cyclononanes: the extensive chemistry of fundamentally simple …378128/uq378128... · 2019. 10....

144
Accepted Manuscript Title: Cyclononanes: The extensive chemistry of fundamentally simple ligands Author: Lawrence R. Gahan PII: S0010-8545(15)30035-7 DOI: http://dx.doi.org/doi:10.1016/j.ccr.2015.11.011 Reference: CCR 112170 To appear in: Coordination Chemistry Reviews Received date: 21-8-2015 Revised date: 12-11-2015 Accepted date: 13-11-2015 Please cite this article as: L.R. Gahan, Cyclononanes: the extensive chemistry of fundamentally simple ligands, Coordination Chemistry Reviews (2015), http://dx.doi.org/10.1016/j.ccr.2015.11.011 This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

Upload: others

Post on 25-Jan-2021

2 views

Category:

Documents


0 download

TRANSCRIPT

  • Accepted Manuscript

    Title: Cyclononanes: The extensive chemistry offundamentally simple ligands

    Author: Lawrence R. Gahan

    PII: S0010-8545(15)30035-7DOI: http://dx.doi.org/doi:10.1016/j.ccr.2015.11.011Reference: CCR 112170

    To appear in: Coordination Chemistry Reviews

    Received date: 21-8-2015Revised date: 12-11-2015Accepted date: 13-11-2015

    Please cite this article as: L.R. Gahan, Cyclononanes: the extensive chemistryof fundamentally simple ligands, Coordination Chemistry Reviews (2015),http://dx.doi.org/10.1016/j.ccr.2015.11.011

    This is a PDF file of an unedited manuscript that has been accepted for publication.As a service to our customers we are providing this early version of the manuscript.The manuscript will undergo copyediting, typesetting, and review of the resulting proofbefore it is published in its final form. Please note that during the production processerrors may be discovered which could affect the content, and all legal disclaimers thatapply to the journal pertain.

    http://dx.doi.org/doi:10.1016/j.ccr.2015.11.011http://dx.doi.org/10.1016/j.ccr.2015.11.011

  • Page 1 of 143

    Acce

    pted

    Man

    uscr

    ipt

    1

    Cyclononanes: the extensive chemistry of fundamentally simple ligands

    Lawrence R Gahan

    School of Chemistry and Molecular Biosciences,

    The University of Queensland,

    Brisbane 4072, Australia

    Keywords:

    Cyclononane

    Ligand

    Complex

    Synthesis

    Structure

    Spectroscopy

  • Page 2 of 143

    Acce

    pted

    Man

    uscr

    ipt

    2

    Contents

    1. Introduction

    1.1 Scope of the review

    1.2 Nomenclature

    1.3 Synthetic methodology

    1.4 Donor Properties

    1.5 Ligand Conformation

    1.6 Geometric Isomers

    2. 1-thia-4,7-diazacyclononane (1,4,7-thiadiazonane; [9]aneN2S) and N,N'-dimethyl-1-

    thia-4,7-diazacyclononane (4,7-dimethyl-1,4,7-thiadiazonane; Me2[9]aneN2S)

    2.1 Synthesis and properties

    2.2 X-ray crystal structure of [9]aneN2S

    2.3 Metal Complexes

    2.3.1 Titanium

    3.3.2 Vanadium(IV)

    2.3.3 Chromium, Molybdenum and Tungsten

    2.3.4 Manganese(II), Rhenium(I)

    2.3.5 Iron(II/III), Ruthenium(II/III)

    2.3.6 Cobalt(III), Rhodium(III),

    3.3.7 Nickel(II/III), Platinum(II), Palladium(II)

    2.3.7 Copper(II), Silver and Gold(III)

    2.3.8 Zinc(II), Mercury(II) and Cadmium(II)

    2.3.9 Indium(III), Thallium(I,III)

    2.3.10 Lead(II)

    2.3.11 Lithium

    2.3.12 Stability constants

    3. 1-Thia-4,7-diazacycyclononane-S-oxide (1,4,7-thiadiazonane 1-oxide; [9]aneN2S(O))

    3.1 Synthesis

    3.2 Metal Complexes

    3.2.1 Cobalt(III)

    3.2.2 Nickel(II), Platinum(II), Palladium(II)

    3.2.3 Copper(II)

  • Page 3 of 143

    Acce

    pted

    Man

    uscr

    ipt

    3

    4. 1,4-dithia-7-azacyclononane (1,4,7-dithiazonane ; [9]aneNS2) and N-methyl-1,4-

    dithia-7-azacyclononane (7-methyl-1,4,7-dithiazonane; Me[9]aneNS2)

    4.1 Synthesis

    4.2 Metal Complexes

    4.2.1 Iron(II), Ruthenium(II)

    4.2.2 Ruthenium(II)

    4.2.3 Rhodium(III)

    4.2.4 Nickel(II/III), Palladium(II/III)

    4.2.5 Copper(II), Silver(I), Gold(I)

    4.2.6 Chemosensor Applications

    5. 4,7-diaza-1-selenocyclononane (1,4,7-selenadiazonane; [9]aneN2Se)

    5.1 Synthesis

    6. 1-aza-4-oxa-7-thiacyclononane (1,4,7-oxathiazonane; [9]aneNOS)

    6.1 Synthesis

    6.2 Metal Complexes

    6.2.1 Nickel(II)

    7. 1-oxa-4,7-diazacyclononane (1,4,7-oxadiazonane; [9]aneN2O) and N,N'-dimethyl-1-

    oxa-4,7-diazacyclononane (4,7-dimethyl-1,4,7-oxadiazonane; Me2[9]aneN2O)

    7.1 Ligand Synthesis and Properties

    7.2 X-ray crystal structure of [9]aneN2O.2HBr

    7.3 Metal Complexes

    7.3.1 Lithium(I)

    7.3.2 Titanium(IV)

    7.3.3 Vanadium(IV)

    7.3.4 Manganese(II/III)

    7.3.5 Iron (III)

    7.3.6 Cobalt(II)

    7.3.7 Nickel(II), Palladium(II)

    7.3.8 Copper(II)

    7.3.9 Zinc(II)

    7.4 Lanthanide Complexes

    7.5 Stability constants

    8. 1,4-dioxa-7-azacyclononane (1,4,7-dioxazonane; [9]aneNO2)

    8.1 Synthesis

    9. 1-oxa-4,7-dithiacyclononane (1,4,7-oxadithionane; [9]aneOS2)

  • Page 4 of 143

    Acce

    pted

    Man

    uscr

    ipt

    4

    9.1 Synthesis

    9.2 Metal Complexes

    9.2.1 Vanadium(III)

    9.2.2 Chromium(III)

    9.2.3 Cobalt(II/III)

    9.2.4 Nickel(II), Platinum(II) and Palladium(II)

    9.2.5 Copper(I/II), Silver(I), Gold(I/II/III)

    9.2.6 Cadmium(II), Mercury(II)

    9.2.7 Antimony(III)

    10. 1-thia-4,7-dioxacyclononane (1,4,7-dioxathionane; [9]aneO2S)

    10.1 Synthesis

    11. 1,4-dithia-7-telluracyclononane (1,4,7-dithiatelluronane; [9]aneS2Te)

    11.1 Synthesis

    11.2 Metal Complexes

    11.2.1 Manganese(I)

    11.2.2 Rhodium(III)

    12. 1-tellura-4,7-dioxoacyclononane (1,4,7-dioxatelluronane; [9]aneO2Te)

