 550 600 0.06 0.04 0.02 0.00 -0.02 300 400 500 600 700 800 Wavelength (nm) Figure 2.38 – UV-Vis spectrum of a 10-5 M solution of 5 in CH3CN. Inset: A section of the UV-Vis spectrum of a 10-4 M solution of 5 in CH3CN. 85 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates Table 2.15 − Crystal data and structural refinement parameters for 5. Compound 5 Empirical formula Molecular mass/g mol C48H82MnMo12N4O62P8 -1 Crystal colour/shape 3 3161.16 Orange / octahedral Crystal size/mm 0.25×0.25×0.10 Crystal system Orthorhombic Space group Pbca a/ Å 19.757(7) b/ Å 17.463(6) c/ Å 30.649(10) /º 90 /º 90 /º 90 3 V/ Å 10574(6) Z 1 Temperature (K) 108(2) -3 Density/Mg m 2.085 -1 Absorp. coef./mm 1.718 F(000) 6404 2 50 max/º Reflections collected 41398 Independent reflections 9259 [R(int) = 0.0438] Data / restraints / parameters 9259 / 0 / 664 2 S on F 1.073 R1, wR2 [I>2 (I)] 0.0763, 0.2084 R1, wR2 (all data) 0.0863, 0.2185 Largest diff. peak and hole/e.Å-3 1.859 and -1.627 86 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates Table 2.16 − Crystal data and structural refinement parameters for 6. Compound 6 Empirical formula Molecular mass/g mol C48H82FeMo12N4O62P8 -1 Crystal colour/shape 3 3162.07 Red brown / polygonal Crystal size/mm 0.15×0.15×0.25 Crystal system Orthorhombic Space group Pbca a/ Å 19.601(8) b/ Å 17.357(6) c/ Å 30.430(12) /º 90 /º 90 /º 90 3 V/ Å 10353(7) Z 1 Temperature (K) 108(2) -3 Density/Mg m 2.071 -1 Absorp. coef./mm 1.767 F(000) 6170 2 50 max/º Reflections collected 58886 Independent reflections 9089 [R(int) = 0.0648] Data / restraints / parameters 9089 / 0 / 646 2 S on F 1.049 R1, wR2 [I>2 (I)] 0.0964, 0.2517 R1, wR2 (all data) 0.1057, 0.2613 Largest diff. peak and hole/e.Å-3 1.807 and -1.966 87 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates 2.3 ESI-MS STUDIES OF COMPLEX REACTION MIXTURES TO INVESTIGATE THE FORMATION OF HYBRID ORGANIC-INORGANIC POLYOXOMOLYBDATES 2.3.1 Investigation of the reaction system that led to the formation of the cubane structures in 1 and 2 Partial reduction of (NH4)6Mo7O24 4H2O in the presence of phenylarsonic acid and acetic acid using N2H4 H2SO4 resulted in a deep blue solution. When the initial reaction mixture is examined by ESI-MS, the spectrum reveals isotopic envelopes in the high molecular mass region centered at m/z = 1315.3 a.m.u. and m/z = 1659.2 a.m.u. (Figure 2.41). Within a time period of three weeks, dark red crystals 1, of (NH4)2H2[MoV4O8(O3AsC6H5)4]·5H2O separate from this reaction mixture (Figure 2.39). Figure 2.39 – The core structure of the tetranuclear Mo complexes in 1 and 2. Colour code: MoV blue, As orange, O red, C grey. The signal at m/z = 1315.3 a.m.u. in the mass spectrum of the reaction mixture that produced 1 is in agreement with the crystallographically determined formula of the cluster anion in 1 (Table 2.17). The signal can be reproduced by recording a spectrum of pristine crystals of 1 dissolved in dimethyl sulfoxide (DMSO), further confirming that the cluster core of 1 is stable in this polar solution as seen in section 2.2.1 of this chapter. Lower 88 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates molecular mass signals in the ESI-MS spectrum centered at m/z = 816.6 a.m.u. and m/z = 646.6 a.m.u. originate from the cubane cluster and can be assigned to the species {H9[MoV4O14(O3AsC6H5)]}- and [MoVI4O16H7]-, respectively. The ESI-MS spectrum of the reaction mixture further contains a high molecular mass signal centered at m/z = 1659.2 a.m.u. We were able to assign this signal to a {Mo6} species with the formula {(NH4)2H4[MoV4MoIV2O12(OH)3(O3AsC6H5)4]}-. A closely related core structure that is stabilised by phosphonates has previously been isolated in a compound {Na[MoV6O12(OH)3(O3PC6H5)4]2}9-.13 It can be crystallised from comparable reaction mixtures that contain alkali metal ions, producing a solid state structure in which two [MoV6O12(OH)3(O3PC6H5)4]5- fragments are linked through monovalent ions. We have successfully isolated the hexanuclear molybdenum fragments in compounds 5-9 and a description of the {Mo6} core structure was presented in section 2.2.2. Table 2.17 – ESI-MS assignment for the reaction mixture that led to the crystallisation of 1. m/z Species attributed 646.6 [MoVI4O16H7]- → {Mo4} Reaction 816.6 {H9[MoV4O14(O3AsC6H5)]}- → {Mo4} mixture of 1 1315.3 {H3[MoV4O8(O3AsC6H5)4]}- → {Mo4} 1659.2 {(NH4)2H4[MoV4MoIV2O12(OH)3(O3AsC6H5)4]}- → {Mo6} {Mo6} Figure 2.40 – Representation of a structural motif that agrees with the constitutional assignment for the {Mo6} species identified in the mass spectrum of the reaction mixture that led to the crystallisation of 1. Colour code: Mo blue, As orange, C grey, O light grey. We successfully simulated the isotopic envelopes for all of the identified species and decomposition products of 1 that originate from the reaction mixture (Figure 2.41), 89 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates and good fits between experimental and theoretical data further confirm our structural and constitutional assignments. The ESI-MS analysis further supports the assignment of the oxidation states and, in addition, suggests the presence of other relatively labile species in Relative ionic abundance Relative ionic abundance Relative ionic abundance Relative ionic abundance Relative ionic abundance the solution. Figure 2.41 – Negative-mode ESI-MS spectra for the reaction mixture that led to the crystallisation of 1 (24 h after preparation). Comparison of the experimental isotopic envelopes (black spectrum) with simulated patterns (red spectrum) for [MoVI4O16H7]- centered at m/z = 646.6 a.m.u., {H9[MoV4O14(O3AsC6H5)]}- centered at m/z = 816.6 a.m.u., {H3[MoV4O8(O3AsC6H5)4]}- centered at m/z = 1315.3 a.m.u. and {(NH4)2H4[MoV4MoIV2O12(OH)3(O3AsC6H5)4]}- centered at m/z = 1659.2 a.m.u. (cone voltage: 30 V). 90 − − Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates Figure 2.42 – Negative-mode ESI-MS spectra for the reaction mixture that led to the crystallisation of 1 at different cone voltages: 10V, 25V, 30V, 50V and 75V, respectively. Cone voltage (CV) variation experiments were performed in order to investigate its effect on the ESI-MS spectra (Figure 2.42). According to literature observations, through (CV) variations one can distinguish between the signals originating from the parent ions that form in solution and the fragment ions produced within the spraying chamber. Upon increasing the CV values during the ESI-MS experiments, the relative ionic abundance of the parent species that form in solution decreases, while the abundance of the fragment ions that form in the gas phase, increases. We and others observe that higher cone voltages result in the defragmentation of the coordination clusters within the spraying chamber.56, 57 91 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates As seen in Figure 2.42, the relative ionic abundances of {Mo4} and {Mo6} species decrease with increasing CV values. These experimental observations suggest that {Mo4} and {Mo6} species form in the same reaction mixture. The {Mo6} species appears to be structurally related to the hexanuclear molybdenum units in 5-9, while the {Mo4} species is structurally related to the cubane structure in 1 which crystallises out from this reaction mixture. The most probable structural arrangement for {Mo6} species is represented in Figure 2.40. When p-aminophenylarsonic acid was used as a stabilising ligand under the same reaction conditions that produced 1, we obtained a blue solution. The ESI-MS spectrum of this reaction mixture (Figure 2.43) only contains one isotopic envelope in the high mass region of the spectrum. The signal is centered at m/z = 1374.4 a.m.u. and corresponds to the cubane structure in 2, (NH4)2H2[MoV4O8(O3AsC6H4NH2)4]·DMF·4H2O, which crystallises from this solution as red rod-shaped crystals. Table 2.18 – ESI-MS assignment for the reaction mixture that led to the crystallisation of 2. Species attributed 1374.4 {H3[MoV4O8(O3AsC6H4NH2)4]}- → {Mo4} Relative ionic abundance Reaction mixture of 2 m/z Figure 2.43 – Negative-mode ESI-MS spectra for the reaction mixture that led to the crystallisation of 2 (24 h after preparation). Inset: Comparison of the experimental isotopic envelopes (black spectrum) with simulated patterns (red spectrum) for {H3[MoV4O8(O3AsC6H4NH2)4]}- centered at m/z = 1374.4 a.m.u. (cone voltage: 30 V). 92 − − Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates Figure 2.44 – Negative-mode ESI-MS spectra for the reaction mixture that led to the crystallisation of 2 at different cone voltages: 10V, 25V, 30V, 50V and 75V, respectively. ESI-MS and NMR analyses of the isolated compound in DMSO, presented in section 2.2.1, further confirm our assignments. ESI-MS cone voltage variation experiments performed on the reaction mixture that led to the crystallisation of 2, show that the signal attributed to the {Mo4} species is independent on CV variations up to 75 V (Figure 2.44), and the relative ionic abundance of this species start to decrease in intensity at CV greater than 100 V. These observations further confirm that the {Mo4} species form in solution and is structurally related to the cubane structure in 2. 93 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates 2.3.2 Investigation of the reaction system that led to the formation of the {Mo5} complex in 3 Compound 3, (NH4)5[MoVI2MoV3O11(O3AsC6H4OH)5]·9H2O crystallises under the same reaction conditions that led to the formation of 1 and 2, but using phydroxyphenylarsonic acid as a stabilising ligand. The ESI-MS spectrum of this reaction mixture is complex and is characterised by four major isotopic envelopes centred at m/z = 1377.3 a.m.u., m/z = 1522.2 a.m.u., m/z = 1704.1 a.m.u. and m/z = 1740.2 a.m.u. in the higher mass region of the spectrum (Figure 2.46). Figure 2.45 – Crystal structure of the pentanuclear Mo complex in 3. Colour code: MoVI lavender blue, MoV blue, As orange, O red, C grey (hydrogen atoms have been omitted for clarity). The spectrum proved to be very interesting, as we were able to identify three different species, a {Mo4}, a {Mo5} and a {Mo6} species, in the high molecular mass region of the spectrum (Table 2.19). The signal at m/z = 1377.3 a.m.u. can be assigned to the cubane structure, also observed in the previously examined reaction mixtures, whilst the high molecular mass signal centered at m/z = 1740.2 a.m.u. originates from a new pentanuclear mixed valent Mo cluster in 3. The signal at m/z = 1522.2 a.m.u. can be attributed to a defragmentation product of 3, in which one organic ligand is abstracted from the cluster anion. The ESI-MS spectrum of the reaction mixture of 3 displays a fourth signal at m/z = 1704.1 a.m.u. which can be assigned to a hexanuclear compound {(NH4)H3[MoV6O12(OH)3(O3AsC6H4OH)4]}–. A similar compound was also observed during analysis of the previously described reaction systems. 94 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates Table 2.19 – ESI-MS assignment for the reaction mixture that led to the crystallisation of 3. Species attributed 1377.3 {H3[MoV4O8(O3AsC6H4OH)4]}- → {Mo4} Reaction 1522.2 {H2[MoVI2MoV3O11(O3AsC6H4OH)4]}- → {Mo5} mixture of 3 1704.1 {(NH4)H3[MoV6O12(OH)3(O3AsC6H4OH)4]}- → {Mo6} 1740.2 {H4[MoVI2MoV3O11(O3AsC6H4OH)5]}- → {Mo5} Relative ionic abundance Relative ionic abundance Relative ionic abundance Relative ionic abundance Relative ionic abundance m/z Figure 2.46 – Negative-mode ESI-MS spectra for the reaction mixture that led to the crystallisation of 3 (24 h after preparation). Comparison of the experimental isotopic envelopes (black spectrum) with simulated patterns (red spectrum) for {H3[MoV4O8(O3AsC6H4OH)4]}- centered at m/z = 1377.3 a.m.u., {H2[MoVI2MoV3O11(O3AsC6H4OH)4]}centered at m/z = 1522.2 a.m.u., {(NH4)H3[MoV6O12(OH)3(O3AsC6H4OH)4]}centered at m/z = 1704.1 a.m.u. and {H4[MoVI2MoV3O11(O3AsC6H4OH)5]}- centered at m/z = 1740.2 a.m.u. (cone voltage: 30 V). 95 − − Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates Figure 2.47 – Negative-mode ESI-MS spectra for the reaction mixture that led to the crystallisation of 3 at different cone voltages: 10V, 25V, 30V, 50V and 75V, respectively. ESI-MS analyses including cone voltage variations clearly confirm that the tetranuclear cubane structure and the pentanuclear complex 3 form in solution in the same reaction mixture. We successfully simulated the isotopic envelopes for all of the identified species and decomposition products of the pentanuclear cluster, and good fits between experimental and theoretical data further substantiate our assignments (Figure 2.46). 96 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates 2.3.3 Investigation of the reaction system that led to the formation of the {Mo4} complex in 4 The same reaction conditions that led to the formation of 1-3 but using a disubstituted arsonate ligand to control the self-assembly process, gave a dark green solution from which compound 4, (NH4)4[MoVI4O10(O3AsC6H3NO2OH)4]·2H2O, was obtained. The ESI-MS spectrum of this reaction mixture is complex, showing the presence of multiple species in the higher mass region of the spectrum. We were able to identify unambiguously the composition of one of these species. Its isotopic envelope is centered at m/z = 1615.4 a.m.u. and corresponds to the formula of a tetranuclear molybdenum complex with the composition {H9[MoV2MoIV2O10(O3AsC6H3NO2OH)4] ·H2O}– (Figure 2.49, Table 2.20). Orange crystals of a closely related compound 4 were obtained from the reaction mixture within a time period of three weeks, and single crystal X-ray studies were performed. Figure 2.48 – Crystal structure of the tetranuclear Mo complex in 4. Colour code: MoVI sky blue, As orange, O red, N light blue, C grey (hydrogen atoms have been omitted for clarity). 97 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates Table 2.20 – ESI-MS assignment for the reaction mixture that led to the crystallisation of 4. Species attributed 1615.4 {H9[MoV2MoIV2O10(O3AsC6H3NO2OH)4]·H2O}– → {Mo4} Relative ionic abundance Relative ionic abundance Reaction mixture of 4 m/z Figure 2.49 – Negative-mode ESI-MS spectra for the reaction mixture that led to the crystallisation of 4 (24 h after preparation). Comparison of the experimental isotopic envelopes (black spectrum) with simulated patterns (red spectrum) for {H9[MoV2MoIV2O10(O3AsC6H3NO2OH)4] H2O}- centered at m/z = 1615.4 a.m.u. (cone voltage: 30 V). 98 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates 2.4 CONCLUSION AND FUTURE WORK In this chapter we present a facile synthetic approach to functionalise polyoxomolybdate clusters. The approach involves the partial reduction of MoVI salts in the presence of organoarsonate and phosphonate ligands. We demonstrate how slight perturbations of the ligand functionalities can be exploited to stabilise unprecedented core structures. Our investigations underline the stability of 1-3 in solution, an essential requirement for potential applications as catalysts. Redox-active transition metals that adopt cubane as observed in 1 and 2 or related structures are of particular interest to scientists because of their resemblance to active sites of enzymes, and it has recently been suggested that such oxo clusters might hold the key to catalysing the splitting of water,58-62 a process that might have an impact on future energy requirements providing a conceptional solution to climate issues. We demonstrate that the cubane structures tolerate amine functional groups in the para position to the arsonate group, a characteristic that might allow the tethering of clusters to surfaces, for instance, electrodes. We isolated an unprecedented pentanuclear ring structure in 3 when phydroxyphenyl arsenic acid is employed as a ligand. The mixed-valent oxo-cluster is characterised by strong antiferromagnetic interactions between the MoV centres. The use of (4-hydroxy-3-nitrophenyl)arsonic acid as a ligand promotes the formation of a rhombic planar arrangement in a tetranuclear complex in 4 which can formally be related to the rearrangement of the cube-type structures in 1 and 2. Another aspect of the work focused on the synthesis of hybrid organic-inorganic heteropolyoxomolybdates that incorporate additional d-block elements (Mn, Fe, Co, Ni). The secondary transition metal ions serve as a bridging linker between two polyoxomolybdate moieties leading to sandwich structures. We synthesised and isolated a series of isostructural transition metal compounds and a main group derivative (Mg2+ compound). Our investigations include the structure determination and some physicochemical characterisations of the compounds. Preliminary studies of the magnetic properties of compounds 5-9 suggest that the S = ½ MoV centres in these compounds are strongly antiferromagnetically coupled at room temperature leading to magnetically silent hexanuclear subunits. The temperature dependence of the susceptibility of the compounds is mainly determined by the presence of the transition metal ions. Based on the results of our synthetic efforts it appears feasible that f-block elements such as lanthanide ions (Tb, 99 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates Dy, Ho) could also be utilised to link the two {Mo6} fragments. The resulting heteropolyoxomolybdates might exhibit interesting photophysical and magnetic properties. Future work will involve the synthesis of hybrid organic-inorganic heteropolyoxomolybdates that incorporate f-block elements (Tb, Dy, Ho) in order to expand the family of clusters 5-9. We also intend to use a series of rigid extended polyphosphonate and –arsonate ligands which could give rise to the formation of porous polyoxometalate metal−organic frameworks (P-MOFs). The electrochemical and catalytic properties of the compounds will also be investigated. The ESI-MS study of complex reaction mixtures presented in the last section of this chapter aims to investigate the formation processes of hybrid organic-inorganic polyoxomolybdates. We have demonstrated that ESI-MS, in combination with X-ray crystallography, provides an extremely powerful tool to identify and characterise new species that form in solution. Our approach allowed us to use simple, common, wellinvestigated, and commercially available ligands and to screen their involvement in condensation reactions, providing us with a time-effective protocol to selectively isolate and structurally characterise a series of novel species (structures described in section 2.2.1). The effectiveness of mass spectrometry in this field of science has only recently been highlighted, and our efforts extend accomplishments to complex hybrid structures. Our results underline the possibility of exploring real-time growth reactions of polyoxometales that emerge in solution, transforming from small or oligonuclear species and aggregating into larger molecular clusters. A summary of the ESI-MS assignment for the reaction mixtures that led to the formation of the cubane structures in 1 and 2, and also to the {Mo5} complex in 3 highlighting related structural motifs is presented in Figure 2.50. We observe the {Mo4} cubane species in the ESI-MS spectra of the reaction mixture that led to the crystallisation of 1, 2 and 3. A {Mo6} species is observed in the ESI-MS spectra of the reaction mixture that led to the crystallisation of 1 and 3 and seems to be structurally related to the hexanuclear molybdenum units in 5-9. The third species observed in the ESI-MS spectrum of the reaction mixture that led to the crystallisation of 3, {Mo5}, forms in solution and structurally relates to the pentanuclear molybdenum cluster that crystallises from this reaction mixture. Our efforts focus on the application of analytical techniques that allow us to monitor and direct complex condensation reactions to produce novel functionalised hybrid materials. 100 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates Reaction mixture that led to the crystallisation of the following species; X-Ray crystal structure ESI-MS assigned species Structural motifs associated with the ESIMS assignment {Mo4} {Mo6} {Mo4} {Mo4} {Mo4} {Mo4} {Mo5} {Mo6} {Mo5} Figure 2.50 – Summary of the ESI-MS assignment for the reaction mixtures that led to the formation of the cubane structures in 1 and 2, and also, to the {Mo5} complex in 3 highlighting related structural motifs. In conclusion, we demonstrated that the topological features of the described hybrid molybdenum phosphonates and –arsonates are strongly dependent on the nature of the employed organic ligand. The described approach provides an efficient protocol to prepare hybrid polyoxomolybdates. Our results provide a solid foundation for research objectives whereby rigid extended poly-phosphonate and –arsonate ligands could give rise to the formation of porous polyoxometalate metal−organic frameworks (P-MOFs). These systems would be interesting for catalysis, gas sorption and separation applications.80 101 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates REFERENCES 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. D.-L. Long, R. Tsunashima and L. Cronin, Angewandte Chemie International Edition, 2010, 49, 1736-1758. A. Proust, R. Thouvenot and P. Gouzerh, Chemical Communications, 2008, 18371852. D.-L. Long and L. Cronin, Chemistry – A European Journal, 2006, 12, 3698-3706. P. Gouzerh and A. Proust, Chemical Reviews, 1998, 98, 77-112. P. Mialane, A. Dolbecq and F. Secheresse, Chemical Communications, 2006, 34773485. W. Qi, Y. Wang, W. Li and L. Wu, Chemistry – A European Journal, 2010, 16, 1068-1078. N. Mizuno and M. Misono, Chemical Reviews, 1998, 98, 199-218. B. J. S. Johnson, S. A. Geers, W. W. Brennessel, J. V. G. Young and A. Stein, Dalton Transactions, 2003, 4678-4681. B. J. S. Johnson, R. C. Schroden, C. Zhu and A. Stein, Inorganic Chemistry, 2001, 40, 5972-5978. C. Sassoye, K. Norton and S. C. Sevov, Inorganic Chemistry, 2003, 42, 1652-1655. Y.-D. Chang and J. Zubieta, Inorganica Chimica Acta, 1996, 245, 177-198. A. Gammersbach, P. Angenault, W. F. Hölderich, W. Clegg and M. R. J. Elsegood, Polyhedron, 1998, 17, 2535-2541. M. I. Khan, Q. Chen and J. Zubieta, Inorganica Chimica Acta, 1995, 235, 135-145. N. Calin and S. C. Sevov, Inorganic Chemistry, 2003, 42, 7304-7308. E. Burkholder and J. Zubieta, Inorganica Chimica Acta, 2004, 357, 301-304. D.-G. Lyxell, D. Bostrom, M. Hashimoto and L. Pettersson, Acta Crystallographica Section B, 1998, 54, 424-430. B. Liu, Z.-T. Yu, J. Yang, W. Hua, Y.-Y. Liu and J.-F. Ma, Inorganic Chemistry, 2011, 50, 8967-8972. J. Hao, J. Zhang, P. Yin, Z. Xiao, F. Xiao and Y. Wei, Chemistry – A European Journal, 2012, 18, 2503-2506. P. DeBurgomaster, A. Aldous, H. Liu, C. J. O’Connor and J. Zubieta, Crystal Growth & Design, 2010, 10, 2209-2218. J. L. Q. Tuoi and E. Muller, Rapid Communications in Mass Spectrometry, 1994, 8, 692-694. M. J. Deery, O. W. Howarth and K. R. Jennings, Journal of the Chemical Society, Dalton Transactions, 1997, 4783-4788. D. K. Walanda, R. C. Burns, G. A. Lawrance and E. I. von Nagy-Felsobuki, Journal of the Chemical Society, Dalton Transactions, 1999, 311-322. M. Bonchio, O. Bortolini, V. Conte and A. Sartorel, European Journal of Inorganic Chemistry, 2003, 699-704. F. Sahureka, R. C. Burns and E. I. von Nagy-Felsobuki, Inorganica Chimica Acta, 2003, 351, 69-78. M. T. Ma, T. Waters, K. Beyer, R. Palamarczuk, P. J. S. Richardt, R. A. J. O’Hair and A. G. Wedd, Inorganic Chemistry, 2008, 48, 598-606. Y.-F. Song, D.-L. Long, S. E. Kelly and L. Cronin, Inorganic Chemistry, 2008, 47, 9137-9139. E. F. Wilson, H. Abbas, B. J. Duncombe, C. Streb, D.-L. Long and L. Cronin, Journal of the American Chemical Society, 2008, 130, 13876-13884. R. Llusar, I. Sorribes and C. Vicent, Journal of Cluster Science, 2009, 20, 177-192. 102 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. H. N. Miras, E. F. Wilson and L. Cronin, Chemical Communications, 2009, 12971311. R. Tsunashima, D.-L. Long, H. N. Miras, D. Gabb, C. P. Pradeep and L. Cronin, Angewandte Chemie International Edition, 2010, 49, 113-116. L. Vila-Nadal, S. G. Mitchell, A. Rodriguez-Fortea, H. N. Miras, L. Cronin and J. M. Poblet, Physical Chemistry Chemical Physics, 2011, 13, 20136-20145. E. F. Wilson, H. N. Miras, M. H. Rosnes and L. Cronin, Angewandte Chemie International Edition, 2011, 50, 3720-3724. H. N. Miras, M. Sorus, J. Hawkett, D. O. Sells, E. J. L. McInnes and L. Cronin, Journal of the American Chemical Society, 2012, 134, 6980-6983. K. Y. Matsumoto, Bulletin of the Chemical Society of Japan, 1978, 51, 492-498. X.-Z. You, H.-L. Li, K.-B. Yu and L. Li, Acta Crystallographica Section C, 1993, 49, 1300-1303. B. J. S. Johnson, C. E. Buss, V. G. Young, Jnr and A. Stein, Acta Crystallographica Section C, 1999, 55, 549-551. E. Burkholder, S. Wright, V. Golub, C. J. O'Connor and J. Zubieta, Inorganic Chemistry, 2003, 42, 7460-7471. S. Tan, M. Hobday, J. Gorman, G. Amiet and C. Rix, Journal of Materials Chemistry, 2003, 13, 1180-1185. A. Dolbecq, J.-D. Compain, P. Mialane, J. Marrot, F. Sécheresse, B. Keita, L. R. B. Holzle, F. Miserque and L. Nadjo, Chemistry – A European Journal, 2009, 15, 733741. H.-X. Guo and S.-X. Liu, Inorganic Chemistry Communications, 2004, 7, 12171220. D. I. Brown, Oxford University Press, 2002, 26. C. A. Hunter and J. K. M. Sanders, Journal of the American Chemical Society, 1990, 112, 5525-5534. A. Jimtaisong, L. Feng, S. Sreehari, C. Bayse and R. Luck, Journal of Cluster Science, 2008, 19, 181-195. B. Modec, J. V. Brencic, E. M. Burkholder and J. Zubieta, Dalton Transactions, 2003, 4618-4625. W. Schirmer, U. Flörke and H. J. Haupt, Zeitschrift für anorganische und allgemeine Chemie, 1989, 574, 239-255. W. Kwak, L. M. Rajkovic, J. K. Stalick, M. T. Pope and C. O. Quicksall, Inorganic Chemistry, 1976, 15, 2778-2783. K. M. Barkigia, L. M. Rajkovic-Blazer, M. T. Pope and C. O. Quicksall, Inorganic Chemistry, 1981, 20, 3318-3323. K. Nakamoto, Infrared and Raman spectra of Inorganic and coordination compounds. Part B, Applications in coordination, organometallic, and bioinorganic chemistry, 5th ed., John Wiley & Sons, New York; Chichester, 1997. X. Zhang, J.-Q. Xu, J.-H. Yu, J. Lu, Y. Xu, Y. Chen, T.-G. Wang, X.-Y. Yu, Q.-F. Yang and Q. Hou, Journal of Solid State Chemistry, 2007, 180, 1949-1956. Y. Ma, Y. Li, E. Wang, Y. Lu, X. Wang, D. Xiao and X. Xu, Inorganica Chimica Acta, 2007, 360, 421-430. A. Hashikawa, M. Fujimoto, Y. Hayashi and H. Miyasaka, Chemical Communications, 2011, 47, 12361-12363. B. Botar, P. Kogerler, A. Muller, R. Garcia-Serres and C. L. Hill, Chemical Communications, 2005, 5621-5623. D. M. Poojary, Y. Zhang, B. Zhang and A. Clearfield, Chemistry of Materials, 1995, 7, 822-827. 103 Chapter 2 – Hybrid Organic-Inorganic Polyoxomolybdates 54. 55. 56. 57. 58. 59. 60. 61. 62. G. Cao, R. C. Haushalter and K. G. Strohmaier, Inorganic Chemistry, 1993, 32, 127-128. S. Gupta, S. Roy, T. N. Mandal, K. Das, S. Ray, R. J. Butcher and S. K. Kar, Journal of Chemical Sciences, 2010, 122, 239-245. W. Henderson and J. S. McIndoe, John Wiley & Sons, 2005, 292. V. A. Pashynska, M. V. Kosevich, H. V. d. Heuvel and M. Claeys, Rapid Communications in Mass Spectrometry, 2006, 20, 755-763. M. W. Kanan, Y. Surendranath and D. G. Nocera, Chemical Society Reviews, 2009, 38, 109-114. R. Brimblecombe, D. R. J. Kolling, A. M. Bond, G. C. Dismukes, G. F. Swiegers and L. Spiccia, Inorganic Chemistry, 2009, 48, 7269-7279. M. W. Kanan and D. G. Nocera, Science, 2008, 321, 1072-1075. J. Barber, Inorganic Chemistry, 2008, 47, 1700-1710. T. G. Carrell, E. Bourles, M. Lin and G. C. Dismukes, Inorganic Chemistry, 2003, 42, 2849-2858. 104 Chapter 3 – Polynuclear Manganese Coordination Complexes 3. POLYNUCLEAR MANGANESE COORDINATION COMPLEXES 105 Chapter 3 – Polynuclear Manganese Coordination Complexes 3.1 INTRODUCTION AND MOTIVATION The chemistry of Mn coordination clusters has gained considerable attention during the last decades due to their relevance to bioinorganic chemistry and potential applications as molecular magnetic materials. Manganese is an essential redox active element in biology, playing an important role in the active sites of many proteins and enzymes responsible for carrying out vital chemical transformations. For example, the photosynthetic water oxidation is catalyzed within photosystem II (PSII) by a {Mn4O4Ca} cluster species.1-5 Mn complexes with various nuclearities and topologies have been intensively studied for their fascinating physical properties and, also for the intrinsic architectural beauty they often possess.6 Depending on the oxidation states, the nuclearity, topology and bridging modes, these coordination clusters can reveal properties of single molecule magnets (SMMs) that exhibit superparamagnetic behaviour below a certain blocking temperature. This type of materials show potential future applications in high-density magnetic storage devices and molecular electronics.6-11 A large number of Mn complexes have been prepared using carboxylate ligands, whereas phosphonates have only rarely been employed.6 However, the versatile nature of these potential ligands and their high binding affinity towards metal ions renders this class of organic molecules to be highly suitable to stablise low-dimenesional molecular species.12-15 Based on these considerations we set out to investigate the influence of different phosphonate ligands on the Mn coordination cluster formation and to explore the properties of the resulting materials. Our aim was to develop new Mn coordination clusters that may display interesting magnetic properties. Thus, we selected a set of three different ligands with different characteristics: phenylphosphonic acid, benzylphosphonic acid and triphenylmethylphosphonic acid. Benzylphosphonic acid contains, in comparison to pheylphosphonic acid, a flexible −CH2− group that can freely rotate to meet the steric requirements imposed by the coordination geometries of aggregating Mn ions. In contrast, triphenylmethylphosphonic acid represents a more bulky ligand, which might restrict oligomerisation and lead to the formation of discrete cages. 106 Chapter 3 – Polynuclear Manganese Coordination Complexes The preparation of new Mn coordination clusters and SMMs is predominantly based on serendipitous synthetic approaches. In order to produce novel materials with interesting magnetic properties, one needs to understand the formation processes of such complex molecular entities and to be able to control the reaction systems to generate targeted species. There is a need to develop rational synthetic approaches for the formation of single molecule magnets.11, 16 Additionally, suitable characterisation tools, that would allow the identification of new and interesting species in solution, or would give the possibility to identify building units and monitor their aggregations, are required.17 NMR spectroscopy is one of the main techniques which can provide detailed information about the structure, reaction state, chemical environment of molecules, etc., but, is of little use when a large number of paramagnetic metal centres are involved. Other techniques must be employed in such cases. Generally, a combination of techniques provides the most complete picture of the chemistry of the reaction system under study.18, 19 Mass spectrometry in combination with X-ray crystallography proved to be an extremely powerful approach to identify and characterize new species that form in solution. It was previously used to study the self-assembly processes of polyoxometalate systems that emerge in solution.20-25 Based on these considerations, we decided to exploit the ESI-MS technique and apply it to characterise manganese-phosphonate coodination clusters system. Surprisingly ESI-MS studies that investigate complex Mn coordination cluster systems have not been reported in the literature. Our aim was to investigate the formation processes of polynuclear manganese clusters. The characterisation of complex reaction mixtures might provide insights into underlying condensation reactions and might lead the development of more rational approaches for the design of single molecule magnets. In this chapter, we demonstrate how electrospray ionisation mass spectrometry can be used to screen complex Mn reaction mixtures in order to identify new polynuclear manganese species that form in solution. Some of these species were isolated and their structures and other physicochemical properties were characterised. 