    12.1 Synthesis

    12.2 Metal Complexes

    12.2.1 Palladium(II) and Platinum(II)

    13. 1-selena-4,7-dioxacyclononane (1,4,7-dioxaselenonane; [9]aneO2Se)

    13.1 Synthesis

    14. 1,4,7-triphenyl-1,4,7-triphosphonane (based on [9]aneP3) and 5,10,15-tris(2-

    fluorophenyl)-10,15-dihydro-5H-tribenzo[b,e,h][1,4,7]triarsonine

    (tribenzo[9]aneAs3Ph,PhF2)

    14.1 Syntheses and properties of metal complexes

    15. 4-benzyldodecahydro-1H-benzo[b][1,4,7]diazaphosphonine ([9]aneN2P)

    15.1 Synthesis and metal complexes

    16. 1-Phenyl-1-phospha-4,7-dithiacyclononane (7-phenyl-1,4,7-dithiaphosphonane;

    [9]aneP(Ph)S2)

    16.1 Synthesis

    16.2 Metal Complexes

    16.2.1 Molybdenum

    16.2.2 Iron(II)

    16.2.3 Nickel(II)

  • Page 5 of 143

    Acce

    pted

    Man

    uscr

    ipt

    5

    16.2.4 Copper(I)

    16.2.5 Mercury(II)

    17. Conclusions

    Abstract

    The cyclononane ligands are fundamental examples of the macrocyclic ligand. With three

    donor atoms in a nine-membered ring they represent the simplest of the macrocyclic class.

    However, this simplicity is deceptive, for within this class the variations are extensive and

    perhaps hitherto unrecognized. As well as the relatively familiar 1,4,7-triaza- ([9]aneN3) and

    1,4,7-trithia-([9]aneS3) cyclononanes whole families of cyclononane ligands are known with

    thioether/nitrogen, nitrogen/oxygen, selenium, tellurium as well as phosphorus and arsenic

    donors. The chemistry of some of these ligands, for example [9]aneN2S, [9]aneNS2 and

    [9]aneN2O has been reported extensively, and some examples have simply been reported as

    having been synthesised (for example [9]aneO2S and [9]aneO2Se) with little other

    information. This review attempts to introduce the extensive chemistry of what are

    fundamentally the simplest examples of macrocyclic ligands.

    1. Introduction

    Macrocyclic ligands are defined as polydentate ligands containing the donor atoms

    incorporated in a cyclic backbone and containing at least three donor atoms, the macrocyclic

    ring consisting of a minimum of nine atoms [1]. As such, the cyclononane ligands are

    representatives of the ultimate definition of the macrocyclic ligand. Of the cyclononanes the

    chemistry of 1,4,7-triazacyclononane ([9]aneN3) and 1,4,7-trithiacyclononane ([9]aneS3) has

    been widely reported and the contribution of these ligands to our understanding of all areas of

    coordination chemistry has been significant. The chemistry and applications of [9]aneN3 have

    been reviewed extensively [2-38], the [9]aneS3 ligand less so [39].

    These cyclononane ligands are of interest for a number of reasons, including:- (i) the

    small cyclononane ring invariably means that complexation with a metal ion occurs on one

    face, rather than the more traditional macrocyclic coordination mode of the metal ions placed

    within the macrocyclic ring; (ii) within this facial arrangement, cyclononane ligands may

    adopt an endodentate (the donor electron pairs point out of the cavity of the macrocycle,

    essentially away from the metal ion) or exodentate conformation of the donors (the electrons

    on the donor atom point into the cavity of the macrocycle towards the metal ion), depending

    on the nature of the donors and the metal ion; (iii) a variety of donor atoms are represented –

    examples containing secondary and tertiary amines, thiaether, oxygen, phosphorus, arsenic,

  • Page 6 of 143

    Acce

    pted

    Man

    uscr

    ipt

    6

    selenium and tellurium donors have been synthesised; and, (iv) given the scope of donor

    atoms available, complexation to many of the metal ions in the periodic table is possible.

    1.1 Scope of the review

    This review focuses on the cyclononane macrocycles containing mixed donors and their

    transition metal complexes; the chemistry of the oxygen analogue [9]aneO3 is not considered

    [40-42, 43 , 44 , 45, 46 , 47-56]. An excellent review, published in 1998, discusses the

    chemistry of the mixed nitrogen and sulfur donor cyclononane and cyclodecane macrocycles

    [57]. That review covered the chemistry of the [9]aneN2S, [9]aneNS2, [10]aneN2S and

    [10]aneNS2 ligands, as well as the known analogues of these systems [57]. In order to allow

    full appreciation of the extensive chemistry of these nitrogen- and sulfur-tridentate ligands,

    and the subsequent extensions of the chemistry of cyclononane ligands, it has been necessary

    to recap some of the examples in that earlier report in order to allow a more complete

    perspective of the developments in this chemistry post-1998. Where figures depicting the

    structures of metal complexes under discussion have been included in the previous review,

    these figures are not duplicated here. Due acknowledgement of the content of that earlier

    review is given at his point [57].

    1.2 Nomenclature

    As the chemistry of the cyclononane ligands developed the nomenclature employed to

    describe them has been systematised, to some extent.

    Initial papers reported the 1,4,7-triazacyclonane ligand as “tacn” [7] although the

    nomenclature [9]aneN3 has now mostly replaced this initial name. The IUPAC nomenclature,

    1,4,7-triazonane, has been employed in some instances [58-67]. The initial papers describing

    1-dithia-4,7-azacyclononane abbreviated the name to “tasn” [68-70] in line with the “tacn”

    nomenclature and this nomenclature is still in use in some reports [71] although the

    abbreviation [9]aneN2S is common, although not strictly correct in relation to the IUPAC

    nomenclature, 1,4,7-thiadiazonane. In this work, as much as is practicable, the [9]aneXYZ

    nomenclature will be applied, except in instances where the report details a more complex

    synthetically elaborated form of the ligand and in that case the nomenclature employed in the

    report will be used. Where possible, the IUPAC ligand name will be introduced.

    The ligands discussed in this review include, 1-thia-4,7-diazazcyclononane

    ([9]aneN2S) and 1,4-dithia-7-azacyclononane ([9]aneNS2), and in more detail the perhaps

    less well known examples of the cyclononane ligands including, 1-oxa-4,7-

    diazacyclononane ([9]aneN2O), 1,4-dioxa-7-azacyclononane ([9]aneNO2), 1-oxa-4,7-

    dithiacyclononane ([9]aneOS2), 4,7-diaza-1-selenocyclononane ([9]aneN2Se), 1-aza-4-oxa-

  • Page 7 of 143

    Acce

    pted

    Man

    uscr

    ipt

    7

    7-thiacyclononane ([9]aneNOS), 1-thia-4,7-dioxacyclononane ([9]aneO2S), 1,4-dithia-7-

    telluracyclononane ([9]aneS2Te), 1-tellura-4,7-dioxoacyclononane ([9]aneO2Te), 1-selena-

    4,7-dioxacyclononane ([9]aneO2Se), [9]aneP3, [9]aneN2P, and [9]aneAs3.

    In addition, examples of structurally enhanced cyclononane ligands including 1-thia-

    4,7-diazacyclononane-N,N’-diacetic acid [72], dimethyl 1-thia-4,7-diaza-4,7-

    cyclononanediacetate, [73] 4,7-bis(2-methylpyridyl)-1-thia-4,7-diazacyclononane [74], 4,7-

    bis(hydroxyethyl)-1-thia-4,7-diazacyclononane [75], 4,7-bis(2-hydroxy-2-methylpropyl)-1-

    thia-4,7-diazacyclononane [75], 4,7-bis(2-cyclohexyl-2-hydroxymethyl)-1-thia-4,7-

    diazacyclononane [75], 4,7-bis(2-cyanoethyl)-1-thia-4,7-diazacyclononane [76-78], 15-thia-

    1,5,8,12-tetraazabicyclo[10.5.2]nonadecane [77], 4,4'-((4,5-dihydro-1H-pyrazole-3,5-

    diyl)bis(methylene))bis(1,4,7-thiadiazonane) [79] and 4,7-bis(3-aminopropyl)-1-thia-4,7-

    diazacyclononane [77, 78, 80] and their metal complexes are discussed throughout. In these

    cases the literature reports employ numerous nomenclature systems, (for example the

    abbreviations (py2[9]aneN2S; bmmpTASN; L1, L2, etc), and in these cases the

    nomenclature employed in the actual report has been used, recognising that this may be

    repetitive and/or ambiguous in some cases.

    In cases where the X-ray crystal structures of metal complex cations or anions are

    shown, the hydrogen atoms and counter ion have been removed for clarity. The structures

    were obtained through ConQuest V1.17 based on data deposited at the Cambridge

    Crystallographic Data Centre [81, 82].

    1.3 Synthetic methodology

    The synthetic methodologies employed for the cyclononane ligands are generally

    relatively straightforward and in the majority of cases in what follows a brief description of

    the synthetic approach is given but not a synthetic scheme; there are some exceptions. The

    syntheses usually involve the cyclic coupling of two linear moieties in the presence of a base

    and most often in dimethylformamide (DMF) solution. Tosylated oligomeric glycols are

    commonly utilized as one component, the synthetic challenge usually being the form of the

    second fragment containing the heteroatoms; for example, in the cases of the syntheses of

    [9]aneO2Te and [9]aneO2Se. The synthetic approaches to the [9]aneP3, [9]aneN2P, and

    [9]aneAs3 ligands are more elaborate and are considered in some detail.

    1.4 Donor Properties

    Nitrogen donors, in contrast to thioethers, are considered to be hard bases, having strong

    affinity for hard acids like Co(III), Fe(III) and Cr(III) [83] as well as alkali and alkali earth

    cations [84]. The lone pair of electrons for amine ligands allows for σ-donation to metal ions

    but not π-interactions. In comparison to donors like nitrogen and phosphorus [85, 86]

  • Page 8 of 143

    Acce

    pted

    Man

    uscr

    ipt

    8

    thiaether donors bind transition metals relatively weakly [85-87], the weak bonding attributed

    to the relatively low σ-donor and π-acceptor abilities of thiaethers. The sulfur donor can also

    act as a π-acceptor where its empty d-orbitals can take part in metal to sulfur π-back donation

    if the symmetry is suitable [85]; there is little evidence for thiaethers acting as π-donors [85].

    There are a number of characteristics of thiaether complexes which can be rationalised by the

    π-acceptor ability of sulfur donors (i) they have a tendency to stabilise lower oxidation states

    of metal ions, in comparison to amine donors [86, 88]; (ii) they usually enforce low spin

    conditions on transition metal states that commonly exhibit high spin behaviour [89-97], a

    property believed to be due to the delocalisation of t2g electron density into ligand π* orbitals

    by π-back donation, thus lowering electron-electron repulsion causing the reduction of the

    spin pairing energy [88]; and, (iii) thioether donors are found lower in the nephelauxetic

    series relative to aqua and amine ligands [98]. Phosphorus-containing ligands such as 1-

    phenyl-1-phospha-4,7-dithiacyclononane are suggested to exert a stronger ligand field than

    [9]aneS3 [99]. The donor properties of selenium and tellurium have not been extensively

    discussed [100, 101].

    1.5 Ligand Conformation

    The classification scheme employed in this review to describe the conformations of the

    chelate rings in the nine-membered macrocycles is largely that of Dale [102-104]; there are

    a number of reports containing excellent explanations of the classification scheme [105,

    106]. Briefly, the classification is based on the torsion-angle sequences between the anti

    and gauche bonds in the cyclononane ring, although it is applicable to larger macrocyclic

    rings. A series of numbers is used to define the conformation, each number designating the

    number of chemical bonds between what are called consecutive genuine corners. A genuine

    corner occurs when the central atom of an anti–gauche–gauche–anti bond has both

    consecutive gauche torsion angles of the same sign; a pseudo corner occurs when they are of

    opposite sign. The starting point and direction followed around the ring are chosen so as to

    give the smallest number [102-106]. For cyclononane rings the three lowest energy

    conformations are [333] , [234] and [12222]; these are illustrated in Figure 1 [102-104].

  • Page 9 of 143

    Acce

    pted

    Man

    uscr

    ipt

    9

    Figure 1: Conformations of cyclononane ligands [102-104]

    Molecular mechanics calculations indicate that the strain energy for complexes of

    cyclononane macrocycles is always less for the [333] conformer than for the [234]

    conformer [107]. For example, for the [Ni([9]aneN3)2]2+ and [Ni([9]aneS3)2]

    2+ complexes

    the energy difference between the [234] and [333] conformations was considerable (35.7

    and 28.1 kcal mol-1 and 21.4 and 12.8 kcal mol-1, respectively) suggesting why the [333]

    conformer was prevalent in these complexes and suggesting a reason for the rarity of the

    [234] conformer [107]. In the case of the mixed donor systems the differences are much

    smaller. For example for [Ni([9]aneN2S)2]2+ calculations show that for the [234] conformer

    the strain energy is 28.6 kcal mol-1 whereas for the [333] conformer it is 25.1 kcal mol-1; for

    the [Ni([9]aneNS2)2]2+ the calculated difference is also small (27.8 compared with 19.2 kcal

    mol-1) [107]. The [234] conformer would therefore be expected to be more common in the

    mixed donor systems. An extensive list of ring conformations for complexes with [9]aneN3

    and [9]aneS3 ligands, in addition to a limited number of examples of [9]aneN2S, and

    [9]aneN2O complexes has been reported [106].

    1.6 Geometric Isomers

    For metal complexes of the [9]aneN3 and [9]aneS3 (and [9]aneP3 and [9]aneAs3 analogues,

    vide infra) ligands, facial η3 coordination of the ligands to the metal ion is commonly, but

    not exclusively, found. In the case of the [9]aneX2Y type ligands complexes with both cis-

    and trans-diastereoisomers are possible, and in some instances have been isolated. The

    trans-, or meso-, isomer has C2h symmetry, whereas the cis-isomer displays C2 symmetry.

    The possible geometric forms are shown diagrammatically in Figure 2.

    -135+55

    +55

    -135

    +55+55

    -135

    +55

    +55+65 -85

    -45

    +150

    -55

    -40+155

    -140

    +50

    [333] [234]

    +110

    -45

    -85

    +75

    -85

    -45

    +110

    [12222]

    -100

    +75

  • Page 10 of 143

    Acce

    pted

    Man

    uscr

    ipt

    10

    Figure 2. Possible geometric isomers for [M([9]aneX2Y)2]n+ complexes [108, 109].