107 Chapter 3 – Polynuclear Manganese Coordination Complexes 3.2 PHOSPHONATE LIGANDS A set of different aromatic phosphonates were used to prepare a series of polynuclear manganese coordination complexes. Phenylphosphonic acid and benzylphosphonic acid, which were commercially available, and triphenylmethylphosphonic acid that was prepared according to literature procedures.26-28 For the synthesis of triphenylmethylphosphonic acid, triphenylcarbinol was treated with phosphorus trichloride to give the corresponding phosphorous acid dichloride according to the equation (i) in Scheme 3.1. Then, the desired phosphonic acid was produced by treating the triphenylmethoxyphosphorous dichloride with an alcoholic solution of potassium hydroxide (Scheme 3.1 (ii)). (i) (ii) OH + PCl3 OPCl2 2KOH + EtOH OPCl2 + HCl PO3H2 -2KCl Scheme 3.1 – Reaction scheme for the synthesis of triphenylmethylphosphonic acid. These phosphonate ligands were then reacted with manganese salts in comproportionation reactions with the aim of generating manganese phosphonate coordination clusters. The geometrical characteristics of the functional group of the organophosphonates and the strong coordination capability of these ligands towards metal ions were expected to promote the formation of unprecedented core structures with novel properties. 108 Chapter 3 – Polynuclear Manganese Coordination Complexes 3.3 POLYNUCLEAR MANGANESE COMPLEXES STABILISED BY ORGANOPHOSPHONATES 3.3.1 Synthesis and characterisation of a pentadecanuclear manganese complex 3.3.1.1 [Μn ΙΙΙ 15 (μ2-H2Ο)2(CΗ3ΟΗ)16(C6Η5PΟ3)20]Cl5·22CH3OH·8H2O (10) Compound 10, [ nΙΙΙ15(μ2-H2 )2(C 3 )16(C6 5P 3)20]Cl5·22CH3OH·8H2O, was prepared by a comproportionation reaction between MnCl2·4H2O and KMnO4, in MeOH, in the presence of phenylphosphonic acid. The reaction mixture was stirred for five hours, filtered and left undisturbed for four days at room temperature. Then, the reaction mixture was transferred into a refrigerator. Rectangular, red-brown crystals of 10 were obtained after keeping the reaction mixture at about 2 °C for another four days. The crystal structure of 10 was determined by single crystal X-ray diffraction measurements. 10 crystallises in the triclinic crystal system in the space group P and contains a pentadecanuclear MnIII complex (Figure 3.1). The manganese cluster core in 10 is centrosymmetric, consisting of two symmetry equivalent {MnIII5O25} units, (Figure 3.1, A), linked by a central {MnIII5O28} unit (Figure 3.1, B) to give the {Mn15} core structure. Within the {MnIII5O25} unit four Mn ions (Mn(1), Mn(3), Mn(4) and Mn(6)) are connected through eight phosphonate ligands to form a twelve membered ring (As ions included). The fifth Mn ion (Mn(2)) caps one side of the ring (Figure 3.2). The “basketlike” {MnIII5O25} unit is further stabilized by five CH3OH ligands. Each Mn ion in A displays a square pyramidal coordination environment in which four O donors are provided by four different phosphonate ligands whilst the remaining O donor arises from a CH3OH molecule. All Mn ions in A adopt the oxidation state +III as calculated using bond valence sum analysis29 based on the bond distances between the Mn centers and the surrounding O donors. The bond distances between the Mn ions and the O donors originating from the organic phosphonate ligands vary between 1.875(7) Å – 1.934(7) Å, whilst the remaining Mn-Omethanol bond lengths are 2.171(7) Å, 2.211(7) Å, 2.139(7) Å, 2.116(8) Å and 2.156(8) 109 Chapter 3 – Polynuclear Manganese Coordination Complexes Å, respectively for Mn(1)-O(33), Mn(2)-O(36), Mn(3)-O(38), Mn(4)-O(34) and Mn(6)O(35) (Table 3.1). (A) (B) (A) Figure 3.1 − Crystal structure of the pentadecanuclear MnIII complex in 10 formally constructed of two {MnIII5O25} units A and a connecting {MnIII5O28} unit B; (schematic representation {MnIII5O25} – blue sphere, {MnIII5O28} – blue rod). Colour code: MnIII blue, P purple, O red, C grey, Cl green (hydrogen atoms have been omitted for clarity). 110 Chapter 3 – Polynuclear Manganese Coordination Complexes a) b) (A) Figure 3.2 − Different views of the {MnIII5O25} unit A. (a) Ball-and-stick and (b) polyhedral representation. Colour code: MnIII blue, P purple, O red, C grey, Cl green (hydrogen atoms have been omitted for clarity). 111 Chapter 3 – Polynuclear Manganese Coordination Complexes Table 3.1 − Selected bond lengths [Å] and bond valence sum (BVS) for compound 10. Atom Bond Bond distances (Å) BVS Assigned oxidation state Mn(1) Mn(1)-O(19) Mn(1)-O(9) Mn(1)-O(20) Mn(1)-O(5) Mn(1)-O(33) 1.888(7) 1.890(7) 1.894(7) 1.928(7) 2.171(7) 3.066 +3 Mn(2) Mn(2)-O(28) Mn(2)-O(22) Mn(2)-O(11) Mn(2)-O(21) Mn(2)-O(36) 1.898(7) 1.904(7) 1.905(7) 1.934(7) 2.211(7) 2.956 +3 Mn(3) Mn(3)-O(30) Mn(3)-O(4) Mn(3)-O(12) Mn(3)-O(6) Mn(3)-O(38) 1.876(7) 1.876(7) 1.892(7) 1.909(7) 2.139(7) 3.182 +3 Mn(4) Mn(4)-O(24) Mn(4)-O(29) Mn(4)-O(23) Mn(4)-O(7) Mn(4)-O(34) 1.875(7) 1.882(7) 1.885(7) 1.902(8) 2.116(8) 3.222 +3 Mn(5) Mn(5)-O(31) Mn(5)-O(10) Mn(5)-O(27) Mn(5)-O(26) Mn(5)-O(37) Mn(5)-O(39) 1.884(7) 1.889(7) 1.893(8) 1.903(7) 2.191(9) 2.364(8) 3.299 +3 Mn(6) Mn(6)-O(25) Mn(6)-O(14) Mn(6)-O(13) Mn(6)-O(8) Mn(6)-O(35) 1.890(7) 1.886(8) 1.896(7) 1.929(7) 2.156(8) 3.077 +3 Mn(7) Mn(7)-O(2) Mn(7)-O(15) Mn(7)-O(3) Mn(7)-O(16) Mn(7)-O(32) Mn(7)-O(39) 1.892(7) 1.900(8) 1.902(7) 1.919(8) 2.210(8) 2.345(7) 3.212 +3 Mn(8) Mn(8)-O(1) Mn(8)-O(1`) Mn(8)-O(17`) Mn(8)-O(17) Mn(8)-O(18`) Mn(8)-O(18) 1.877(7) 1.877(7) 1.912(8) 1.912(7) 2.243(8) 2.243(8) 3.320 +3 Mn(1) ··· Cl(1) Mn(2) ··· Cl(1) Mn(6) ··· Cl(1) 2.644(3) 2.651(3) 2.640(3) The {MnIII5O28} unit B (Figure 3.3) consists of two {MnIII2O11} symmetry equivalent subunits connected to each other through a central {MnO6} unit involving four 112 Chapter 3 – Polynuclear Manganese Coordination Complexes bridging phosphonate ligands. The {MnIII5O28} unit is further stabilized by six CH3OH ligands. The position of the central Mn(8) atom coincides with the position of an inversion symmetry centre. (B) Figure 3.3 − Polyhedral representation of the {MnIII5O28} unit B. Colour code: MnIII blue, P purple, O red, C grey (hydrogen atoms have been omitted for clarity). Within the {MnIII2O11} subunit two Mn ions (Mn(5) and Mn(7)) share a common vertex (O(39)) of their distorted octahedral coordination environment. The O(39) donor arises from a bridging H2O ligand, whilst the remaining coordination sites are occupied by one Omethanol and four O donors originating from four distinct phosphonate ligands. The distorted nature of the octahedra can be observed from the bond angles and bond lengths of the MnIII metal centres. The following bond angles: O(37)-Mn(5)-O(39), O(31)-Mn(5)O(10) and O(27)-Mn(5)-O(26) of 178.8(3)°, 178.0(3)° and 176.6(3)° respectively, deviate from the ideal octahedral angle of 180°, whilst the bond angles O(31)-Mn(5)-O(39), O(26)Mn(5)-O(39), O(10)-Mn(5)-O(26), O(10)-Mn(5)-O(39), O(10)-Mn(5)-O(27) and O(27)Mn(5)-O(39) of 91.6(3)°, 91.1(3)°, 91.0(3)°, 90.3(3)°, 87.8(3)° and 85.7(3)° respectively, deviate from the ideal angle of 90° (Table 3.2). Similar values can be found for the bond angles of the Mn(7) ion, which are summarized in Table 3.2. The four Mn-O bond distances between Mn ions and O donors originating from the organic ligands, vary between 1.884(7) Å – 1.919(8) Å, whilst the remaining two bond lengths involving trans located MeOH and H2O ligands of the octahedral coordination environment are slightly elongated. Thus, the Mn(5) and Mn(7) ions display Jahn-Teller (JT) axial elongation 113 Chapter 3 – Polynuclear Manganese Coordination Complexes typical for MnIII ions.30 Mn-Omethanol bond distances, Mn(5)-O(37) and Mn(7)-O(32) are 2.191(9) Å and 2.210(8) Å, respectively, and the bond lengths between Mn(5), Mn(7) and the μ2-H2O ligand, O(39), are 2.364(8) Å and 2.345(7) Å (Table 3.1). The central Mn(8) ion located on an inversion center displays a more regular octahedral coordination polyhedron compared to those of Mn(5) and Mn(7) ions. The bond angles of the Mn(8) ion are very close to the ideal octahedral angle of 180° and 90°, as seen in Table 3.2. However, substantial distortion can be observed due to the elongated Mn-Omethanol bonds of 2.243(8) Å occupying the apical positions of the Jahn-Teller distorted MnIII ion. Table 3.2 − Selected bond angles [º] for compound 10. Bond Angle (º) O(37)-Mn(5)-O(39) O(31)-Mn(5)-O(10) O(27)-Mn(5)-O(26) O(31)-Mn(5)-O(39) O(26)-Mn(5)-O(39) O(10)-Mn(5)-O(26) O(10)-Mn(5)-O(39) O(10)-Mn(5)-O(27) O(27)-Mn(5)-O(39) 178.8(3) 178.0(3) 176.6(3) 91.6(3) 91.1(3) 91.0(3) 90.3(3) 87.8(3) 85.7(3) O(17)`-Mn(8)-O(17) O(1)-Mn(8)-O(1`) O(18`)-Mn(8)-O(18) O(1)-Mn(8)-O(17) O(17`)-Mn(8)-O(18) O(1)-Mn(8)-O(18) O(1)-Mn(8)-O(18`) O(17)-Mn(8)-O(18) O(1)-Mn(8)-O(17`) 180.0(2) 179.999(4) 179.997(1) 93.9(3) 93.4(3) 92.2(3) 87.8(3) 86.6(3) 86.1(3) Bond Angle (º) O(2)-Mn(7)-O(3) O(15)-Mn(7)-O(16) O(32)-Mn(7)-O(39) O(3)-Mn(7)-O(39) O(15)-Mn(7)-O(3) O(2)-Mn(7)-O(39) O(15)-Mn(7)-O(39) O(3)-Mn(7)-O(16) O(16)-Mn(7)-O(39) 175.8(3) 175.5(3) 174.3(3) 93.8(3) 91.3(3) 90.4(3) 89.5(3) 88.3(3) 86.1(3) Each of the twenty PhPO32- moieties in 10 bridge between three Mn atoms in a η1:η1:η1:μ3 bridging mode. The oxidation states of the Mn ions and the protonation of the bridging donor species, i.e. H2O ligands were established by Mn and O bond valence sum calculations (Table 3.1), close examination of the geometric parameters and detection of MnIII Jahn-Teller (JT) elongation axes. The Mn – (μ2-H2O) distances in 10 are comparable to those observed in other reported aqua-bridged dimanganese complexes (2.25 Å – 2.18 Å).31-34 These distances are found to be significantly longer than the Mn – (μ2-OH-) distances of 2.05 Å – 2.09 Å and the Mn – (μ2-O2-) distances of 1.78 Å – 1.81 Å. The overall charge of the cluster is compensated by five chlorine ions. Two of the chlorine ions are encapsulated within the {MnIII5O25} units and may serve as templates promoting the 114 Chapter 3 – Polynuclear Manganese Coordination Complexes cluster formation. The distances between the chlorine ion Cl(1) and Mn(1), Mn(2) and Mn(6) are 2.644(3) Å, 2.651(3) Å and 2.640(3) Å, respectively (Table 3.1). In the solid state, the packing arrangement of the clusters is stabilized by weak hydrogen bonding interactions between the coordinated methanol molecules and constitutional solvent molecules. A detailed analysis of these interactions was not possible due to the large number of disordered solvent molecules and the complexity of the cluster entity. Weak offset π-π interactions35 between two adjacent clusters can also be observed. These involve the phenyl rings of the phosphonate ligands from the coordination sphere of Mn(2) and its symmetry related partners, resulting in an interplanar separation distance of 4.144(2) Å. The clusters are further linked through halogen bonding interactions between Cl(3) and two hydrogen atoms of two phosphonate rings of two adjacent coordination clusters (2.763(5) Å and 2.805(5) Å, respectively). In addition some weak interaction between Cl(3), Cl(4), Cl(5) and hydrogen atoms of some constitutional solvent molecules preveil. The crystal structure displays small intercluster cavities which are filled with solvent molecules. These channels extend in the direction of the crystallographic b-axis as can be seen in Figure 3.4 (b). a) b) Figure 3.4 – Packing arrangement of the pentadecanuclear MnIII clusters in 10, viewed in the direction of the crystallographic: (a) a-axis and (b) b-axis. Colour code: MnIII blue, P purple, O red, C grey, Cl green (crystallization solvent molecules and hydrogen atoms have been omitted for clarity). 115 Chapter 3 – Polynuclear Manganese Coordination Complexes - PHYSICOCHEMICAL CHARACTERISATION - FT-IR spectroscopy Compound 10 was characterized by infrared spectroscopy. The IR spectrum of 10 is shown in Figure 3.5 and reveals some characteristic stretches of the organophosphonate ligands. The set of bands observed between 3000 – 2800 cm-1 can be attributed to C–H stretching vibrations, while the bands between 1500 – 1430 cm-1 are most likely due to C– C skeletal vibrations of the phenyl rings. Typical C–H out-of-plane bending vibrations of the aromatic rings can be found between 800 – 650 cm-1 as sharp bands. The set of bands between 1200 – 900 cm-1 are due to the different P–O stretching vibrations of the phosphonate groups. Finally, the corresponding O–H stretching vibrations and H–O–H bending vibrations of the crystallization water molecules engaged in H-bonds appear as broad bands centered at ca. 3266 cm-1 and ca. 1628 cm-1, respectively.36-40 95 90 85 1628 Transmittance (%) 80 1487 2835 2944 75 70 3266 65 1438 60 55 50 45 40 750 722 35 30 25 693 1124 1032 20 15 10 4000 983 3600 3200 2800 2400 2000 1800 Wavenumber 1600 1400 1200 1000 800 650 (cm-1) Figure 3.5 – Infrared spectrum of 10. 116 Chapter 3 – Polynuclear Manganese Coordination Complexes - Thermogravimetric analysis The thermal stability of compound 10 was investigated by thermogravimetric analysis (TGA) using a freshly prepared crystalline sample. The analysis was carried out in the temperature range between 30 and 900 °C, in an N2 atmosphere. The TGA curve of 10 shown in Figure 3.6 exhibits an initial weight loss of 10.9 % between 30 – 200 °C that can be attributed to the removal of sixteen coordination CH3OH molecules (calcd: 11.0 %). The other crystallization solvent molecules of 10 may have been lost prior to the TGA during the storage of the sample at room temperature. A further increase of the temperature affected a gradual weight loss between 200 – 500 °C corresponding to the decomposition of the organic ligands, followed by further cluster degradation above 500 °C. 100 Weight % (%) 90 80 70 60 50 0 200 400 600 800 1000 Temperature (°C) Figure 3.6 – Thermogravimetric analysis of 10. 117 Chapter 3 – Polynuclear Manganese Coordination Complexes - Magnetism The magnetic properties for compound 10 were studied on a powdered microcrystalline sample. The temperature dependence of the magnetic susceptibility of 10 was measured between 298 and 1.8 K (Figure 3.7). The χT value of 22.4 cm3 K mol-1 at room temperature is significantly lower than the expected value of 45 cm3 K mol-1 for the presence of fifteen S = 2 MnIII carriers (C = S(S+1) with = 0.12505 cm3 K mol-1 and g = 2).41 This result is characteristic for predominatly strong antiferromagnetic (AF) exchange parameters between spins carriers. Upon lowering the temperature, the χT product decreases down to a minimum of ca. 0.8 cm3 K mol-1 at 1.8 K. This thermal behavior confirms the presence of dominant AF interactions between spin carriers and the low temperature χT value suggests an ST = 0 ground state. This result does not exclude the presence of ferromagnetic interactions but only suggests that the AF interactions are dominant. a) b) Figure 3.7 – (a) Temperature dependence of the χT product of 10 at 0.1 and 1 T. (b) A magnified view of the χT product in (a) between 1 and 100 K. Unfortunately the experimental data can not be fitted to a Curie-Weiss law probably because the magnetic exchange parameters within the cluster core are so large and thus a mean-field model can not be used for this molecular system. The field dependence of the magnetisation for this compound has been measured at low temperatures between 1.8 and 8 K (Figure 3.8). The M vs H plot is typical of a complex with diamagnetic ground state chacterised by weak magnetisation values, almost linear field dependence and no saturation even at high field. 118 Chapter 3 – Polynuclear Manganese Coordination Complexes 8 7 6 5 4 1.8 K 3K 5K 8K 3 2 1 0 0 20000 40000 60000 H / Oe Figure 3.8 – Field dependence of the magnetisation at and below 8 K. It is worth noting that no hysteresis effects upon measureing the field-dependence of the magnetisation (M vs H, 100 – 200 Oe/min) and no ac susceptibility in zero dc field (at 1000 Hz) have been observed above 1.8 K. 119 Chapter 3 – Polynuclear Manganese Coordination Complexes - Mass spectrometry The stability of compound 10 in solution was investigated by electrospray ionization mass spectrometry which provides a very powerful analytical tool for the characterization of coordination compounds in solution.42-48 Negative mode ESI-MS spectra of compound 10 dissolved in a suitable solvent (CH3CN, DMF and DMSO, respectively), were recorded (Table 3.3, Figure 3.9). The mass spectra of these solutions reveal only one major isotopic envelope in the high molecular mass region, centered at m/z = 1684.3 a.m.u. This signal can be assigned to the {H17[MnIV6MnIIIO15(C6H5PO3)6]Cl3}species, which was modelled (Figure 3.9, inset) and good fits between the experimental and simulated patterns could be found. The entire {Mn15} complex could not be identified in the ESI-MS spectrum, presumably due to a facile dissociation involving the partially solvated, central Mn(8) ion (Figure 3.3) . Table 3.3 – ESI-MS assignment for compound 10. Crystals of m/z Species attributed CH3CN 1684.3 {H17[MnIV6MnIIIO15(C6H5PO3)6]Cl3}- → {Mn7} DMF 1684.3 {H17[MnIV6MnIIIO15(C6H5PO3)6]Cl3}- → {Mn7} DMSO 1684.3 {H17[MnIV6MnIIIO15(C6H5PO3)6]Cl3}- → {Mn7} Relative ionic abundance 10 in: Solvent Figure 3.9 – Negative-mode ESI-MS spectra for crystals of 10 dissolved in DMF. Inset: Comparison of the experimental isotopic envelopes (black spectrum) with simulated patterns (red spectrum) for {H17[MnIV6MnIIIO15(C6H5PO3)6]Cl3}- centered at m/z = 1684.3 a.m.u. (cone voltage: 30 V). 120 Chapter 3 – Polynuclear Manganese Coordination Complexes - UV-Vis spectroscopy The UV-Vis absorption spectrum of 10 recorded in CH3CN can be seen in Figure 3.10. The shoulder observed at ca. 280 nm ( = 34000 L mol-1 cm-1) can be attributed to π − π* transitions of the phosphonate ligands, while the weak band at ca. 465 nm ( = 6400 L mol-1 cm-1) involve a d – d transition that can be assigned to a 5T2g III octahedral Mn ions within 10. 5 Eg transition for the 49-52 0.7 Absorbance (a.u.) 0.6 0.6 0.5 0.5 0.4 0.4 0.3 0.3 400 450 500 550 600 0.2 0.1 0.0 300 400 500 600 700 800 Wavelength (nm) Figure 3.10 – UV-Vis spectrum of a 10-5 M solution of 10 in CH3CN. Inset: A section of the UV-Vis spectrum of a 10-4 M solution of 10 in CH3CN. 121 Chapter 3 – Polynuclear Manganese Coordination Complexes Table 3.4 − Crystal data and structural refinement parameters for 10. Compound 10 Empirical formula a Molecular mass/g mol C136H168Cl5Mn15O78P20 -1 a Crystal colour/shape 3 Red brown / rectangular block Crystal size/mm 0.50×0.20×0.20 Crystal system Triclinic Space group P a/ Å 14.5661(5) b/ Å 16.5025(5) c/ Å 51.7573(16) /º 95.1300(10) /º 97.2120(10) /º 102.500(2) 3 V/ Å 11962.6(7) Z 1 Temperature (K) 150(2) -3 Density/Mg m 1.379 -1 Absorp. coef./mm 0.974 F(000) 5044 2 50 max/º Reflections collected 93851 Independent reflections 41381 [R(int) = 0.0329] Data / restraints / parameters 41381 / 0 / 2782 2 a 4671.55 S on F 1.079 R1, wR2 [I>2 (I)] 0.1236, 0.2785 R1, wR2 (all data) 0.1286, 0.2813 Largest diff. peak and hole/e.Å-3 2.823 and -2.275 Excluding solvate molecules 122 Chapter 3 – Polynuclear Manganese Coordination Complexes 3.3.2 Synthesis and characterisation of tridecanuclear manganese complexes 3.3.2.1 [Μn ΙΙΙ 13 (μ4-Ο)2(μ3-Ο)4(μ2-ΟH)2(μ2-CΗ3Ο)4(C6Η5PΟ3)10(C5Η5Ν)5Cl]·3H2O (11) Compound 11 was prepared using a similar synthetic procedure that led to the formation of 10. MnCl2·4H2O and KMnO4 were reacted with phenylphosphonic acid in a MeOH/CH3CN solution and pyridine was added as an organic base. After stirring the reaction mixture for five hours, it was filtered and left at room temperature for slow evaporation of the solvent. During a time period of one week small rhombic red-brown crystals of [ n 13( 4- )2( 3- )4( 2- H)2( 2-C 3 )4(C6 5P 3)10(C5 5 )5Cl]·3H2O (11) were obtained and characterized by single crystal X-ray diffraction measurements. 11 crystallises in the orthorhombic crystal system in the space group P212121. It contains a tridecanuclear MnIII complex that is shown in Figure 3.11. The core structure of 11 features a three layered metal-centered distorted cuboctahedron (Figure 3.11, Figure 3.12). The triangular units A and A` are staggered in a trigonal antiprismatic fashion. They consist of three MnIII ions (Mn(9), Mn(10) and Mn(11), and Mn(4), Mn(8) and Mn(12), respectively) connected through a 2- H- moiety and three phosphonate ligands (Figure 3.13). Unit A is further stabilised by two pyridine molecules, one chloride ion and two methanol molecules, while unit A` is further stabilised by three pyridine and two methanol molecules. Two of the three Mn ions in A and A` (Mn(9) and Mn(10), Mn(8) and Mn(12), respectively) adopt distorted square pyramidal geometries that share a common vertex O(33) and O(10), respectively. The coordination spheres of Mn(9), Mn(10), Mn(8) and Mn(12), each consist of one 2- H- group, three O donors from three distinct phosphonate ligands and a N atom from a pyridine molecule. The bond distances between Mn ions and the 2- H- are 1.899(13) Å, 1.960(13) Å, 1.897(13) Å, and 1.953(13) Å for Mn(9)-O(33), Mn(10)-O(33), Mn(8)-O(10) and Mn(12)O(10), respectively. The bond distances between these latter Mn ions and the O donors originating from the organic phosphonate ligands range between 1.868(11) Å – 1.934(12) Å, while the bond distances between Mn(9)-N(2), Mn(10)-N(3), Mn(8)-N(5) and Mn(12)N(6) are 2.195(14) Å, 2.243(17) Å, 2.255(17) Å and 2.231(15) Å, respectively (Table 3.5). 123 Chapter 3 – Polynuclear Manganese Coordination Complexes (A) (B) (A`) (A) (B) (A`) Figure 3.11 − Crystal structure of the tridecanuclear manganese complex in 11 showing the triangular units A and the hexagonal unit B. Colour code: MnIII blue, P purple, O red, Cl green, C grey (hydrogen atoms have been omitted for clarity). 124 Chapter 3 – Polynuclear Manganese Coordination Complexes Figure 3.12 – The core structure of the tridecanuclear manganese complex in 11. Colour code: MnIII blue, P purple, O red, Cl green, C grey. The third MnIII ion in the A and A` moieties, Mn(11) and Mn(4) respectively, displays a distorted octahedral coordination environment. The distortion arises from the geometrical restrictions of the binding ligands and the nature of the MnIII Jahn-Teller ion, and can be exemplified by examining the bond angles and bond lengths that deviate from the ideal octahedral geometry. Selected bond angles for Mn(11) and Mn(4) ions are listed in Table 3.7. The distorted octahedral geometry of Mn(11) and Mn(4) ions consists of two Ophosphonate donors forming Mn-O bond lengths that range between 1.905(11) Å – 2.092(13) Å, two Omethanol donors which engage in Mn-O bonds that range between 1.943(10) Å – 2.141(13) Å, and a 4- 2- ion forming Mn-O bond lengths of 1.959(10) Å and 1.928(10) Å, for Mn(11)-O(34) and Mn(4)-O(24), respectively. The coordination environment of Mn(11) is completed by a Cl ion that forms a Mn(11)-Cl(1) bond of 2.281(7) Å, while the coordination environment of Mn(4) is completed by a N atom from a pyridine molecule that forms a Mn(4)-N(1) bond of 2.208(11) Å (Table 3.5). The Mn(11)-Cl(1) and Mn(4)N(1) bonds are about 0.1 – 0.3 Å longer than the other bonds of the coordination environments of Mn(11) and Mn(4), respectively. In addition it is worth noting that the bond angles O(6)-Mn(11)-O(17), O(16)-Mn(11)-O(6), O(22)-Mn(4)-O(8) and O(22)125 Chapter 3 – Polynuclear Manganese Coordination Complexes Mn(4)-N(1) of 168.6(5)°, 95.5(5)°, 167.2(5)° and 97.9(4)°, respectively, show the greatest deviation from the ideal octahedral geometry. (A) (A`) Figure 3.13 − Polyhedral representation of the triangular units A and A` in 11. Colour code: MnIII blue, P purple, O red, Cl green, C grey. The hexagonal unit B (Figure 3.14) consists of six Mn atoms (Mn(1), Mn(2), Mn(3), Mn(5), Mn(6), Mn(7)) and one Mn atom located in the centre (Mn(13)). All Mn ions in B adopt a near-octahedral geometry, with an obvious Jahn−Teller (JT) distortion (axial elongation) further supporting the assignment of the +III oxidation state of the Mn ions.30 The distorted octahedral geometry of Mn(1), Mn(2), Mn(5), and Mn(7) consists of three Ophosphonate donors, one Omethanol donor, one 4- 2- ion and a 3- 2- ion. The Mn- Ophosphonate bond distances range between 1.913(10) Å – 2.253(11) Å, the Mn-Omethanol bond distances range between 1.871(12) Å – 1.932(11) Å, the Mn-( 4between 1.901(9) Å – 1.946(9) Å, while the Mn-( 3- 2- ) bond distances range 2- ) bond distances range between 1.908(10) Å – 1.971(12) Å. The distorted octahedral geometry of Mn(3) and Mn(6) consists of four Ophosphonate donors and two 3- 2- ions. The Mn-Ophosphonate bond distances range between 1.906(10) Å – 2.240(11) Å, while the Mn-( 3- 2- ) bond distances range between 1.904(11) Å – 1.937(9) Å. Finally, the octahedral coordination sphere of Mn(13) 126 Chapter 3 – Polynuclear Manganese Coordination Complexes 4- compress two 2- ions and four 3- 2- ions. The Mn-( 4- 2- ) bond distances are 2.369(10) Å and 2.423(10) Å for Mn(13)-O(34) and Mn(13)-O(24), respectively, while the Mn-( 3- 2- ) bond distances range between 1.935(9) Å – 1.967(9) Å. The distorted nature of the octahedral geometry of the metal ions within the hexagonal sub-unit can be observed from the bond lengths and the bond angles of the MnIII metal centres that deviate from the ideal octahedral geometry. Selected bond lengths and bond angles for the Mn ions in 11 are listed in Table 3.5 and Table 3.7. (B) Figure 3.14 – Ball-and-stick representation of the hexagonal unit B in 11. Colour code: MnIII blue, P purple, O red, C grey. The hexagonal unit B is stabilized to the outside by six phosphonate ligands and four methanol molecules. The units A and B are connected to each other by a combination of O2-, CH3O- and phosphonate ligands to give the {Mn13} core structure. All four methanol molecules are deprotonated and bridge between two Mn ions (Mn(1) and Mn(4), Mn(2) and Mn(4), Mn(5) and Mn(11), Mn(7) and Mn(11), respectively). 127 Chapter 3 – Polynuclear Manganese Coordination Complexes Table 3.5 − Selected bond lengths [Å] and bond valence sum (BVS) for compound 11. Atom Bond Bond distances (Å) BVS Assigned oxidation state Mn(1) Mn(1)-O(37) Mn(1)-O(39) Mn(1)-O(28) Mn(1)-O(24) Mn(1)-O(36) Mn(1)-O(23) 1.908(10) 1.909(12) 1.921(11) 1.946(9) 2.156(10) 2.162(11) 3.271 +3 Mn(2) Mn(2)-O(22) Mn(2)-O(24) Mn(2)-O(21) Mn(2)-O(25) Mn(2)-O(13) Mn(2)-O(23) 1.882(12) 1.901(9) 1.913(10) 1.926(11) 2.209(9) 2.229(10) 3.281 +3 Mn(3) Mn(3)-O(26) Mn(3)-O(4) Mn(3)-O(30) Mn(3)-O(37) Mn(3)-O(7) Mn(3)-O(36) 1.904(11) 1.906(10) 1.909(12) 1.928(10) 2.194(11) 2.240(11) 3.238 +3 Mn(4) Mn(4)-O(24) Mn(4)-O(14) Mn(4)-O(22) Mn(4)-O(39) Mn(4)-O(8) Mn(4)-N(1) 1.928(10) 2.020(12) 2.072(11) 2.073(11) 2.092(13) 2.208(11) 2.799 +3 Mn(5) Mn(5)-O(34) Mn(5)-O(3) Mn(5)-O(43) Mn(5)-O(26) Mn(5)-O(11) Mn(5)-O(7) 1.913(9) 1.926(9) 1.932(11) 1.971(12) 2.154(11) 2.220(10) 3.127 +3 Mn(6) Mn(6)-O(27) Mn(6)-O(20) Mn(6)-O(19) Mn(6)-O(25) Mn(6)-O(13) Mn(6)-O(35) 1.920(10) 1.924(9) 1.926(11) 1.937(9) 2.226(10) 2.226(12) 3.117 +3 Mn(7) Mn(7)-O(17) Mn(7)-O(18) Mn(7)-O(34) Mn(7)-O(27) Mn(7)-O(11) Mn(7)-O(35) 1.871(12) 1.930(12) 1.934(10) 1.967(11) 2.181(10) 2.253(11) 3.153 +3 Mn(8) Mn(8)-O(12) Mn(8)-O(10) Mn(8)-O(2) Mn(8)-O(38) Mn(8)-N(5) 1.868(11) 1.897(13) 1.910(15) 1.911(12) 2.255(17) 3.122 +3 Mn(9) Mn(9)-O(1) Mn(9)-O(32) Mn(9)-O(33) Mn(9)-O(40) Mn(9)-N(2) 1.880(13) 1.887(11) 1.899(13) 1.911(12) 2.195(14) 3.200 +3 128 Chapter 3 – Polynuclear Manganese Coordination Complexes Atom Bond Bond distances (Å) BVS Assigned oxidation state Mn(10) Mn(10)-O(29) Mn(10)-O(31) Mn(10)-O(41) Mn(10)-O(33) Mn(10)-N(3) 1.881(13) 1.884(10) 1.893(13) 1.960(13) 2.243(17) 3.082 +3 Mn(11) Mn(11)-O(16) Mn(11)-O(43) Mn(11)-O(34) Mn(11)-O(6) Mn(11)-O(17) Mn(11)-Cl(1) 1.905(11) 1.943(10) 1.959(10) 2.057(12) 2.141(13) 2.281(7) 3.340 +3 Mn(12) Mn(12)-O(9) Mn(12)-O(42) Mn(12)-O(5) Mn(12)-O(10) Mn(12)-N(6) 1.871(12) 1.901(11) 1.934(12) 1.953(13) 2.231(15) 3.019 +3 Mn(13) Mn(13)-O(26) Mn(13)-O(25) Mn(13)-O(27) Mn(13)-O(37) Mn(13)-O(34) Mn(13)-O(24) 1.935(9) 1.945(9) 1.946(11) 1.967(9) 2.369(10) 2.423(10) 2.766 +3 Mn(1) ··· Mn(4) Mn(1) ··· Mn(3) Mn(1) ··· Mn(2) Mn(1) ··· Mn(13) Mn(2) ··· Mn(4) Mn(2) ··· Mn(6) Mn(2) ··· Mn(13) Mn(3) ··· Mn(13) 2.940(4) 3.023(4) 3.099(3) 3.119(4) 2.944(4) 3.023(4) 3.147(4) 2.905(3) Mn(3) ··· Mn(5) Mn(5) ··· Mn(11) Mn(5) ··· Mn(7) Mn(5) ··· Mn(13) Mn(6) ··· Mn(13) Mn(6) ··· Mn(7) Mn(7) ··· Mn(11) Mn(7) ··· Mn(13) 3.031(4) 2.929(4) 3.072(3) 3.087(4) 2.895(4) 3.033(4) 2.993(4) 3.106(4) Table 3.6 − Bond valence sum (BVS) calculations for some O atoms in 11. Atom O(24) O(34) O(26) O(37) O(27) O(25) O(10) O(33) BVS 2.090 2.063 1.866 1.877 1.825 1.865 1.284 1.269 Assignment a μ4-O2μ4-O2μ3-O2μ3-O2μ3-O2μ3-O2μ2-OHμ2-OH- An oxygen BVS in the ∼1.8-2.0, ∼1.0-1.2, and ∼0.2-0.4 ranges is indicative of non-, single- and double protonation, respectively.7, 53 a 129 Chapter 3 – Polynuclear Manganese Coordination Complexes Table 3.7 − Selected bond angles [º] for compound 11. Bond Angle (º) Bond Angle (º) O(28)-Mn(1)-O(24) O(39)-Mn(1)-O(37) O(23)-Mn(1)-O(36) O(28)-Mn(1)-O(23) O(39)-Mn(1)-O(28) O(39)-Mn(1)-O(23) O(39)-Mn(1)-O(36) O(37)-Mn(1)-O(23) O(39)-Mn(1)-O(24) O(24)-Mn(1)-O(23) 178.1(5) 171.2(5) 169.1(4) 97.5(4) 94.0(5) 93.5(5) 91.8(5) 89.5(4) 84.3(5) 81.8(4) O(4)-Mn(3)-O(37) O(30)-Mn(3)-O(26) O(7)-Mn(3)-O(36) O(4)-Mn(3)-O(36) O(30)-Mn(3)-O(7) O(30)-Mn(3)-O(37) O(26)-Mn(3)-O(36) O(30)-Mn(3)-O(4) O(30)-Mn(3)-O(36) O(37)-Mn(3)-O(36) 178.2(4) 176.8(4) 170.8(4) 98.3(5) 97.2(4) 93.3(4) 90.2(4) 88.4(4) 89.2(4) 81.2(4) O(14)-Mn(4)-O(39) O(24)-Mn(4)-N(1) O(22)-Mn(4)-O(8) O(22)-Mn(4)-N(1) O(24)-Mn(4)-O(14) O(8)-Mn(4)-N(1) O(39)-Mn(4)-N(1) O(14)-Mn(4)-O(8) O(14)-Mn(4)-N(1) O(14)-Mn(4)-O(22) 175.3(5) 171.4(4) 167.2(5) 97.9(4) 96.6(4) 94.9(4) 92.0(4) 93.1(5) 90.6(4) 87.4(5) O(16)-Mn(11)-O(43) O(34)-Mn(11)-Cl(1) O(6)-Mn(11)-O(17) O(16)-Mn(11)-O(6) O(17)-Mn(11)-Cl(1) O(43)-Mn(11)-Cl(1) O(6)-Mn(11)-Cl(1) O(16)-Mn(11)-O(34) O(16)-Mn(11)-Cl(1) O(16)-Mn(11)-O(17) 172.4(5) 171.3(4) 168.6(5) 95.5(5) 95.4(4) 95.0(4) 95.0(4) 92.4(4) 91.8(4) 88.8(5) O(34)-Mn(13)-O(24) O(26)-Mn(13)-O(25) O(37)-Mn(13)-O(27) O(26)-Mn(13)-O(24) O(27)-Mn(13)-O(24) O(26)-Mn(13)-O(27) O(26)-Mn(13)-O(37) O(26)-Mn(13)-O(34) O(37)-Mn(13)-O(24) O(25)-Mn(13)-O(24) 178.0 (4) 177.4(6) 176.6(6) 105.2(4) 102.4(4) 98.9(4) 81.6(4) 76.4(4) 74.2(4) 72.5(4) 130 Chapter 3 – Polynuclear Manganese Coordination Complexes The phosphonate ligands bridge between the hexagonal unit B and the triangular units A in η1:η1:η1:μ3 (P(4), P(5), P(7) and P(8)), and η1:η1:η2:μ4 (P(1), P(2), P(3), P(6), P(9), and P(10)) bridging modes (Figure 3.15). The oxidation states of the Mn ions and the protonation states of O ions were assigned by bond valence sum analysis (Table 3.5, Table 3.6). In the solid state the tridecanuclear clusters are intermolecularly linked by weak hydrogen bonds involving constitution solvent molecules resulting in a relatively densely packed structure that can be seen in Figure 3.16. Figure 3.15 – Ball-and-stick representation of the {Mn13} core structure in 11, showing the phosphonates’ bridging modes. Colour code: MnIII blue, P purple, O red, Cl green, C grey. 131 Chapter 3 – Polynuclear Manganese Coordination Complexes Figure 3.16 – Packing arrangement of the tridecanuclear manganese clusters in 11 viewed in the direction of the crystallographic a- and b-axis. Colour code: Mn blue, P purple, O red, C grey, (crystallization solvent molecules and hydrogen atoms have been omitted for clarity). The metal-centered cuboctahedral core structure of 11 is unprecedented but can be regarded as a member of a family of {Mn13} complexes stabilised by carboxylate and phosphonate ligands, such as: [ nΙV nΙΙΙ6 nΙΙ6O8(O2CPh)12(OEt)6], [ nΙV nΙΙΙ6 nΙΙ6O8 (OH)6(ndc)6], [ nΙV nΙΙΙ6 nΙΙ6O8(OEt)5(OH)(ndc)6], [ nΙV nΙΙΙ6 nΙΙ6O8(OMe)6(ndc)6], (ndcH2 = 1,8-naphthalenedicarboxylic [ nΙΙ nΙΙΙ12O6( H)6( Cl2, [ nIII13O6(t-BuP 3PC6 3)10( 11)10(py)6], acid), [ n13O8( Et)6(O2CC6H4OPh)12], [ nΙΙ nΙΙΙ12O8Cl6(t-BuP H)2(N3)6(MeCOOH)2(H2O)2]-.12, {Mn13} complexes the neutral [ nΙΙ nΙΙΙ12O6( H)6( 48, 3PC6 3)8][ 53-55 nΙΙ(CH3CN)6] Amongst these 11)10(py)6] compound reported by Zheng et al. reveals the closest similarity to 11.12 Zheng used cyclohexyl phosphonic acid as a stabilising ligand and obtained a mixed-valent {Mn13} cluster in which the central Mn ion adopts the oxidation state +II. There are a number of other elements that differentiate Zheng`s compound from compound 11. The triangular units in Zheng`s compound consist of one square pyramidally and two octahedrally coordinated 132 Chapter 3 – Polynuclear Manganese Coordination Complexes MnIII ions. In contrast to 11 in the reported compound four 4- 2- and two bridge between the triangular and hexagonal units. In addition the four 11 are replaced with four 2- The [ nIII13O6(t-BuP 2- 3- 2-CH3 ligands ligands in H- ligands in Zheng`s compound. H)2(N3)6(MeCOOH)2(H2O)2]- coordination cluster 3)10( reported by Chen et al. also shows similarities to the here presented complex. It comprises of thirteen octahedrally coordinated MnIII ions connected through six 4- 2- oxo bridges, ten t-BuPO32-, two OH-, six N3-, two CH3COOH and two H2O ligands.48, 55 Another {Mn13} complex with metal centred cuboctahedral cluster-cores stabilised by tertbutylphosphonates, [ nΙΙ nΙΙΙ12( 4-O8)( 4-Cl6)(t-BuP 3)8], was reported by Schmitt et al. In this compound the central Mn ion adopts the oxidation state +II and is coordinated by eight 4- 2- oxo ligands. The remaining twelve MnIII atoms adopt tetragonally distorted octahedral coordination environments, and the peripheral ligation is provided by six Clligands and eight t-BuPO32- ligands.48 Structurally related compounds were also reported by Christou et al.53 and Murrie et al.54 in which the metal centred cuboctahedral atom arrangement is stabilised by carboxylate ligands. The core structures of these reported mixed valent species comprise of one central MnIV ion, six MnIII ions forming the triangular units and six MnII ions forming the hexagonal unit. In his paper, Christou reports the structure and properties [ nΙV nΙΙΙ6 nΙΙ6O8(OH)6(ndc)6], [ nΙV nΙΙΙ6 nΙΙ6O8(OMe)6(ndc)6] of the following {Mn13} complexes: [ nΙV nΙΙΙ6 nΙΙ6O8(OEt)5(OH)(ndc)6], and [Mn13O8(O2CPh)12(OEt)6] (ndcH2 = 1,8- naphthalenedicarboxylic acid). All the Mn atoms in these compounds adopt a near5- octahedral coordination geometry and are connected by six Additional bridging and peripheral ligation is provided by 2- 3- and two H-, 3-Et 3- 2- or ligands. 