    2. 1-thia-4,7-diazacyclononane (1,4,7-thiadiazonane; [9]aneN2S) and N,N'-dimethyl-

    1-thia-4,7-diazacyclononane (4,7-dimethyl-1,4,7-thiadiazonane; Me2[9]aneN2S)

    2.1 Synthesis and properties

    The synthesis of [9]aneN2S (the ligand was named “tasn” in that paper) was reported

    in 1982; at the time no synthetic details were given, the method described as similar to that

    employed for the [9]aneN3 analogue using the Richman-Atkins procedure [68, 110]. The

    manuscript described the preparation and isolation of the [Co([9]aneN2S)2](ClO4)3 complex

    [68].

    Subsequently, the syntheses of the [9]aneN2S ligand, and derivatives, have been

    reported a number of times [75, 111-117]. The synthetic strategy typically involved the use of

    bis(2-aminoethyl)sulfide, the synthesis of which traditionally employed the reaction of

    ethyleneimine and H2S [118-121]. Other syntheses of bis(2-aminoethyl)sulfide have

    employed NH2CH2CH2OSO3H with NaSH and S or Na2S and S [122], and the treatment of

    cysteamine hydrochloride with chloroethylamine hydrochloride in the presence of NaOH

    [123] or reaction of 2-bromoethylamine hydrobromide with sodium hydroxide and sodium

    sulfide nonahydrate [124]. Reaction of the bis(2-aminoethyl)sulfide with toluenesulfonyl

    chloride, resulting in N,N'-(thiobis(ethane-2,1-diyl))bis(4-methylbenzenesulfonamide), and

    subsequent reaction of the disodium salt with ethane-1,2-diyl bis(4-methylbenzenesulfonate)

    in dimethylformamide resulted in 4,7-bis(tolyl-p-sulfonyl)-1-thia-4,7-diazacyclononane.

    Removal of the protecting groups with, for example, hydrobromic acid/acetic acid [68, 75],

    X Y X

    XYX

    X Y X

    YXX

    M M

    trans- or meso-diastereoisomer cis-diastereoisomer

    S

    NH

    HN

    [9]aneN2SS

    N

    N

    Me2[9]aneN2S

  • Page 11 of 143

    Acce

    pted

    Man

    uscr

    ipt

    11

    or lithium in ammonia [75], resulted in the desired ligand, in varying yields. The ligand

    displays two pKa values of 9.67(2) and 3.98(2) [125]. The analogue Me2[9]aneN2S was

    prepared by refluxing [9]aneN2S.2HBr with formic acid and formaldehyde [126].

    Analogues such as N,N'-(((1,4,7-thiadiazonane-4,7-diyl)bis(methylene))bis(1-methyl-

    1H-imidazole-2,4-diyl))bis(2,2-dimethylpropanamide) and 4,7-bis(2-thiophenoyl)-1-thia-4,7-

    diazacyclononane have been prepared by reaction of the [9]aneN2S with N-(2-formyl-1-

    methyl-1H-imidazol-4-yl)pivalamide [127], and 2-chlorocarbonyl)thiophene [128],

    respectively (Figure 3).

    (a) (b)

    Figure 3. (a) N,N'-(((1,4,7-thiadiazonane-4,7-diyl)bis(methylene))bis(1-methyl-1H-

    imidazole-2,4-diyl))bis(2,2-dimethylpropanamide); (b) 4,7-bis(2-thiophenoyl)-1-thia-4,7-

    diazacyclononane

    The salt 4-(2-bromoacetyl)-8-oxo-1-thionia-4,7-diazabicyclo[5.2.2]undecane bromide

    resulted after reaction of [9]aneN2S with bromoacetyl bromide in chloroform (Figure 4). The

    crystal structure shows that two salt moieties are linked by S···Br contacts about a

    crystallographic inversion centre, forming dimers linked by halide-halide contacts into

    extended ribbons [129].

    Figure 4. 4-(2-bromoacetyl)-8-oxo-1-thionia-4,7-diazabicyclo[5.2.2]undecane bromide

    Further examples of derivatised [9]aneN2S ligands and the metal complexes formed

    from them, are discussed in the following text.

    N

    N

    OHN

    S

    NN

    NHO

    N

    N

    S

    S

    S

    N

    N

    S+

    N

    N OO

    Br

    Br-

  • Page 12 of 143

    Acce

    pted

    Man

    uscr

    ipt

    12

    2.2 X-ray crystal structure of [9]aneN2S

    The X-ray crystal structure of [9]aneN2S [130] has been reported. The hydrogen-

    bond networks in mono- and diprotonated [9]aneN2S have been investigated [130]. The

    structure of [9]aneN2S·HCl consists of a monoprotonated [9]aneN2S ring and a chloride anion

    (Figure 5).

    Figure 5. Crystal structure of [9]aneN2S·HCl [130]

    The protonated nitrogen serves as a hydrogen-bond donor to the free amine via an

    intramolecular interaction. In addition to the intramolecular H-bonding interaction, each

    [9]aneN2S·HCl unit serves as both an H-bond donor and an H-bond acceptor via

    intermolecular interactions [130]. The stronger of the two interactions occurs between the

    protonated nitrogen and chloride. The chloride anions of [9]aneN2S⋅HCl align in interlocking

    columns in the a- and c-direction. The structure of [9]aneN2S·2HBr contains a doubly

    protonated [9]aneN2S ring and two bromide anions. The diprotonated [9]aneN2S⋅2HBr also

    exhibits the ribbon-like network with strong intermolecular and weak intramolecular N-

    H⋅⋅⋅Br hydrogen bonds. The intramolecular and intermolecular interactions result in an

    extended array in the b-direction. These ribbons align in the crystal in an interlocking or

    “zipper”-like arrangement, although no notable inter-strand contacts apparent. The extended

  • Page 13 of 143

    Acce

    pted

    Man

    uscr

    ipt

    13

    H-bonding network yields a columnar array in the b-direction. The bromine columns fill a

    channel that runs through the crystal along the c-direction [130].

    2.3 Metal Complexes

    2.3.1 Titanium

    Reaction of [Ti(NBut)Cl2(py)3] with Me2[9]aneN2S in dichloromethane at room temperature

    resulted in the air and moisture sensitive orange complex [Ti(NBut)Cl2(Me2[9]aneN2S)] [131,

    132]. The Ti(IV) ion is coordinated by the facially arranged Me2[9]aneN2S macrocycle, two

    mutually cis-chlorido ligands and the multiply bonded tert-butylimido ligand (Figure 6).

    Figure 6. Crystal structure of [Ti(NBut)Cl2(Me2[9]aneN2S)] [131,

    132]

    The Ti-N(cyclononane) bonds cis- (2.285(9) Ǻ) and trans- (2.498(8) Ǻ) show the trans-effect of

    the tert-butylimido ligand with the Ti-Nimide 1.708(8) Ǻ; the Ti-S bond distance is 2.561(4) Ǻ

    and Ti-Cl distances 2.379(4) and 2.397(4) [131, 132]. The 13C NMR of the complex

    displayed a ten line spectrum consistent with a C1 symmetric arrangement of the macrocycle

    with the thiaether donor cis to the imido group; the sharp 13C and 1H NMR resonances

    suggested that the complex was conformationally rigid in solution [132].

    3.3.2 Vanadium(IV)

    The light blue vanadium(IV) complex [VOCl2([9]aneN2S)].CH3CN was obtained after

    reaction of VCl3 and [9]aneN2S in refluxing acetonitrile under nitrogen, the oxidation to

    V(IV) attributed to trace amounts of oxygen during crystallisation of the complex [133].