3-Me - ions, and ndc2- or benzoate ligands, respectively. The {Mn13} core in Murrie`s compound [ n13O8( Et)6(O2CC6H4OPh)12] is stabilised by six 4- 2- , two 3- 2- , six 3-Et - and twelve 2-phenoxybenzoate ligands. The central MnIV and the six MnIII ions in Murrie`s compound display a distorted octahedral coordination environment, whilst the remaining six MnII ions adopt a distorted square pyramidal geometry. 133 Chapter 3 – Polynuclear Manganese Coordination Complexes - PHYSICOCHEMICAL CHARACTERISATION - FT-IR spectroscopy The infrared spectrum of 11 presented in Figure 3.17 is very similar to that of 10. Characteristic C–C skeletal vibrations of the aromatic rings appear in the 1600 – 1430 cm-1 region, while the C–H out-of-plane bending vibrations can be observed in the 800 – 650 cm-1 region. Compared with 10 the IR spectrum of 11 displays some extra bands in the 1200 – 900 cm-1 region due to C–C stretching vibrations of the pyridine rings, along with the different P–O stretching vibrations of the phosphonate groups present in this region. Some O–H stretching vibrations appear as a broad band at ca. 3500 cm-1.36-40, 56, 57 Figure 3.17 – Infrared spectrum of 11. 134 Chapter 3 – Polynuclear Manganese Coordination Complexes - Thermogravimetric analysis The thermogravimetric analysis for compound 11 was performed using a freshly prepared crystalline sample, in the temperature range 30 – 900 °C, in a N2 atmosphere (Figure 3.18). Upon thermolysis 11 undergoes a weight loss of 6.8 % in the temperature range 30 – 200 °C which may be assigned to the loss of three crystallization H2O molecules and four coordination CH3OH molecules (calcd: 6.1 %). This assignment is in agremment with composition determined by elemental analysis. The next step in the TGA curve that corresponds to a weight loss of 12.7 % between 200 – 300 °C can be attributed to the loss of five pyridine molecules (calcd: 13.1 %). A further weight loss that occurs in two distinct steps can be observed above 300 °C. This can be associated with the degradation of the organic ligands within 11 and cluster degradation processes resulting in the probable formation of metal oxide materials. Weight % (%) 100 90 80 70 60 50 0 200 400 600 800 1000 Temperature (ºC) Figure 3.18 – Thermogravimetric analysis of 11. 135 Chapter 3 – Polynuclear Manganese Coordination Complexes - Magnetism The temperature dependence of the magnetic susceptibility of 11 was measured between 298 and 1.8 K (Figure 3.19). The χT value of 30.7 cm3 K mol-1 at room temperature is significantly lower than the expected value of 39 cm3 K mol-1 for the presence of thirteen S = 2 MnIII carriers (C = S(S+1) with = 0.12505 cm3 K mol-1 and g = 2).41 This result highlights dominant antiferromagnetic (AF) couplings between spin carriers. Upon lowering the temperature, the χT product decreases down to a minimum of ca. 18.3 cm3 K mol-1 at about 20 K. This thermal behavior confirms the presence of dominant AF interactions between spin carriers. The result does not exclude the presence of ferromagnetic interactions but only suggests that the AF interactions are dominant. Experimental data can be fitted to a Curie-Weiss law for a temperature above 120 K, with C = 38.3(5) cm3 K mol-1 and = -68(2) K. The Curie constant is in good agreement with the expected value (39 cm3 K/mol) arising from thirteen MnIII S = 2 spins. Below 20 K, at 1000 Oe the χT product increases to reach 19.3 cm3 K mol-1 at 1.8 K. This value suggests an ST = 6 ground state. At 1 T, below 10 K, the χT product decreases to reach 7.9 cm3 K mol-1 at 1.8 K due to field saturation effects of the magnetisation. a) b) Figure 3.19 – (a) Temperature dependence of the χT product of 11 at 0.1 and 1 T. (b) A magnified view of the χT product in (a) between 1 and 100 K. The green solid line corresponds to the best fit of the experimental data with the Curie-Weiss law (C = 38.3(5) cm3 K mol-1 and = -68(2) K). The field dependence of the magnetisation for this compound has been measured at low temperatures between 1.8 and 8 K (Figure 3.20, a). The magnetisation at low field displays a rapid increase without inflexion point confirming the absence of weak antiferromagnetic interactions and the presence of a well-defined ground state. The high 136 Chapter 3 – Polynuclear Manganese Coordination Complexes field behavior that displays a non-linear increase without clear saturation even at 1.8 K at 7 T, suggests the presence of magnetic anisotropy. It is worth noting that the presence of low-lying excited states and also inter-complex magnetic interactions could contribute to this M vs H data even if these two effects are not reflected in χT vs T data at 0.1 T. The data represented in M vs H/T plots at different temperatures (Figure 3.20, b) confirm the presence of anisotropy intrinsic to MnIII metal ions as the data are not superposed on a single master-curve as expected for an isotropic system with a well defined spin ground state. At 1.8 K, the magnetisation reaches 12.2 B at 7 T which is in agreement with an ST = 6 ground state. a) b) Figure 3.20 – (a) M vs H and (b) M vs H/T data at and below 8 K. Both measurements, the field dependence of the magnetisation at 1.8 K and the temperature dependence of the χT product at 1000 Oe, are coherent with an ST = 6 ground state. This experimental value can be explained with a configuration of eight MnIII centres whose spin vectors are oriented in an opposite direction to those of the five remaining Mn centres, as a possible result of competing interactions between spin carriers. The M vs H data at 1.8 K do not show any sign of slow relaxation i.e. hysteresis effects. Nevertheless, the ac susceptibility in zero dc field has been measured to probe possible slow dynamics of the magnetisation for this compound. Clearly at temperatures below 4 K (for frequency around 10000 Hz), slow relaxation of the magnetisation is observed based on the appearance of an out-of-phase signal. 137 Chapter 3 – Polynuclear Manganese Coordination Complexes b) 10 Hz 30 Hz 60 Hz 100 Hz 150 Hz 200 Hz 300 Hz 400 Hz 600 Hz 800 Hz 10 χ' / cm3 mol-1 8 6 1000 Hz 1200 Hz 1500 Hz 2000 Hz 3000 Hz 4000 Hz 5000 Hz 6000 Hz 8000 Hz 10000 Hz 4 10 Hz 30 Hz 60 Hz 100 Hz 150 Hz 200 Hz 300 Hz 400 Hz 600 Hz 800 Hz 1.6 χ"/ cm3 mol-1 a) 1.2 0.8 0.4 2 Hdc = 0 Oe Hdc = 0 Oe 0 1000 Hz 1200 Hz 1500 Hz 2000 Hz 3000 Hz 4000 Hz 5000 Hz 6000 Hz 8000 Hz 10000 Hz 0 1 2 3 4 5 6 7 8 1 1.5 2 2.5 3 3.5 4 4.5 5 T/K T/K Figure 3.21 – Temperature dependence of the (a) in-phase (χ') and (b) out-of-phase (χ") ac susceptibility under zero dc field for 11. In consecutive experiments we measured the ac susceptibility as a function of the frequency at different temperatures. This measurement was performed in zero dc field with the objective to more accurately estimate the relaxation time. The observed frequency dependence and resulting shape of the curve almost by itself demonstrates that this compound is consistent with an SMM.7 b) 12 χ' / cm3 mol-1 10 8 6 4 2 0 10 2 Hdc = 0 Oe 1.8 K 1.9 K 2K 2.1 K 2.2 2.3 2.4 2.5 K K K K 100 2.6 2.7 2.8 2.9 K K K K 1000 3K 3.2 K 3.4 K 10000 χ'' / cm3 mol-1 a) 1.5 1.8 K 1.9 K 2K 2.1 K 2.2 K 2.3 K 2.4 K 2.5 K 2.6 K 2.7 K 2.8 K 2.9 K 3K 3.2 K 3.4 K 1 Hdc = 0 Oe 0.5 0 10 100 1000 10000 ν / Hz ν / Hz Figure 3.22 – Frequency dependence of the (a) in-phase (χ') and (b) out-of-phase (χ") ac susceptibility from 1.8 to 3.4 K under zero dc field for 11. Using a scaling technique, the relaxation time of the compound was determined up to 3.4 K. 138 Chapter 3 – Polynuclear Manganese Coordination Complexes a) b) 0.2 10-4 τ = τ0 exp( χ"/χ'0 0.15 0.1 Hdc = 0 Oe 0.05 0 10 100 1000 10 τ /s 1.8 K 1.9 K 2K 2.1 K 2.2 K 2.3 K 2.4 K 2.5 K 2.6 K 2.7 K 2.8 K 2.9 K 3K 3.2 K 3.4 K eff /kBT) -5 10-6 Hdc = 0 Oe 10 -7 10 -8 10000 τ0 = 8.8 10-10 s 0 0.1 0.2 = 19.2 K 0.4 0.5 0.3 T-1 αν / Hz eff/kB 0.6 /K-1 Figure 3.23 – (a) χ"/χ' vs from 1.8 to 3.4 K under zero dc field; (b) Magnetisation relaxation time ( ) vs T-1plot for 11 under zero dc field (the solid line corresponds to the Arrhenius law). From these data, the relaxation time can be deduced between 1.9 and 3 K and fitted to an Arrhenius law. The exponential increase of the relaxation time allowed us to determine the energy barrier of 19.2 K of the thermally activated regime while 8.8 × 10 -10 0 is about s. The ac susceptibility has been measured at dc fields at 1.8 K in order to induce possible quantum relaxation pathways in zero-field and to probe the quantum contribution to the observed relaxation above 1.8 K. a) 0 Oe 200 Oe 400 Oe 600 Oe 1000 Oe 1500 Oe 15 Oe Oe Oe Oe Oe Oe b) 2.5 2 χ' / cm3 mol-1 χ' / cm3 mol-1 T = 1.8 K 2000 2500 3000 4000 5000 6000 10 5 1.5 1 0 Oe 200 Oe 400 Oe 600 Oe 1000 Oe 1500 Oe 2000 Oe 2500 Oe 3000 Oe 4000 Oe 5000 Oe 6000 Oe T = 1.8 K 0.5 0 10 100 1000 10000 0 10 100 1000 10000 ν / Hz ν / Hz Figure 3.24 – Frequency dependence of the (a) in-phase (χ') and (b) out-of-phase (χ") ac susceptibility with dc fields at 1.8 K for 11. With a small dc field, the characteristic frequency increases which demonstrate that the quantum pathway of magnetisation relaxation is only negligible above 1.8 K. Therefore, the ac susceptibility measurements under dc field have not been performed. 139 Chapter 3 – Polynuclear Manganese Coordination Complexes - Mass spectrometry Negative mode ESI-MS spectra of compound 11 dissolved in a suitable solvent have been recorded (Table 3.8, Figure 3.25) in order to investigate the stability of the tridecanuclear manganese cluster in solution.42-48 The mass spectra of 11 dissolved in DMF and DMSO reveal the presence of two isotopic envelopes centered at m/z = 1279.1 a.m.u and m/z = 2559.1 a.m.u. These signals were attributed to a -2 charged and a -1 charged {Mn13} species as shown in Table 3.8. The mass spectrum of 11 in CH3CN exhibits a major signal centered at m/z = 1278.5 a.m.u which corresponds to the doubly charged species {H[MnIV2MnIII11 10(C 3 )4(C6 5P 3)10]} 2- . The ESI-MS studies using the crystalline material confirmed the stability of the {Mn13} cluster in solution. The signals of the identified species were modelled and good fits between the experimental and simulated isotopic envelopes confirm the assignments. Table 3.8 – ESI-MS assignment for compound 11. Crystals of Solvent m/z CH3CN 1278.5 {H[MnIV2MnIII11 10(C 3 )4(C6 5P 23)10]} → {Mn13} 1279.1 {H2[MnIVMnIII12 10(C 3 )4(C6 5P 23)10]} → {Mn13} 2559.1 {H[MnIV3MnIII10 10(C 3 )4(C6 5P 3)10]} → {Mn13} IV 2- DMF 11 in: 1279.1 {H2[Mn MnIII12 10(C 3 )4(C6 5P 3)10]} → {Mn13} 2559.1 {H[MnIV3MnIII10 10(C 3 )4(C6 5P 3)10]} → {Mn13} Relative ionic abundance DMSO Species attributed Figure 3.25 – Negative-mode ESI-MS spectra for crystals of 11 dissolved in DMF. Inset: Comparison of the experimental isotopic envelopes (black spectrum) with simulated patterns (red spectrum) for {H2[MnIVMnIII12 10(C 3 )4(C6 5P 3)10]}2- and {H[MnIV3MnIII10 10(C 3 )4 (C6 5P 3)10]}- centered at m/z = 1279.1 a.m.u. and m/z = 2559.1 a.m.u., respectively (cone voltage: 30 V). 140 Chapter 3 – Polynuclear Manganese Coordination Complexes - UV-Vis spectroscopy The UV-Vis absorption spectrum of 11 recorded in CH3CN is presented in Figure 3.26. The π − π* transitions of the organic ligands can be observed by a band centred at ca. 270 nm ( = 162000 L mol-1 cm-1), while the d – d transition of the octahedrally coordinated MnIII ions is displayed as a weaker band at ca. 446 nm ( = 3400 L mol-1 cm-1). The signal corresponds to a 5T2g 5 Eg transition for the octahedrally coordinated MnIII ions within 11.49-52 Absorbance (a.u.) 1.8 1.6 0.35 1.4 0.30 1.2 0.25 0.20 1.0 0.15 0.8 0.10 400 0.6 450 500 550 600 0.4 0.2 0.0 300 400 500 600 700 800 Wavelength (nm) Figure 3.26 – UV-Vis spectrum of a 10-5 M solution of 11 in CH3CN. Inset: A section of the UV-Vis spectrum of a 10-4 M solution of 11 in CH3CN. 141 Chapter 3 – Polynuclear Manganese Coordination Complexes Table 3.9 − Crystal data and structural refinement parameters for 11. Compound 11 Empirical formula a Molecular mass/g mol C89H89ClMn13N5O42P10 -1 a Crystal colour/shape 3 Red brown / rhombic Crystal size/mm 0.30×0.30×0.15 Crystal system Orthorhombic Space group P212121 a/ Å 18.8873(6) b/ Å 23.1480(7) c/ Å 25.3591(7) /º 90 /º 90 /º 90 3 V/ Å 11087(6) Z 4 Temperature (K) 100(2) -3 Density/Mg m 1.778 -1 Absorp. coef./mm 13.959 F(000) 5920 2 100 max/º Reflections collected 42875 Independent reflections 11337 [R(int)=0.1006] Data / restraints / parameters 11337 / 0 / 1075 2 a 2960.05 S on F 1.028 R1, wR2 [I>2 (I)] 0.0753, 0.1869 R1, wR2 (all data) 0.0989, 0.2047 Largest diff. peak and hole/e.Å-3 1.353 and -0.712 Excluding solvate molecules 142 Chapter 3 – Polynuclear Manganese Coordination Complexes 3.3.2.2 [Μn Compound (C5 5 ΙΙΙ 13 (μ4-Ο)2(μ3-Ο)4(μ2-ΟH)2(μ2-CΗ3Ο)4(C6Η5CΗ2PΟ3)10(C5Η5Ν)6]Cl·5H2O (12) 12, [ n 13( 4- )2( 3- )4( 2- H)2( 2-C 3 )4(C6 5C 2P 3)10 )6]Cl·5H2O was obtained by the replacement of the phenylphosphonic acid ligand with benzylphosphonic acid during the preparation procedure that led to the formation of 11. Rectangular brown crystals of 12 separate from the reaction mixture in a time period of about four days and were characterized by single crystal X-ray diffraction measurements. 12 crystallises in the triclinic crystal system in the space group P and is structurally related with 11. The core structure within 12 is centrosymetric, revealing a similar metal-centred, distorted cuboctahedral arrangement as observed in 11. However, here the atom positions are slightly displaced leading to a lower symmetry compared to 11 (Figure 3.27). The triangular units A in 12 appear slightly moved and are not centrally located above and below the hexagonal plane. (A) (B) (A) Figure 3.27 − Crystal structure of the tridecanuclear manganese complex in 12. Colour code: MnIII blue, P purple, O red, C grey (hydrogen atoms have been omitted for clarity). 143 Chapter 3 – Polynuclear Manganese Coordination Complexes Compound 11 Compound 12 a) 7.785(9) Å 7.639(4) Å b) 83.61(1)ᵒ 79.43(3)⁰ ~ 0.6 Å 1.227(1) Å 6.352(8) Å 89.13(1)ᵒ 6.461(4) Å 6.359(1) Å 6.377(1) Å c) 89.61(3)ᵒ 1.228(2) Å Figure 3.28 – (a) Comparison of the metal-centred distorted cuboctahedral arrangement of the {Mn13} clusters in 11 and 12; (b) The oblique triangular antiprisms constructed using the two triangular units A within 11 and 12, respectively; (c) Schematic representation showing the position of the centroid of one base relative to the other base in the triangular antiprism representation of units A within 11 and 12, respectively. 144 Chapter 3 – Polynuclear Manganese Coordination Complexes Compound 11 Compound 12 a) 5.482(4) Å 5.375(4) Å 5.408(4) Å 5.565(6) Å 3.640(4) Å 3.656(1) Å b) 3.031(4) Å 2.988(5) Å 3.023(4) Å 3.072(4) Å 3.158(4) Å 3.091(1) Å Figure 3.29 – Comparison of the structural parameters of the triangular units A and hexagonal unit B within 11 and 12. (a) Top view of the oblique triangular antiprisms constructed using the two triangular units A within 11 and 12, respectively; (b) Hexagonal unit B within 11 and 12, respectively. A close examination of complexes 11 and 12 reveals the structural differences between the compounds (Figure 3.28, Figure 3.29). The first one is represented by the peripheral ligation which is accomplished by ten phenylphosphonates, five pyridine molecules and a chloride ion for 11 and ten benzylphosphonates and six pyridine molecules for 12. Then, the two triangular units in 11 and 12 form oblique triangular antiprisms that have edges of 7.639(4) Å and 7.785(9) Å (distances measured between Mn(8) ··· Mn(9) in 11 and Mn(5) ··· Mn(7) in 12), respectively and show an angle of 83.61(1)⁰ and 79.43(3)⁰, respectively between one base and the centroid of the other triangular base (Figure 3.28). Thus, the triangular units A in 12 appear moved aside with about 0.6 Å compared to those in 11. Also, the structural parameters within the triangular and hexagonal units in 12 differ slightly from those in 11. The geometry of the units in 12 are characterised by interatomic Mn-Mn distances of 3.656(1) Å for Mn(5) ··· Mn(7`), 5.408(4) Å for Mn(5) ··· Mn(6) and 5.565(6) Å for Mn(6) ··· Mn(7`), compared with 3.640(4) Å for Mn(9) ··· Mn(10), 5.375(4) Å for Mn(9) ··· Mn(11) and 5.482(4) Å for Mn(10) ··· Mn(11) in 11 (Figure 3.29). 145 Chapter 3 – Polynuclear Manganese Coordination Complexes (A) Figure 3.30 − Polyhedral representation of the triangular unit A in 12. Colour code: MnIII blue, P purple, O red, C grey. (B) Figure 3.31 – Ball-and-stick representation of the hexagonal unit B in 12. Colour code: MnIII blue, P purple, O red, C grey. 146 Chapter 3 – Polynuclear Manganese Coordination Complexes Table 3.10 − Selected bond lengths [Å] and bond valence sum (BVS) for compound 12. Atom Bond Bond distances (Å) BVS Assigned oxidation state Mn(1) Mn(1)-O(19) Mn(1)-O(21) Mn(1)-O(17) Mn(1)-O(9) Mn(1)-O(3) Mn(1)-O(13) 1.947(4) 1.962(4) 1.961(5) 1.963(4) 2.174(4) 2.239(4) 2.939 +3 Mn(2) Mn(2)-O(5) Mn(2)-O(12) Mn(2)-O(21) Mn(2)-O(18) Mn(2)-O(3) Mn(2)-O(14) 1.898(4) 1.918(5) 1.911(4) 1.960(4) 2.273(4) 2.299(4) 3.069 +3 Mn(3) Mn(3)-O(19`) Mn(3)-O(18) Mn(3)-O(6) Mn(3)-O(7`) Mn(3)-O(14) Mn(3)-O(13`) 1.948(4) 1.965(4) 1.964(5) 1.964(4) 2.237(4) 2.276(4) 2.849 +3 Mn(4) Mn(4)-O(18) Mn(4)-O(18`) Mn(4)-O(19) Mn(4)-O(19`) Mn(4)-O(21`) Mn(4)-O(21) 2.008(5) 2.008(5) 2.017(4) 2.017(4) 2.292(5) 2.292(5) 2.494 +3 Mn(5) Mn(5)-O(8`) Mn(5)-O(16) Mn(5)-O(15) Mn(5)-O(2`) Mn(5)-N(1) 1.819(5) 1.890(5) 1.942(5) 1.949(4) 2.303(6) 3.067 +3 Mn(6) Mn(6)-O(11) Mn(6)-O(10) Mn(6)-O(12) Mn(6)-O(17) Mn(6)-O(21) Mn(6)-N(3) 1.975(5) 1.985(6) 2.032(5) 2.023(5) 2.107(5) 2.419(4) 2.685 +3 Mn(7) Mn(7)-O(20`) Mn(7)-O(1) Mn(7)-O(4) Mn(7)-O(16`) Mn(7)-N(2) 1.852(5) 1.873(4) 1.970(6) 1.988(5) 2.253(6) 2.967 +3 Mn(2) ··· Mn(3) Mn(2) ··· Mn(6) Mn(4) ··· Mn(2) Mn(4) ··· Mn(3) 2.9877(16) 2.8701(18) 3.2364(13) 2.9675(11) Mn(1) ··· Mn(2) Mn(1) ··· Mn(3`) Mn(1) ··· Mn(4) Mn(1) ··· Mn(6) 3.0913(14) 3.1575(16) 3.0260(13) 3.1309(15) The angles within each triangular unit in 12 measure 38.9(1)° for Mn(5)-Mn(6)Mn(7`), 68.20(1)° for Mn(5)-Mn(7`)-Mn(6) and 72.90(1)° for Mn(6)-Mn(5)-Mn(7`), while for 11 the corresponding angles are 39.64(5)° for Mn(8)-Mn(4)-Mn(12), 66.73(7)° for 147 Chapter 3 – Polynuclear Manganese Coordination Complexes Mn(4)-Mn(12)-Mn(8) and 73.63(8)° for Mn(4)-Mn(8)-Mn(12). The interatomic distances within the hexagonal brucite plane in 12, Mn(1) ··· Mn(2), Mn(2) ··· Mn(3) and Mn(3) ··· Mn(1`) are 3.091(1) Å, 2.988(5) Å and 3.158(4) Å, respectively, whereas in 11 the corresponding interatomic distances Mn(1) ··· Mn(3), Mn(3) ··· Mn(5) and Mn(5) ··· Mn(7) are 3.023(4) Å, 3.031(4) Å and 3.072(4) Å, respectively. The hexagonal unit in 12 displays angles of 113.83(1)°, 121.19(1)° and 124.98(1)° for Mn(1)-Mn(2)-Mn(3), Mn(2)Mn(1)-Mn(3`) and Mn(1`)-Mn(3)-Mn(2), respectively, whereas the hexagonal unit in 11 displays coressponding angles of 116.16(11)°, 117.22(11)° and 126.07(12)° for Mn(5)Mn(7)-Mn(6), Mn(3)-Mn(5)-Mn(7) and Mn(1)-Mn(3)-Mn(5), respectively. However, despite the slight differences, the bond distances and bond angles within 12 are comparable with those found in 11. Selected bond lengths and bond angles for compound 12 are listed in Table 3.10 and Table 3.11. Table 3.11 − Selected bond angles [º] for compound 12. Bond Angle (º) Bond Angle (º) O(5)-Mn(2)-O(21) O(12)-Mn(2)-O(18) O(3)-Mn(2)-O(14) O(5)-Mn(2)-O(3) O(5)-Mn(2)-O(12) O(12)-Mn(2)-O(3) O(21)-Mn(2)-O(12) O(18)-Mn(2)-O(3) O(12)-Mn(2)-O(14) O(21)-Mn(2)-O(3) 178.5(2) 170.3(2) 168.03(15) 97.89(18) 89.9(2) 93.5(2) 91.6(2) 89.91(17) 88.44(19) 82.27(16) O(18)-Mn(3)-O(7`) O(19`)-Mn(3)-O(6) O(14)-Mn(3)-O(13`) O(6)-Mn(3)-O(13`) O(7`)-Mn(3)-O(14) O(19`)-Mn(3)-O(14) O(6)-Mn(3)-O(18) O(6)-Mn(3)-O(7`) O(18)-Mn(3)-O(14) O(6)-Mn(3)-O(14) 176.10(19) 175.94(18) 169.83(15) 104.22(18) 95.01(18) 93.92(17) 92.93(19) 89.93(19) 87.99(16) 82.72(18) O(18)-Mn(4)-O(18`) O(19)-Mn(4)-O(19`) O(21`)-Mn(4)-O(21) O(18`)-Mn(4)-O(21) O(19`)-Mn(4)-O(21) O(18`)-Mn(4)-O(19`) O(18)-Mn(4)-O(19`) O(19`)-Mn(4)-O(21`) O(19)-Mn(4)-O(21) O(18)-Mn(4)-O(21) 180.000(1) 180.000(1) 180.000(1) 109.66(15) 100.23(16) 98.19(17) 81.81(17) 79.77(16) 79.77(16) 70.34(15) N(3)-Mn(6)-O(21) O(10)-Mn(6)-O(12) O(11)-Mn(6)-O(17) N(3)-Mn(6)-O(17) N(3)-Mn(6)-O(10) O(11)-Mn(6)-O(10) N(3)-Mn(6)-O(12) O(10)-Mn(6)-O(21) N(3)-Mn(6)-O(11) O(10)-Mn(6)-O(17) 172.65(15) 171.8(2) 170.41(19) 102.21(16) 97.8(2) 93.6(3) 90.18(17) 88.9(2) 86.92(19) 88.1(2) The phosphonate ligands within 12 adopt two different bridging modes η1:η1:η1:μ3 and η1:η1:η2:μ4, similar as in 11. The oxidation states of the Mn and O ions in 12 were established by bond valence sum analysis and charge considerations (Table 3.10 and Table 3.12). 148 Chapter 3 – Polynuclear Manganese Coordination Complexes Table 3.12 − Bond valence sum (BVS) calculations for some O atoms in 12. Atom O(21) O(19) O(18) O(16) BVS 1.871 1.677 1.667 1.242 Assignment a μ4-O2μ3-O2μ3-O2μ2-OH- An oxygen BVS in the ∼1.8-2.0, ∼1.0-1.2, and ∼0.2-0.4 ranges is indicative of non-, single- and double protonation, respectively.7, 53 a The packing arrangement of the {Mn13} clusters in 12 displays small channels that extend in the direction of the crystallographic b-axis (Figure 3.32). These channels are filed with constitutional solvent molecules. Figure 3.32 – Packing arrangement of the tridecanuclear manganese clusters in 12 viewed in the direction of the crystallographic a- and b-axis. Colour code: Mn blue, P purple, O red, C grey, (crystallization solvent molecules and hydrogen atoms have been omitted for clarity). 149 Chapter 3 – Polynuclear Manganese Coordination Complexes - PHYSICOCHEMICAL CHARACTERISATION - FT-IR spectroscopy In comparison to 11, the IR spectrum of 12 presented in Figure 3.33 displays some extra bands in the 1500 – 1150 cm-1 and 850 – 690 cm-1 region due to the C–H bending and rocking vibrations of the –CH2 groups of the benzylphosphonic acid ligands. The bands located in the 3040 – 2800 cm-1 region are more intense due to the overlapping C–H stretching vibrations of the aromatic rings and aliphatic –CH2 groups.36-40, 56, 57 Figure 3.33 – Infrared spectrum of 12. 150 Chapter 3 – Polynuclear Manganese Coordination Complexes - Thermogravimetric analysis Thermogravimetric analysis was performed using a freshly prepared crystalline sample of 12, in the temperature range 30 – 900 °C, in an N2 atmosphere. The compound reveals a similar decomposition behaviour as 11. The first thermogravimetric step of about 6.5 % occurs in a temperature range between 30 – 195 °C as a result of the proposed loss of five crystallization H2O molecules and four coordinated CH3OH molecules (calcd: 6.7 %). This constitutional assignemnet is in agreement with the elemental analysis of 12. The next step in the TGA curve corresponds to a weight loss of 14.3 % between 195 – 270 °C, which is associated with the loss of six pyridine molecules (calcd: 14.5 %). Further cluster degradation and decomposition of the organic ligands can be observed above 270 °C. Weight % (%) 100 90 80 70 60 50 0 200 400 600 800 1000 Temperature (ºC) Figure 3.34 – Thermogravimetric analysis of 12. 151 Chapter 3 – Polynuclear Manganese Coordination Complexes - Magnetism The temperature dependence of the magnetic susceptibility of 12 was measured between 298 and 1.8 K (Figure 3.35). At room temperature, the experimental χT product reaches 38.1 cm3 K mol-1. That is approximately in agreement with the expected value of 39 cm3 K mol-1 for the presence of thirteen S = 2 MnIII metal ions (C = 3 cm3 K mol-1 with g = 2). Upon lowering the temperature, the χT product decreases down to 34.7 at 100 K, then increases to 35.1 cm3 K mol-1 at 50 K and then finally decreases down to 32.6 cm3 K mol-1 at 1.8 K. This thermal behavior suggests the presence of dominant antiferromagnetic (AF) interactions between the spin carriers. This result does not exclude the presence of ferromagnetic interactions but only suggests that the AF exchange pathways are dominant. It is worth mentioning that this magnetic behavior is significantly different from that observed for 11 showing that the intra-complex magnetic interactions must be different. Despite the fact that the spin ground state is very difficult to establish only based on the χT vs T data. The value of about 35 cm3 K mol-1 at low temperature suggests a spin ground state of 8. a) b) Figure 3.35 – (a) Temperature dependence of the χT product of 12 at 0.1 T. (b) A magnified view of the χT product in (a) between 1 and 100 K. The field dependence of magnetisation for this compound has been measured at low temperatures between 1.8 and 8 K (Figure 3.36, a). The magnetisation at low field displays a rapid increase without inflexion point confirming the absence of weak antiferromagnetic interactions. The high field behavior displays a non-linear and nonmonotone increase without clear saturation even at 1.8 K at 7 T, suggesting the presence of magnetic anisotropy and also low lying excited states. The data shown as a M vs H/T plot (Figure 3.36, b) confirms the presence of both magnetic anisotropy intrinsic to MnIII metal ions and low-lying excited states as the data are not superposed on a single master-curve as 152 Chapter 3 – Polynuclear Manganese Coordination Complexes expected for an isotropic system with a well defined spin ground state. At 1.8 K, the magnetisation reaches 18.9 B at 7 T which is above the 16 B expected for an ST = 8 ground state confirming the field induce population of low lying excited states. a) b) Figure 3.36 – (a) M vs H and (b) M vs H/T data at and below 8 K. Both measurements, the field dependence of the magnetisation at 1.8 K and the temperature dependence of the χT product at 1000 Oe, are coherent with an ST = 8 ground state even if this conclusion is not as clear as for 11. The M vs H data at 1.8 K do not show any sign of slow relaxation i.e. hysteresis effects. Nevertheless, the ac susceptibility in zero dc field has been measured to probe possible slow dynamics of the magnetisation in this compound. Clearly at temperatures below 4 K (for frequency around 10000 Hz), slow relaxation of the magnetisation is observed in agreement with the appearance of an out-of-phase signal. a) b) Figure 3.37 – Temperature dependence of the (a) in-phase (χ') and (b) out-of-phase (χ") ac susceptibility under zero dc field for 12. 153 Chapter 3 – Polynuclear Manganese Coordination Complexes Cosecutively, we measured the ac susceptibility as a function of the frequency at different temperatures in zero dc field in order to estimate the relaxation time. The observed frequency dependence and shape of the curve idicates that this compound is consistent with an SMM.7 a) b) Figure 3.38 – Frequency dependence of the (a) in-phase (χ') and (b) out-of-phase (χ") ac susceptibility from 1.8 to 2.8 K under zero dc field for 12. Using a scaling technique, the relaxation time of the compound was determined up to 2.8 K. a) b) Figure 3.39 – (a) χ"/χ' vs from 1.8 to 2.8 K under zero dc field; (b) Magnetisation relaxation time ( ) vs T-1plot for 12 under zero dc field (the solid line correspond to the Arrhenius law). From these data, the relaxation time can be deduced between 1.9 and 2.8 K and fitted to an Arrhenius law. The exponential increase of the relaxation allowed us to determine the energy barrier of 20.8 K of the thermally activated regime while 0 is about 8.5 × 10-10 s. 154 Chapter 3 – Polynuclear Manganese Coordination Complexes The ac susceptibility has been measured in dc fields at 1.8 K in order to investigate possible quantum relaxation pathways in zero-field and to probe the quantum contribution to the observed relaxation above 1.8 K. a) b) Figure 3.40 – Frequency dependence of the (a) in-phase (χ') and (b) out-of-phase (χ") ac susceptibility with dc fields at 1.8 K for 12. Figure 3.41 – vs H plot at 1.8 K. The shift of the characteristic frequency (with an optimum field of 300 Oe) is very small highlighting that the quantum pathway only provides a minor contribution to the relaxation of the magnetisation at 1.8 K. Therefore, the ac susceptibility measurements at 300 Oe have not been performed as the resulting data would be extremely similar to that obtained at zero-dc field. 155 Chapter 3 – Polynuclear Manganese Coordination Complexes - Mass spectrometry The stability of the {Mn13} cluster in solution was investigated by electrospray ionisation mass spectrometry. Negative mode ESI-MS spectra of 12 in CH3CN, DMF (Figure 3.42) and DMSO reveal the presence of an isotopic envelope centred at m/z = 2699.1 {H[ a.m.u. nIV3MnIII10 (Table 10(CH3 3.13). )4(C6 This 5CH2P signal 3)10]} - can be attributed species, confirming that to the the cluster is stable in CH3CN, DMF or DMSO environments. Table 3.13 – ESI-MS assignment for compound 12. Crystals of m/z Species attributed CH3CN 2699.1 {H[MnIV3MnIII10 10(C 3 )4(C6 5C 2P 3)10]} → {Mn13} DMF 2699.1 {H[MnIV3MnIII10 10(C 3 )4(C6 5C 2P 3)10]} → {Mn13} DMSO 2699.1 {H[MnIV3MnIII10 10(C 3 )4(C6 5C 2P 3)10]} → {Mn13} Relative ionic abundance 12 in: Solvent Figure 3.42 – Negative-mode ESI-MS spectra for crystals of 12 dissolved in DMF. Inset: Comparison of the experimental isotopic envelopes (black spectrum) with simulated patterns (red spectrum) for {H[ nIV3MnIII10 10(CH3 )4(C6 5CH2P 3)10]}- centered at m/z = 2699.1 a.m.u. (cone voltage: 30 V). 156 Chapter 3 – Polynuclear Manganese Coordination Complexes - UV-Vis spectroscopy The UV-Vis absorption spectrum of 12 recorded in CH3CN (Figure 3.43) displays a shoulder at ca. 257 nm ( = 61000 L mol-1 cm-1) attributed to π − π* transitions of the organic ligands. A very weak absorption band is also displayed at ca. 443 nm ( = 4500 L mol-1 cm-1) which corresponds to a 5T2g III coordinated Mn ions. 5 Eg transition within the octahedrally 49-52 Absorbance (a.u.) 0.8 0.7 0.5 0.6 0.4 0.5 0.3 0.2 0.4 0.1 0.3 400 450 500 550 600 0.2 0.1 0.0 300 400 500 600 700 800 Wavelength (nm) Figure 3.43 – UV-Vis spectrum of a 10-5 M solution of 12 in CH3CN. Inset: A section of the UV-Vis spectrum of a 10-4 M solution of 12 in CH3CN. 157 Chapter 3 – Polynuclear Manganese Coordination Complexes Table 3.14 − Crystal data and structural refinement parameters for 12. Compound 12 Empirical formula a Molecular mass/g mol C104H114ClMn13N6O42P10 -1 a Crystal colour/shape 3 Brown / rectangular block Crystal size/mm 0.50×0.15×0.10 Crystal system Triclinic Space group P a/ Å 14.990(3) b/ Å 15.833(3) c/ Å 16.410(3) /º 73.19(3) /º 83.00(3) /º 64.78(3) 3 V/ Å 3372.9(11) Z 1 Temperature (K) 150(2) -3 Density/Mg m 1.530 -1 Absorp. coef./mm 1.366 F(000) 1555 2 60 max/º Reflections collected 44311 Independent reflections 15941 [R(int)=0.0527] Data / restraints / parameters 15941 / 6 / 751 2 a 3179.41 S on F 1.035 R1, wR2 [I>2 (I)] 0.0814, 0.2364 R1, wR2 (all data) 0.1200, 0.2619 Largest diff. peak and hole/e.Å-3 3.320 and -0.925 Excluding solvate molecules 158 Chapter 3 – Polynuclear Manganese Coordination Complexes 3.3.2.3 [Μn ΙΙΙ 13 (μ4-Ο)2(μ3-Ο)4(μ2-ΟH)2(μ2-CΗ3Ο)4(C6Η5CΗ2PΟ3)10(C6Η5-C3Η6-C5Η4Ν)6]Cl ·5H2O (13) Compound 13, [ n C3 6-C5 4 13( 4- )2( 3- )4( 2- H)2( 2-C 3 )4(C6 5C 2P 3)10(C6 5- )6]Cl·5H2O was the result of replacing the pyridine ligands in the reaction mixture of 12 with 4-(3-phenylpropyl)pyridine ligands. The crystallisation process that occurs over a week affords rectangular brown crystals of 13 which were characterized by single crystal X-ray diffraction measurements. 13 crystallises in the triclinic crystal system in the space group P and is structurally related to 11 and 12. The metal-centered, distorted cuboctahedral arrangement of the {Mn13} cluster in 13 is very similar to the highly symmetric arrangement in 11 (Figure 3.44). The bond distances and bond angles within 13 are comparable with those found in 11 and 12 as can be seen in Table 3.15 and Table 3.16. (A) (B) (A) Figure 3.44 − Crystal structure of the tridecanuclear manganese complex in 13. Colour code: MnIII blue, P purple, O red, C grey (hydrogen atoms have been omitted for clarity). 159 Chapter 3 – Polynuclear Manganese Coordination Complexes Table 3.15 − Selected bond lengths [Å] and bond valence sum (BVS) for compound 13. Atom Bond Bond distances (Å) BVS Assigned oxidation state Mn(1) Mn(1)-O(17) Mn(1)-O(17`) Mn(1)-O(16) Mn(1)-O(16`) Mn(1)-O(10) Mn(1)-O(10`) 1.947(4) 1.947(4) 1.954(5) 1.954(5) 2.404(5) 2.404(5) 2.738 +3 Mn(2) Mn(2)-O(21) Mn(2)-O(10`) Mn(2)-O(19) Mn(2)-O(17) Mn(2)-O(8`) Mn(2)-O(20) 1.909(5) 1.926(5) 1.926(5) 1.944(5) 2.185(5) 2.220(5) 3.156 +3 Mn(3) Mn(3)-O(11) Mn(3)-O(18) Mn(3)-O(16) Mn(3)-O(17) Mn(3)-O(3) Mn(3)-O(20) 1.925(5) 1.926(5) 1.929(5) 1.929(5) 2.228(5) 2.228(5) 3.105 +3 Mn(4) Mn(4)-O(9) Mn(4)-O(5) Mn(4)-O(16) Mn(4)-O(10) Mn(4)-O(8) Mn(4)-O(3) 1.878(6) 1.915(5) 1.946(5) 1.957(5) 2.195(5) 2.229(5) 3.162 +3 Mn(5) Mn(5)-O(10) Mn(5)-O(2) Mn(5)-O(19) Mn(5)-N(1) Mn(5)-O(1) Mn(5)-O(9) 1.902(5) 1.958(5) 1.992(5) 2.033(6) 2.079(6) 2.132(6) 3.215 +3 Mn(6) Mn(6)-O(23) Mn(6)-O(7) Mn(6)-O(6) Mn(6)-O(12) Mn(6)-N(3) 1.873(6) 1.887(5) 1.898(6) 1.928(5) 2.229(7) 3.138 +3 Mn(7) Mn(7)-O(22) Mn(7)-O(4) Mn(7)-O(14) Mn(7)-O(12) Mn(7)-N(2) 1.871(6) 1.871(6) 1.891(5) 1.908(6) 2.216(8) 3.234 +3 Mn(1)-Mn(3) Mn(1)-Mn(2) Mn(1)-Mn(4) Mn(2)-Mn(5`) 2.9125(11) 3.1145(11) 3.1551(11) 2.8865(16) Mn(2)-Mn(3) Mn(2)-Mn(4`) Mn(3)-Mn(4) Mn(4)-Mn(5) 3.0407(15) 3.1115(15) 3.0415(15) 2.9452(17) 160 Chapter 3 – Polynuclear Manganese Coordination Complexes Table 3.16 − Selected bond angles [º] for compound 13. Bond Angle (º) Bond Angle (º) O(17`)-Mn(1)-O(17) O(16`)-Mn(1)-O(16) O(10`)-Mn(1)-O(10) O(16`)-Mn(1)-O(10) O(17)-Mn(1)-O(10) O(17)-Mn(1)-O(16`) O(17)-Mn(1)-O(16) O(17)-Mn(1)-O(10`) O(17`)-Mn(1)-O(10) O(16)-Mn(1)-O(10) 180.000(1) 180.000(1) 180.000(1) 105.82(18) 105.04(17) 97.87(19) 82.13(19) 74.96(17) 74.95(17) 74.18(18) O(21)-Mn(2)-O(10`) O(19)-Mn(2)-O(17) O(8`)-Mn(2)-O(20) O(21)-Mn(2)-O(8`) O(21)-Mn(2)-O(19) O(21)-Mn(2)-O(20) O(21)-Mn(2)-O(17) O(19)-Mn(2)-O(20) O(19)-Mn(2)-O(8`) O(10`)-Mn(2)-O(19) 179.9(2) 169.5(2) 167.82(19) 98.1(2) 96.2(2) 93.6(2) 92.8(2) 91.3(2) 90.6(2) 83.7(2) O(11)-Mn(3)-O(17) O(18)-Mn(3)-O(16) O(3)-Mn(3)-O(20) O(18)-Mn(3)-O(3) O(18)-Mn(3)-O(17) O(17)-Mn(3)-O(3) O(11)-Mn(3)-O(3) O(16)-Mn(3)-O(17) O(17)-Mn(3)-O(20) O(16)-Mn(3)-O(3) 177.7(2) 177.5(2) 170.45(18) 98.5(2) 94.5(2) 89.8(2) 89.7(2) 83.3(2) 82.87(19) 82.59(19) O(10)-Mn(5)-N(1) O(2)-Mn(5)-O(19) O(1)-Mn(5)-O(9) N(1)-Mn(5)-O(9) O(10)-Mn(5)-O(1) O(19)-Mn(5)-N(1) O(2)-Mn(5)-O(1) O(2)-Mn(5)-N(1) O(19)-Mn(5)-O(1) N(1)-Mn(5)-O(1) 175.7(2) 175.2(2) 174.9(2) 97.4(2) 95.0(2) 94.2(2) 94.5(2) 90.3(2) 87.1(2) 87.6(2) A close inspection of the three complexes 11, 12 and 13 reveals only a few structural differences, the most important ones being the peripheral ligation and the symmetry of the core structure. The structural parameters within 13 are closer to that of 11 than 12. In accordance to this observation the triangular units in 13 display lateral Mn-Mn distances of 3.616(3) Å for Mn(6) ··· Mn(7), 5.387(3) Å for Mn(5) ··· Mn(6) and 5.408(4) Å for Mn(5) ··· Mn(7), and angles of 39.15(2)° for Mn(6)-Mn(5)-Mn(7), 70.11(3)° for Mn(5)-Mn(7)-Mn(6) and 70.75(3)° for Mn(5)-Mn(6)-Mn(7). Further, the interatomic distances in the hexagonal unit in 13 are 3.041(3) Å, 3.042(3) Å and 3.111(3) Å for Mn(2) ··· Mn(3), Mn(3) ··· Mn(4) and Mn(4) ··· Mn(2), respectively. Also, the angles within the hexagonal unit in 13 are closer to those in 11. The Mn(2)-Mn(4)-Mn(3), Mn(3)-Mn(2)Mn(4), and Mn(2)-Mn(3)-Mn(4) angles measure 115.63(5)°, 117.36(5)° and 127.00(5)°, respectively. 161 Chapter 3 – Polynuclear Manganese Coordination Complexes (A) Figure 3.45 − Polyhedral representation of the triangular unit A in 13. Colour code: MnIII blue, P purple, O red, C grey. (B) Figure 3.46 – Ball-and-stick representation of the hexagonal unit B in 13. Colour code: MnIII blue, P purple, O red, C grey. 162 Chapter 3 – Polynuclear Manganese Coordination Complexes All phosphonate ligands within 13 bridge between three MnIII ions in a η1:η1:η1:μ3 and η1:η1:η2:μ4 bridging modes. The oxidation states of the Mn and O ions in 13 were established by bond valence sum analysis and charge considerations (Table 3.15 and Table 3.17). Table 3.17 − Bond valence sum (BVS) calculations for some O atoms in 13. Atom O(10) O(16) O(17) O(12) BVS 2.079 1.827 1.842 1.270 Assignment a μ4-O2μ3-O2μ3-O2μ2-OH- An oxygen BVS in the ∼1.8-2.0, ∼1.0-1.2, and ∼0.2-0.4 ranges is indicative of non-, single- and double protonation, respectively.7, 53 a In the solid state the tridecanuclear manganese clusters are linked by weak hydrogen bonds between the Mn clusters and constitutional solvent molecules resulting in a lamellar arragement with alternating organic hydrophobic and polar inorganic areas (Figure 3.47). In addition dispersion forces between the peripheral, hydrophobic 4-(3phenylpropyl)pyridine ligands stabilise the observed packing arrangement. The coordination clusters arrange in the ac-plane and the latter forces involving the hydrophobic moieties stabilise the packing in [010]. Figure 3.47 – Packing arrangement of the tridecanuclear manganese clusters in 13 viewed in the direction of the crystallographic c-axis. Colour code: Mn blue, P purple, O red, C grey, (crystallization solvent molecules and hydrogen atoms have been omitted for clarity). 163 Chapter 3 – Polynuclear Manganese Coordination Complexes - PHYSICOCHEMICAL CHARACTERISATION - FT-IR spectroscopy The IR spectrum of 13 (Figure 3.48) is comparable with that of 12. Some dissimilarities can be observed due to the contribution of the C–H vibrations of the –CH2 groups of the 4-(3-phenylpropyl)pyridine ligands. This has an effect on the C–H vibration bands which appear in the spectrum as more intense and better resolved bands compared with those of 12.36-40, 56, 57 Figure 3.48 – Infrared spectrum of 13. 164 Chapter 3 – Polynuclear Manganese Coordination Complexes - Thermogravimetric analysis 13 displays a similar thermal decomposition behaviour in N2 as 11 and 12 (Figure 3.49). An initial weight loss of 5.6 % occurs between 30 – 150 °C as a possible result of the loss of five crystallization H2O molecules and four coordinated CH3OH molecules (calcd: 5.3 %). This assignment is also in agreement with the elemental analysis of the 13. The following step in the TGA curve corresponds to a weight loss of 27.3 % between 150 – 550 °C, which is associated with the loss of six 4-(3-phenylpropyl)pyridine molecules (calcd: 29.7 %). Further degradation of the organic ligands occours above 550 °C. Weight % (%) 100 90 80 70 60 0 200 400 600 800 1000 Temperature (ºC) Figure 3.49 – Thermogravimetric analysis of 13. 165 Chapter 3 – Polynuclear Manganese Coordination Complexes - Magnetism Compound 13 shows similar magnetic behaviour as 11. The room temperature χT value of 30.3 cm3 K mol-1 (Figure 3.50) is significantly lower than the expected value of 39 cm3 K mol-1 for the presence of thirteen S = 2 MnIII carriers (C = 3 cm3 K mol-1 with g = 2). This result indicates dominant antiferromagnetic (AF) exchange couplings between spins carriers. Upon lowering the temperature, the χT product decreases down to a minimum of ca. 16.5 cm3 K mol-1 at about 20 K. The experimental data can be fitted to a Curie-Weiss law for temperatures above 70 K, with C = 39.3(9) cm3 K mol-1 and =- 84(5) K. The Curie constant is in good agreement with the expected value (39 cm3 K/mol) agreeing well with the presence of the thirteen MnIII S = 2 spin centres. Below 20 K, at 1000 Oe the χT product increases to reach 17.8 cm3 K mol-1 at 1.8 K. This value suggests an ST = 6 ground state as for 11. At 1 T, below 10 K, the χT product decreases to reach 7.2 cm3 K mol-1 at 1.8 K due to field saturation effects. a) b) Figure 3.50 – (a) Temperature dependence of the χT product of 13 at 0.1 and 1 T. (b) A magnified view of the χT product in (a) between 1 and 100 K. The green solid line corresponds to the best fit of the experimental data with the Curie-Weiss law (C = 39.3(9) cm3 K mol-1 and = - 84(5) K). The field dependence of magnetisation for this compound has also been measured below 8 K (Figure 3.51, a). The magnetisation at low field displays a rapid increase without inflexion point confirming the absence of weak antiferromagnetic interactions and the presence of a well-defined ground state. The high field behavior that displays no clear saturation even at 1.8 K at 7 T, suggests the presence of magnetic anisotropy. It is worth noting that the presence of low-lying excited states and inter-complex magnetic interactions could contribute to this M vs H data even if these two effects have not been 166 Chapter 3 – Polynuclear Manganese Coordination Complexes observed for χT vs T data at 0.1 T. The data presented as a M vs H/T plot (Figure 3.51, b) confirms the presence of magnetic anisotropy intrinsic to MnIII metal ions and the cluster as the data are not superposed on a single master-curve as expected for an isotropic system with a well defined spin ground state. At 1.8 K, the magnetisation reaches 11.3 B at 7 T in agreement with an ST = 6 ground state. a) b) Figure 3.51 – (a) M vs H and (b) M vs H/T data at and below 8 K. Both measurements, the field dependence of the magnetisation at 1.8 K and the temperature dependence of the χT product at 1000 Oe, are coherent with a ground state ST = 6. This experimental value can be explained by a configuration in which the spin vectors of eight MnIII centres are oriented in opposite direction than those of the five remaing centre, due to probable competing interactions between spin carriers. The M vs H data at 1.8 K do not show any sign of slow relaxation i.e. hysteresis effects. Nevertheless as for 11 and 12, the ac susceptibility in zero dc field has been measured to probe possible slow dynamics of the magnetisation. Clearly below 4 K (for frequency around 10000 Hz), slow relaxation of the magnetisation is observed in agreement with the appearance of an out-of-phase signal. 167 Chapter 3 – Polynuclear Manganese Coordination Complexes 10 Hz 30 Hz 60 Hz 100 Hz 150 Hz a) 200 300 400 600 800 Hz Hz Hz Hz Hz 1000 1200 1500 2000 3000 Hz Hz Hz Hz Hz 4000 Hz 5000 Hz 6000 Hz 8000 Hz 10000 Hz 10 Hz 30 Hz 60 Hz 100 Hz 150 Hz b) 1000 1200 1500 2000 3000 Hz Hz Hz Hz Hz 4000 Hz 5000 Hz 6000 Hz 8000 Hz 10000 Hz Hdc = 0 Oe Hdc = 0 Oe 10 2 χ" / cm3 mol-1 χ' / cm3 mol-1 Hz Hz Hz Hz Hz 2.5 12 8 6 4 1.5 1 0.5 2 0 200 300 400 600 800 1 2 3 4 5 6 7 T/K 8 0 1 1.5 2 2.5 3 3.5 4 T/K Figure 3.52 – Temperature dependence of the (a) in-phase (χ') and (b) out-of-phase (χ") ac susceptibility under zero dc field for 13. Frequency-dependent out-of-phase (χ") ac susceptibility signals were also observed at temperatures below 3.2 K, suggesting that 13 is consistent with an SMM.7 For this compound the relaxation time can be deduced between 1.9 and 3.2 K and fitted to an Arrhenius law. The fit of the thermally activated region gives the energy barrier of 17.8 K while 0 is about 2.3 × 10-9 s. 168 Chapter 3 – Polynuclear Manganese Coordination Complexes - Mass spectrometry The {Mn13} cluster in 13 was also found to be stable in CH3CN, DMF and DMSO environments. The {H[ nIV3MnIII10 10(CH3 )4(C6 5CH2P 3)10]} - species could be identified in the ESI-MS spectrum of pristine crystals dissolved in CH3CN, DMF or DMSO (Table 3.18, Figure 3.53) as a signal centered at m/z = 2699.1 a.m.u. Furthermore, the ESI-MS spectrum of 13 dissolved in CH3CN reveals the presence of a signal centered at (C6 m/z = 5CH2P 2753.4 3)10]} a.m.u. that was assigned to {H7[ nIV3MnIII10 13(CH3 )4 - . This latter species is related to the former identified species but has three additional H2O molecules associated with the coordination cluster. Both species were modelled and good fits between the experimental and simulated isotopic envelopes could be found. Table 3.18 – ESI-MS assignment for compound 13. Solvent Crystals of Species attributed 2699.1 {H[MnIV3MnIII10 IV 10(C 3 )4(C6 5C 2P 3)10]} → {Mn13} - 2753.4 {H7[Mn III 3Mn 10 13(C 3 )4(C6 5C 2P 3)10]} → {Mn13} DMF 2699.1 {H[MnIV3MnIII10 10(C 3 )4(C6 5C 2P 3)10]} → {Mn13} DMSO 2699.1 {H[MnIV3MnIII10 10(C 3 )4(C6 5C 2P 3)10]} → {Mn13} Relative ionic abundance 13 in: CH3CN m/z Figure 3.53 – Negative-mode ESI-MS spectra for crystals of 13 dissolved in DMF. Inset: Comparison of the experimental isotopic envelopes (black spectrum) with simulated patterns (red spectrum) for {H[ nIV3MnIII10 10(CH3 )4(C6 5CH2P 3)10]}- centered at m/z = 2699.1 a.m.u. (cone voltage: 30 V). 169 Chapter 3 – Polynuclear Manganese Coordination Complexes - UV-Vis spectroscopy Similar to the previous three compounds the UV-Vis absorption spectrum of 13 recorded in CH3CN (Figure 3.54) displays two absorption bands, an intense band at ca. 256 nm ( = 41000 L mol-1 cm-1) and a weak band at ca. 446 nm ( = 3600 L mol-1 cm-1). These corresponds to a π − π* transition of the organic ligands and a 5T2g III within the octahedrally coordinated Mn ions. 5 Eg transition 49-52 Absorbance (a.u.) 0.45 0.40 0.4 0.35 0.3 0.30 0.2 0.25 0.1 0.20 400 450 500 550 600 0.15 0.10 0.05 0.00 300 400 500 600 700 800 Wavelength (nm) Figure 3.54 – UV-Vis spectrum of a 10-5 M solution of 13 in CH3CN. Inset: A section of the UV-Vis spectrum of a 10-4 M solution of 13 in CH3CN. 170 Chapter 3 – Polynuclear Manganese Coordination Complexes Table 3.19 − Crystal data and structural refinement parameters for 13. Compound 13 Empirical formula a Molecular mass/g mol C158H174ClMn13N6O42P10 -1 a Crystal colour/shape 3 Brown / rectangular block Crystal size/mm 0.11×0.06×0.05 Crystal system Triclinic Space group P a/ Å 14.8581(5) b/ Å 16.0216(6) c/ Å 19.3531(7) /º 102.721(2) /º 97.439(2) /º 90.847(2) 3 V/ Å 4451.6(3) Z 1 Temperature (K) 100(2) -3 Density/Mg m 1.498 -1 Absorp. coef./mm 8.863 F(000) 2064 2 60 max/º Reflections collected 25996 Independent reflections 12843 [R(int)=0.0571] Data / restraints / parameters 12843 / 2151 / 1012 2 a 3888.47 S on F 0.981 R1, wR2 [I>2 (I)] 0.0960, 0.2609 R1, wR2 (all data) 0.1120, 0.2775 Largest diff. peak and hole/e.Å-3 2.005 and -1.021 Excluding solvate molecules 171 Chapter 3 – Polynuclear Manganese Coordination Complexes 3.3.3 Synthesis and characterisation of dodecanuclear manganese complexes 3.3.3.1 K(H2O)4[MnIII12(μ3-O)6(CH3OH)6(C6H5CH2PO3H)7(C6H5CH2PO3)8]Cl2·3CH3OH ·4H2O (14) The labile nature of the pyridine ligands in Jahn-Teller sites in 12 allowed the replacement of these ligands with 4-(3-phenylpropyl)pyridine in 13. This observation prompted us to investigate if we could use a bipyridine ligand to connect the polynuclear manganese complexes into polymers. Therefore, 4,4'-trimethylenedipyridine and 4,4'bipyridine were used as organic bases in similar reaction mixtures that led to the formation of the tridecanuclear manganese complexes. Unfortunately, no crystalline material was obtained when 4,4'-trimethylenedipyridine was used. However, when 4,4'-bipyridine was employed, rectangular brown crystals of K(H2O)4[MnIII12( 3-O)6(CH3OH)6 (C6H5CH2PO3H)7(C6H5CH2PO3)8]Cl2·3CH3OH·4H2O (14) separated from the reaction mixture within a time period of a week. Single crystal X-ray diffraction measurements revealed that 14 contains a dodecanuclear manganese complex. 14 crystallises in the orthorhombic crystal system in the space group Pbca. The exceptional ring-shaped cluster core in 14 possesses some structural features that resemble those of the complexes in 11-13. The cluster core can be visualised as constructed of two triangular units A and three dinuclear units B (Figure 3.55). However, in this case the triangular units A are structurally different to those in 1113. They adopt an eclipsed arrangement, are directly linked through O-donar atoms, and are further stabilised by three dimer moijeties B. The two triangular units A in 14 each consist of three MnIII ions (Mn(2), Mn(3) and Mn(5), and Mn(1), Mn(4) and Mn(6), respectively) connected through three deprotonated, bridging phosphonate ligands to form a ring motif (Figure 3.56). Each Mn ion in the each triangular unit adopts a square pyramidal geometry that shares a common edge with its partner from the second triangular unit to generate a {Mn2O8}10− subunit (Figure 3.56). The coordination environments of the MnIII ions in A comprise of two bridging 3- 2- oxo ligands involved in the {Mn2O8}10− subunits and three phosphonate O donors that derive from three deprotonated organophosphonate ligands. The short Mn ··· Mn contact in the {Mn2O8}10− subunit (2.8043(14) Å – 2.8281(14) Å) is a result of the short Mn – ( 3- 2- ) distances (1.871(4) Å 172 Chapter 3 – Polynuclear Manganese Coordination Complexes – 1.891(4) Å). The other three Mn–O bond lengths involving the phosphonate ligands vary between 1.873(4) Å – 2.114(4) Å (Table 3.20). (A) (A) (B) (A) Figure 3.55 − Crystal structure of the dodecanuclear manganese complex in 14 showing the triangular units A and the dinuclear units B. Colour code: MnIII blue, P purple, O red, C grey, Cl green, K yellow (hydrogen atoms have been omitted for clarity). 173 Chapter 3 – Polynuclear Manganese Coordination Complexes (A) Figure 3.56 − Polyhedral representation of the triangular units A in 14. Colour code: MnIII blue, P purple, O red, C grey, Cl green. The remaining six MnIII ions in 14 (Mn(7) – Mn(12)) are organised into three {Mn2O10}14− dimer units (B) (Figure 3.57). In these dinuclear units the Mn ions display slightly distorted octahedral coordination environments. The polyhedra in the dimeric subunit share a common edge that involves two phosphonate O donors. The remaining coordination sites of the polyhedra are occupied by two other phosphonate O donors (Mn – Ophosphonate distances 1.885(4) Å – 2.374(4) Å), one terminal methanol O-donor (Mn – Omethanol distances 2.185(4) Å – 2.220(5) Å) and a 1.857(4) Å – 1.870(4) Å). The 3- 2- 3- 2- oxo ligand (Mn – ( 3- ) distances oxo ligands bridge between the three {Mn2O10}14− 174 Chapter 3 – Polynuclear Manganese Coordination Complexes dimer units B and the triangular units A to give the {Mn12} core structure. The distorted nature of the octahedral coordination environment of the Mn ions in B arises from the JT axial elongation and sterical restraints. The distorsion is reflected in various bond lenths and angles that deviate from the ideal octahedral geometrical parameters. Thus, the atoms in the elongated JT axes show bond lengths to the Mn centres that range between 2.185(4) Å – 2.374(4) Å and employ one terminal methanol O-donor and one phosphonate O donor. The bond angle O(38)-Mn(11)-O(41) of 165.74(16)° shows the greatest deviation from the ideal octahedral angle of 180°, while the bond angle O(43)-Mn(9)-O(49) of 96.44(18)° shows the greatest deviation from the ideal octahedral angle of 90°. Selected bond angles for the Mn ions in B are listed in Table 3.21. The cluster is stabilized by altogether fifteen phosphonate ligands that bridge between the A and B units. Nine ligands are situated on the periphery of the {Mn12} ring, with six of them bridging between an octahedron and a square pyramidal polyhedron in a η1:η1:μ2 mode (P(8), P(9), P(10), P(12), P(14), P(15)). The other three perifical organic ligands bridge between two octahedra: one phosphonate moijety adopts a η1:η1:μ2 mode (P(11)) and the two other phosphonate groups act as η1:η1:η1:μ3 ligands, (P(6), P(7)). The remaining six stabilising organophosphonate ligands can be visualised as pairs of three, situated on opposite sides of the ring moijety. Each of these ligands bridge between two square pyramidal polyhedra and a common vertex of two octahedra in a η1:η1:η1:μ3 mode (P(1) – P(5) and P(13)) (Figure 3.58). The {Mn12} cluster in 14 is further stabilised by intramolecular hydrogen bonding between the protonated phosphonate ligands, giving rise to O-O distances of 2.447(0) Å – 2.564(0) Å (Figure 3.58). All Mn atoms in 14 adopt the oxidation state +III. The assignment of the oxidation states was confirmed by BVS calculations (Table 3.20). The cluster is surrounded by two Cl− counterions, one situated in the central cavity of the ring. The ion is located 2.835(2) Å, 2.8124(19) Å and 2.830(2) Å from Mn(2), Mn(3) and Mn(5), respectively. The other Cl− counterion, Cl(2), resides in the vicinity of the cluster core, being situated at 3.612(0) Å from the Cl(1) atom of a neighbouring cluster. The charge of the cluster and the Cl- ions is further compensated by partially hydrated potassium ions {K(H2O)4}+ that link the cluster entities into 1D zigzag chains. The resulting packing arrangement with views in the direction of the crystallographic a- and c- axes is shown in Figure 3.59. Intermolecular hydrogen bonding between the crystallisation solvent molecules and the core structure within 14, together with weak - interactions (shortest contact 3.6 Å) 175 Chapter 3 – Polynuclear Manganese Coordination Complexes invoving the aromatic ring moieties of phosphonate ligands further contribute to the stability of 14 in the solid state. Table 3.20 − Selected bond lengths [Å] and bond valence sum (BVS) for compound 14. Atom Bond Bond distances (Å) BVS Assigned oxidation state Mn(1) Mn(1)-O(37) Mn(1)-O(26) Mn(1)-O(3) Mn(1)-O(12) Mn(1)-O(32) 1.873(4) 1.883(4) 1.887(4) 1.916(4) 2.074(4) 3.244 +3 Mn(2) Mn(2)-O(3) Mn(2)-O(26) Mn(2)-O(18) Mn(2)-O(24) Mn(2)-O(50) 1.879(4) 1.879(4) 1.919(4) 1.936(4) 2.103(4) 3.114 +3 Mn(3) Mn(3)-O(39) Mn(3)-O(6) Mn(3)-O(22) Mn(3)-O(19) Mn(3)-O(42) 1.871(4) 1.885(4) 1.921(4) 1.931(4) 2.101(4) 3.125 +3 Mn(4) Mn(4)-O(6) Mn(4)-O(39) Mn(4)-O(27) Mn(4)-O(46) Mn(4)-O(14) 1.875(4) 1.888(4) 1.900(4) 1.919(4) 2.060(4) 3.044 +3 Mn(5) Mn(5)-O(61) Mn(5)-O(17) Mn(5)-O(25) Mn(5)-O(35) Mn(5)-O(21) 1.882(4) 1.891(4) 1.915(4) 1.921(4) 2.114(4) 3.106 +3 Mn(6) Mn(6)-O(17) Mn(6)-O(61) Mn(6)-O(40) Mn(6)-O(53) Mn(6)-O(48) 1.877(4) 1.884(4) 1.891(4) 1.922(4) 2.078(4) 3.434 +3 Mn(7) Mn(7)-O(17) Mn(7)-O(51) Mn(7)-O(28) Mn(7)-O(7) Mn(7)-O(45) Mn(7)-O(15) 1.858(4) 1.895(4) 1.903(4) 1.963(4) 2.211(5) 2.324(4) 2.886 +3 Mn(8) Mn(8)-O(3) Mn(8)-O(60) Mn(8)-O(31) Mn(8)-O(15) Mn(8)-O(30) Mn(8)-O(7) 1.862(4) 1.893(4) 1.921(4) 1.986(4) 2.186(4) 2.293(4) 2.670 +3 Mn(9) Mn(9)-O(26) Mn(9)-O(52) Mn(9)-O(43) Mn(9)-O(13) Mn(9)-O(49) Mn(9)-O(20) 1.858(4) 1.894(4) 1.899(4) 1.975(4) 2.202(5) 2.313(4) 3.232 +3 176 Chapter 3 – Polynuclear Manganese Coordination Complexes Atom Bond Mn(10) Mn(10)-O(39) Mn(10)-O(16) Mn(10)-O(29) Mn(10)-O(20) Mn(10)-O(11) Mn(10)-O(13) 1.870(4) 1.913(4) 1.914(4) 1.970(4) 2.210(4) 2.313(4) 3.148 +3 Mn(11) Mn(11)-O(6) Mn(11)-O(33) Mn(11)-O(59) Mn(11)-O(44) Mn(11)-O(38) Mn(11)-O(41) 1.857(4) 1.885(4) 1.919(5) 1.984(4) 2.220(5) 2.374(4) 3.155 +3 Mn(12) Mn(12)-O(61) Mn(12)-O(34) Mn(12)-O(47) Mn(12)-O(41) Mn(12)-O(54) Mn(12)-O(44) 1.861(4) 1.906(4) 1.916(4) 1.965(4) 2.185(4) 2.298(4) 3.212 +3 Mn(1) ··· Mn(2) Mn(3) ··· Mn(4) Mn(5) ··· Mn(6) 2.8135(14) 2.8281(14) 2.8043(14) Bond distances (Å) BVS Mn(2) ··· Cl(1) Mn(3) ··· Cl(1) Mn(5) ··· Cl(1) Cl(1) ··· Cl(2) Assigned oxidation state 2.835(2) 2.8124(19) 2.830(2) 3.612(0) Figure 3.57 – Polyhedral representation of the {Mn2O10}14− units in 14. Colour code: MnIII blue, P purple, O red, C grey, Cl green. 177 Chapter 3 – Polynuclear Manganese Coordination Complexes Figure 3.58 – Polyhedral representation of the {Mn12} core structure in 14 showing the intramolecular hydrogen bonding. Colour code: MnIII blue, P purple, O red, C grey, Cl green, K yellow. 178 Chapter 3 – Polynuclear Manganese Coordination Complexes Table 3.21 − Selected bond angles [º] for compound 14. Bond Angle (º) Bond Angle (º) O(17)-Mn(7)-O(28) O(51)-Mn(7)-O(7) O(45)-Mn(7)-O(15) O(51)-Mn(7)-O(45) O(17)-Mn(7)-O(51) O(17)-Mn(7)-O(45) O(7)-Mn(7)-O(45) O(28)-Mn(7)-O(45) O(17)-Mn(7)-O(15) O(17)-Mn(7)-O(7) 177.8(2) 174.97(19) 169.03(17) 94.40(19) 92.81(18) 91.61(17) 90.41(17) 88.96(17) 88.93(16) 88.52(17) O(3)-Mn(8)-O(31) O(60)-Mn(8)-O(15) O(30)-Mn(8)-O(7) O(60)-Mn(8)-O(30) O(3)-Mn(8)-O(30) O(3)-Mn(8)-O(60) O(15)-Mn(8)-O(30) O(31)-Mn(8)-O(30) O(3)-Mn(8)-O(15) O(3)-Mn(8)-O(7) 176.06(19) 172.88(19) 170.59(16) 95.09(19) 93.78(17) 92.01(18) 91.86(17) 89.69(17) 89.08(17) 88.19(16) O(26)-Mn(9)-O(52) O(43)-Mn(9)-O(13) O(49)-Mn(9)-O(20) O(43)-Mn(9)-O(49) O(26)-Mn(9)-O(43) O(26)-Mn(9)-O(49) O(52)-Mn(9)-O(49) O(13)-Mn(9)-O(49) O(26)-Mn(9)-O(13) O(26)-Mn(9)-O(20) 177.48(19) 174.67(18) 166.69(16) 96.44(18) 92.75(18) 91.90(17) 89.15(18) 88.65(17) 88.64(17) 88.77(16) O(39)-Mn(10)-O(16) O(29)-Mn(10)-O(20) O(11)-Mn(10)-O(13) O(39)-Mn(10)-O(11) O(20)-Mn(10)-O(11) O(29)-Mn(10)-O(11) O(39)-Mn(10)-O(29) O(39)-Mn(10)-O(20) O(16)-Mn(10)-O(11) O(39)-Mn(10)-O(13) 176.30(18) 172.45(19) 171.53(16) 94.00(17) 93.77(17) 93.51(18) 92.34(18) 89.16(17) 88.87(18) 88.50(16) O(33)-Mn(11)-O(44) O(6)-Mn(11)-O(59) O(38)-Mn(11)-O(41) O(33)-Mn(11)-O(38) O(6)-Mn(11)-O(38) O(6)-Mn(11)-O(33) O(44)-Mn(11)-O(38) O(59)-Mn(11)-O(38) O(6)-Mn(11)-O(44) O(6)-Mn(11)-O(41) 176.0(2) 175.65(19) 165.74(16) 94.84(19) 93.76(17) 93.07(18) 88.92(17) 88.91(19) 88.14(18) 86.79(16) O(61)-Mn(12)-O(47) O(34)-Mn(12)-O(41) O(54)-Mn(12)-O(44) O(34)-Mn(12)-O(54) O(61)-Mn(12)-O(54) O(61)-Mn(12)-O(34) O(41)-Mn(12)-O(54) O(47)-Mn(12)-O(54) O(61)-Mn(12)-O(41) O(61)-Mn(12)-O(44) 176.62(19) 174.30(19) 169.67(17) 94.58(19) 92.97(18) 92.57(19) 90.64(18) 90.04(18) 89.46(18) 87.53(17) 179 Chapter 3 – Polynuclear Manganese Coordination Complexes Figure 3.59 – Packing arrangement of the dodecanuclear manganese clusters in 14 viewed in the direction of the crystallographic a- and c-axis. Colour code: Mn blue, P purple, O red, C grey, Cl green, (crystallization solvent molecules and hydrogen atoms have been omitted for clarity). 180 Chapter 3 – Polynuclear Manganese Coordination Complexes - PHYSICOCHEMICAL CHARACTERISATION - FT-IR spectroscopy The IR spectrum of 14 is presented in Figure 3.60. Characteristic C–H out-of-plane bending vibrations of the benzylphosphonic acid ligands can be found in the range 800 – 650 cm-1. C–C skeletal vibrations of the phenyl rings and C–H bending and rocking vibrations of the –CH2 groups of the benzylphosphonic acid ligands appear in the 1500 – 1200 cm-1 region. Different P–O stretching vibrations of the phosphonate groups can be observed in the range 1200 – 900 cm-1. In addition, the corresponding O–H stretching vibrations and H–O–H bending vibrations of the crystallization water molecules engaged in H-bonds appear as broad bands at ca. 3200 cm-1 and ca. 1600 cm-1, respectively.36-40 Figure 3.60 – Infrared spectrum of 14. 181 Chapter 3 – Polynuclear Manganese Coordination Complexes - Thermogravimetric analysis The thermal stability of a freshly prepared crystalline sample of 14 was investigated, in an N2 atmosphere, in the temperature range 30 – 900 °C (Figure 3.61). The first thermogravimetric step observed in the TGA curve of 14 corresponds to a weight loss of 4.7 % and occurs in the temperature range between 30 – 100 °C. This event corresponds to the loss of three crystallization CH3OH molecules and four crystallization H2O molecules (calcd: 4.4 %). The following step in the TGA curve corresponds to a weight loss of 6.9 % between 100 – 210 °C, which is associated with the loss of six coordinated CH3OH molecules and four coordinated H2O molecules (calcd: 6.8 %). The compound then undergoes a gradual decomposition between 210 – 500 °C involving the organic ligands. The final thermogravimetric step is centered at 550 °C . Weight % (%) 100 90 80 70 60 50 0 200 400 600 800 1000 Temperature (ºC) Figure 3.61 – Thermogravimetric analysis of 14. 