    The X-ray structure shows that in [VOCl2([9]aneN2S)].CH3CN the metal ion is coordinated

    facially by cyclononane ligand, the thiaether and the vanadyl oxygen arranged trans to each

    other, the chloride ligands mutually cis and the metal ion is six coordinate. The metal ion

    sits 0.32 Å above the plane of the nitrogen and chloride donors. The V-O bond is short

    (1.632(2) Ǻ) and the V-S bond trans is elongated (2.689(1) Ǻ) suggesting a strong trans

    influence; the same effect is seen in the [VOCl2([9]aneS3)] analogue [134]. The V-N bond

    lengths (2.153(2) Ǻ, 2.150(2) Ǻ) are similar to those observed for other oxovanadium

    complexes, for example [V2O2(µ-OH)2([9]aneN3)2]Br2 (2.151(5)-2.303(6) Ǻ) [135]. The V-

    Cl bond lengths (2.346(1) Ǻ, 2.337(1) Å) are longer than in the related [VOCl2([9]aneS3)]

    complex (2.295(5)average Ǻ) [134]. The IR spectrum of the complex shows sharp ν(NH)

    absorptions at 3240 and 3200 cm-1, the ν(VO) vibration shifted to 985 cm-1 as compared with

  • Page 14 of 143

    Acce

    pted

    Man

    uscr

    ipt

    14

    the [VOCl2([9]aneS3)] complex (962 cm-1) as a result of the weak bonding of the thiaether in

    [VOCl2([9]aneN2S)].CH3CN [133, 134]. For the d1 system, three vanadyl(IV) d–d

    transitions at 27624 cm-1 (ε = 67 M-1 cm-1), 19800 cm-1 (ε = 21 M-1 cm-1) and 16474 cm-1 (ε

    = 63 M-1 cm-1) were observed with Dq = 19 800 cm-1 [133]. The higher than expected

    ligand field exerted by [9]aneN2S was confirmed by the EPR parameters, with 51V

    hyperfine coupling constants Aiso = 91.2 × 10-4 cm-1 at g = 1.986 and A|| = 161.1 × 10

    -4

    cm-1 at g|| = 1.965 [133].

    2.3.3 Chromium, Molybdenum and Tungsten

    The chromium(III) complex [Cr2(OH)(O2CMe)2(Me2[9]aneN2S)2](ClO4)3 was

    prepared following the procedure employed for the analogous

    [Cr2(OH)(O2CMe)2(Me3[9]aneN3)2](ClO4)3 complex [136, 137]. The X-ray crystal structure

    of [Cr2(OH)(O2CMe)2(Me2[9]aneN2S)2](ClO4)3 shows Cr-N bonds lengths (2.15(2) Ǻ) longer

    than those for the Me3[9]aneN3 analogue (2.102(6) Ǻ) [136, 137]. Fitting of the magnetic

    susceptibility data (4.2 – 300 K) for exchange coupled pairs of chromium(III) ions (S1 =3/2,

    S2 =3/2; H = -2JS1·S2) for [Cr2(OH)(O2CMe)2(Me2[9]aneN2S)2](ClO4)3 showed that the metal

    ions were antiferromagnetically coupled (J = -15 cm-1), the coupling very similar to that

    observed for the Me3[9]aneN3 analogue (J = -15.5 cm-1) [137].

    Reaction of Mo(CO)6 with [9]aneN2S in ethanol resulted in isolation of

    [Mo([9]aneN2S)(CO)3]; a similar approach resulted in [W([9]aneN2S)(CO)3] [113]. The

    infrared spectrum of the molybdenum complex shows bands at 1913, 1783 and 1713 cm-1

    assigned to the carbonyl group. The X-ray crystal structure of [Mo([9]aneN2S)(CO)3] shows

    the facial arrangement of the cyclononane macrocycle [113], the ring having the [333]

    conformation [102] with Mo-N distances of 2.317(5) Ǻ and 2.292(5) Ǻ and Mo-S 2.526(2) Ǻ

    [113]. The three Mo-C distances were 1.924(6), 1.926(6) and 1.952(6) Ǻ, the shorter bond

    trans to the thiaether donor [113]. Both [Mo([9]aneN2S)(CO)3] and [W([9]aneN2S)(CO)3]

    were employed to prepare nitrosyl-, halogeno, and for the former, oxomolybdenum

    complexes [138]. Complexes with M = Mo or W were obtained commencing with the

    appropriate [M([9]aneN2S)(CO)3] complex. Thus, reaction of [M([9]aneN2S)(CO)3] with an

    aqueous solution of sodium nitrite, with subsequent reaction with HCl, resulted in isolation of

    [M([9]aneN2S)(CO)2(NO)](PF6). When [M([9]aneN2S)(CO)3] was dissolved in 1 M HCl

    and reacted with aqueous sodium nitrite the product obtained was

    [M([9]aneN2S)(NO)2Cl](PF6) (M = Mo, W); a similar approach with HBr resulted in the

    analogous [W([9]aneN2S)(NO)2Br](PF6) complex [138]. The infrared spectrum of

    [Mo([9]aneN2S)(CO)2(NO)](PF6) showed carbonyl stretching frequencies at 2010 and 1940

    cm-1 and an NO stretch at 1670 cm-1; the nitrosyl stretching frequencies for

  • Page 15 of 143

    Acce

    pted

    Man

    uscr

    ipt

    15

    [Mo([9]aneN2S)(NO)2Cl](PF6) were observed at 1790 and 1690 cm-1 reportedly typical for

    dinitrosyl compounds with linear M-N-O groups. The complex [Mo([9]aneN2S)Cl3] was

    prepared after reaction of [Mo([9]aneN2S)(CO)3] with concentrated HCl [138]. Reaction of

    [Mo([9]aneN2S)(CO)3] in trichloromethane with bromine resulted in the isolation of

    [Mo([9]aneN2S)(CO)3Br](PF6); the analogous tungsten complex was not reported [138].

    [Mo2([9]aneN2S)2O4](ClO4)2 was isolated after reaction of [Mo([9]aneN2S)(CO)3] with

    HClO4; reaction of this product with HCl resulted in the green complex

    [Mo2([9]aneN2S)2O3Cl2](ClO4)2 [138]. The [Mo2([9]aneN2S)2O4](ClO4)2 complex

    displayed infrared and Raman bands at 965, 940, 735, 715, 450 and 445 cm-1 typical of a syn-

    Mo2O4 structure [138]. For [Mo2([9]aneN2S)2O3Cl2](ClO4)2 the infrared spectrum displayed

    bands at 905 and 775 cm-1 assigned to υMo=O and υas(Mo-O-Mo) of the Mo2O3 core [139, 140].

    The X-ray structures of the complexes [Mo([9]aneN2S)Cl3], [Mo([9]aneN2S)(NO)2Cl](PF6),

    [Mo([9]aneN2S)(CO)2(NO)](PF6) and [Mo2([9]aneN2S)2O4](ZnCl4).H2O have been reported

    [138]. The structures of [Mo([9]aneN2S)(CO)2(NO)](PF6) and

    [Mo([9]aneN2S)(NO)2Cl](PF6) [138] show that in both complexes the metal ions are six-

    coordinate with the cyclononane ligand coordinated facially. In the former complex ion, the

    strong π-acid ligand NO is trans to a nitrogen donor of the macrocycle, underlining the π-

    acceptor properties of the thiaether. In [Mo([9]aneN2S)(NO)2Cl](PF6) the NO ligands are

    again trans to the cyclononane amine donor, with the thiaether then trans to the weakly π-

    accepting chlorido ligand. In the two complexes the Mo-NO bond distances differ

    considerably ([Mo([9]aneN2S)(CO)2(NO)](PF6), 1.874(3) Ǻ;

    [Mo([9]aneN2S)(NO)2Cl](PF6), 1.819(7) Ǻ) the difference attributed to the different number

    of strong π-acid ligands, which share the back-donation [138]. The mixed donor sets on the

    [Mo([9]aneN2S)(CO)2(NO)](PF6) and [Mo([9]aneN2S)(NO)2Cl](PF6) complexes offer the

    possibly of isomeric forms. Based on the spectroscopic properties and the structural

    parameters for the [Mo([9]aneN2S)(CO)3] complex, the CO ligand trans to the thiaether is

    less strongly bound than the CO ligands in the cis positions [138]. The expectation would

    have been therefore that the former CO group would be replaced by the NO ligand. In

    contrast, the entering ligand NO is a better π-acceptor than CO and may preferentially bind

    trans to the N donor atoms, if the thiaether sulfur atom possesses some π-acidity.