182 Chapter 3 – Polynuclear Manganese Coordination Complexes - Magnetism The temperature dependence of the magnetic susceptibility of 14 was measured between 298 and 1.8 K (Figure 3.62). The χT value of 29.5 cm3 K mol-1 at room temperature is significantly lower than the expected value of 36 cm3 K mol-1 of twelve S = 2 MnIII spin carriers (C = S(S+1) with = 0.12505 cm3 K mol-1 and g = 2).41 This result indicates that antiferromagnetic interactions dominate between the spin centres. The gradually decrease of the χT product down to a minimum of ca. 18.5 cm3 K mol-1 at ca. 40 K further confirm the presence of AF interactions. Below 40 K, at 1000 Oe the χT product decreases rapidely down to 6 cm3 K mol-1 at 1.8 K suggesting an ST = 3 spin ground state. a) b) Figure 3.62 – (a) Temperature dependence of the χT product of 14 at 0.1 and 1 T. (b) A magnified view of the χT product in (a) between 1 and 100 K. The field dependence of the magnetisation measured at temperatures between 1.84 and 8 K is shown in Figure 3.63, a. The M vs H plot reveals a continuous increase of the magnetisation without clear saturation even at 1.84 K at 7 T. This observation suggests the presence of magnetic anisotropy. The data represented as a M vs H/T plot (Figure 3.63, b) confirms the presence of anisotropy intrinsic to MnIII metal ions and the cluster as the data are not superposed on a single master-curve as expected for isotropic systems. However, the complex is not expected to have a significant anisotropy as the MnIII Jahn-Teller axes are approximately aligned with the plane of the {Mn12} ring (see crystal structure in Figure 3.58). 183 Chapter 3 – Polynuclear Manganese Coordination Complexes a) b) Figure 3.63 – (a) M vs H and (b) M vs H/T data at and below 8 K. - Mass spectrometry ESI-MS studies on the crystalline material revealed that the {Mn12} cluster in 14 is stable in CH3CN, DMF and DMSO environments (Table 3.22). The mass spectrum of 14 dissolved in DMSO is presented in Figure 3.64 and reveals the presence of one major isotopic envelop in the high molecular mass region. This signal is centered at m/z = 1673.4 a.m.u. and was assigned to a doubly charged species {H7[ nIII10MnII2 6(C6 5CH2P 3)15] Cl}2-. A summary of the ESI-MS assignment is given in Table 3.22. The envelopes of all Relative ionic abundance assigned species were modelled. Figure 3.64 – Negative-mode ESI-MS spectra for crystals of 14 dissolved in DMSO. Inset: Comparison of the experimental isotopic envelopes (black spectrum) with simulated patterns (red spectrum) for {H7[ nIII10MnII2 6(C6 5CH2P 3)15]Cl}2- centered at m/z = 1673.4 a.m.u. (cone voltage: 30 V). 184 Chapter 3 – Polynuclear Manganese Coordination Complexes Table 3.22 – ESI-MS assignment for compound 14. Crystals of Solvent m/z Species attributed CH3CN 1673.4 {H7[MnIII10MnII2O6(C6H5CH2PO3)15]Cl}2- → {Mn12} DMF 1673.4 {H7[MnIII10MnII2O6(C6H5CH2PO3)15]Cl}2- → {Mn12} 1673.4 {H7[MnIII10MnII2O6(C6H5CH2PO3)15]Cl}2- → {Mn12} 14 in: - DMSO UV-Vis spectroscopy The UV-Vis absorption spectrum of 14 recorded in CH3CN (Figure 3.65) displays an intense absorption band at ca. 257 nm ( = 36000 L mol-1 cm-1) which can be attributed to π − π* transitions of the aromatic phosphonate ligands. The weak band observed at ca. 488 nm ( = 3200 L mol-1 cm-1) may arise from d – d transitions that can be assigned to T2g 5 Eg transitions of octahedrally coordinatied MnIII ions within 14.49-52 0.8 0.35 0.30 0.6 Absorbance (a.u.) 5 0.25 0.20 0.15 0.4 0.10 400 450 500 550 600 0.2 0.0 300 400 500 600 700 800 Wavelength (nm) Figure 3.65 – UV-Vis spectrum of a 10-5 M solution of 14 in CH3CN. Inset: A section of the UV-Vis spectrum of a 10-4 M solution of 14 in CH3CN. 185 Chapter 3 – Polynuclear Manganese Coordination Complexes Table 3.23 − Crystal data and structural refinement parameters for 14. Compound 14 Empirical formula a Molecular mass/g mol C111H144Cl2KMn12O61P15 -1 a Crystal colour/shape 3 Dark brown / rectangular block Crystal size/mm 0.50×0.20×0.15 Crystal system Orthorhombic Space group Pbca a/ Å 27.2527(17) b/ Å 29.947(2) c/ Å 41.980(3) /º 90 /º 90 /º 90 3 V/ Å 34261(4) Z 1 Temperature (K) 150(2) -3 Density/Mg m 1.402 -1 Absorp. coef./mm 1.125 F(000) 14144 2 50 max/º Reflections collected 198099 Independent reflections 30136 [R(int)=0.1519] Data / restraints / parameters 30136 / 0 / 1880 2 a 3688.16 S on F 1.024 R1, wR2 [I>2 (I)] 0.0602, 0.1412 R1, wR2 (all data) 0.1146, 0.1514 Largest diff. peak and hole/e.Å-3 2.299 and -1.597 Excluding solvate molecules 186 Chapter 3 – Polynuclear Manganese Coordination Complexes 3.3.3.2 K(H2O)4[MnIII12(μ3-O)6(CH3OH)6(C6H5CH2PO3H)7(C6H5CH2PO3)8]Br2·2CH3OH·2H2O (15) Compound 15, K(H2O)4[MnIII12( 3-O)6(CH3OH)6(C6H5CH2PO3H)7 (C6H5CH2PO3)8]Br2·2CH3OH·2H2O was prepared in a similar manner to 14 using MnBr2·4H2O instead of MnCl2·4H2O, in order to investigate the role of the halide ion as a potential templating agent. X-ray structural analysis revealed that 15 (Figure 3.66) is almost isostructural to 14, having only slightly different structural and geometrical parameters. Selected interatomic distances and angles for compound 15 are given in Table 3.24 and Table 3.25, and agree very well with those observed for 14. Bond valence sum analysis confirms again that all the Mn ions in 15 are in the oxidation state +III; the overall charge of the cluster is compensated by one K+ ion and two Br− ions. As before, one Br− ion is situated in the central cavity of the ring, at 2.954(1) Å, 2.973(1) Å and 2.981(1) Å to Mn(1), Mn(3) and Mn(5) respectively. The other Br− ion, Br(2) is situated just above the cavity of the ring, residing 3.879(2) Å from Br(1). As for 14, the potassium ions link the clster entities into 1D zigzag chains. Figure 3.66 − Crystal structure of the dodecanuclear manganese complex in 15. Colour code: MnIII blue, P purple, O red, C grey, Br dark green, K yellow (hydrogen atoms have been omitted for clarity). 187 Chapter 3 – Polynuclear Manganese Coordination Complexes (A) Figure 3.67 − Polyhedral representation of the triangular units A found in 15. Colour code: MnIII blue, P purple, O red, C grey, Br dark green. (B) Figure 3.68 – Polyhedral representation of the {Mn2O10}14− units in 15. Colour code: MnIII blue, P purple, O red, C grey, Br dark green. 188 Chapter 3 – Polynuclear Manganese Coordination Complexes Table 3.24 − Selected bond lengths [Å] and bond valence sum (BVS) for compound 15. Atom Bond Bond distances (Å) BVS Assigned oxidation state Mn(1) Mn(1)-O(46) Mn(1)-O(47) Mn(1)-O(39) Mn(1)-O(45) Mn(1)-O(8) 1.871(5) 1.877(5) 1.916(5) 1.938(5) 2.104(5) 3.135 +3 Mn(2) Mn(2)-O(46) Mn(2)-O(47) Mn(2)-O(23) Mn(2)-O(41) Mn(2)-O(48) 1.874(5) 1.891(5) 1.899(5) 1.914(5) 2.084(5) 3.198 +3 Mn(3) Mn(3)-O(33) Mn(3)-O(32) Mn(3)-O(38) Mn(3)-O(60) Mn(3)-O(31) 1.873(5) 1.877(5) 1.911(5) 1.943(5) 2.077(5) 3.161 +3 Mn(4) Mn(4)-O(32) Mn(4)-O(16) Mn(4)-O(33) Mn(4)-O(24) Mn(4)-O(26) 1.873(5) 1.884(5) 1.891(4) 1.920(5) 2.076(5) 3.224 +3 Mn(5) Mn(5)-O(51) Mn(5)-O(61) Mn(5)-O(59) Mn(5)-O(43) Mn(5)-O(63) 1.879(5) 1.893(5) 1.909(5) 1.930(5) 2.094(5) 3.125 +3 Mn(6) Mn(6)-O(61) Mn(6)-O(51) Mn(6)-O(40) Mn(6)-O(14) Mn(6)-O(22) 1.873(5) 1.882(5) 1.897(5) 1.925(5) 2.080(5) 3.204 +3 Mn(7) Mn(7)-O(51) Mn(7)-O(52) Mn(7)-O(53) Mn(7)-O(42) Mn(7)-O(56) Mn(7)-O(44) 1.863(4) 1.911(5) 1.939(5) 1.974(5) 2.188(5) 2.324(5) 3.127 +3 Mn(8) Mn(8)-O(46) Mn(8)-O(50) Mn(8)-O(54) Mn(8)-O(44) Mn(8)-O(49) Mn(8)-O(42) 1.865(4) 1.892(5) 1.912(5) 1.989(5) 2.218(5) 2.366(5) 3.135 +3 Mn(9) Mn(9)-O(47) Mn(9)-O(9) Mn(9)-O(7) Mn(9)-O(25) Mn(9)-O(36) Mn(9)-O(37) 1.862(4) 1.903(5) 1.923(5) 1.968(5) 2.214(5) 2.313(5) 3.166 +3 189 Chapter 3 – Polynuclear Manganese Coordination Complexes Atom Bond Mn(10) Mn(10)-O(32) Mn(10)-O(34) Mn(10)-O(28) Mn(10)-O(37) Mn(10)-O(35) Mn(10)-O(25) 1.866(4) 1.890(5) 1.902(5) 1.992(5) 2.217(5) 2.316(5) 3.178 +3 Mn(11) Mn(11)-O(33) Mn(11)-O(29) Mn(11)-O(1) Mn(11)-O(12) Mn(11)-O(2W) Mn(11)-O(58) 1.863(4) 1.903(5) 1.917(4) 1.991(5) 2.197(5) 2.303(5) 3.161 +3 Mn(12) Mn(12)-O(61) Mn(12)-O(18) Mn(12)-O(5) Mn(12)-O(58) Mn(12)-O(4W) Mn(12)-O(12) 1.861(5) 1.895(5) 1.907(5) 1.978(5) 2.212(5) 2.332(5) 3.185 +3 Mn(1) ··· Br(1) Mn(3) ··· Brl(1) Mn(5) ··· Br(1) Br(1) ··· Br(2) 2.954(1) 2.973(1) 2.981(1) 3.879(2) Mn(1) ··· Mn(2) Mn(3) ··· Mn(4) Mn(5) ··· Mn(6) 2.8232(15) 2.8098(15) 2.8029(16) Bond distances (Å) BVS Assigned oxidation state 190 Chapter 3 – Polynuclear Manganese Coordination Complexes Table 3.25 − Selected bond angles [º] for compound 15. Bond Angle (º) Bond Angle (º) O(51)-Mn(7)-O(53) O(52)-Mn(7)-O(42) O(56)-Mn(7)-O(44) O(52)-Mn(7)-O(56) O(51)-Mn(7)-O(56) O(51)-Mn(7)-O(52) O(42)-Mn(7)-O(56) O(53)-Mn(7)-O(56) O(51)-Mn(7)-O(42) O(51)-Mn(7)-O(44) 177.3(2) 173.5(2) 169.23(19) 94.9(2) 91.9(2) 91.8(2) 91.4(2) 90.4(2) 89.8(2) 88.27(19) O(46)-Mn(8)-O(54) O(50)-Mn(8)-O(44) O(49)-Mn(8)-O(42) O(50)-Mn(8)-O(49) O(46)-Mn(8)-O(50) O(46)-Mn(8)-O(49) O(44)-Mn(8)-O(49) O(54)-Mn(8)-O(49) O(46)-Mn(8)-O(44) O(46)-Mn(8)-O(42) 176.6(2) 175.8(2) 165.46(19) 95.0(2) 92.6(2) 92.5(2) 88.9(2) 88.7(2) 88.6(2) 87.76(19) O(47)-Mn(9)-O(9) O(7)-Mn(9)-O(25) O(36)-Mn(9)-O(37) O(7)-Mn(9)-O(36) O(25)-Mn(9)-O(36) O(47)-Mn(9)-O(36) O(47)-Mn(9)-O(7) O(9)-Mn(9)-O(36) O(47)-Mn(9)-O(25) O(47)-Mn(9)-O(37) 177.2(2) 171.8(2) 171.41(19) 94.1(2) 94.0(2) 92.8(2) 91.8(2) 89.3(2) 89.05(19) 88.64(18) O(32)-Mn(10)-O(34) O(28)-Mn(10)-O(37) O(35)-Mn(10)-O(25) O(28)-Mn(10)-O(35) O(32)-Mn(10)-O(28) O(32)-Mn(10)-O(35) O(34)-Mn(10)-O(35) O(32)-Mn(10)-O(25) O(32)-Mn(10)-O(37) O(37)-Mn(10)-O(35) 178.0(2) 174.5(2) 165.67(18) 96.6(2) 92.7(2) 91.51(19) 89.3(2) 88.90(18) 88.82(19) 88.63(19) O(33)-Mn(11)-O(1) O(29)-Mn(11)-O(12) O(2)-Mn(11)-O(58) O(29)-Mn(11)-O(2) O(33)-Mn(11)-O(2) O(12)-Mn(11)-O(2) O(33)-Mn(11)-O(29) O(1)-Mn(11)-O(2) O(33)-Mn(11)-O(12) O(33)-Mn(11)-O(58) 176.7(2) 172.1(2) 170.60(19) 95.8(2) 92.83(19) 91.95(19) 91.7(2) 90.09(19) 89.46(19) 88.39(18) O(61)-Mn(12)-O(5) O(18)-Mn(12)-O(58) O(4)-Mn(12)-O(12) O(18)-Mn(12)-O(4) O(61)-Mn(12)-O(18) O(61)-Mn(12)-O(4) O(58)-Mn(12)-O(4) O(61)-Mn(12)-O(12) O(5)-Mn(12)-O(4) O(61)-Mn(12)-O(58) 178.8(2) 175.0(2) 168.66(19) 94.4(2) 92.3(2) 91.13(19) 90.4(2) 89.52(18) 89.37(19) 88.9(2) 191 Chapter 3 – Polynuclear Manganese Coordination Complexes - PHYSICOCHEMICAL CHARACTERISATION The IR spectrum of compound 15 exhibits the same general characteristics than that of 14. The UV-Vis absorption spectrum of 15 in CH3CN also displays idential features to the spectrum of 14, showing an intense absorption band at ca. 258 nm ( = 52000 L mol-1 cm-1) (π − π* transitions of the phosphonate ligands), and a weak absorption band at ca. 489 nm ( = 4000 L mol-1 cm-1) (5T2g 5 Eg most likely resulting from d-d transtions of octahedrally coordinated MnIII centers). - Thermogravimetric analysis Weight % (%) 100 90 80 70 60 50 0 200 400 600 800 1000 Temperature (ºC) Figure 3.69 – Thermogravimetric analysis of 15. The TGA analysis in an N2 atmosphere reveals that 15 also undergoes a similar decomposition behaviour as 14. The first termogravimetric step observed in the TGA curve of 15 (Figure 3.69) corresponds to a weight loss of 2.6 % between 30 – 100 °C. This step was attributed to the loss of two crystallization CH3OH molecules and two crystallization H2O molecules (calcd: 2.6 %). The next termogravimetric step observed in the TGA occurs in the temperature range between 100 – 200 °C and corresponds to a weight loss of 6.3 %. 192 Chapter 3 – Polynuclear Manganese Coordination Complexes This weight loss can be associated with the loss of six coordinated CH3OH molecules and four coordinated H2O molecules (calcd: 6.8 %). Decompositions of the organic ligands and thermolysis processes between 200 – 600 °C destroy the cluster entities. - Magnetism Magnetisation data for compound 15 were collected at high magnetic field strengths and low temperatures. The magnetic properties of 15 are almost identical with those of 14, beeing caracterised by dominant antiferromagnetic interactions between the spin centres and an ST = 3 ground state. As expected, exchange of the halide ions has no significant effect on the structure and thus on magnetic properties of the cluster core. 15 14 Figure 3.70 – Comparison of the χT vs T plot of 14 and 15 at 0.1 T. 193 Chapter 3 – Polynuclear Manganese Coordination Complexes - Mass spectrometry The solution stability of the {Mn12} cluster in 15 was again investigated by electrospray ionisation mass spectrometry. The ESI-MS spectrum of pristine crystals of 15 dissolved in CH3CN is presented in Figure 3.71 and reveals the presence of three major isotopic envelopes in the high mass region of the spectrum. Two of the signals were assigned to {H6[ nIII10MnII2 {H6[ nIII11MnII 6(C6 5CH2P 6(C6 5CH2P 3)15]Br} 2- 3)15]} 2- centered at m/z = 1655.0 a.m.u. and centered at m/z = 1695.4 a.m.u., while the third signal could not be assigned, most likely due to overlapping species. ESI-MS studies revealed that the {Mn12} cluster in 15 is stable in CH3CN, DMF and DMSO environments. A summary of the ESI-MS assignment is presented in Table 3.26. The isotopic envelopes of all assigned species were modelled. Table 3.26 – ESI-MS assignment for compound 15. Solvent CH3CN Crystals of 15 in: DMF Species attributed 1655.0 {H6[MnIII10MnII2O6(C6H5CH2PO3)15]}2- → {Mn12} 1695.4 {H6[MnIII11MnIIO6(C6H5CH2PO3)15]Br}2- → {Mn12} 1655.0 {H6[MnIII10MnII2O6(C6H5CH2PO3)15]}2- → {Mn12} 1655.0 {H6[MnIII10MnII2O6(C6H5CH2PO3)15]}2- → {Mn12} 1695.4 {H6[MnIII11MnIIO6(C6H5CH2PO3)15]Br}2- → {Mn12} Relative ionic abundance DMSO m/z Figure 3.71 – Negative-mode ESI-MS spectra for crystals of 15 dissolved in CH3CN. Inset: Comparison of the experimental isotopic envelopes (black spectrum) with simulated patterns (red spectrum) for {H6[ nIII10MnII2 6(C6 5CH2P 3)15]}2- centered at m/z = 1655.0 a.m.u. and {H6[ nIII11MnII 6(C6 5CH2P 3)15]Br}2- centered at m/z = 1695.4 a.m.u. (cone voltage: 30 V). 194 Chapter 3 – Polynuclear Manganese Coordination Complexes Table 3.27 − Crystal data and structural refinement parameters for 15. Compound 15 Empirical formula a Molecular mass/g mol C111H144Br2KMn12O61P15 -1 a Crystal colour/shape 3 Brown / rectangular block Crystal size/mm 0.50×0.10×0.10 Crystal system Orthorhombic Space group Pbca a/ Å 27.112(2) b/ Å 29.957(2) c/ Å 41.972(3) /º 90 /º 90 /º 90 3 V/ Å 34089(4) Z 1 Temperature (K) 150(2) -3 Density/Mg m 1.458 -1 Absorp. coef./mm 1.569 F(000) 14584 2 50 max/º Reflections collected 194198 Independent reflections 29987 [R(int)=0.1221] Data / restraints / parameters 29987 / 0 / 1859 2 a 3777.06 S on F 0.965 R1, wR2 [I>2 (I)] 0.0708, 0.2030 R1, wR2 (all data) 0.1095, 0.2196 Largest diff. peak and hole/e.Å-3 2.661 and -3.278 Excluding solvate molecules 195 Chapter 3 – Polynuclear Manganese Coordination Complexes 3.3.4 Synthesis and characterisation of a hexanuclear manganese complex 3.3.4.1 (H3O)4[MnIII2MnII4(μ4-O)2(H2O)2(CH3CN)2{(C6H5)3CPO3}6]Cl2·2CH3CN·4H2O (16) Compound 16, (H3O)4[MnIII2MnII4( 4-O)2(H2O)2(CH3CN)2{(C6H5)3CPO3}6]Cl2 ·2CH3CN·4H2O, (Figure 3.72) was prepared by a comproportionation reaction between MnCl2·4H2O and KMnO4, using triphenylmethylphosphonic acid as a stabilising ligand and triethylamine as an organic base. The reaction mixture was stirred for five hours, filtered and kept undisturbed at 25 °C for slow evaporation of the solvent. Rectangular redbrown crystals of 16 separated from the reaction mixture within a time period of five days. These were characterised by single crystal X-ray diffraction measurements. 16 crystallises in the monoclinic crystal system in the space group C2/c. This compound contains a mixed-valent, hexanuclear manganese cluster in which the six Mn ions are arranged to form two edge-sharing tetrahedra, with a μ4-O2- ion at the center of each tetrahedron (Figure 3.73). The symmetry of the cluster core is characterised by an inversion centre located at the mid-point of the common edge of the two terahedra. Figure 3.72 − Crystal structure of the hexanuclear manganese complex in 16. Colour code: MnIII blue, MnII cyan, P purple, O red, C grey, N dark blue, Cl green (hydrogen atoms have been omitted for clarity). The resultant {MnIII2MnII4O2}10+ core structure is related to previously reported coordination clusters that are stabilised by carboxylate ligands.58-64 However, in 16, the 196 Chapter 3 – Polynuclear Manganese Coordination Complexes two edge-sharing tetrahedra are further connected by six fully deprotonated phosphonate ligands, that bridge between three Mn atoms in a η1:η1:η1:μ3 bridging mode. The cluster is further stabilised by two CH3CN and two H2O molecules that complete the trigonal bipyramidal environments of the four peripheral MnII ions (Mn(3) – O(6) bond length 2.250(4) Å and Mn(2) – N(1) bond length 2.312(5) Å). The μ4-O2- oxo ligand and its symmetry equivalent are situated in trans positions to the H2O/CH3CN ligands giving rise to Mn(2) – O(3) and Mn(3) – O(3) distances of 2.347(3) Å and 2.319(4) Å, respectively. The remaining coordination sites of the trigonal bipyramidal coordination environments of these outer Mn centres, Mn(2), Mn(3) and symmetry equivalents, are occupied by three phosphonate O donors. The resulting Mn – O distances range between 2.034(4) Å – 2.073(4) Å. The two central Mn ions, Mn(1) and Mn(1`), adopt square pyramidal coordination environments, with the polyhedra facing in opposite directions with respect to each other. The square pyramidal polyhedra consist of two μ4-O2- oxo ligands (Mn(1) – O(3) and Mn(1) – O(3`) distances 1.896(3) Å and 1.880(3) Å, respectively), and three phosphonate O donors (Mn – O distances range between 1.952(3) Å – 2.143(4) Å). Bond valence sum analysis (Table 3.28) confirms that the two Mn ions Mn(1) an Mn(1’) are in the oxidation state +III. The remaning four Mn ions are in the oxidation state +II. The short Mn(1) ··· Mn(1`) contact of 2.7950(16) Å is in agreement with the assigned +III oxidation state of the Mn ions, and compares well with bond distances observed in structurally related compounds [Mn6O2(O2CPh)10(py)2(MeCN)2]·2MeCN (2.820(3) Å), [Mn6O2(O2CPh)10(4hmpH)3 (MeCN)] (2.817(19) Å) and [Mn6O2(O2CPh)10(pym)2 (MeCN)2] (2.823(3) Å) previously reported in the literature.58, 59 16 is further stabilised by two symmetry equivalent chloride counterions that reside in the vicinity of the cluster core. Four H3O+ ions were assigned in order to balance the overall charge of the cluster. Hydronium counterions have previously been reported to stabilise related Mn coordination clusters, e.g. in {(H3O)2[Mn4-(4-Haba)2(4- aba)6(SCN)4(H2O)2]} (4-Haba = 4-aminobenzoic acid), (H3O)[Na2MnIII6MnII2(μ4-O)2(μ1,1N3)7(μ1,3-N3)(H2L)6Cl]·3H2O, (H3O)[Na4MnIII12MnII4(μ4-O)4(μ1,1-N3)9(μ1,1,3-N3)2(H2L)12 (CH3O)4(H2O)4]·2ClO4·2(CH3OH)2·4H2O, (H4L = 2-{[(2-hydroxy-3-methoxyphenyl) methylene]amino}-2-(hydroxymet-hyl)-1,3-propanediol).65, 66 197 Chapter 3 – Polynuclear Manganese Coordination Complexes Figure 3.73 – Polyhedral representation of the {Mn6} core structure in 16. Colour code: MnIII blue, MnII cyan, P purple, O red, C grey. Table 3.28 − Selected bond lengths [Å] and bond valence sum (BVS) for compound 16. Atom Bond Bond distances (Å) BVS Assigned oxidation state Mn(1) Mn(1)-O(3`) Mn(1)-O(3) Mn(1)-O(8) Mn(1)-O(7) Mn(1)-O(9) 1.880(3) 1.896(3) 1.952(4) 1.954(4) 2.143(4) 2.955 +3 Mn(2) Mn(2)-O(2`) Mn(2)-O(10) Mn(2)-O(1`) Mn(2)-N(1) Mn(2)-O(3) 2.034(4) 2.037(4) 2.064(4) 2.312(5) 2.347(3) 1.971 +2 Mn(3) Mn(3)-O(11`) Mn(3)-O(4) Mn(3)-O(5) Mn(3)-O(6) Mn(3)-O(3) 2.048(4) 2.055(4) 2.073(4) 2.250(4) 2.319(4) 1.977 +2 Mn(1) ··· Mn(1`) 2.7950(16) 198 Chapter 3 – Polynuclear Manganese Coordination Complexes The {MnIII2MnII4O2}10+ core structure in 16 can alternatively be described as consisting of two triangular units with Mn(1), Mn(2) and Mn(3), and Mn(1`), Mn(2`) and Mn(3`) as vertices (Mn(1) ··· Mn(2), Mn(2) ··· Mn(3) and Mn(3) ··· Mn(1) are 3.499(55) Å, 3.664(3) Å and 3.598(2) Å, respectively). These triangular units are connected to each other by four phosphonate ligands and two μ4-O2- ions. The crystal structure of 16 is stabilized by weak intermolecular hydrogen bonds and π-π interactions, resulting in a grid-like packing arrangement that can be seen in the direction of the crystallographic b-axis. A number of hydrogen bonds involve the two coordinated H2O molecules of the Mn cluster and the N atoms of the CH3CN solvent molecules (O(6) – N(2) distance 2.764(31) Å). Figure 3.74 – Packing arrangement of the hexanuclear manganese clusters in 16 viewed in the direction of the crystallographic a- and b-axis. Colour code: Mn blue, P purple, O red, C grey, Cl green, (crystallization solvent molecules and hydrogen atoms have been omitted for clarity). 199 Chapter 3 – Polynuclear Manganese Coordination Complexes - PHYSICOCHEMICAL CHARACTERISATION - FT-IR spectroscopy The IR spectrum of 16 presented in Figure 3.75 displays some characteristic bands of the aromatic rings of the phosphoate ligands, such as C–C skeletal vibrations (1600 – 1400 cm-1) and C–H out-of-plane bending vibrations (900 – 650 cm-1). Different P–O stretching vibrations of the phosphonate groups can be observed in the range 1200 – 900 cm-1, while the corresponding O–H stretching vibrations and H–O–H bending vibrations of the crystallization water molecules engaged in H-bonds appear as broad bands at ca. 3200 cm-1 and ca. 1600 cm-1, respectively.36-40 Figure 3.75 – Infrared spectrum of 16. 200 Chapter 3 – Polynuclear Manganese Coordination Complexes - Thermogravimetric analysis The thermal stability of the compound in an N2 atmosphere in the temperature range 30 – 900 °C was investigated using a freshly prepared crystalline sample of 16 (Figure 3.76). The compound exhibits a weight loss of 6.3 % in the temperature range between 30 – 150 °C that can be attributed to the loss of two crystallization CH3CN molecules and four crystallization H2O molecules (calcd: 5.7 %). A further temperature increase results in a gradual weight loss of 4.8 % between 150 – 280 °C, which could be associated with the loss of two coordinated CH3CN molecules and two coordinated H2O molecules (calcd: 4.4 %). Between 280 – 600 °C the oxidative degradation of the organic ligands takes place and further thermolysis processes result in the cluster degradation. 100 Weight % (%) 90 80 70 60 50 40 0 200 400 600 800 1000 Temperature (ºC) Figure 3.76 – Thermogravimetric analysis of 16. 201 Chapter 3 – Polynuclear Manganese Coordination Complexes - Magnetism The magnetic studies for compound 16 were carried out on a polycrystalline sample at high magnetic field strengths and low temperatures. The temperature dependence of the χT product (Figure 3.77) reveals dominant AF interactions between the spin centres; a Curie-Weiss fit was applied to analyse the data. The Curie constant of 23.2 cm3 K mol-1 is comparable to the expected value of 23.5 cm3 K mol-1 of two S = 2 MnIII and four S = MnII spin carriers (C = S(S+1) with = 0.12505 cm3 K mol-1 and g = 2).41 The extrapolation of the χT vs T data at low temperature implies that the ground state of 16 is ST = 0. Figure 3.77 – Temperature dependence of the χT product of 16 at 0.1 T. The red solid line corresponds to the best fit of the experimental data with the Curie-Weiss law (C = 23.2 cm3 K mol-1 and = 34 K). 202 Chapter 3 – Polynuclear Manganese Coordination Complexes - Mass spectrometry The stability of the {Mn6} cluster in solution was investigated by electrospray ionisation mass spectrometry. A major isotopic envelop centered at m/z = 2296.1 a.m.u is observed in the ESI-MS spectrum of the pristine crystals dissolved in CH3CN and DMF and it can be assigned to a {H nIII2MnII4 2[(C6 5)3CP 3]6} - species. The corresponding doubly charged species could also be identified in the ESI-MS spectrum of the pristine crystals dissolved in CH3CN (m/z = 1147.4, {H2 nII6 2[(C6 5)3CP 23]6} ). A summary of the ESI-MS signal assignment for compound 16 is presented in Table 3.29. The mass spectrum of 16 dissolved in DMSO does not show any signals in the high molecular mass Relative ionic abundance Relative ionic abundance Relative ionic abundance Relative ionic abundance region, suggesting that 16 is not stable in DMSO. Figure 3.78 – Negative-mode ESI-MS spectra for crystals of 16 dissolved in CH3CN. Comparison of the experimental isotopic envelopes (black spectrum) with simulated patterns (red spectrum) for {H2MnII6 2[(C6 5)3CP 3]6]}2- centered at m/z = 1147.4 a.m.u., {H nIII2MnII4 2[(C6 5)3CP 3]6]}centered at m/z = 2296.1 a.m.u. and {H7MnII6 4[(C6 5)3CP 3]6]}- centered at m/z = 2334.0 a.m.u., (cone voltage: 30 V). 203 Chapter 3 – Polynuclear Manganese Coordination Complexes Table 3.29 – ESI-MS assignment for compound 16. Solvent Crystals of CH3CN 16 in: DMF - m/z Species attributed 1147.4 {H2MnII6O2[(C6H5)3CPO3]6]}2- → {Mn6} 2296.1 {HMnIII2MnII4O2[(C6H5)3CPO3]6]}- → {Mn6} 2334.0 {H7MnII6O4[(C6H5)3CPO3]6]}- → {Mn6} 2296.1 {HMnIII2MnII4O2[(C6H5)3CPO3]6]}- → {Mn6} UV-Vis spectroscopy The UV-Vis absorption spectrum of 16 recorded in CH3CN is presented in Figure 3.79. The intense absorption band observed at ca. 263 nm ( = 87000 L mol-1 cm-1) can be attributed to π − π* transitions of the phosphonate ligands, while the weak band observed at ca. 487 nm ( = 900 L mol-1 cm-1) involves d – d transitions.49-52 Absorbance (a.u.) 0.90 0.85 0.12 0.80 0.10 0.08 0.75 0.06 0.70 0.04 0.65 400 450 500 550 600 0.60 0.55 0.50 300 400 500 600 700 800 Wavelength (nm) Figure 3.79 – UV-Vis spectrum of a 10-5 M solution of 16 in CH3CN. Inset: A section of the UV-Vis spectrum of a 10-4 M solution of 16 in CH3CN. 204 Chapter 3 – Polynuclear Manganese Coordination Complexes Table 3.30 − Crystal data and structural refinement parameters for 16. Compound 16 Empirical formula a Molecular mass/g mol C118H112Cl2Mn6N2O26P6 -1 a Crystal colour/shape 3 Red brown / rectangular block Crystal size/mm 0.60×0.20×0.10 Crystal system Monoclinic Space group C 2/c a/ Å 31.374(6) b/ Å 14.709(3) c/ Å 28.425(6) /º 90 /º 93.46(3) /º 90 3 V/ Å 13094(5) Z 1 Temperature (K) 150(2) -3 Density/Mg m 1.336 -1 Absorp. coef./mm 0.743 F(000) 5385 2 50 max/º Reflections collected 49746 Independent reflections 11477 [R(int)=0.0473] Data / restraints / parameters 11477/ 0 / 743 2 a 2560.52 S on F 1.081 R1, wR2 [I>2 (I)] 0.0804, 0.2205 R1, wR2 (all data) 0.0832, 0.2233 Largest diff. peak and hole/e.Å-3 1.834 and -0.714 Excluding solvate molecules 205 Chapter 3 – Polynuclear Manganese Coordination Complexes 3.4 ESI-MS STUDIES OF COMPLEX REACTION MIXTURES TO INVESTIGATE THE FORMATION OF POLYNUCLEAR MANGANESE COORDINATION COMPLEXES 3.4.1 Investigation of the reaction system that led to the formation of the {Mn15} complex in 10 The comproportionation reaction between MnCl2·4H2O and KMnO4, in MeOH, in the presence of phenylphosphonic acid resulted in a deep brown solution, from which redbrown crystals of 10, [ nΙΙΙ15(μ2-H2 )2(C 3 )16(C6 5P 3)20]Cl5·22CH3OH·8H2O, separate within a time period of approximately one week (Figure 3.80). Figure 3.80 – Crystal structure of the pentadecanuclear manganese complex in 10. Colour code: MnIII blue, P purple, O red, C grey, Cl green (hydrogen atoms have been omitted for clarity). The initial reaction mixture was examined by ESI-MS and a rather complex mass spectrum was obtained (Figure 3.82). The spectrum revealed to be very interesting as three different species, {Mn7}, {Mn8} and {Mn13}, could be identified in the high molecular mass region of the spectrum (Table 3.31). The most intense signal in the high molecular mass region of the spectrum is centered at m/z = 1684.3 a.m.u. and corresponds to a {Mn7} species, {H17[MnIV6MnIIIO15(C6H5PO3)6]Cl3}-. The signals centered at m/z = 1016.7 a.m.u. and m/z = 2034.3 a.m.u. can be attributed to a {Mn8} species, the {H5[MnIII8O5(C6H5PO3)9]Cl3}2and {H4[MnIV2MnIII6O5(C6H5PO3)9]Cl3}-, respectively. The fourth isotopic envelope that 206 Chapter 3 – Polynuclear Manganese Coordination Complexes can be assigned in the mass spectrum of the reaction mixture of 10 is centered at m/z = 2122.0 a.m.u. and corresponds to the formula of a tridecanuclear manganese complex, {H[MnIV8MnIII5O15(CH3O)4(C6H5PO3)6]Cl3}-. The {Mn15} complex itself could not be identified in the mass spectrum of the reaction mixture that led to the crystallisation of 10 as there is no signal present in the spectrum above 3000 a.m.u. The species could also not be identified in the low molecular mass region of the spectrum as this part of the spectrum contains crowded large number of signals due to fragmentation and arising from multiple charged species. All species identified in the mass spectrum were modelled and good fits between the experimental and simulated isotopic envelopes confirm the assignments (Figure 3.82). Table 3.31 – ESI-MS assignment for the reaction mixture that led to the crystallisation of 10. m/z Species attributed 1016.7 {H5[MnIII8O5(C6H5PO3)9]Cl3}2IV III → {Mn8} - Reaction 1684.3 {H17[Mn 6Mn O15(C6H5PO3)6]Cl3} → {Mn7} mixture of 10 2034.3 {H4[MnIV2MnIII6O5(C6H5PO3)9]Cl3}- → {Mn8} 2122.0 {H[MnIV8MnIII5O15(CH3O)4(C6H5PO3)6]Cl3}- → {Mn13} a) b) {Mn7} {Mn8} Figure 3.81 – Representation of structural motifs that agree with the constitutional assignment: (a) {Mn7} and (b) {Mn8} species identified in the mass spectrum of the reaction mixture that led to the crystallisation of 10. Colour code: Mn blue, P purple, C grey, O and Cl light grey. 207 Relative ionic abundance Relative ionic abundance Relative ionic abundance Relative ionic abundance Relative ionic abundance Chapter 3 – Polynuclear Manganese Coordination Complexes Figure 3.82 – Negative-mode ESI-MS spectra for the reaction mixture that led to the crystallisation of 10 (24 h after preparation). Comparison of the experimental isotopic envelopes (black spectrum) with simulated patterns (red spectrum) for {H5[MnIII8O5(C6H5PO3)9]Cl3}2- centered at m/z = 1016.7 a.m.u., {H17[MnIV6MnIIIO15(C6H5PO3)6]Cl3}centered at m/z = 1684.3 a.m.u., IV {H4[Mn 2MnIII6O5(C6H5PO3)9]Cl3}centered at m/z = 2034.3 a.m.u. and {H[MnIV8MnIII5O15(CH3O)4(C6H5PO3)6]Cl3}- centered at m/z = 2122.0 a.m.u. (cone voltage: 30 V). 208 − − Chapter 3 – Polynuclear Manganese Coordination Complexes Figure 3.83 – Negative-mode ESI-MS spectra for the reaction mixture that led to the crystallisation of 10 at different cone voltages: 10V, 20V, 30V, 50V and 75V, respectively. In an effort to confirm that the assigned species coexist in solution and that they are not a result of a primary processes occurring in the spraying chamber, we set out to investigate the effect of cone voltage (CV) variations on the obtained ESI-MS spectrum (Figure 3.83). Usually the species observed in the gas phase are closely related to the actual species in solution, and lower CV values favour preservation of these species. In addition, it was observed in the literature that the ionic abundances of the fragment ions 209 Chapter 3 – Polynuclear Manganese Coordination Complexes produced in the gas phase grow with increasing CV values, while the abundance of the parent ions decreases with increasing CV values.67, 68 The signal attributed to the {Mn7} species in the mass spectrum of the reaction mixture that led to the crystallisation of 10 is independent on CV variations up to 75 V (Figure 3.83), and the relative ionic abundance of the species start to decrease in intensity at CV greater than 100 V. The signal attributed to the {Mn8} species shows similar behaviour and the relative ionic abundance of this species starts to decrease in intensity at 75 V, due to the induced defragmentation processes. These suggests that {Mn7}, and {Mn8} species exist in solution, and we believe that these species represent building units of the pentadecanuclear manganese complex which crystallises out from this reaction mixture. The most probable structural arrangement for {Mn7} and {Mn8} species are represented in Figure 3.81. These structural motifs that agree with the constitutional assignment correspond to subunits of the {Mn15} cluster, and are further supported by the experimental observation of the same {Mn7} species in the mass spectra of pristine crystals of 10 dissolved in CH3CN, DMF and DMSO. As earlier mentioned in section 3.3.1 of this chapter the {Mn15} cluster is not stable in solution and break into two {Mn7} units as a consequence of the labile methanol groups from the JT sites of the central Mn(8) ion, (see section 3.3.1, Figure 3.3). Another remarkable feature observed in the mass spectrum of the reaction mixture of 10 are signals that are undoubtably attributable to a {Mn13} species. Surprisingly CV variations experiments show that the ionic abundance of this species increases when CV values were increased from 10 to 75 V. This behaviour suggests that this species could indeed form in the spaying chamber and may not be structurally related to the {Mn15} structure in 10. Alternatively, it could be related to the {Mn13} clusters that crystallise form other reaction mixtures, such those of compounds 11, 12 and 13. Considerable fragmentation can be observed at higher cone voltages. The {Mn7}, {Mn8} and {Mn13} species cannot be identified at CV greater than 150 V, suggesting that the species no longer retain their integrity under these harsh ionisation conditions. 210 Chapter 3 – Polynuclear Manganese Coordination Complexes 3.4.2 Investigation of the reaction system that led to the formation of {Mn13} complexes observed in 11-13 Using a similar synthetic procedure as for 10 and adding pyridine as an organic base, a dark brown solution was obtained. The ESI-MS spectrum of this reaction mixture is much simpler (Figure 3.85) than the original mixture that led to the crystallisation of 10. It reveals the presence of only one major isotopic envelope in the high mass region centered at m/z = 2594.0 a.m.u. This signal was assigned to the phenylphosphonic acid-stabilised {H2[MnIV4MnIII9O11(C6H5PO3)10]Cl4}- species, which corresponds to the tridecanuclear manganese C 3 )4(C6 complex 5P 3)10(C5 characterised 5 in 11, [ n 13( 4- )2( 3- )4( 2- H)2( 2- )5Cl]·3H2O, that indeed crystallizes out from this reaction mixture (Table 3.32). Figure 3.84 – Crystal structure of the tridecanuclear manganese complex in 11. Colour code: MnIII blue, P purple, O red, Cl green, C grey (hydrogen atoms have been omitted for clarity). Table 3.32 – ESI-MS assignment for the reaction mixture that led to the crystallisation of 11. Reaction mixture of 11 m/z Species attributed 2594.0 {H2[MnIV4MnIII9O11(C6H5PO3)10]Cl4}- → {Mn13} 211 Relative ionic abundance Chapter 3 – Polynuclear Manganese Coordination Complexes Figure 3.85 – Negative-mode ESI-MS spectra for the reaction mixture that led to the crystallisation of 11 (24 h after preparation). Inset: Comparison of the experimental isotopic envelopes (black spectrum) with simulated patterns (red spectrum) for {H2[MnIV4MnIII9O11(C6H5PO3)10]Cl4}- centered at m/z = 2594.0 a.m.u. (cone voltage: 30 V). The ESI-MS technique proved to be very efficient in identifying new polynuclear manganese species that form under certain conditions in the reaction mixture, and therefore we decided to exploit the analytic screening technique further and investigate the influence of different ligands on the formation of manganese coordination clusters. When benzylphosphonic acid was used instead of phenylphosphonic acid under the same reaction conditions that led to 11, we obtained a reaction mixture from which compound 12, [ n 13( 4- )2( 3- )4( 2- H)2( 2-C 3 )4(C6 5C 2P 3)10(C5 5 )6]Cl ·5H2O separates over a time period of four days. Table 3.33 – ESI-MS assignment for the reaction mixture that led to the crystallisation of 12. m/z 1086.6 Reaction mixture of 12 1348.6 2138.3 2212.3 2699.1 Species attributed {H2[MnIV2MnIII11 {H[MnIV2MnIII11 13(C 10(C )4(C6 3 3 )4(C6 5C 5C 2P 2P 23)6]Cl3} → {Mn13} 23)10]} → {Mn13} {H9[MnIV2MnIII11O18(C6H5CH2PO3)6]Cl3}{H7[MnIV2MnIII11 {H[MnIV3MnIII10 15(C 10(C 3 3 )4(C6 )4(C6 5C 5C 2P 2P → {Mn13} 3)6]Cl3} → {Mn13} 3)10]} → {Mn13} 212 Relative ionic abundance Relative ionic abundance Relative ionic abundance Relative ionic abundance Relative ionic abundance Relative ionic abundance Chapter 3 – Polynuclear Manganese Coordination Complexes Figure 3.86 – Negative-mode ESI-MS spectra for the reaction mixture that led to the crystallisation of 12 (24 h after preparation). Comparison of the experimental isotopic envelopes (black spectrum) with simulated patterns (red spectrum) for {H2[ nIV2MnIII11 13(CH3 )4(C6 5CH2P 3)6]Cl3}2- centered at m/z = 1086.6 a.m.u., {H[ nIV2MnIII11 10(CH3 )4(C6 5CH2P 3)10]}2- centered at m/z = 1348.6 a.m.u., {H9[MnIV2MnIII11O18(C6H5CH2PO3)6]Cl3}centered at m/z = 2138.3 a.m.u., {H7[ nIV2MnIII11 15(CH3 )4(C6 5CH2P 3)6]Cl3}- centered at m/z = 2212.3 a.m.u. and {H[ nIV3MnIII10 10(CH3 )4(C6 5CH2P 3)10]}- centered at m/z = 2699.1 a.m.u. (cone voltage: 30 V). The ESI-MS spectrum of this reaction mixture is presented in Figure 3.86 showing an intense peak centered at m/z = 2138.3 a.m.u. which corresponds to the parent {H9[MnIV2MnIII11O18(C6H5CH2PO3)6]Cl3}- species. Two other singly charged and two other doubly charged {Mn13} species stabilized by a different number of phosphonate ligands, water molecules and chloride ions could be identified in the high mass region of the spectrum (Table 3.33). We successfully simulated the isotopic envelopes for these 213 Chapter 3 – Polynuclear Manganese Coordination Complexes identified species and good fits between experimental and theoretical data further substantiate our assignments (Figure 3.86). The same reaction conditions as for 12 but replacing the pyridine ligands with 4-(3phenylpropyl)pyridine gave a dark brown solution from which compound 13, [ n )2( 3- )4( 2- H)2( 2-C 3 )4(C6 5C 2P 3)10(C6 5-C3 6-C5 4 13( 4- )6]Cl·5H2O, was obtained. The mass spectrum of this reaction mixture recorded in negative mode is simple (Figure 3.87). It reveals the presence of an intense signal corresponding to a singly charged {Mn13} species and a much lower intensity signal arising from a doubly charged species. The ESI-MS peak assignment for the reaction mixture of 13 is presented in Table 3.34. Table 3.34 – ESI-MS assignment for the reaction mixture that led to the crystallisation of 13. m/z 1086.6 2138.3 {H2[MnIV2MnIII11 13(C 3 )4(C6 5C 2P 23)6]Cl3} {H9[MnIV2MnIII11O18(C6H5CH2PO3)6]Cl3}- → {Mn13} → {Mn13} Relative ionic abundance Reaction mixture of 13 Species attributed Figure 3.87 – Negative-mode ESI-MS spectra for the reaction mixture that led to the crystallisation of 13 (24 h after preparation). Inset: Comparison of the experimental isotopic envelopes (black spectrum) with simulated patterns (red spectrum) for {H2[ nIV2MnIII11 13(CH3 )4(C6 5CH2P 3)6] Cl3}2- centered at m/z = 1086.6 a.m.u. and {H9[MnIV2MnIII11O18(C6H5CH2PO3)6]Cl3}- centered at m/z = 2138.3 a.m.u. (cone voltage: 30 V). 214 Chapter 3 – Polynuclear Manganese Coordination Complexes 3.4.3 Investigation of the reaction system that led to the formation of {Mn12} complexes observed in 14-15 Preliminary studies on the coordination clusters 11 – 13 demonstrate that the ligands in apical position of the MnIII Jahn Teller sites are labile and can be successively replaced. This observation prompted us to explore if the tridecanuclear manganese clusters could be connected using bifunctional pyridine ligands. Therefore, 4,4`-bipyridine was employed as an organic base in a similar reaction to that used for the synthesis of the tridecanuclear manganese complexes. However, this approach did not result in the linkage of the tridecanuclear manganese clusters. However, under these conditions a {Mn12} coordination cluster was obtained instead. The ESI-MS spectrum of the reaction mixture that led to the crystallisation of 14, K(H2O)4[MnIII12( 3-O)6(CH3OH)6(C6H5CH2PO3H)7 (C6H5CH2PO3)8]Cl2·3CH3OH·4H2O (Figure 3.88), is relatively simple. Nevertheless, the resulting spectrum is interesting, as we were able to identify three different species that exist in the same reaction mixture: a {Mn7}, a {Mn12} and a {Mn13} species (Table 3.35, Figure 3.89). Figure 3.88 – Crystal structure of the dodecanuclear manganese complex in 14. Colour code: MnIII blue, P purple, O red, C grey, Cl green, K yellow (hydrogen atoms have been omitted for clarity). The most intense isotopic envelope in the high molecular mass region of the spectrum is centered at m/z = 1796.4 a.m.u. and corresponds to the {Mn7} species {H29[MnIII3MnII4O16(C6H5CH2PO3)6]Cl3}-. Two other significantly less intensive signals centered at m/z = 915.7 a.m.u. and m/z = 1920.1 a.m.u. could be observed in the high molecular mass region of the spectrum. These signals were attributed to a {Mn12} species, 215 Chapter 3 – Polynuclear Manganese Coordination Complexes {H6[MnIII4MnII8O11(C6H5CH2PO3)5]Cl4}2- and a {Mn13} species, {H6[MnIII7MnII6O13(C6H5CH2PO3)5]Cl4}-, respectively. The former species relates to the cluster of 14 that crystallises under these conditions, the latter is expected to be related to the cluster core structures of 11-13. Table 3.35 – ESI-MS assignment for the reaction mixture that led to the crystallisation of 14. Reaction Species attributed 915.7 {H6[MnIII4MnII8O11(C6H5CH2PO3)5]Cl4}2- → {Mn12} 1796.4 {H29[MnIII3MnII4O16(C6H5CH2PO3)6]Cl3}- → {Mn7} 1920.1 {H6[MnIII7MnII6O13(C6H5CH2PO3)5]Cl4}- → {Mn13} Relative ionic abundance Relative ionic abundance Relative ionic abundance Relative ionic abundance mixture of 14 m/z Figure 3.89 – Negative-mode ESI-MS spectra for the reaction mixture that led to the crystallisation of 14 (24 h after preparation). Comparison of the experimental isotopic envelopes (black spectrum) with simulated patterns (red spectrum) for {H6[MnIII4MnII8O11(C6H5CH2PO3)5]Cl4}2- centered at m/z = 915.7 a.m.u., {H29[MnIII3MnII4O16(C6H5CH2PO3)6]Cl3}- centered at m/z = 1796.4 a.m.u. and {H6[MnIII7MnII6O13(C6H5CH2PO3)5]Cl4}- centered at m/z = 1920.1 a.m.u. (cone voltage: 30 V). A summary of the ESI-MS assignment for the reaction mixture that led to the crystallisation of 14 is presented in Table 3.35, and a comparison of the experimental isotopic envelopes together with the simulated patterns for all identified species can be found in Figure 3.89. 216 − − Chapter 3 – Polynuclear Manganese Coordination Complexes Figure 3.90 – Negative-mode ESI-MS spectra for the reaction mixture that led to the crystallisation of 14 at different cone voltages: 10V, 20V, 30V, 50V and 75V, respectively. Cone voltage variations experiments show that the relative ionic abundance of {Mn7}, {Mn12} and {Mn13} species decreases with increasing CV values. These experimental observations indicating that the assigned species form in solution in the same reaction mixture (Figure 3.90). Furthermore, we believe that the {Mn7} species is structurally related to the {Mn13} core structure, adopting a hexagonal brucite arrangement as shown in Figure 3.91. 217 Chapter 3 – Polynuclear Manganese Coordination Complexes Figure 3.91 – Representation of a structural motif that agrees with the constitutional assignment for the {Mn7} species identified in the mass spectrum of the reaction mixture that led to the crystallisation of 14. Colour code: Mn blue, P purple, C grey, O light grey. 15, Compound K(H2O)4[MnIII12( 3-O)6(CH3OH)6(C6H5CH2PO3H)7 (C6H5CH2PO3)8]Br2·2CH3OH·2H2O was obtained by replacing the MnCl2·4H2O with MnBr2·4H2O in the reaction mixture of 14 in order to investigate the role of the halide ions as a potential templating agent. The ESI-MS study on the reaction mixture that led to the crystallisation of 15 is essentially analogous to that of 14. The mass spectrum displays characteristic signals at m/z = 740.0 a.m.u. and m/z = 1930.3 a.m.u. which correspond to {Mn3} and (Mn7} species, {HMnIV3O4(C6H5CH2PO3)3]}- and {H29[MnIII3MnII4O16(C6H5CH2PO3)6]Br3}-, respectively. Two other significantly less intense isotopic envelopes centered at m/z = 2032.3 a.m.u. and m/z = 2062.9 a.m.u. were attributed to the {Mn12} and {Mn13} species, {H[MnIII12O12(C6H5CH2PO3)6]Br2}- and {H3[MnIII6MnII7O11(C6H5CH2PO3)5]Br4}- (Table 3.36, Figure 3.92). Table 3.36 – ESI-MS assignment for the reaction mixture that led to the crystallisation of 15. m/z Species attributed 740.0 {H[MnIV3O4(C6H5CH2PO3)3]}- → {Mn3} Reaction 1930.3 {H29[MnIII3MnII4O16(C6H5CH2PO3)6]Br3}- → {Mn7} mixture of 15 2032.3 {H[MnIII12O12(C6H5CH2PO3)6]Br2}- → {Mn12} 2062.9 {H3[MnIII6MnII7O11(C6H5CH2PO3)5]Br4}- → {Mn13} 218 Relative ionic abundance Relative ionic abundance Relative ionic abundance Relative ionic abundance Relative ionic abundance Chapter 3 – Polynuclear Manganese Coordination Complexes Figure 3.92 – Negative-mode ESI-MS spectra for the reaction mixture that led to the crystallisation of 15 (24 h after preparation). Comparison of the experimental isotopic envelopes (black spectrum) with simulated patterns (red spectrum) for {H[MnIV3O4(C6H5CH2PO3)3]}- centered at m/z = 740.0 a.m.u., {H29[MnIII3MnII4O16(C6H5CH2PO3)6]Br3}- centered at m/z = 1930.3 a.m.u., {H[MnIII12O12(C6H5CH2PO3)6]Br2}centered at m/z = 2032.3 a.m.u. and {H3[MnIII6MnII7O11(C6H5CH2PO3)5]Br4}- centered at m/z = 2062.9 a.m.u. (cone voltage: 30 V). 219 Chapter 3 – Polynuclear Manganese Coordination Complexes 3.4.4 Investigation of the reaction system that led to the formation of the {Mn6} complex in 16 When the reaction mixture of compound 16, (H3O)4{MnIII2MnII4( 4-O)2 (H2O)2(CH3CN)2[(C6H5)3CPO3]6}Cl2·2CH3CN·4H2O, was examined by ESI-MS we obtained a spectrum in which one major isotopic isotope could be observed. Unfortunately, we were not able to assign this signal, but we succeeded to identify a {Mn6} species which relates to a considerably less intense signal centered at m/z = 2296.1 a.m.u. The signal corresponds to the hexanuclear complex in 16 which crystallizes from this reaction mixture (Table 3.37, Figure 3.94). Figure 3.93 – Crystal structure of the hexanuclear manganese complex in 16. Colour code: MnIII blue, MnII cyan, P purple, O red, C grey, N dark blue, Cl green (hydrogen atoms have been omitted for clarity). Table 3.37 – ESI-MS assignment for the reaction mixture that led to the crystallisation of 16. Reaction mixture of 16 m/z Species attributed 2296.1 {HMnIII2MnII4O2[(C6H5) 3CPO3]6}- → {Mn6} 220 Relative ionic abundance Chapter 3 – Polynuclear Manganese Coordination Complexes Figure 3.94 – Negative-mode ESI-MS spectra for the reaction mixture that led to the crystallisation of 16 (24 h after preparation). Inset: Comparison of the experimental isotopic envelopes (black spectrum) with simulated patterns (red spectrum) for {HMnIII2MnII4O2[(C6H5)3CPO3]6}- centered at m/z = 2296.1 a.m.u. (cone voltage: 30 V). 221 Chapter 3 – Polynuclear Manganese Coordination Complexes 3.5 CONCLUSION AND FUTURE WORK A facile synthetic approach to polynuclear manganese coordination complexes that involves comproportionation reactions between MnII and MnVII salts in the presence of organophosphonate ligands was developed. This synthetic approach led to the formation of three types of unprecedented manganese cluster compounds in 10, 14 and 15, three structurally related compounds 11, 12 and 13 that belong to a family of {Mn13} complexes containing a metal-centered cuboctahedral core structures and a hexanuclear, mixed valent coordination complex (16). We found that small changes to the reaction conditions result in drastic changes of the cluster nuclearity and topology. The nature of the phosphonate ligand and organic base are most influencial and determine the product formation. Our approach allowed us to utilize organophosphonate as organic, stabilizing ligands. In comparison to carboxylate ligands, there are only a few polynuclear manganese complexes stabilized by phenylphosphonic acid ligands reported in the literature, which makes the here reported class of compounds interesting. 10 contains a remarkable {Mn15} oxo-cluster with characteristic dunbell-type arrangement of its sub-units. Compound 11 was the result of adding pyridine to the reaction mixture of 10, while compound 12 was obtained by replacing the phenylphosphonic acid ligands in 11 with benzylphosphonic acid ligands. We noticed that the {Mn13} clusters contain active sites provided by ligands that reside in tetragonally elongated positions of the Jahn-Teller distorted, octahedral coordination environments. The kinetic lability of these sites gives rise to ligand exchange reactions and functionalisation of the compounds (e.g. introduction of organic moieties in these positions). In consecutive experiments, we further aimed to exploit this feature and connect the {Mn13} clusters using bi-functional ligands, e.g. bipyridine-type ligands, however, our attempts were unsuccessful. Studies of the magnetic properties reveal dominant antiferromagnetic interactions between the spin centres and a ground state spin values of ST = 0 for 10 and 16, ST = 3 for 14 and 15, ST = 6 for 11 and 13 and ST = 8 for 12. The ac susceptibility studies in zero dc field for compounds 11-13 reveal slow relaxation of the magnetisation highlighted by appearance of out-of-phase signals below 4 K (for frequency around 10000 Hz), indicating that these compounds are SMMs. 222 Chapter 3 – Polynuclear Manganese Coordination Complexes The thermal stability of the compounds was investigated by thermogravimeric analysis. Solution stability studies for 10 – 16 were performed using ESI-MS and UV-Vis spectroscopy. In particular, the ESI-MS technique proved to be highly valuble to charactterise the stability of the complexes. It is surprising that this technique has not previously been utilized to investigate the solution behavior of larger, polynuclear Mn complexes. Indeed we also applied ESI-MS to screen reaction mixtures prior crystallisation attempts of 10-16. These approaches aimed to investigate the real-time growth reactions of polynuclear Mn complexes and analyses of the magnetic properties of 10-16. The {Mn7} and the {Mn13} species were observed in the ESI-MS spectra of the reaction mixture that led to the crystallisation of 10, 14 and 15. The {Mn7} species may be a brucite-type, disk-like building unit which appears in several clusters reported in the literature and in tridecanuclear clusters in 11 - 13.69-72 The {Mn13} species identified by ESI-MS relates to the latter family of tridecanuclear manganese complexes which contain a metal-centered cuboctahedral topology. Another noticeable feature observed in the mass spectra of the reaction mixture that led to the crystallisation of 10, 14 and 15 is that the signals attributed to the {Mn7} species appear as the most intense signals in the high mass region of the spectra. This observation in combination with cone voltage variations suggest that this {Mn7} species is the predominant high molecular mass species present in these reaction mixtures. Consequently, one expects several cluster topologies to incorporate this building unit and forming larger assemblies involving for instance hydroxo- and oxogroups or donor atoms of multidentate ligands. This observation is substantiated by the crystal structures of 11-13 and several structures reported in the literature that contain the said brucite-type core.12, 48, 53-55 In addition the identification of trinuclear Mn species support this proposal. The {Mn12} species identified in the reaction mixture that led to the crystallisation of 14 and 15 seems to be in equilibrium with the {Mn13} species. This equilibrium can be manipulated in order to crystallise either the {Mn12} or the {Mn13} species by slight changes of the reaction conditions. In the case of the pentadecanuclear manganese complex in 10, we found that the {Mn15} cluster is not stable in solution since this species could not be identified by ESI-MS studies. Moreover, we were able to identify building units of the {Mn15} cluster as {Mn7} and {Mn8} species present in the reaction mixture that led to the crystallisation of 10. The fact that the {Mn7} species was also present in the mass spectra of pristine crystals of 10 dissolved in CH3CN, DMF and DMSO suggest that the {Mn15} cluster break into two {Mn7} units, most likely due to the labile methanol groups from the JT sites of the central Mn ion. 223 Chapter 3 – Polynuclear Manganese Coordination Complexes A summary of the ESI-MS assignment for the reaction mixtures of the {Mn15}, {Mn13} and {Mn12} systems, highlighting related structural motifs and possible building units, is presented in Figure 3.95. Future work aims to exploit the kinetic lability of the Jahn-Teller sites of the {Mn13} clusters for further functionalisation of these compounds. It could also be possible to connect the {Mn13} clusters using bi-functional ligands, which could lead to the formation of porous structures. Additionally, ESI-MS technique will be further exploited to characterise new polynuclear manganese coordination complexes. The extensive research in the field of molecular magnetism is not only concerned with the understanding of the magnetic properties of these materials, and providing structure-property relationships but also with the real need for the development of rational synthetic approaches to complex polynuclear species. Currently it is rather challenging to predict what kind of cluster topologies will be obtained and which would be desired to achieve exciting magnetic properties. Under this prevue mass spectrometry analyses might provide a step towards more rational approaches to more predictable, structurally related materials that will help to understand and improve the properties of molecule magnetic materials. 224 Chapter 3 – Polynuclear Manganese Coordination Complexes Reaction mixture that led to the crystallisation of the following species; X-Ray crystal structure ESI-MS assigned species Structural motifs associated with the ESIMS assignment {Mn7} {Mn8} {Mn15} {Mn13} {Mn13} {Mn13} {Mn3} {Mn7} {Mn12} {Mn13} {Mn12} Figure 3.95 – Summary of the ESI-MS assignment for the reaction mixture of the {Mn15}, {Mn13} and {Mn12} systems, showing some related structures and possible building units. 225 Chapter 3 – Polynuclear Manganese Coordination Complexes REFERENCES 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. Y. Umena, K. Kawakami, J.-R. Shen and N. Kamiya, Nature, 2011, 473, 55-60. J. S. Kanady, E. Y. Tsui, M. W. Day and T. Agapie, Science, 2011, 333, 733-736. A. W. Rutherford and A. Boussac, Science, 2004, 303, 1782-1784. R. K. Hocking, R. Brimblecombe, L.-Y. S. Chang, Archana, M. H. Cheah, C. Glover, W. H. Casey and L. Spiccia, Nature Chemistry, 2011, 6, 461-466. K. N. Ferreira, T. M. Iverson, K. Maghlaoui, J. Barber and S. Iwata, Science, 2004, 303, 1831-1838. G. E. Kostakis, A. M. Ako and A. K. Powell, Chemical Society Reviews, 2010, 39, 2238-2271. T. C. Stamatatos, V. Nastopoulos, A. J. Tasiopoulos, E. E. Moushi, W. Wernsdorfer, G. Christou and S. P. Perlepes, Inorganic Chemistry, 2008, 47, 10081-10089. T. C. Stamatatos, K. A. Abboud, W. Wernsdorfer and G. Christou, Angewandte Chemie International Edition, 2007, 46, 884-888. M. Murugesu, M. Habrych, W. Wernsdorfer, K. A. Abboud and G. Christou, Journal of the American Chemical Society, 2004, 126, 4766-4767. J. R. Long, Chemistry of Nanostructured Materials, 2003, 291-315. G. Christou, Polyhedron, 2005, 24, 2065-2075. Y.-S. Ma, Y. Song, Y.-Z. Li and L.-M. Zheng, Inorganic Chemistry, 2007, 46, 5459-5461. A. Clearfield and K. Demadis, Royal Society of Chemistry, 2011. M. Shanmugam, M. Shanmugam, G. Chastanet, R. Sessoli, T. Mallah, W. Wernsdorfer and R. E. P. Winpenny, Journal of Materials Chemistry, 2006, 16, 2576-2578. M. Wang, C. Ma, H. Wen and C. Chen, Dalton Transactions, 2009, 994-1003. T. Glaser, Chemical Communications, 2011, 47, 116-130. R. Jirásko and M. Hol apek, Mass Spectrometry Reviews, 2011, 30, 1013-1036. J. R. Johansson and B. Norden, Proceedings of the National Academy of Sciences of the United States of America, 2012, 109, 2186-2187. D. Schaumlöffel and A. Tholey, Analytical and Bioanalytical Chemistry, 2011, 400, 1645-1652. C. I. Onet, L. Zhang, R. Clérac, J. B. Jean-Denis, M. Feeney, T. McCabe and W. Schmitt, Inorganic Chemistry, 2010, 50, 604-613. E. F. Wilson, H. Abbas, B. J. Duncombe, C. Streb, D.-L. Long and L. Cronin, Journal of the American Chemical Society, 2008, 130, 13876-13884. E. F. Wilson, H. N. Miras, M. H. Rosnes and L. Cronin, Angewandte Chemie International Edition, 2011, 50, 3720-3724. L. Vila-Nadal, S. G. Mitchell, A. Rodriguez-Fortea, H. N. Miras, L. Cronin and J. M. Poblet, Physical Chemistry Chemical Physics, 2011, 13, 20136-20145. L. Vilà-Nadal, A. Rodríguez-Fortea, L.-K. Yan, E. F. Wilson, L. Cronin and J. M. Poblet, Angewandte Chemie International Edition, 2009, 48, 5452-5456. C. Ritchie, E. M. Burkholder, D.-L. Long, D. Adam, P. Kogerler and L. Cronin, Chemical Communications, 2007, 468-470. D. R. Boyd and G. Chignell, Journal of the Chemical Society, 1923, 123, 813-817. D. R. Boyd and F. J. Smith, Journal of the Chemical Society, Transactions, 1924, 125, 1477-1480. V. Chandrasekhar and P. Sasikumar, Dalton Transactions, 2008, 6475-6480. 226 Chapter 3 – Polynuclear Manganese Coordination Complexes 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. D. I. Brown, Oxford University Press, 2002, 26. I. B. Bersuker, Chambridge University Press, 2006, 616. B.-H. Ye, I. D. Williams and X.-Y. Li, Journal of Inorganic Biochemistry, 2002, 92, 128-136. S. B. Yu, S. J. Lippard, I. Shweky and A. Bino, Inorganic Chemistry, 1992, 31, 3502-3504. D. Coucouvanis, R. A. Reynolds and W. R. Dunham, Journal of the American Chemical Society, 1995, 117, 7570-7571. B.-H. Ye, T. Mak, I. D. Williams and X.-y. Li, Chemical Communications, 1997, 1813-1814. C. A. Hunter and J. K. M. Sanders, Journal of the American Chemical Society, 1990, 112, 5525-5534. R. Lei, X. Chai, H. Mei, H. Zhang, Y. Chen and Y. Sun, Journal of Solid State Chemistry, 2010, 183, 1510-1520. M. M. Gómez-Alcántara, A. Cabeza, M. A. G. Aranda, A. Guagliardi, J. G. Mao and A. Clearfield, Solid State Sciences, 2004, 6, 479-487. N. Stock and T. Bein, Journal of Solid State Chemistry, 2002, 167, 330-336. A. Cabeza, M. A. G. Aranda, S. Bruque, D. M. Poojary and A. Clearfield, Materials Research Bulletin, 1998, 33, 1265-1274. N. Stock and T. Bein, Journal of Materials Chemistry, 2005, 15. O. Kahn, VCH: New York, 1993, 380. R. K. Seidler-Egdal, A. Nielsen, A. D. Bond, M. J. Bjerrum and C. J. McKenzie, Dalton Transactions, 2011, 40. J. Yoon, M. S. Seo, Y. Kim, S.-J. Kim, S. Yoon, H. G. Jang and W. Nam, Bulletin of the Korean Chemical Society, 2009, 30, 679-682. J. A. Lessa, A. Horn Jr, C. B. Pinheiro, L. L. Farah, M. N. Eberlin, M. Benassi, R. R. Catharino and C. Fernandes, Inorganic Chemistry Communications, 2007, 10, 863-866. W.-Y. Hsieh and S. Liu, Inorganic Chemistry, 2006, 45, 5034-5043. D. Ruiz-Molina, P. Gerbier, E. Rumberger, D. B. Amabilino, I. A. Guzei, K. Folting, J. C. Huffman, A. Rheingold, G. Christou, J. Veciana and D. N. Hendrickson, Journal of Materials Chemistry, 2002, 12. V. Pavlishchuk, F. Birkelbach, T. Weyhermüller, K. Wieghardt and P. Chaudhuri, Inorganic Chemistry, 2002, 41, 4405-4416. L. Zhang, R. Clérac, P. Heijboer and W. Schmitt, Angewandte Chemie International Edition, 2012, 51, 3007-3011. A. Eichhöfer, O. Hampe, S. Lebedkin and F. Weigend, Inorganic Chemistry, 2010, 49, 7331-7339. T. G. Carrell, E. Bourles, M. Lin and G. C. Dismukes, Inorganic Chemistry, 2003, 42, 2849-2858. W. S. Kijlstra, E. K. Poels, A. Bliek, B. M. Weckhuysen and R. A. Schoonheydt, The Journal of Physical Chemistry B, 1997, 101, 309-316. S. Kunju and G. Krishnan, PHI Learning Private Limited, 2010. C. Lampropoulos, C. Koo, S. O. Hill, K. Abboud and G. Christou, Inorganic Chemistry, 2008, 47, 11180-11190. A. Ferguson, K. Thomson, A. Parkin, P. Cooper, C. J. Milios, E. K. Brechin and M. Murrie, Dalton Transactions, 2007. M. Wang, C. Ma and C. Chen, Dalton Transactions, 2008. P. J. Larkin, Elsevier Inc, 2011. S. Bayari and S. Yurdakul, Spectroscopy Letters, 2000, 33, 475-483. 227 Chapter 3 – Polynuclear Manganese Coordination Complexes 58. 59. 60. 61. 62. 63. 64. 65. 66. 67. 68. 69. 70. 71. 72. T. C. Stamatatos, D. Foguet-Albiol, S. P. Perlepes, C. P. Raptopoulou, A. Terzis, C. S. Patrickios, G. Christou and A. J. Tasiopoulos, Polyhedron, 2006, 25, 1737-1746. A. R. Schake, J. B. Vincent, Q. Li, P. D. W. Boyd, K. Folting, J. C. Huffman, D. N. Hendrickson and G. Christou, Inorganic Chemistry, 1989, 28, 1915-1923. V. Ovcharenko, E. Fursova, G. Romanenko and V. Ikorskii, Inorganic Chemistry, 2004, 43, 3332-3334. M. A. Kiskin, I. G. Fomina, G. G. Aleksandrov, A. A. Sidorov, V. M. Novotortsev, Y. V. Rakitin, Z. V. Dobrokhotova, V. N. Ikorskii, Y. G. Shvedenkov, I. L. Eremenko and I. I. Moiseev, Inorganic Chemistry Communications, 2005, 8, 89-93. P. Karsten and J. Strahle, Acta Crystallographica Section C, 1998, 54, 1403-1406. M. A. Halcrow, W. E. Streib, K. Folting and G. Christou, Acta Crystallographica Section C, 1995, 51, 1263-1267. A. R. E. Baikie, A. J. Howes, M. B. Hursthouse, A. B. Quick and P. Thornton, Journal of the Chemical Society, Chemical Communications, 1986, 1587-1587. R. Wang, D. Yuan, F. Jiang, L. Han, S. Gao and M. Hong, European Journal of Inorganic Chemistry, 2006, 8, 1649-1656. D. Liu, Q. Zhou, Y. Chen, F. Yang, Y. Yu, Z. Shi and S. Feng, Dalton Transactions, 2010, 39, 5504-5508. W. Henderson and J. S. McIndoe, John Wiley & Sons, 2005, 292. V. A. Pashynska, M. V. Kosevich, H. V. d. Heuvel and M. Claeys, Rapid Communications in Mass Spectrometry, 2006, 20, 755-763. M. A. Bolcar, S. M. J. Aubin, K. Folting, D. N. Hendrickson and G. Christou, Chemical Communications, 1997, 1485 - 1486. T. Liu, B.-W. Wang, Y.-H. Chen, Z.-M. Wang and S. Gao, Zeitschrift für anorganische und allgemeine Chemie, 2008, 634, 778-783. T. C. Stamatatos, K. M. Poole, D. Foguet-Albiol, K. A. Abboud, T. A. O’Brien and G. Christou, Inorganic Chemistry, 2008, 47, 6593-6595. S.-Y. Chen, C. C. Beedle, P.-R. Gan, G.-H. Lee, S. Hill and E.-C. Yang, Inorganic Chemistry, 2012, 51, 4448-4457. 