    The molecule [Mo([9]aneN2S)C13] is chiral; the five-membered chelate rings formed

    by the cyclononane ligand have either the δδδ or λλλ conformation. The complex crystallises

    in an acentric space group (Pn21a) which contains both enantiomers [138]. The Mo-N bond

    distances are short (2.193(3) and 2.209(4) Ǻ) reflecting the 3+ oxidation state of the Mo. The

    Mo-S bond (2.467(1) Ǻ) is longer than in [Mo([9]aneN2S)(NO)2Cl](PF6) (2.447(2) Ǻ) and the

  • Page 16 of 143

    Acce

    pted

    Man

    uscr

    ipt

    16

    trans Mo-Cl bond in the latter complex (2.440(2) Ǻ) is longer than those in

    [Mo([9]aneN2S)C13] (2.425(1), 2.408(1), 2.426(1) Ǻ), the difference related to π-acceptor

    properties of the thiaether group. Interestingly, reaction of [Mo([9]aneN2S)C13] with nitric

    acid led to isolation of a colourless solid, the infrared spectrum of which contained a strong

    band at 1150 cm-1, assigned as an S=O stretching frequency [138]. Previously it has been

    reported that reaction of [9]aneN2S with sodium bromite resulted in clean oxidation of the

    thiaether to a sulfoxide with formation of 1-thia-4,7-diazacyclononane-S-oxide [115, 141].

    The X-ray crystal structure of the orange [Mo2([9]aneN2S)2O4]2+ complex consists of

    two [(1-thia-4,7-diazacyclononane)oxomolybdenum(V)] cations linked through a bis(µ-oxo)

    moiety, the Mo=O groups in the syn position, and the Mo-Mo distance 2.549(1) Ǻ [138].

    The [9]aneN2S ligands were rotationally disordered at both molybdenum sites such that each

    macrocycle was attached to the metal atom statistically in two different orientations. This

    rotational disorder had the effect that only one coordination site in the equatorial planes at the

    two molybdenum atoms is occupied exclusively by nitrogen. The other equatorial sites and

    the axial sites trans to the terminal oxo ligands are occupied statistically by the thiaether and

    nitrogen donor atoms. The site occupation factors at the metal atoms showed that the trans

    position was preferred by the thiaether group with a probability of 70-80% [138]. The

    conformation of the five-membered chelate rings in [Mo2([9]aneN2S)2O4]2+ was δλλ at one

    Mo site and λδδ at the other [138].

    2.3.4 Manganese(II), Rhenium(I)

    Reaction of manganese(II) perchlorate and [9]aneN2S in methanol solution resulted in

    isolation of [Mn([9]aneN2S)2](ClO4)2 [124, 133]. The X-ray crystal structure shows that all

    donors are coordinated to the metal ion, the thiaether donors are trans and the nitrogen donors

    in the equatorial plane of the molecule. At 298 K the magnetic moment for the complex was

    5.82 µB, indicative of a high spin (S = 5/2) configuration. The magnetic moment remained

    constant over the range 300 to 80 K and upon lowering the temperature to 4.2 K the magnetic

    moment was reduced to 5.65 µB. The decrease in magnetic moment was ascribed to second

    order spin orbit coupling, giving rise to zero-field splitting of the 6A1g state [124]. Computer

    simulation of the Q-band EPR spectrum for the complex yielded g = 1.99±0.01, ׀D׀ =

    0.19±0.005 cm-1 and E/D = 0.04±0.02 [124]. These values were compared with those for the

    analogous [Mn([9]aneN3)2](ClO4)2 complex, with g = 1.98±0.01, ׀D0.090±0.003 = ׀ cm-1 and

    E/D = 0.10±0.01, the larger value of D for the former complex reflecting the larger axial

    distortion induced by the axial thiaether donors [124].

    The bimetallic complex [Mn2(Me2[9]aneN2S)2(µ-OH)(µ-CH3COO)2](ClO4) has been

    prepared but not structurally characterised [126]. Variable temperature (300-2K) magnetic

  • Page 17 of 143

    Acce

    pted

    Man

    uscr

    ipt

    17

    susceptibility of a powered sample exhibited a decrease in the effective magnetic moment

    from 7.63 µB/molecule at 300 K to 1.36 µB/molecule at 4.2 K, indicative of antiferromagnetic

    coupling between the two manganese(II) sites. A least squares fit of the molar

    susceptibilities using the general isotropic exchange Hamiltonian H = -2JS1·S2, where S1 = S2

    = 5/2 produced J = -3.6 cm-1 with g = 1.92, Ɵ = -0.87 K and p = 1.10 (percent monomeric

    paramagnetic impurity) [126]. In solution the magnetic moment was determined to be 7.4

    µB/molecule, in agreement with the solid state measurement at 300 K indicating that the

    complex retained its integrity in solution [126]. For the [Mn2(Me3[9]aneN3)2(µ-OH)(µ-

    CH3COO)2](ClO4) analogue, J = -9 cm-1 with g = 1.98 [142]. The solid state X-band EPR

    spectrum of [Mn2(Me2[9]aneN2S)2(µ-OH)(µ-CH3COO)2](ClO4) at 130 K displayed a group

    of intense transitions at g = 2, a group at half field showing fine structure at g = 4 and a broad

    peak at g = 16. The hyperfine structure observed on the intense peak and the peak at half

    field (0.0570 T) exhibited a hyperfine coupling constant of A = 42 x 10-4 cm-1, a value half

    that for mononuclear Mn(II) complexes (typically A = 90 x 10-4 cm-1) and confirming that the

    complex is dinuclear. The intense transitions at g = 2 were assigned to ∆Ms = ±1 transitions

    and the weaker half field transitions at g = 4 to the ∆Ms = ±2 transitions. The presence of the

    ∆Ms = ±2 transitions indicate coupling of the two Mn(II) ions. The broad peak at g = 16 was

    tentatively assigned to the ∆Ms = ±4 transitions [126]. Variable temperature (110-10 K)

    studies on the solid and frozen solution (acetonitrile:toluene glass) were used in order to

    corroborate the magnetic susceptibility studies. Analysis of the data fitted to an excited S = 2

    manifold resulted in a coupling of -3.2 to -3.4 cm-1, in good agreement with the data obtained

    from the solid state susceptibility measurements [126]. Analysis of the axial field splitting,

    cm-1, in terms of the magnetic dipole interaction between the two Mn(II) ions 0.073 = ׀D׀

    resulted in a calculated metal-metal separation of 3.28 Ǻ [126].