228 Chapter 4 – Supramolecular Coordination Networks 4. SUPRAMOLECULAR COORDINATION NETWORKS EMPLOYING TRIPHOSPHONATE LINKERS 229 Chapter 4 – Supramolecular Coordination Networks 4.1 INTRODUCTION The massive increase in energy demand, the depletion of the world`s fossil fuels combined with the undesired products of burning fossil fuels, such as carbon dioxide, has led to a need for alternative energy sources. A promising alternative fuel which has attracted much attention in the last few years is hydrogen, which is a clean energy carrier having the highest energy content per unit of weight of any known fuel.1, 2 Thus, hydrogen is shown to be the fuel of the future. But, before being used at large scale there are a few problems that need to be solved, such as finding a safe and cost effective storage medium for hydrogen.3-5 Considerable research efforts have been devoted to the development of innovative materials including high surface area adsorbents such as metal-organic frameworks for reversible hydrogen storage and carbon dioxide capture.6-8 Metal-organic frameworks (MOFs) represent a new class of porous materials with high surface areas and tunable pore sizes, which can result in the selective uptake of small molecules. These unique properties make MOFs interesting materials for potential applications in catalysis, gas storage and separation.7, 9-11 The most studied class of MOFs is represented by the carboxylate-based compounds, while the use of phosphonte ligands for the formation of framework structures is less common. The reason for this is that many metal phosphonates tend to form densely packed layered structures that are not porous; growth of single crystals with phosphonates represents a real challenge as they often precipitate rapidly as less ordered, insoluble phases; phosphonates have more possible ligating modes compared to carboxylates, making the coordination chemistry of phosphonates less predictable. Even though, carboxylate-based MOFs have high surface area and uniform pore size distribution, the large scale applications are hampered by their lack of stability. Phosphonates have shown to form stronger bonds with metal ions than carboxylates and therefore have great potential to form robust porous solids.12-14 In this chapter we describe the synthesis and characterisation of coordination networks employing mononuclear SBUs and triphosphonate linkers. 230 Chapter 4 – Supramolecular Coordination Networks 4.2 EXTENDED TRIPHOSPHONATE LIGANDS The triphosphonate ligands were synthesised by a modified Michaelis-Arbuzov reaction (Scheme 4.1, b).15, 16 The synthetic procedure involves a catalyzed phosphonylation of aryl halides by trialkyl phosphites in the presence of 1,3diisopropylbenzene as a solvent and at high temperature. Nickel(II) salts are usually used as catalysts.17 a) (RO)3P b) (RO)3P: + R' R1 (RO)2(R')P+ - O - R + X- P O NiBr2 (RO)3P+R' + X- X OR RO X R1 NiL2 (RO)2(R')PO + RX c) OR OH OR HO RO RO P P O P O Ni R1 Ester Hydrolysis RO OR P O OR Ni X R1 L L H+ R1 RX R1 RO OR P X = Cl, Br, I L = P(OEt)3 L Ni R1 L OR X L Scheme 4.1 – a) Mechanism of the Michaelis – Arbuzov reaction for the preparation of alkylphosphonates (R` = alkyl); b) Modified Michaelis – Arbuzov reaction for the preparation of arylphosphonates; c) Ester hydrolysis under acidic conditions.15-17 The reaction mechanism involves in situ reduction of the nickel(II) halide to a nickel(0) phosphite complex, followed by the formation of an aryl nickel(II) intermediate. This intermediate is formed by an oxidative insertion of the nickel(0) phosphite complex into the carbon-halogen bond of the aryl halide. Then, an aryl(alkoxy)phosphonium salt is formed which rearranges to give the desired phosphonate ester. The major by-product of the reaction is an alkyl halide. It was observed that in order to obtain better yields one needs to remove the generated alkyl halide continuously from the reaction mixture until the reaction is complete. The phosphonate ester is then converted to the corresponding phosphonic acid by ester hydrolysis under acidic conditions.15, 16 The synthetic approach used to obtain phosphonic acids is presented in Scheme 4.1, a - c. 231 Chapter 4 – Supramolecular Coordination Networks The halogenated triarylbenzenes, used as starting materials for the preparation of the triphosphonate ligands, were synthesised through a triple condensation reaction according to literature procedures.18, 19 A reaction mechanism for the formation of triarylbenzenes was proposed by A. Goel et al.20 As shown in Scheme 4.2, the protonation of acetophenone results in the formation of intermediates (a) and (b). Then subsequent reactions between the intermediates, and further dehydration and 6 -electrocyclisation processes afford the desired triarylbenzene ligand. (A) (a) (b) (a) (b) (a) (A) (A) Scheme 4.2 – Plausible reaction mechanism for the formation of triarylbenzene.20 232 Chapter 4 – Supramolecular Coordination Networks 4.2.1 Synthesis of 1,3,5-Tris(4-phosphonophenyl)benzene (P-TPB) 1,3,5-Tris(4-phosphonophenyl)benzene (P-TPB) was prepared by a modified Arbuzov reaction using the aryl halide tris(4-bromophenyl)benzene (Br-TPB) as an intermediate.17 The synthetic route to Br-TPB involves a triple condensation reaction using 4-bromoacetophenone as the starting material, in the presence of H2SO4 and K2S2O7 (Scheme 4.3).18 Br Br H2SO4, K2S2O7 3 - 3H2O O Br Br - 3EtBr 3P(OEt)3 NiBr2 PO3Et2 PO3H2 6HCl - 6EtCl H2O3P PO3H2 Et2O3P PO3Et2 Scheme 4.3 – Schematic representation for the formation of P-TPB. Br-TPB formed in high yields (80%), while the P-TPB ligand formed in a lower yield of 58%. The ligands were characterised by FT-IR and NMR spectroscopy, and mass spectrometry. The 1H NMR spectra of the Br-TPB and P-TPB ligands, including the assignment of the aromatic H-atoms of the triphenyl-benzene backbone moieties, can be seen in Figure 4.1 and Figure 4.2. The conversion of Br-TPB into P-TPB was confirmed by the presence of an intense absorption band at 1139 cm-1 in the IR spectrum, due to P-O 233 Chapter 4 – Supramolecular Coordination Networks stretching vibrations. In addition the signal at 13.7 ppm in the 31 P NMR spectrum is characteristic for the desired phosphonate ligand. A detailed description of the preparation procedures together with the analytical data of the ligands is presented in the experimental section of chapter 8. 3 1 Br 2 1 2 Br 3 Br Figure 4.1 – A section of the 1H NMR (400MHz, CDCl3) spectrum showing the assignment for Br-TPB. 1 H2O DMSO PO3H2 2 1 3 H2O3P 2 3 PO3H2 Figure 4.2 – The 1H NMR (400MHz, CDCl3) spectrum and assignment for P-TPB. 234 Chapter 4 – Supramolecular Coordination Networks 4.2.2 Synthesis of 1,3,5-Tris(4`-phosphonobiphenyl-4-yl)benzene (PTBB) 1,3,5-Tris(4`-phosphonobiphenyl-4-yl)benzene (P-TBB) is a novel triphosphonate ligand which was prepared by a similar method to that that led to P-TPB. In this case, 1,3,5-tris(4`-bromobiphenyl-4-yl)benzene (Br-TBB) was used as aryl halide reactant. It was synthesised from commercially available 4-acetyl-4`-bromobiphenyl via a triple condensation reaction.19 Br-TBB formed in about 50% yield, while in the next synthetic step P-TBB ligand formed in a yield of 66%. The synthetic route used for the preparation of the P-TBB ligand is presented in Scheme 4.4. Br Br CF3SO3H 3 - 3H2O O Br Br - 3EtBr 3P(OEt)3 NiBr2 PO3Et2 PO3H2 6HCl - 6EtCl H2O3P PO3H2 Et2O3P PO3Et2 Scheme 4.4 – Schematic representation for the formation of P-TBB. The Br-TBB and P-TBB ligands were characterised by FT-IR and NMR spectroscopy, and mass spectrometry. The conversion of Br-TBB into P-TBB was 235 Chapter 4 – Supramolecular Coordination Networks confirmed by FT-IR and 31 P NMR spectroscopy. The IR spectrum of the P-TBB ligand displays an intense absorption band at 1137 cm-1 due to P-O stretching vibrations, while the 31 P NMR spectrum reveals a characteristic signal at 12.6 ppm. The identity and structure of the P-TBB ligand was further characterised by X-ray diffraction measurements. Colourless crystals were obtained after re-crystallisation, upon dissolving the P-TBB ligand in DMF and layering the resulting solution with CH3CN (43% yield). The ligand crystallises in the monoclinic crystal system in the space group C 2/c. The crystal structure of the ligand is presented in Figure 4.3. The benzene rings of the PTBB molecule are tilted with respect to each other. The tilt angles of the aryl rings denoted as 2, 3, and 4 to the central benzene ring 1 (Figure 4.3) are 37.21(99)°, 42.59(99)° and 28.76(87)°, respectively, while the tilt angles of the aryl rings 5, 6, and 7 to the central benzene ring 1 are 86.85(94)°, 6.51(78)°, 57.86(77)°, respectively. In the solid state the P-TBB molecules are linked by weak - stacking interactions and hydrogen bonding to generate a honeycomb-like supramolecular structure that supports voids of about 2.5 nm in diameter. The solvent molecules within these voids were highly disordered and could not be modelled properly, thus the SQUEEZE routine of the PLATON program was applied to remove the contributions to the scattering from the solvent molecules. The honeycomb-like structure can be seen in the direction of the crystallographic c-axis and is presented in Figure 4.4. In the crystal structure the unit cell comprises of eight P-TBB molecules organised into dimers. The molecules in these dimer assemblies exhibit a face-to-face (A) and a head-to-tail (B) arrangement, creating an interdigitating packing arrangement which can be seen in the direction of the crystallographic a-axis (Figure 4.5). The dimers are stabilised by hydrogen bonding and interactions. Strong intermolecular H-bonding between O(1) and O(2) of two adjacent molecules of the dimer unit A are characterised by an interatomic distance of 2.466(16) Å.21 Face-to-face - interactions between rings 1, 3, 6 of two adjacent molecules in dimer A range from 3.722(12) Å to 3.891(12) Å and off-set - interactions of 3.663(10) Å and 3.828(13) Å can also be observed in dimer A.22 Dimer B is stabilised by off-set interactions of 3.958(12) Å and 3.977(13) Å. 236 Chapter 4 – Supramolecular Coordination Networks Figure 4.3 – The crystal structure of the P-TBB ligand. Colour code: P purple, O red, C black (H atoms have been removed for clarity). ~25 Å Figure 4.4 – The packing arrangement of the P-TBB ligand viewed in the direction of the crystallographic c-axis. Colour code: P purple, O red, C black, H grey. 237 Chapter 4 – Supramolecular Coordination Networks a) B B A A b) Figure 4.5 – (a) The unit cell content showing the face-to-face (A) and head-to-tail (B) arrangement of the molecules; (b) The packing arrangement of the P-TBB ligand viewed in the direction of the crystallographic a-axis. Colour code: P purple, O red, C black, H grey. 238 Chapter 4 – Supramolecular Coordination Networks Table 4.1 − Crystal data and structural refinement parameters for P-TBB. P-TBB Ligand Empirical formula a Molecular mass/g mol C42H33O9P3 -1 a Crystal colour/shape 3 Colourless / needle Crystal size/mm 0.30×0.15×0.15 Crystal system Monoclinic Space group C 2/c a/ Å 24.378(14) b/ Å 50.24(3) c/ Å 12.132(7) /º 90 /º 90.761(10) /º 90 3 V/ Å 14856(14) Z 8 Temperature (K) 150(2) -3 Density/Mg m 0.693 -1 Absorp. coef./mm 0.109 F(000) 3216 2 40 max/º Reflections collected 17967 Independent reflections 6933 [R(int) = 0.1149] Data / restraints / parameters 6933 / 27 / 328 2 a 774.62 S on F 0.809 R1, wR2 [I>2 (I)] 0.1303, 0.3031 R1, wR2 (all data) 0.2345, 0.3528 Largest diff. peak and hole/e.Å-3 0.590 and -0.431 Excluding solvate molecules 239 Chapter 4 – Supramolecular Coordination Networks 4.3 SYNTHESIS AND CHARACTERISATION OF COORDINATION NETWORKS EMPLOYING MONONUCLEAR INORGANIC SBUS AND TRIPHOSPHONATE LINKERS 4.3.1 (H3O){ΜnII(H2Ο)(CΗ3ΟΗ)2[C6H3(C6Η4PΟ3Η)3]}·xCH3OH·yH2O (17) Compound 17, (H3O){ n(H2 )(C 3 )2[C6H3(C6 4P 3 )3]}·xCH3OH· yH2O was prepared by the reaction of 1,3,5-tris(4-phosphonophenyl)benzene (P-TPB) with MnCl2·4H2O in MeOH. Rectangular yellow crystals of 17 were obtained during a time period of four days and characterised by single crystal X-ray diffraction measurements. 17 crystallises in the triclinic crystal system in the space group P , and consists of a coordination network containing a mononuclear MnII SBU. The Mn ion in 17 displays a distorted octahedral coordination environment formed by one oxygen atom from a coordinating water molecule, two oxygen donors arising from two CH3OH ligands and the remaining coordination sites are occupied by oxygen donors originating from three distinct phosphonate ligands (Figure 4.6). The bond distances between the Mn ion and the O donors originating from the organic phosphonate ligands vary between 2.105(4) Å – 2.185(4) Å, the Mn-Omethanol bond lengths for Mn(1)-O(2) and Mn(1)-O(6) are 2.216(4) Å and 2.225(4) Å, respectively, while the bond distance between the Mn ion and the O atom of the water molecule, Mn(1)-O(1), is 2.222(4) Å. The distorted nature of the octahedral coordination geometry of the MnII ion can be observed from the bond angles involving the metal centre. Indeed, the bond angles, O(3)Mn(1)-O(6), O(5)-Mn(1)-O(2) and O(4)-Mn(1)-O(1) of 174.53(16)°, 170.51(16)° and 165.38(16)°, respectively, deviate significantly from the ideal octahedral angle of 180°, whilst the bond angles O(3)-Mn(1)-O(1) and O(4)-Mn(1)-O(6), of 98.07(16)° and 81.57(16)°, respectively, show the greatest deviation from the ideal angle of 90°. Selected bond distances and bond angles for compound 17 are summarised in Table 4.2 and Table 4.3, respectively. 240 Chapter 4 – Supramolecular Coordination Networks a) b) ~19 Å c) d) e) f) Figure 4.6 – (a) The hexagonal packing arrangement of the mononuclear MnII SBUs in 17 as seen in the direction of the crystallographic a-axis; (b) The coordination environment of the MnII ion; (c) Packing arrangement showing two layers with hexagonal motif viewed in the direction of the crystallographic aaxis; (d) - stacking of the organic moieties; (e) and (f) Packing arrangement of the 2D networks in 17 viewed in the direction of the crystallographic a- and c-axis, respectively. Colour code: Mn blue, P purple, O red, C dark grey, H light grey (crystallization solvent molecules have been omitted for clarity). 241 Chapter 4 – Supramolecular Coordination Networks Table 4.2 − Selected bond lengths [Å] for compound 17. Bond Bond distances (Å) Mn(1)-O(3) Mn(1)-O(4) Mn(1)-O(5) Mn(1)-O(2) Mn(1)-O(1) Mn(1)-O(6) P(2)-O(3) P(2)-O(9) P(2)-O(10) P(3)-O(4) P(3)-O(11) P(3)-O(12) P(4)-O(8) P(4)-O(5) P(4)-O(7) 2.105(4) 2.136(4) 2.185(4) 2.216(4) 2.222(4) 2.225(4) 1.490(4) 1.522(4) 1.575(4) 1.488(4) 1.542(4) 1.575(4) 1.516(5) 1.519(4) 1.578(5) Table 4.3 − Selected bond angles [º] for compound 17. Bond Angle (º) O(3)-Mn(1)-O(6) O(5)-Mn(1)-O(2) O(4)-Mn(1)-O(1) O(3)-Mn(1)-O(1) O(3)-Mn(1)-O(4) O(4)-Mn(1)-O(5) O(5)-Mn(1)-O(6) O(4)-Mn(1)-O(2) 174.53(16) 170.51(16) 165.38(16) 98.07(16) 96.52(16) 95.82(15) 94.57(17) 93.26(16) Bond Angle (º) O(3)-Mn(1)-O(5) O(2)-Mn(1)-O(6) O(2)-Mn(1)-O(1) O(3)-Mn(1)-O(2) O(5)-Mn(1)-O(1) O(1)-Mn(1)-O(6) O(4)-Mn(1)-O(6) 90.72(15) 89.50(18) 86.95(16) 85.48(18) 84.97(15) 83.82(16) 81.57(16) All three phosphonate groups of the P-TPB ligand are monodentate and singly protonated. The protonation state of the phosphonate groups was established by examining the P-O bond lengths (Table 4.2). Each phosphonate group of the organic ligands display an elongated P-O bond that range between 1.575(4) Å – 1.578(5) Å, and which is consistent with the values found in the literature for P-OH bonds.23-25 Thus each anionic ligand carries three negative charges which are compensated by one MnII ion and a H3O+ ion. Several coordination networks stabilised by hydronium counterions have previously been reported, e.g. in [MII6(C4O7)(C3HO5)3(H2O)6(H3O)] 8H2O (M = Zn or Mn), {(H3O)2[Mn4(4-Haba)2(4-aba)6(SCN)4(H2O)2]} (4-Haba = 4-aminobenzoic acid), 242 Chapter 4 – Supramolecular Coordination Networks [H3O]2x[M(pyzdc)2]x (M = Mn, Cd; pyzdcH2 = 2,3-pyrazinedicarboxilic acid), (H3O)2[Cu3(tbip)3.5(H2O)2] H2O (H2tbip = 5-tert-butyl isophtalic acid).26-29 The connectivity of the organic ligands and the mononuclear MnII SBUs in 17 results a layered architecture, whereby the hexagonal topology extends parallel to the (011) plane (Figure 4.6, a). The layers stack in an ABAB fashion and can be described using the Schläfli symbols as having a honeycomb {6,3} topology.30 The two-dimensional honeycomb-like net results voids of almost 2 nm in diameter and is stabilised in the crystal structure by week hydrogen bonding and - interactions between different layers that stack off-set in the direction of the crystallographic a-axis. The hydrogen bonding interactions involve the coordinating and constitutional solvent molecules, but a detailed analysis of these interactions was not possible due to the disordered nature of the constitutional solvent molecules and the low quality of the obtained data-set of the crystal structure analysis. The - interactions observed in the crystal structure of 17 involve the organic moieties of different layers and range between 3.795(25) Å – 4.095(54) Å. The packing arrangement of the mononuclear MnII SBUs in 17 and the network topology viewed in the direction of the crystallographic a- and c-axis are shown in Figure 4.6. 243 Chapter 4 – Supramolecular Coordination Networks - PHYSICOCHEMICAL CHARACTERISATION - FT-IR spectroscopy The IR spectrum of 17 shown in Figure 4.7 reveals some characteristic stretches of the organophosphonate ligands. The set of bands in the region 1600 – 1350 cm-1 arise from C–C skeletal vibrations of the phenyl rings, while the bands in the region 1200 – 900 cm-1 are attributed to different P–O stretching vibrations of the phosphonate groups. The two sharp bands observed between 820 – 650 cm-1 arise from C–H out-of-plane bending vibrations of the aromatic rings. Then, the corresponding O–H stretching vibrations and H– O–H bending vibrations of the crystallization water molecules engaged in H-bonds appear as broad bands at ca. 3200 cm-1 and ca. 1650 cm-1, respectively.31-35 Figure 4.7 – Infrared spectrum of 17. 244 Chapter 4 – Supramolecular Coordination Networks - Thermogravimetric analysis The thermal stability of compound 17 was investigated by thermogravimetric analysis (TGA) using a freshly prepared crystalline sample. The analysis was carried out in the temperature range between 30 and 900 °C, in an N2 atmosphere. Upon heating 17 undergoes a weight loss of 25.1 % below 200 °C. This weight loss can be attributed to the loss of two crystallization CH3OH molecules, three crystallization H2O molecules, two coordinating CH3OH molecules and one coordinating H2O molecule (calcd: 24.5 %). Other crystallization solvent molecules of 17 may have been lost prior to the TGA. The compound then undergoes a gradual thermal decomposition between 200 – 500 °C that can be associated with the degradation of the organic ligands. The thermogravimetric step centered at 550 °C is, most likely, being associated with the formation of a metal oxide material. In addition it is expected that under the applied N2 atmosphere, amorphous carbon is formed. 100 Weight % (%) 90 80 70 60 50 40 30 0 200 400 600 800 1000 Temperature (ºC) Figure 4.8 – Thermogravimetric analysis of 17. 245 Chapter 4 – Supramolecular Coordination Networks Table 4.4 − Crystal data and structural refinement parameters for 17. Compound 17 Empirical formula a Molecular mass/g mol C26H31MnO13P3 -1 a Crystal colour/shape Yellow / rectangular Crystal size/mm3 0.50×0.20×0.10 Crystal system Triclinic Space group P a/ Å 7.7458(15) b/ Å 16.352(3) c/ Å 16.975(5) /º 112.58(3) /º 100.65(3) /º 100.90(3) 3 V/ Å 1869.2(7) Z 1 Temperature (K) 150(2) -3 Density/Mg m 1.358 -1 Absorp. coef./mm 0.547 F(000) 756 2 50 max/º Reflections collected 28825 Independent reflections 6578 [R(int) = 0.2102] Data / restraints / parameters 6578 / 0 / 464 2 a 699.37 S on F 1.027 R1, wR2 [I>2 (I)] 0.0802, 0.1962 R1, wR2 (all data) 0.1260, 0.2678 Largest diff. peak and hole/e.Å-3 0.948 and -1.329 Excluding solvate molecules 246 Chapter 4 – Supramolecular Coordination Networks 4.3.2 (H3O){Cu(H2Ο)2(CΗ3ΟΗ)[C6H3(C6Η4PΟ3Η)3]}·xCH3OH·yH2O (18) Compound 18, (H3O){Cu(H2 )2(C 3 )[C6H3(C6 4P 3 )3]}·xCH3OH· yH2O was prepared by the reaction of P-TPB with CuCl2·4H2O in MeOH. Needle-like green crystals of 18 were obtained during a time period of five days and characterised by single crystal X-ray diffraction measurements. 18 crystallises in the triclinic crystal system in the space group P , and is structurally closely related to 17. It consists of a coordination network containing mononuclear CuII SBUs, in which the CuII ion displays a distorted octahedral coordination environment. The coordination sphere of the Cu ion consists of two oxygen atoms originating from two coordinating water molecules, one oxygen donor arising from a CH3OH ligand and the remaining coordination sites are occupied by monodentate oxygen donors originating from three distinct phosphonate ligands (Figure 4.9). The deviation of the CuII ion from the ideal octahedral coordination geometry can be visualised by examining the bond angles and bond lengths of the Cu(II) centre. The bond distances between the Cu ion and the O donors originating from the organic phosphonate ligands vary between 1.897(6) Å – 1.962(5) Å. The three O donors originating from the organic ligands are located in the equatorial plane of the octahedra, along with a forth O donor of a water molecule (Cu(1)-O(1`) of 1.961(6) Å). The apical binding sites of the octahedral coordination polyhedron are occupied by one O atom from a water molecule and one O atom from a methanol molecule. The bond distance between the Cu ion and the O atom of the water molecule Cu(1)-O(8`) is 2.201(15) Å, and the Cu-Omethanol bond length, Cu(1)O(7`) is 2.209(11) Å. These two bonds involving the O atoms located in the apical position of the octahedra are slightly elongated when compared to those in the equatorial plane and which might be explained by the Jahn-Teller (JT) effect of d9 ion in octahedral coordination environment. The bond angles of the Cu ion also deviate slightly from those expected of an ideal octahedral geometry. The bond angles O(8`)-Cu(1)-O(7`), O(1`)Cu(1)-O(4) and O(1)-Cu(1)-O(7) of 178.1(7)°, 171.5(3)° and 168.5(3)°, respectively, deviate from the ideal octahedral angle of 180°, whilst the bond angles O(4)-Cu(1)-O(8`) and O(4)-Cu(1)-O(7`) of 101.9(5)° and 79.3(3)°, respectively, show the greatest deviation from the ideal angle of 90°. Selected bond distances and bond angles for compound 18 are summarised in Table 4.5 and Table 4.6, respectively. 247 Chapter 4 – Supramolecular Coordination Networks a) b) ~19 Å c) d) Figure 4.9 – (a) The hexagonal packing arrangement of the mononuclear CuII SBUs in 18 as seen in the direction of the crystallographic a-axis; (b) The coordination environment of the CuII ion; (c) and (d) Packing arrangement of the 2D networks in 18 viewed in the direction of the crystallographic aand b-axis, respectively. Colour code: Mn blue, P purple, O red, C dark grey, H light grey (crystallization solvent molecules have been omitted for clarity). Table 4.5 − Selected bond lengths [Å] for compound 18. Bond Cu(1)-O(1) Cu(1)-O(7) Cu(1)-O(1`) Cu(1)-O(4) Cu(1)-O(8`) Cu(1)-O(7`) P(1)-O(7) P(1)-O(9) P(1)-O(8) Bond distances (Å) 1.897(6) 1.926(6) 1.961(6) 1.962(5) 2.201(15) 2.209(11) 1.505(7) 1.510(9) 1.559(11) 248 Chapter 4 – Supramolecular Coordination Networks Table 4.6 − Selected bond angles [º] for compound 18. Bond Angle (º) O(8`)-Cu(1)-O(7`) O(1`)-Cu(1)-O(4) O(1)-Cu(1)-O(7) O(4)-Cu(1)-O(8`) O(7)-Cu(1)-O(4) O(7)-Cu(1)-O(7`) O(1)-Cu(1)-O(7`) O(1)-Cu(1)-O(7`) 178.1(7) 171.5(3) 168.5(3) 101.9(5) 94.8(3) 94.5(7) 94.4(7) 93.9(2) Bond Angle (º) O(1`)-Cu(1)-O(7`) O(1)-Cu(1)-O(8`) O(1`)-Cu(1)-O(8`) O(7)-Cu(1)-O(1`) O(1)-Cu(1)-O(1`) O(7)-Cu(1)-O(8`) O(4)-Cu(1)-O(7`) 92.2(4) 87.0(6) 86.6(6) 86.3(3) 86.2(3) 83.9(7) 79.3(3) Similar with 17, all three phosphonate groups of the organic ligands in 18 are monodentate and singly protonated, and the overall charge of the network is compensated by one H3O+ ion. In the solid state 18 displays a similar packing arrangement with 17, and builds up into a two-dimensional honeycomb-like net that adopts a {6,3} topology, to form a layered structure. The packing arrangement of the mononuclear CuII complexes in 18 viewed in the direction of the crystallographic a- and b-axis is shown in Figure 4.9. 249 Chapter 4 – Supramolecular Coordination Networks - PHYSICOCHMICAL CHARACTERISATION - FT-IR spectroscopy The IR spectrum of 18 shown in Figure 4.10 is almost identical with that of 17. Characteristic C–C skeletal vibrations and C–H out-of-plane bending vibrations of the phenyl rings can be observed in the IR spectrum between 1600 – 1350 cm-1 and 850 – 650 cm-1, respectively. Different P–O stretching vibrations of the phosphonate groups can be observed in the range 1200 – 900 cm-1, while the O–H stretching vibrations and H–O–H bending vibrations of the crystallization water molecules engaged in H-bonds appear as broad bands at ca. 3200 cm-1 and ca. 1650 cm-1, respectively.31-35 Figure 4.10 – Infrared spectrum of 18. 250 Chapter 4 – Supramolecular Coordination Networks - Thermogravimetric analysis The thermal stability of a freshly prepared crystalline sample of 18 was investigated in an N2 atmosphere, in the temperature range 30 – 900 °C (Figure 4.11). TGA reveals that 18 undergoes a similar decomposition pathway as 17. The first thermogravimetric step corresponding to 18.9 % is observed in the temperature range between 30 – 200 °C. This weight loss can be attributed to the loss of four crystallization H2O molecules, one coordinating CH3OH molecule and two coordinating H2O molecules (calcd: 18.3 %). Other crystallization solvent molecules of 18 may have been lost prior to the TGA. The oxidation of the organic ligands can be observed above 200 °C resulting, most likely, in the formation of a metal oxide material and amorphous carbon. 100 Weight % (%) 90 80 70 60 50 40 0 200 400 600 800 1000 Temperature (ºC) Figure 4.11 – Thermogravimetric analysis of 18. 251 Chapter 4 – Supramolecular Coordination Networks Table 4.7 − Crystal data and structural refinement parameters for 18. Compound 18 Empirical formula a Molecular mass/g mol C25H29CuO13P3 -1 a Crystal colour/shape Light green / needle Crystal size/mm3 0.60×0.20×0.20 Crystal system Triclinic Space group P a/ Å 7.4116(15) b/ Å 15.240(3) c/ Å 15.942(3) /º 105.45(3) /º 98.15(3) /º 95.68(3) 3 V/ Å 1700.4(6) Z 2 Temperature (K) 150(2) -3 Density/Mg m 1.443 -1 Absorp. coef./mm 0.849 F(000) 748 2 50 max/º Reflections collected 18080 Independent reflections 5964 [R(int) = 0.0287] Data / restraints / parameters 5964 / 0 / 409 2 a 693.95 S on F 1.029 R1, wR2 [I>2 (I)] 0.1171, 0.4288 R1, wR2 (all data) 0.1237, 0.4555 Largest diff. peak and hole/e.Å-3 1.749 and -1.129 Excluding solvate molecules 252 Chapter 4 – Supramolecular Coordination Networks 4.4 CONCLUSION AND FUTURE WORK In this chapter we describe that room temperature synthetic methods can be used to produce coordination networks that employ mononuclear, partially solvated SBUs and triphosphonate ligands as linkers. One of the aims of this project was to synthesise organophosphonate-based openframeworks. To accomplish this, extended triphosphonate ligands, 1,3,5-tris(4phosphonophenyl)benzene (P-TPB) and 1,3,5-tris(4`-phosphonobiphenyl-4-yl)benzene (PTBB) were employed. Metal organic frameworks in 17 and 18, containing mononuclear Mn(II) and Cu(II) SBUs assemble in the presence of the P-TPB ligand. The two structurally related compounds exhibit a 2D layered architecture with hexagonal topologies. These 2D nets pack in the solid state in an ABAB fashion. Due to the fact that the metal complexes are partially solvated one expects that the stability of the two-dimensional honeycomb-like net is limited. This is substantiated by the thermogravimetric analysis of the compounds. The packing structure of 17 and 18 is further characterised by week hydrogen bonding and interactions. BET analyses of 17 and 18 revealed almost negligible surface areas of samples that were treated under vacuum. The materials did not show any permanent porosity. During the course of the project, we designed an even more extended tri-functional phosphonate ligand, P-TBB. The ligand crystallises to generate a honeycomb-like supramolecular structure that has cross-sectional diameters of about 2.5 nm. We believe that this ligand has the potential to generate MOFs. However, up to now we were not able to unambiguously characterise new coordination compounds using this ligand. The low solubility of this ligand severely hampered the isolation of crystalline materials. In addition, synthetic approaches often resulted in precipitations. Future research efforts will be devoted to the construction of rigid threedimensional networks using this rather unique ligand and solvothermal synthesis approaches in less polar solutions might allow the crystallisation of the MOFs. 253 Chapter 4 – Supramolecular Coordination Networks REFERENCES 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. J. R. Bartels, M. B. Pate and N. K. Olson, International Journal of Hydrogen Energy, 2010, 35, 8371-8384. L. Peraldo Bicelli, International Journal of Hydrogen Energy, 1986, 11, 555-562. L. Zhou, Renewable and Sustainable Energy Reviews, 2005, 9, 395-408. http://www.hydrogenenergycenter.org/content.aspx?page_id=22&club_id=108367 &module_id=8616. A. Zuttel, Naturwissenschaften, 2004, 91, 157-172. J. Yang, A. Sudik, C. Wolverton and D. J. Siegel, Chemical Society Reviews, 2010, 39, 656-675. H.-C. Zhou, J. R. Long and O. M. Yaghi, Chemical Reviews, 2012, 112, 673-674. T. Y. Ma and Z. Y. Yuan, CORD Conference Proceedings, 2011, 4, 1407-1419. J. L. C. Rowsell and O. M. Yaghi, Microporous and Mesoporous Materials, 2004, 73, 3-14. S. T. Meek, J. A. Greathouse and M. D. Allendorf, Advanced Materials, 2011, 23, 249-267. R. J. Kuppler, D. J. Timmons, Q.-R. Fang, J.-R. Li, T. A. Makal, M. D. Young, D. Yuan, D. Zhao, W. Zhuang and H.-C. Zhou, Coordination Chemistry Reviews, 2009, 253, 3042-3066. S. Natarajan and P. Mahata, Current Opinion in Solid State and Materials Science, 2009, 13, 46-53. K. J. Gagnon, H. P. Perry and A. Clearfield, Chemical Reviews, 2012, 112, 10341054. V. Chandrasekhar, T. Senapati, A. Dey and S. Hossain, Dalton Transactions, 2011, 40, 5394-5418. Q. Yao and S. Levchik, Tetrahedron Letters, 2006, 47, 277-281. A. K. Bhattacharya and G. Thyagarajan, Chemical Reviews, 1981, 81, 415-430. Z. Wang, J. M. Heising and A. Clearfield, Journal of the American Chemical Society, 2003, 125, 10375-10383. J. Palomero, J. A. Mata, F. Gonzalez and E. Peris, New Journal of Chemistry, 2002, 26, 291-297. J. Lu, Y. Tao, M. D'Iorio, Y. Li, J. Ding and M. Day, Macromolecules, 2004, 37, 2442-2449. A. Kumar, M. Dixit, S. P. Singh, R. Raghunandan, P. R. Maulik and A. Goel, Tetrahedron Letters, 2009, 50, 4335-4339. T. Steiner, Angewandte Chemie International Edition, 2002, 41, 48-76. C. A. Hunter and J. K. M. Sanders, Journal of the American Chemical Society, 1990, 112, 5525-5534. J.-L. Song and J.-G. Mao, Journal of Solid State Chemistry, 2005, 178, 3514-3521. H. Li, L. Zhang, G. Li, Y. Yu, Q. Huo and Y. Liu, Microporous and Mesoporous Materials, 2010, 131, 186-191. J.-L. Song and J.-G. Mao, Journal of Molecular Structure, 2005, 740, 181-186. B. F. Abrahams, T. A. Hudson and R. Robson, Journal of Molecular Structure, 2006, 796, 2-8. R. Wang, D. Yuan, F. Jiang, L. Han, S. Gao and M. Hong, European Journal of Inorganic Chemistry, 2006, 2006, 1649-1656. 254 Chapter 4 – Supramolecular Coordination Networks 28. 29. 30. 31. 32. 33. 34. 35. L. Mao, S. J. Rettig, R. C. Thompson, J. Trotter and S. Xia, Canadian Journal of Chemistry, 1996, 74, 2413-2423. X.-F. Wang, L. Li, Y.-M. Kong and Y. Liu, Inorganic Chemistry Communications, 2012, 21, 72-75. V. A. Blatov, M. O'Keeffe and D. M. Proserpio, CrystEngComm, 2010, 12, 44-48. N. Stock and T. Bein, Journal of Solid State Chemistry, 2002, 167, 330-336. N. Stock and T. Bein, Journal of Materials Chemistry, 2005, 15. R. Lei, X. Chai, H. Mei, H. Zhang, Y. Chen and Y. Sun, Journal of Solid State Chemistry, 2010, 183, 1510-1520. M. M. Gómez-Alcántara, A. Cabeza, M. A. G. Aranda, A. Guagliardi, J. G. Mao and A. Clearfield, Solid State Sciences, 2004, 6, 479-487. A. Cabeza, M. A. G. Aranda, S. Bruque, D. M. Poojary and A. Clearfield, Materials Research Bulletin, 1998, 33, 1265-1274. 