    Reaction of [9]aneN2S with picolyl chloride hydrochloride in the presence of base

    resulted in isolation of the pentadentate ligand 4,7-bis(2-pyridylmethyl)-1-thia-4,7-

    diazacyclononane (py2[9]aneN2S) (Figure 7(a)) [117]. The Mn(II) complex

    [Mn(py2[9]aneN2S)Cl](CF3SO3) was isolated and structurally characterised (Figure 7(b)).

    N

    N

    S

    N

    N

  • Page 18 of 143

    Acce

    pted

    Man

    uscr

    ipt

    18

    (a) (b)

    Figure 7. (a) 4,7-bis(2-pyridylmethyl)-1-thia-4,7-diazacyclononane (py2[9]aneN2S), and (b)

    Crystal structure of [Mn(py2[9]aneN2S)Cl]+ [117]

    Of particular interest were the relative conformations of the pendant arms on the

    diazacyclononane ring. Two coordinating conformers are possible, one with the pendant

    arms opposite to (α) or on the same side (β) of the metal ion. The β conformer leads to a

    trigonal prismatic geometry at the metal centre whereas the α conformer leads to a distorted

    octahedral geometry. The [Mn([9]N2Spy2)Cl](CF3SO3) complex displayed a coordination

    geometry closer to trigonal prismatic. The complex exhibited a magnetic moment of 5.90 µB,

    indicative of a high spin d5 configuration. The cyclic voltammogram displayed a Mn(II/III)

    couple at 0.92 V (NHE) with a peak separation of 80 mV [117].

    The rhenium complex [Re([9]aneN2S)(CO)3]Br has been isolated after reaction of the

    ligand with [Re(CO)5Br] in DMF [113]. The complex shows IR bands at 2030, and 1980-

    1860 cm-1 assigned to carbonyl vibrational modes. The X-ray crystal structure of the

    complex shows Re-N distances of 2.209(6) Ǻ, 2.196(6) Ǻ and 2.441(2) Ǻ, shorter than the

    Mo analogue (2.317(5) Ǻ and 2.292(5) Ǻ and Mo-S 2.526(2) Ǻ) [113]. The three Re-C

    distances 1.900(7) Ǻ, 1.927(7) Ǻ and 1.912(7) Ǻ, do not follow the trend seen for the Mo

    analogue, with the bond trans to the thiaether donor (1.912(7) Ǻ) being in the middle of the

    range seen [113].

    Reaction of dirhenium heptaoxide with [9]aneN2S in tetrahydrofuran resulted in a

    rhenium(VII) half-sandwich complex formulated as [ReO3([9]aneN2S)][ReO4] [133]. The

    Raman spectrum showed two strong, well resolved bands of equal intensity at 972 and 964

    cm-1 assigned to the symmetric stretching vibrations of ReO3+ and ReO4

    -. A vibration at 910

    cm-1 and a shoulder at 930 cm-1 in the IR spectrum were assigned to the asymmetric

    stretching vibrations of ReO3+ and ReO4

    - after comparison with similar IR bands seen for

    the analogous complexes[ReO3([9]aneN3)][ReO4] and [ReO3([9]aneS3)][ReO4] (935 and 909,

    921 and 912 cm-1, respectively) [133, 143, 144].

    2.3.5 Iron(II/III), Ruthenium(II/III)

    The synthesis and characterisation of the iron(II) complex [Fe([9]aneN2S)2](ClO4)2 has been

    reported [94, 133]. The metal ion has the trans-N4S2 coordination although only one of the

    possible geometric isomers has been characterised. At 81 K the complex crystallised in the

    monoclinic space group P21/c with the Fe atom on an inversion centre [133]. At room

  • Page 19 of 143

    Acce

    pted

    Man

    uscr

    ipt

    19

    temperature the orthorhombic space group Pmcb was found and the complex was disordered

    in order to accommodate the symmetry [94]. The Fe-S and Fe-N bond lengths in both

    structures were indistinguishable [94, 133]. The structure of the corresponding iron(III)

    complex [Fe([9]aneN2S)2](ClO4)3 shows a centrosymmetric cation, the iron lying on a 2/m

    symmetry site, and the two [9]aneN2S ligands facially coordinated [94]. The thiaether donors

    were in axial positions (S-Fe-S 180(1)°) with the amine donors in the equatorial positions

    completing the coordination sphere [94]. Comparison of the reported structures of

    [Fe([9]aneN2S)2](ClO4)2 and [Fe([9]aneN2S)2](ClO4)3 shows a contraction in the Fe-N

    bond lengths (Fe(II), 2.072(2), 2.063(7); Fe(III) 2.006(3) Å), also observed in the

    complexes [Fe([9]aneN3)2]Cl2 and [Fe([9]aneN3)2]Cl3 (2.03(1) and 1.99(2) Å,

    respectively), and expected on the basis of the decrease in ionic radius from Fe(II) to

    Fe(III). The Fe-S bond distance in [Fe([9]aneN2S)2](ClO4)2 is also longer than that

    observed for [Fe([9]aneN2S)2](ClO4)3 (Fe(II)-S 2.337(1) Å, Fe(III)-S 2.272(1) Å). This

    difference has been advanced as structural evidence for the π-acceptor properties of the

    thiaether donor atoms [145].

    Variable-temperature susceptibility measurements for [Fe([9]aneN2S)2](ClO4)2 in the

    range 2–300 K revealed a temperature-dependent magnetic moment [94]. Between 4.2 and

    150 K the magnetic moment was 0.5 µB and this gradually increased to 2.95 µB at 300 K. For

    a high spin d6 configuration the expected spin-only magnetic moment is 4.9 µB while the low

    spin state would be diamagnetic. The behaviour exhibited by this complex was typical of a

    thermally induced high-low spin transition between the high spin 5T2g and low-spin 1A1g

    ground states, although the data indicate that the transition is not complete at 300 K. In

    solution, [Fe([9]aneN2S)2]2+ displays the same temperature magnetic behaviour. Assuming

    spin-only values of 4.90 and 0 µB, and from plots of lnKeq versus 1/T, the variable

    temperature solution susceptibility data were fitted such that ∆H 0 = 20 kJ mol–1 and ∆S 0 =

    53 J mol–1 K–1; these data were similar to those reported for analogous thermally induced

    high/low spin transitions in iron(II) complexes [146]. For the iron(III) complex,

    [Fe([9]aneN2S)2](ClO4)3 the solid state susceptibility data followed a Curie-Weiss law in the

    range 300 K (2.3 µB) – 4.2 K (1.9 µB). The solid-state X-band EPR spectrum of

    [Fe([9]aneN2S)2](ClO4)3 revealed axial symmetry (g⊥ = 2.607, g|| = 1.599). At Q-band

    frequencies the perpendicular resonance was resolved into two components characteristic of a

  • Page 20 of 143

    Acce

    pted

    Man

    uscr

    ipt

    20

    rhombically distorted low spin Fe(III) centre. Computer simulation gave gx = 2.687, gy =

    2.526 and gz = 1.599. Based on the EPR analysis, the susceptibility data were fit to a simple

    model (H = –λL·S + µLz) to give the spin–orbit coupling constant (λ) of –260 ± 10 cm–1 and

    the axial ligand field parameter µ ~ 760 cm-1. The Mössbauer spectra (isomer shift/mm s–1,

    quadrupole splitting/mm s–1, 4.2 K) for [Fe([9]aneN2S)2][ClO4]2 (0.52, 0.57),

    [Fe([9]aneN2S)2][ClO4]3 (0.25, 2.72) and [Fe([9]aneNS2)2][ClO4]2 (0.43, 0.28) were typical

    for iron(II) and iron(III) complexes [94].