255 Chapter 5 – Experimental 5. EXPERIMENTAL 256 Chapter 5 – Experimental 5.1 MATERIALS AND METHODS 5.1.1 Reagents All chemicals and solvents were of reagent grade and purchased from SigmaAldrich Ltd., Fluka Chemica-Biochemica (U.K.), ABCR GmbH & Co. KG (Germany) or local solvent suppliers and used as received, unless otherwise stated. Water was deionised before use. 5.1.2 Elemental Analysis Elemental analyses were performed using an Exeter Analytical CE 440 housed at the analytical laboratory, UCD Belfield. 5.1.3 Infrared Spectroscopy Infrared spectra were recorded on a PerkinElmer Spectrum One FT-IR spectrometer using a universal Attenuated Total Reflectance (ATR) sampling accessory. Data were collected and processed using Spectrum v5.0.1 (2002 PerkinElmer Instrument LLC) software. The scan rate was 16 scans per minute with a resolution of 4 scans in the range 4000-600 cm-1. The following abbreviations were used to describe the intensities: vs, very strong; s, strong; m, medium; w, weak; vw, very weak; sh, shoulder; br, broad and vbr, very broad. 5.1.4 1 Nuclear Magnetic Resonance Spectroscopy H NMR, 13 C NMR and 31 P NMR spectra were recorded on a Bruker DPX 400 spectrometer operating at 400.13 MHz, 100.14 MHz and 161.98 MHz, respectively, by either Dr. John O’Brien or Dr. Manuel Rüther. Samples were carried out in deuterated solvents and are listed for each spectrum. Chemical shifts are reported in ppm and coupling constants in Hertz. Standard abbreviations for spectra: s, singlet; d, doublet; dd, doublet of doublets; t, triplet; qt quartet; q, quaternary; m, multiplet; br broad, J, coupling constant. 257 Chapter 5 – Experimental 5.1.5 Ultraviolet – Visible Spectroscopy UV/vis spectra were recorded in the range 200-800 nm on a Cary 300 Scan spectrophotometer at 20 ºC using quartz cells of 1 cm path length. 5.1.6 Thermogravimetric Analysis Thermogravimetric analysis was carried out using a Perkin Elmer Pyriss 1TGA, in an N2 atmosphere, using a ceramic crucible (ca. 2 mg sample; heating rate of 10 ºC/min; range 25-900 ºC). The instrument was calibrated to Ni and Fe standards in N2 atmosphere. 5.1.7 Single Crystal X-ray Diffraction X-ray analyses for crystals described in this report were performed by Dr. Tom McCabe, Dr. Lei Zhang, Dr. Nianyong Zhu or Dr. Wolfgang Schmitt with a Bruker SMART APEX CCD diffractometer, a Rigaku Saturn-724 diffractometer or a Bruker APEX2 Duo diffractometer. The Bruker SMART APEX CCD and the Rigaku Saturn-724 diffractometers utilise a graphite-monochromated Mo-K radiation ( = 0.71073 ) source, while the Bruker APEX2 Duo utilises two radiation sources, a Mo-K source and a high intensity Cu-K source ( = 1.5418 ) generated from a microfocus anode. This Cu source was especially useful for the analysis of small crystal samples. The omega and phi scans method was used to collect either a full sphere or hemisphere of data for each crystal with a detector to crystal distance of either 5 or 6 cm. The data sets collected from the Bruker SMART APEX CCD diffractometer were processed and corrected for Lorentz and polarisation effects using SMART1 and SAINT-PLUS2 software. The structures were solved using direct methods with the SHELXTL3 program package. The data sets from the Rigaku Saturn-724 diffractometer were collected using Crystalclear-SM 1.4.0 software. Data integration, reduction and correction for absorption and polarisation effects were all performed using the Crystalclear-SM 1.4.0 software.4 Space group determination, structure solution and refinement were obtained using the Crystalstructure ver.3.8 and the Bruker SHELXTL3 software. Datasets collected on the Bruker APEX2 Duo were processed and the structures solved using Bruker APEX v2011.8-0 software.5 258 Chapter 5 – Experimental All non-hydrogen atoms were refined anisotropically. Hydrogen atoms (excluding water) were assigned to calculated positions using a riding model with appropriately fixed isotropic thermal parameters. The SQUEEZE/PLATON program was used in the structural refinement in the case where there were a large number of solvent molecules disordered within the structure.6 5.1.8 X-Ray Powder Diffraction X-ray powder diffraction was performed using a Siemens D500 diffractometer with Cu-K 1 radiation 5.1.9 with a wavelength of = 1.54056 Å. Magnetic Measurements All magnetic measurements were performed and modelled by Prof. Rodolphe Clérac at the Centre de Recherche Paul Pascal (CRPP), Pessac, France. The magnetic susceptibility measurements were obtained with the use of a Quantum Design SQUID magnetometer MPMS-XL and susceptometer PPMS-9. The MPMS-XL magnetometer operates between 1.8 and 400 K for dc applied fields ranging from -7 to 7 T. Measurements were performed on polycrystalline samples. ac-Susceptibility measurements were performed with an oscillating ac field of 3 Oe with a frequency between 1 to 1500 Hz or 1 Oe with a frequency between 10 to 10000 Hz (PPMS-9). The magnetic data were corrected for the sample holder and the diamagnetic contributions. 5.1.10 Mass Spectrometry Electrospray ionization (ESI) mass spectra were collected by Dr. J. Bernard JeanDenis and Dr. Martin Feeney, using a TOF-MS (Time-of-Flight – Mass Spectrometer, LCT Premier) instrument supplied by Waters Corp. All data, including simulated isotope patterns and molecular weight calculations, were processed using Masslynx v 4.0 (Waters Corp.) data analysis software.7 An injection of a standard sodium formate (Na[CHO2Na]n, 10% formic acid/0.1M NaOH/acetonitrile, 1:1:8, v/v/v) solution was used, for a mass calibration between m/z 100-3000. Samples were introduced into the MS via a waters alliance 2690 HPLC at a solvent flow rate of 200 L min-1, whilst a Leu-Enk solution (10 g mL-1 in acetonitrile/0.1 TFA in water, 50:50, v/v) was co-injected via a micropump at 2 259 Chapter 5 – Experimental L min-1 and used as an internal lock mass. The ESI settings were set with the nebuliser gas and desolvatation gas at 60 and 500 L h-1, respectively. The ESI gas used was nitrogen. Source and desolvatation gas temperature were set at 100 and 200 °C, respectively. The ion polarity for all MS scans were recorded in negative mode with voltage of the capillary tip set between 2.5-3 kV, sample cone at 30 V, extraction cone at 3 V, RF value set between 100-1000, m/z range set between 100-3000, scan time at 0.9 sec and inter-scan delay set at 0.1 sec. 5.2 LIGAND SYNTHESIS 5.2.1 Synthesis of Triphenylmethylphosphonic acid (P-TPM) Triphenylmethylphosphonic acid was prepared according to literature procedures.8-10 Triphenylcarbinol (10 g, 38 mmol) and PO3H2 phosphorus trichloride (10 g, 74 mmol) were added to an ice-cooled flask and kept in the ice bath for 30 min. After removal from the ice bath the reaction mixture was refluxed for one hour. A small quantity of diethyl ether was added to the resulting pale-yellow solid, followed by the addition of some ice-cooled deionised H2O. The formed product was collected by filtration, washed with aqueous ammonia and water and dried in a vacuum desiccator. In order to remove the triphenylchloromethane, which is produced along with triphenylmethoxyphosphorus dichloride, the crude product was boiled for 2 min in ethanol. The triphenylmethoxyphosphorus dichloride was further purified by recrystallizing from chloroform to which acetone was added (2:1 v:v mixture). Triphenylmethoxyphosphorus dichloride (4 g, 11 mmol) and 60 mL of 8% potassium hydroxide solution in ethanol was heated under reflux for 45 min. The reaction mixture was treated with 30 mL of water and the mixture was evaporated to dryness on a water bath. The solid residue was treated with 1 L of water and the mixture was made alkaline by the addition of a small quantity of potassium hydroxide. After this, the mixture was boiled, cooled and filtered. Then, some hydrochloric acid was added to the clear filtrate (till the solution became slightly acidic) and a white precipitate of 260 Chapter 5 – Experimental triphenylmethylphosphonic acid formed. After standing for 2 h, the precipitate was filtered, washed with 100 mL of water containing 1 mL of HCl and dried in a vacuum desiccator. The product was recrystallized from ethanol. Yield: ~ 55%. Mp: 235 °C. FT-IR (cm–1) νmax: 1597(w), 1492(m), 1445(m), 1178(m), 1158(m), 1086(w), 1031(s), 988(s), 968(s), 930(sh), 880(w), 758(s), 743(s), 698(s). 1H NMR (400 MHz, DMSO): 15H, C6H5). 13 C NMR (100 MHz, DMSO): (ppm) = 7.28 (m, (ppm) = 142.5 (3Cq, C6H5), 130.1 (6C, C6H5), 127.6 (6C, C6H5), 126.5 (3C, C6H5), 61.8 (Cq, C-Ph3). 31 P NMR (162 MHz, DMSO): (ppm) = 25.1. ESI-MS (DMSO): 347.0 m/z [C19H17O3PNa]+. 5.2.2 Synthesis of Tris(4-bromophenyl)benzene (Br-TPB) The Br Br-TPB ligand was prepared by a condensation reaction, according to literature methods.11 4-Bromoacetophenone (20 g, 10.050 mmol), 1 mL of H2SO4(c) and K2S2O7 (30 g, 118 mmol) were added to a round-bottom flask and heated at 180 °C for 14 h. The Br Br resulting crude solid was cooled to room temperature and refluxed in 100 mL of EtOH for 1 h. Then, the yellow solid was collected by filtration and refluxed again in 100 mL of H2O. The resulting product was dried under vacuum and purified by recrystallizing from chloroform. Yield: ~ 80%. Mp: 260 °C. FT-IR (cm–1) νmax: 3090(w), 1626(m), 1584(m), 1509(m), 1488(s), 1440(m), 1378(m), 1244(m), 1072(s), 1004(s), 885(w), 809(s), 698(m). 1H-NMR (400 MHz, CDCl3): (ppm) = 7.72 (s, 3H, C6H3), 7.63 (d, 6H, 3JHH = 8.5 Hz, C6H4), 7.55 (d, 6H, 3JHH = 8.5 Hz, C6H4). 13 C-NMR (100 MHz, CDCl3): (ppm) = 141.0 (3Cq, C6H3), 139.1 (3Cq, C6H4), 131.6 (6C, C6H4), 128.4 (6C, C6H4), 124.5 (3C, C6H3), 121.6 (3Cq, C6H4–Br). 261 Chapter 5 – Experimental 5.2.3 Synthesis of 1,3,5-Tris(4-phosphonophenyl)benzene (P-TPB) A slightly modified Arbuzov reaction was PO3H2 used for the diethylester of synthesis the of desired the phosphonate acid.12 Tris(4- bromophenyl)benzene (10.860 g, 20 mmol) in 50 mL of 1,3-diisopropylbenzene was heated to 185 °C H2O3P PO3H2 in a 250-mL three-necked round-bottom flask, fitted with a Dean Stark apparatus used in combination with a reflux condenser, N2 inlet, and an additional funnel. The flask was purged with N2, and 0.50 g of NiBr2 was added as a catalyst. Then, 10 mL triethyl phosphite was added over a 6 h period under a gentle stream of N2. The generated ethyl bromide was continuously removed from the reaction mixture until the reaction was complete. The mixture was further refluxed for 24 h followed by the addition of 0.25 g of NiBr2 and 5 mL of triethyl phosphite. After refluxing for another 24 h, the cooled black solution was distilled under vacuum to remove the solvent and unreacted triethyl phosphite. The resulting black solid was dissolved in CHCl3, and the product was extracted with petroleum ether to afford a viscous yellow oil after removing the solvent. This phosphonate diethyl ester was then converted to the corresponding triphosphonic acid by refluxing in HCl. The viscous residue resulted from removing the solvent was dried at high vacuum to obtain the product as an off-white, foam-like solid. Yield: ~ 58%. Mp: 300 °C. FT-IR (cm– 1 ) νmax: 2976(vbr), 1600(m), 1558(vw), 1387(w), 1139(s), 1042(s), 1016(s), 923(s), 818(s), 755(m), 680(s). 1H NMR (400 MHz, DMSO): 3 (ppm) = 8.02 (s, 3H, C6H3), 7.99 (m, 6H, JHH = 8.0 Hz, 4JPH = 2.5 Hz, C6H4), 7.80 (dd, 6H, 3JHH = 8.0 Hz, 3JPH = 12.5 Hz, C6H4- PO3H2). 13C NMR (100 MHz, DMSO): (ppm) = 142.0 (3Cq, C6H4), 141.0 (3Cq, C6H3), 134.2 (3Cq, C6H4-PO3H2), 131.1 (6C, C6H4), 127.0 (6C, C6H4), 125.1 (3C, C6H3). 31 P NMR (162 MHz, DMSO): (ppm) = 13.7. ESI-MS (H2O): 545.0 m/z [C24H20O9P3]-. 262 Chapter 5 – Experimental 5.2.4 Synthesis of 1,3,5-Tris(4`-bromobiphenyl-4-yl)benzene (Br-TBB) The Br-TBB ligand was prepared by a Br condensation procedure. 13 reaction, according to literature 4-Acetyl-4`-bromobiphenyl (4.400 g, 0.016 mol), trifluoromethanesulfonic acid (0.8 mL), and toluene (35 mL) were added to a 100 mL threenecked round-bottom flask fitted with a magnetic Br Br stirrer, a condenser, and a nitrogen inlet. The reaction solution was refluxed under nitrogen for 15 h at 120 °C. After cooling, the crude product was collected by filtration, washed with methanol, and purified by recrystallizing twice from chloroform. The product was washed again with plenty of methanol and the off white (slightly orange) powder was dried in the oven at 60 °C. Yield: ~ 48%. Mp: 280 °C. FT-IR (cm–1) νmax: 3033(w), 1650(m), 1584(m), 1479(m), 1441(w), 1386(m), 1274(m), 1219(m), 1195(m), 1075(m), 999(m), 952(w), 848(w), 807(s), 700(w), 669(w). 1H NMR (400 MHz, CDCl3): (ppm) = 7.89 (s, 3H, C6H3), 7.82 (d, 6H, 3JHH = 8.0 Hz, C6H4), 7.72 (d, 6H, 3JHH = 8.0 Hz, C6H4), 7.62 (d, 6H, 3JHH = 8.0 Hz, C6H4), 7.56 (d, 6H, 3JHH = 8.0 Hz, C6H4). 13C NMR (100 MHz, CDCl3): (ppm) = 141.4 (3Cq, C6H3), 138.8 (9Cq, C6H4), 131.5 (6C, C6H4-Br), 128.2 (6C, C6H4), 127.4 (6C, C6H4), 126.9 (6C, C6H4), 124.6 (3C, C6H3), 121.3 (3Cq, C6H4-Br). MALDI-TOF MS (DMSO): 767.9 m/z [C42H27Br3]. 5.2.5 Synthesis of 1,3,5-Tris(4`-phosphonobiphenyl-4-yl)benzene (P-TBB) The phosphonate diethylester of the desired PO3H2 acid was prepared using a slightly modified Arbusov reaction, and then converted to the corresponding triphosphonic acid by ester hydrolysis under acidic conditions.12 H2O3P PO3H2 1,3,5-Tris(4`-bromobiphenyl-4- yl)benzene (7.827 g, 10.104 mmol) and 35 mL of 1,3diisopropylbenzene were added to a 250 mL three263 Chapter 5 – Experimental necked round-bottom flask, fitted with a Dean-Stark apparatus used in combination with a reflux condenser, N2 inlet and an addition funnel. The solution was heated to 185°C, and 0.25 g of NiBr2 was added as a catalyst. The flask was purged with N2, and 7 mL triethyl phosphite was added over a 4 h period under a gentle stream of N2. The generated ethyl bromide was continuously removed from the reaction mixture until the reaction was complete. The mixture was further refluxed for 24 h followed by the addition of 0.125 g of NiBr2 and 3 mL of triethyl phosphite. After refluxing for another 24 h, the cooled black solution was distilled under vacuum to remove the solvent and unreacted triethyl phosphite. The resulted grey solid was then refluxed in HCl to convert the ester into the corresponding triphosphonic acid. CH3OH was added to aid dissolution. The solvent was then removed under vacuum, and the viscous residue was dried at high vacuum, which afforded the product as a brown solid. Yield: ~ 66%. Mp: 310 °C. FT-IR (cm–1) νmax: 3334(br), 2247(w), 1599(m), 1492(w), 1441(w), 1387(w), 1139(m), 989(s), 922(s), 810(s), 757(s), 659(s). 1H NMR (400 MHz, DMSO): NMR (100 MHz, DMSO): (ppm) = 8.05 (m, 9H), 7.87 (m, 18H). 13C (ppm) = 141.6 (3Cq, C6H4), 140.9 (3Cq, C6H3), 139.5 (6Cq, C6H4), 138.5 (3Cq, C6H4-PO3H2), 131.2 (6C, C6H4), 127.7 (12C, C6H4), 127.2 (6C, C6H4), 126.2 (3C, C6H3). 31P NMR (162 MHz, DMSO): (ppm) = 12.6. ESI-MS (CH3CN): 773.1 m/z [C42H32O9P3]-. 1,3,5-Tris(4`-phosphonobiphenyl-4-yl)benzene (P-TBB) (0.116 g, 0.150 mmol) was dissolved under stirring in DMF (5 mL), at room temperature, and 2 mL of CH3CN were layered to obtain needle-like colourless crystals overnight. Yield: ~ 43%. Anal. Calc. for C50H65N4O19P3, Expected: C% 53.67, H% 5.85, N% 5.01. Found: C% 53.40, H% 5.55, N% 5.15. FTIR (cm–1) νmax: 2929(w), 2321(br), 1655(s), 1596(m), 1493(m), 1437(m), 1410(m), 1386(s), 1253(m), 1137(s), 1092(s), 1000(s), 918(s), 814(s), 757(s), 658(s). 264 Chapter 5 – Experimental 5.3 SYNTHESIS OF THE METAL COMPLEXES 5.3.1 Synthesis of (NH4)2H2[MoV4O8(O3AsC6H5)4]∙5H2O (1) (NH4)6Mo7O24 4H2O (0.224 g, 0.180 mmol) and CH3COONH4 (0.500 g, 6.488 mmol) were dissolved under stirring in H2O (10 mL) at room temperature and N2H4 H2SO4 (0.032 g, 0.244 mmol) was added as a reducing agent. Then phenylarsonic acid (0.612 g, 3.030 mmol) was added and the resulting reaction mixture was stirred for 10 minutes. Subsequently, the pH was adjusted to pH = 3.9 (20°C) through the addition of 3.32 mL of aqueous CH3COOH solution (50%, v/v). After three weeks red-brown crystals were collected from a dark blue solution, and washed with cold water. Yield: ~ 57%. Anal. Calc. for As4C24H40Mo4N2O25, Expected: C% 20.01, H% 2.79, N% 1.94. Found: C% 19.65, H% 2.15, N% 1.69. FT-IR (cm–1) νmax: 2981(b), 1641(w), 1439(m), 1092(m), 966(vs), 743(vs). 1 H NMR (400 MHz, DMSO): C6H5). UV/Vis (DMF): 5.3.2 max (ppm) = 7.67 (m, 12H, C6H5), 7.97 (d, 3JHH = 7.0 Hz, 8H, ( ) = 457 nm (1000 L mol-1 cm-1). Synthesis of (NH4)2H2[MoV4O8(O3AsC6H4NH2)4]∙DMF∙4H2O (2) N2H4 H2SO4 (0.032 g, 0.244 mmol) was added to a mixture of (NH4)6Mo7O24 4H2O (0.224 g, 0.180 mmol), CH3COONH4 (0.500 g, 6.488 mmol), H2O (5 mL) and DMF (5 mL). The solution was vigorously stirred for 5 min and (4-aminophenyl)arsonic acid (0.659 g, 3.030 mmol) was added. The resulting reaction mixture was further stirred for 10 min and subsequently the pH was adjusted to pH = 4.9 (20°C) through addition of an aqueous CH3COOH solution (50%, v/v). Red-brown crystals were obtained after 5-7 days. Yield: ~ 53%. Anal. Calc. for As4C27H49Mo4N7O25, Expected: C% 20.85, H% 3.17, N% 6.30. Found: C% 21.14, H% 2.61, N% 6.23. FT-IR (cm–1) νmax: 3329(m), 3209(m), 1714(w), 1628(m), 1594(s), 1502(m), 1425(m), 1295(w), 1182(w), 1095(m), 958(s), 768(vs), 692(s). 1H NMR (400 MHz, DMSO): (ppm) = 5.80 (s, 8H, NH2), 6.73 (d, 3JHH = 8.5 Hz, 8H, C6H5), 7.56 (d, 3JHH = 8.5 Hz, 8H, C6H5). UV/Vis (DMF): max ( ) = 460 nm (1000 L mol-1 cm-1). 265 Chapter 5 – Experimental 5.3.3 Synthesis of (NH4)5∙[MoVI2MoV3O11(O3AsC6H4OH)5]∙9H2O (3) (NH4)6Mo7O24 4H2O (0.224 g, 0.180 mmol) and CH3COONH4 (0.500 g, 6.488 mmol) were dissolved under stirring in H2O (10 mL) at room temperature and N2H4 H2SO4 (0.032 g, 0.244 mmol) was added as a reducing agent. Then (4-hydroxyphenyl)arsonic acid (0.660 g, 3.030 mmol) was added and the resulting reaction mixture was stirred for 10 minutes. Subsequently the pH was adjusted to pH = 4.0 (20°C) through addition of 3.32 mL of an aqueous CH3COOH solution (50%, v/v). Dark blue crystals were collected after a week, and washed with cold water. Yield: ~ 53%. Anal. Calc. for As5C30H63Mo5N5O40, Expected: C% 18.12, H% 3.19, N% 3.52. Found: C% 19.03, H% 2.70, N% 3.20. FT-IR (cm–1) νmax: 2998(vbr), 1582(m), 1500(w), 1425(s), 1253(m), 1171(w), 1092(s), 944(s), 729(s). UV/Vis (DMF): 5.3.4 max ( ) = 531 nm (2700 L mol-1 cm-1). Synthesis of (NH4)4∙[MoVI4O10(O3AsC6H3NO2OH)4]∙2H2O (4) Synthetic procedure as for 1, using (4-hydroxy-3-nitrophenyl) arsonic acid (0.797 g, 3.030 mmol) as ligand. The pH was adjusted to pH = 3.9 (20°C) through addition of 3.32 mL of an aqueous CH3COOH solution (50%, v/v) and orange crystals of (NH4)4[Mo4O10(O3AsC6H3NO2OH)4] 2H2O were obtained within three weeks. Yield: ~ 75%. Anal. Calc. for As4C24H36Mo4N8O36, Expected: C% 17.00, H% 2.14, N% 6.61. Found: C% 16.52, H% 2.06, N% 5.61. FT-IR (cm–1) νmax: 3189(br), 1614(m), 1570(w), 1532(w), 1407(m), 1355(w), 1327(m), 1247(w), 1158(m), 1099(m), 940(w), 810 (vs), 762(s). 5.3.5 Synthesis of (NH4)4∙H4{Mn [MoV6O12(OH)3(O3PC6H5)4]2}∙8H2O (5) MoO3 (0.135 g, 0.930 mmol), H3BO3 (0.096 g, 1.550 mmol) and H2O (10 mL) were heated to 90ºC and N2H4 H2O (0.075 mL) were added to obtain a blue clear solution. A second solution of phenylphosphonic acid (0.700 g, 4.420 mmol) in H2O (10 mL) was heated to 90 ºC and added to the first solution. The mixture was then stirred at 70 ºC for 10 min, and MnCl2 4H2O (0.140 g, 0.700 mmol) was added. After stirring for another 30 min at 70 ºC the red brown solution was kept at 45ºC and orange crystals were obtained within a time period of three days. Yield: ~ 30%. Anal. Calc. for C48H82MnMo12N4O62P8, Expected: C% 18.23, H% 2.61, N% 1.77. Found: C% 19.14, H% 2.23, N% 1.50. FT-IR 266 Chapter 5 – Experimental (cm–1) νmax: 3211(vbr), 1594(m), 1438(m), 1142(m), 1024(s), 937(vs), 741(s), 721(s), 691(s). UV/Vis (CH3CN): max ( ) = 305 nm (11000 L mol-1 cm-1), 443 nm (1050 L mol-1 cm-1). 5.3.6 Synthesis of (NH4)4∙H4{Fe[MoV6O12(OH)3(O3PC6H5)4]2}∙8H2O (6) Synthetic procedure as for 5, using FeCl3 6H2O (0.189 g, 0.700 mmol). Yield: ~ 40%. Anal. Calc. for C48H82FeMo12N4O62P8, Expected: C% 18.23, H% 2.61, N% 1.77. Found: C% 18.42, H% 2.02, N% 1.36. FT-IR (cm–1) νmax: 3216(vbr), 1595(w), 1438(m), 1142(s), 1022(s), 960(vs), 744(s), 720(s), 689(vs). UV/Vis (CH3CN): -1 -1 -1 max ( ) = 274 nm -1 (88000 L mol cm ), 440 nm (1300 L mol cm ). 5.3.7 Synthesis of (NH4)4∙H4{Co[MoV6O12(OH)3(O3PC6H5)4]2}∙8H2O (7) Synthetic procedure as for 5, using CoCl2 6H2O (0.166 g, 0.700 mmol). Yield: ~ 20%. Anal. Calc. for C48H82CoMo12N4O62P8, Expected: C% 18.21, H% 2.61, N% 1.77. Found: C% 18.42, H% 2.13, N% 1.69. FT-IR (cm–1) νmax: 3209(vbr), 1617(m), 1441(w), 1142(m), 1093(w), 1039(s), 930(vs), 761(w), 739(m), 721 (s), 701 (w), 686 (m). UV/Vis (H2O): max ( ) = 252 nm (30000 L mol-1 cm-1), 450 nm (2900 L mol-1 cm-1). 5.3.8 Synthesis of (NH4)4∙H4{Ni[MoV6O12(OH)3(O3PC6H5)4]2}∙8H2O (8) Synthetic procedure as for 5, using NiCl2 6H2O (0.166 g, 0.700 mmol). Yield: ~ 20%. Anal. Calc. for C48H82NiMo12N4O62P8, Expected: C% 18.21, H% 2.61, N% 1.77. Found: C% 17.83, H% 1.99, N% 1.52. FT-IR (cm–1) νmax: 3215(vbr), 1618(m), 1439(m), 1142(m), 1093(w), 1038(s), 935(ws), 739(s), 721(s), 700 (w), 687 (s). UV/Vis (H2O): max ( ) = 301 nm (14000 L mol-1 cm-1), 440 nm (3900 L mol-1 cm-1). 5.3.9 Synthesis of (NH4)4∙H4{Mg[MoV6O12(OH)3(O3PC6H5)4]2}∙8H2O (9) Synthetic procedure as for 5, using MgCl2 (0.028 g, 0.300 mmol). Yield: ~ 22%. Anal. Calc. for C48H82MgMo12N4O62P8, Expected: C% 18.41, H% 2.64, N% 1.78. Found: C% 17.71, H% 1.99, N% 1.55. FT-IR (cm–1) νmax: 2981(w), 1617(w), 1438(m), 1142(m), 267 Chapter 5 – Experimental 1021(s), 940(vs), 745(s), 718(s), 691(vs). UV/Vis (CH3CN): -1 -1 -1 max ( ) = 297 nm (12000 L -1 mol cm ), 450 nm (800 L mol cm ). 5.3.10 Synthesis of [ nΙΙΙ15(μ2-H2 )2(C 3 )16(C6 5 P ) ]Cl5·22CH3OH·8H2O (10) 3 20 MnCl2·4H2O (0.198 g, 1.000 mmol), KMnO4 (0.039 g, 0.250 mmol) and phenylphosphonic acid (0.158 g, 1.000 mmol) were dissolved in CH3OH (10 mL), stirred for 5h, filtered and left undisturbed for four days at room temperature. Then the reaction mixture was kept at ca. 2°C in a refrigerator for another four days, to obtain rectangular red-brown crystals. Yield: ~ 30%. Anal. Calc. for a dried sample with the expected formula C120H136Cl5Mn15O78P20, Expected: C% 32.41, H% 3.08. Found: C% 32.04, H% 2.68. FTIR (cm–1) νmax: 3266(vbr), 2944(w), 2835(w), 1628(br), 1487(w), 1438(m), 1124(s), 1032(s), 983(s), 750(s), 722(s), 693(s). UV/Vis (CH3CN): 1 max ( ) = 280 nm (34000 L mol- cm-1), 465 nm (6400 L mol-1 cm-1). 5.3.11 Synthesis (C5 5 of [ n 13( 4- )2( 3- )4( 2- H)2( 2-C 3 )4(C6 5P 3)10 )5 Cl]·3H2O (11) MnCl2·4H2O (0.198 g, 1.000 mmol), KMnO4 (0.039 g, 0.250 mmol) and phenylphosphonic acid (0.094 g, 0.600 mmol) were dissolved under stirring in a mixture of CH3OH/CH3CN (5/5 mL), at room temperature, and 1.2 mL of pyridine were added. The reaction mixture was stirred for 5h, filtered and left undisturbed for slow evaporation. Rhombic, red-brown crystals were obtained during a time period of one week. Yield: ~ 20%. Anal. Calc. for a dried sample with the expected formula C85H87ClMn13N5O45P10, Expected: C% 34.51, H% 2.96, N% 2.37, Cl% 1.20. Found: C% 34.74, H% 2.49, N% 2.17, Cl% 0.94. FT-IR (cm–1) νmax: 3503(w), 3055 (w), 1601(m), 1485(s), 1438(m), 1120(s), 1084(s), 1032 (sh), 1003 (sh), 971(s), 753(s), 721(m), 695(s). UV/Vis (CH3CN): max ( )= 270 nm (162000 L mol-1 cm-1), 446 nm (3400 L mol-1 cm-1). 268 Chapter 5 – Experimental 5.3.12 Synthesis of [ (C5 5 n 13( 4- )2( 3- )4( 2- H)2( 2-C 3 )4(C6 5C 2P 3)10 )6]Cl·5H2O (12) MnCl2·4H2O (0.198 g, 1.000 mmol), KMnO4 (0.039 g, 0.250 mmol) and benzylphosphonic acid (0.102 g, 0.600 mmol) were dissolved under stirring in a mixture of CH3OH/CH3CN (5/5 mL), at room temperature, and 0.5 mL of pyridine were added. The reaction mixture was stirred for 5h, filtered and left undisturbed for slow evaporation. Rectangular, brown crystals were obtained within four days. Yield: ~ 23%. Anal. Calc. for a dried sample with the expected formula C100H116ClMn13N6O47P10, Expected: C% 37.38, H% 3.64, N% 2.62. Found: C% 36.95, H% 3.13, N% 2.17. FT-IR (cm–1) νmax: 3423(br), 3028(w), 2918(w), 2815(w), 1601(m), 1495(m), 1446(m), 1409(w), 1240(m), 1192(w), 1133(sh), 1116(sh), 1066(sh), 982(s), 830(m), 788(m), 730(m), 693(s). UV/Vis (CH3CN): max ( ) = 257 nm (61000 L mol-1 cm-1), 443 nm (4500 L mol-1 cm-1). 5.3.13 Synthesis of [ (C6 5-C3 6-C5 4 n 13( 4- )2( 3- )4( 2- H)2( 2-C )6]Cl·5H2O (13) 3 )4(C6 5C 2P 3)10 MnCl2·4H2O (0.198 g, 1.000 mmol), KMnO4 (0.039 g, 0.250 mmol) and benzylphosphonic acid (0.102 g, 0.600 mmol) were dissolved under stirring in CH3OH (20 mL), at room temperature, and 1 mL of 4-(3-phenylpropyl)-pyridine was added. The reaction mixture was stirred for 5h, filtered and left undisturbed for slow evaporation. Rectangular, brown crystals were obtained within one week. Yield: ~ 37%. Anal. Calc. for a dried sample with the expected formula C154H176ClMn13N6O47P10, Expected: C% 47.16, H% 4.52, N% 2.14, Cl% 0.90. Found: C% 47.19, H% 4.16, N% 2.03, Cl% 1.08. FT-IR (cm–1) νmax: 3471(br), 3026(w), 2921(w), 1952(w), 1615(m), 1558(w), 1495(m), 1453(m), 1426(m), 1240(w), 1193(w), 1118(sh), 1067(s), 984(s), 828(m), 789(m), 730(m), 695(s). UV/Vis (CH3CN): max ( ) = 256 nm (41000 L mol-1 cm-1), 446 nm (3600 L mol-1 cm-1). 5.3.14 Synthesis of K(H2O)4[MnIII12( 3-O)6(CH3OH)6(C6H5CH2PO3H)7(C6H5 CH2PO3)8]Cl2·3CH3OH·4H2O (14) MnCl2·4H2O (0.198 g, 1.000 mmol), KMnO4 (0.019 g, 0.125 mmol) and benzylphosphonic acid (0.079 g, 0.450 mmol) were dissolved under stirring in CH3OH (20 mL), at room temperature, and 4,4`-bipyridine (0.015 g, 0.096 mmol) was added. The 269 Chapter 5 – Experimental reaction mixture was stirred for 5h, filtered and left undisturbed for slow evaporation. Rectangular, brown crystals were obtained within one week. Yield: ~ 38%. Anal. Calc. for a dried sample with the expected formula C105H140Cl2KMn12O65P15, Expected: C% 34.31, H% 3.84. Found: C% 34.29, H% 3.49. FT-IR (cm–1) νmax: 3029(vbr), 1602(w), 1495(m), 1454(m), 1406(w), 1245(w), 951(s), 781(m), 727(m), 675(s). UV/Vis (CH3CN): -1 -1 -1 max ( )= -1 257 nm (36000 L mol cm ), 488 nm (3200 L mol cm ). 5.3.15 Synthesis of K(H2O)4[Mn III 12( 3-O)6(CH3OH)6(C6H5CH2PO3H)7(C6H5 CH2PO3)8]Br2·2CH3OH·2H2O (15) MnBr2·4H2O (0.143 g, 0.500 mmol), KMnO4 (0.019 g, 0.125 mmol) and benzylphosphonic acid (0.158 g, 0.918 mmol) were dissolved in CH3OH (20 mL), at room temperature, and stirred for 5h. The resulting orange - brown solution was filtered and left undisturbed for slow evaporation. Small, brown crystals were obtained during a time period of ten days. Yield: ~ 26%. Anal. Calc. for a dried sample with the expected formula C105H136Br2KMn12O63P15, Expected: C% 33.82, H% 3.68. Found: C% 33.74, H% 3.32. FTIR (cm–1) νmax: 3417(br), 3028(w), 1602(m), 1495(m), 1453(m), 1407(w), 1256(w), 1071(s), 960(s), 831(m), 786(s), 730(s), 679(s). UV/Vis (CH3CN): max ( ) = 258 nm (52000 L mol-1 cm-1), 489 nm (4000 L mol-1 cm-1). (H3O)4[MnIII2MnII4( 4-O)2(H2O)2(CH3CN)2{(C6H5)3 CPO3}6]Cl2·2CH3CN·4H2O (16) 5.3.16 Synthesis of MnCl2·4H2O (0.099 g, 0.500 mmol), KMnO4 (0.039 g, 0.250 mmol) and triphenylmethylphosphonic acid (0.324 g, 1.000 mmol) were dissolved under stirring in a mixture of CH3CN/CH2Cl2 (10/10 mL), at room temperature. 1.2 mL of triethylamine was added and the reaction mixture was further stirred for 5h, filtered and left undisturbed for slow evaporation. Rectangular red-brown crystals were obtained within five days. Yield: ~ 16%. Anal. Calc. for a dried sample with the expected formula C116H113Cl2Mn6NO28P6, Expected: C% 54.52, H% 4.46, N% 0.55. Found: C% 54.14, H% 3.98, N% 0.25. FT-IR (cm–1) νmax: 3590(w), 3054(w), 1594(m), 1491(m), 1444(m), 1188(w), 1121(s), 1076(s), 1037(s), 1005(s), 959(s), 931(s), 894(w), 851(w), 745(s), 699(s). UV/Vis (CH3CN): max ( ) = 263 nm (87000 L mol-1 cm-1), 487 nm (900 L mol-1 cm-1). 270 Chapter 5 – Experimental 5.3.17 Synthesis of (H3O){ n(H2 )(C 3 )2[C6H3(C6 4P 3 )3]}·xCH3OH·yH2O (17) 1,3,5-Tris(4-phosphonophenyl)benzene (P-TPB) (0.081 g, 0.150 mmol) was reacted with MnCl2·4H2O (0.029 g, 0.150 mmol) in CH3OH (10 mL). The resultant pale orange solution was stirred at room temperature for 1 hour, and then it was left undisturbed for slow evaporation. During a time period of four days rectangular yellow crystals of 17 were separated from the solution. Yield: ~ 33%. Anal. Calc. for a dried sample with the expected formula C32H63MnO23P3, Expected: C% 39.88, H% 6.59. Found: C% 40.19, H% 6.94. FT-IR (cm–1) νmax: 3189(br), 2278(w), 1647(w), 1597(m), 1385(m), 1135(s), 1036(s), 908(s), 818(s), 689(s). 5.3.18 Synthesis of (H3O){Cu(H2 )2(C 3 )[C6H3(C6 4P 3 )3]}·xCH3OH·yH2O (18) 1,3,5-Tris(4-phosphonophenyl)benzene (P-TPB) (0.081 g, 0.150 mmol) was reacted with CuCl2·2H2O (0.050 g, 0.300 mmol) in CH3OH (10 mL). The resultant pale green solution was stirred at room temperature for 1 hour, and then it was left undisturbed for slow evaporation. During a time period of five days pale green crystals were separated from the solution. Yield: ~ 34%. Anal. Calc. for a dried sample with the expected formula C24H28CuO13.5P3, Expected: C% 41.84, H% 4.10. Found: C% 41.86, H% 3.62. FT-IR (cm– 1 ) νmax: 2943(vbr), 2285(w), 1597(m), 1557(w), 1505(w), 1385(w), 1135(s), 1003(s), 914(s), 817(s), 688(s). 271 Chapter 5 – Experimental REFERENCES 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. Bruker SMART, Version 5.629, 1997-2003, Bruker-Axs Inc. Bruker Saint-Plus, Version 6.22, 1997-2003, Bruker-Axs Inc. G. M. Sheldrick, Version 5.1, 1999, Bruker Axs-Inc. Rigaku Americas Corporation, Crystalclear-SM 1.4.0 software, 9009 New Trails Drive, The Woodlands, TX 77381, United States. Software Reference Manual, version 5.625; Bruker Analytical X-Ray Systems Inc.: Madison, WI, 2001. A. L. Spek, Journal of Applied Crystallography (Wiley-Blackwell), 2003, 36, 7. Masslynx, version 4.0, 2002, Waters Corp. D. R. Boyd and G. Chignell, J. Chem. Soc., 1923, 123, 813-817. D. R. Boyd and F. J. Smith, J. Chem. Soc., Trans., 1924, 125, 1477-1480. V. Chandrasekhar and P. Sasikumar, Dalton Transactions, 2008, 6475-6480. J. Palomero, J. A. Mata, F. Gonzalez and E. Peris, New Journal of Chemistry, 2002, 26, 291-297. Z. Wang, J. M. Heising and A. Clearfield, Journal of the American Chemical Society, 2003, 125, 10375-10383. J. Lu, Y. Tao, M. D'Iorio, Y. Li, J. Ding and M. Day, Macromolecules, 2004, 37, 2442-2449. 272 Appendix APPENDIX 273 Attached CD-ROM The attached CD-ROM contains electronic versions of various files for the structures presented in this thesis. The CIF folder contains crystallographic information files (*.cif) for the compounds. The file names correspond to the name of the compounds as presented in the thesis (e.g. file 1 corresponds to compound 1). The CIF files can be viewed using the Mercury program. The Mercury program is available to download for free from the CCDC website (www.ccdc.cam.ac.uk). The PDF folder contains a PDF file of the entire thesis. PDF files can be viewed using the Adobe Acrobat Reader program. The Adobe Acrobat Reader program is available to download for free from the Adobe website (www.adobe.com/products/acrobat/readermain.html).