    The low-temperature single-crystal absorption spectra of [Fe([9]aneN2S)2](ClO4)2

    exhibited additional bands which resembled pseudo-tetragonal low-symmetry splitting of the

    parent octahedral 1A1g → 1T2g and 1A1g → 1T1g transitions [94]. However, the magnitude of

    this splitting was too large, requiring 10Dq for the thiaether donors to be significantly larger

    than for the amine donors. Instead, these bands were tentatively assigned to weak, low-energy

    S → FeII charge-transfer transitions. Above 200 K, thermal occupation of the high-spin 5T2g

    ground state resulted in observation of the 5T2g → 5Eg transition in the crystal spectrum of

    [Fe([9]aneN2S)2][ClO4]2. From a temperature-dependence study, the separation of the low-

    spin 1A1g and high-spin 5T2g ground states was approximately 1700 cm

    -1. The spectrum of

    the iron(III) complex [Fe([9]aneN2S)2][ClO4]3 was consistent with a low-spin d5

    configuration.

    Reaction of [9]aneN2S with an aqueous mixture of formic acid and formaldehyde

    resulted in the isolation in high yield of the N-methylated ligand N,N ′-dimethyl-1,4-diaza-7-

    thiacyclononane (Me2[9]aneN2S) [147]. Reaction of this ligand with ferric chloride in

    methanol resulted in the isolation of [Fe(Me2[9]aneN2S)Cl3]; a similar procedure with

    [9]aneN2S resulted in [Fe([9]aneN2S)Cl3]. Both complexes were characterised with X-ray

    crystallography. The addition of the N-methyl substituents made very little difference to the

    N-Fe-N and N-Fe-S bond angles with a small elongation observed (0.08 Ǻ) in the Fe-N bond

    length on going from [Fe([9]aneN2S)Cl3] to [Fe(Me2[9]aneN2S)Cl3] (Figures 8(a) and 8(b))

    [147].

  • Page 21 of 143

    Acce

    pted

    Man

    uscr

    ipt

    21

    (a) (b)

    Figure 8. Crystal structures of (a) [Fe([9]aneN2S)Cl3] and (b) [Fe(Me2[9]aneN2S)Cl3] [147]

    The µ-oxo-bis(µ-acetato)diiron(III) complex [Fe2O(O2CMe)2([9]aneN2S)2](PF6)2 was

    prepared by reaction of [Fe([9]aneN2S)Cl3] with sodium acetate and addition of ammonium

    hexafluorophosphate; an analogous reaction with [Fe(Me2[9]aneN2S)Cl3] gave

    [Fe2O(O2CMe)2(Me2[9]aneN2S)2](PF6)2 [147]. The iron(II) dimer

    [Fe2(OH)(O2CMe)2(Me2[9]aneN2S)2](ClO4) was prepared under anaerobic conditions. The

    two iron(III) complexes were characterised by crystal structural studies (Figures 9(a) and

    9(b)).

    (a) (b)

    Figure 9. Crystal structures of (a) [Fe2O(O2CMe)2([9]aneN2S)2]2+ and (b)

    [Fe2O(O2CMe)2(Me2[9]aneN2S)2]2+ [147]

    In these dimer complexes the Me2[9]aneN2S ligand offered less steric repulsion than the

    Me3[9]aneN3 analogue. A slight lengthening (~0.02 Ǻ) of the Fe…Fe distances between

    [Fe2O(O2CMe)2([9]aneN2S)2](PF6)2 and [Fe2O(O2CMe)2(Me2[9]aneN2S)2](PF6)2 was

    observed, reflecting an effect of addition of the N-methyl groups. The effect was not as

    pronounced as for the [9]aneN3 analogues where the increase in Fe….Fe distance

    between [Fe2O(O2CMe)2([9]aneN3)2])(PF6)2 and [Fe2O(O2CMe)2(Me3[9]aneN3)2](PF6)2

    was approximately 0.06 Å. As well, a potential distribution of products with the mixed

    sulfur–nitrogen ligands was possible, with the thioether donors being cis and/or trans

    with respect to the bridging oxo moiety, and in a gauche, anti or syn arrangement with

    respect to the Fe-O-Fe projection. However, for the [Fe2O(O2CMe)2([9]aneN2S)2]2+

    complex the crystal structure indicated that the product isolated displayed the

    thiaethers trans with respect to the bridging oxo unit, with S-Fe-O bond angles of 175.5(2)

    and 178.2(2)° and in a syn configuration with respect to the Fe-O-Fe plane. For the

    analogous [Fe2O(O2CMe)2(Me2[9]aneN2S)2]2+ complex, however, the structural analysis

  • Page 22 of 143

    Acce

    pted

    Man

    uscr

    ipt

    22

    indicated that the thiaethers were located cis to the bridging oxo group, and in a gauche

    configuration with respect to the Fe-O-Fe plane. The solid-state structure, of course, does

    not necessarily reflect that which exists in solution. The Mössbauer spectra of the two

    Fe(III) complexes [Fe2O(O2CMe)2([9]aneN2S)2](PF6)2 and [Fe2O(O2CMe)2(Me2[9]aneN2

    S)2](PF6)2 at 4.2 K and zero field consisted of a symmetric quadrupole doublet with isomer

    shifts of 0.48 and 0.49 mm s-1, respectively; the quadrupole splittings were 1.23 and 1.52

    mm s-1 and these data are consistent with similar high spin Fe-O-Fe type complexes. For the

    Fe(II) complex [Fe2(OH)(O2CMe)2(Me2[9]aneN2S)2](ClO4) the isomer shift (1.19 mm s-1)

    and quadrupole splitting (2.67 mm s-1) were consistent with other high spin binuclear Fe(II)

    complexes. The temperature dependence of the magnetic susceptibility, measured from

    300–4.2 K, for [Fe2O(O2CMe)2([9]aneN2S)2](PF6)2 (H = -2JS1·S2) indicated that the iron(III)

    sites were antiferromagnetically coupled with J = -125 cm-1 and g = 2.078; the exchange

    coupling for this complex has also been explored using density functional theory [148]. For

    the iron(II) complex [Fe2(OH)(O2CMe)2(Me2[9]aneN2S)2](ClO4) the least squares fit to the

    susceptibility data (H = -2JS1·S2) gave J = -7.4 cm-1 with g = 2.23 [147]. For

    [Fe2(OH)(O2CMe)2(Me3[9]aneN3)2](ClO4) J = -13 cm-1 [149].

    The Fe(II) complex of py2[9]aneN2S, [Fe(py2[9]aneN2S)Cl](CF3SO3) has been

    reported (Figure 10) [117].

    Figure 10. Crystal structure of the [Fe(py2[9]aneN2S)Cl]+ complex [117]

    The iron(II) complex, like the Mn(II) analogue, adopted the β-conformation and the complex

    was high spin with 5.17 µB, slightly higher than the spin only value of 4.90 µB [117]. The

    electronic spectrum in acetonitrile displayed a band at 25000 cm-1 (ε = 1710 M-1 cm-1)

    assigned to a metal-ligand charge transfer; on dissolution in water this band shifted to 27027

    cm-1 (ε 1460 M-1 cm-1), the species in solution proposed to be [Fe(py2[9]aneN2S)(H2O)]2+.

    Addition of base to form a complex presumed to be [Fe(py2[9]aneN2S)(OH)]+ and resulted in

  • Page 23 of 143

    Acce

    pted

    Man

    uscr

    ipt

    23

    a redshift of the absorption band to 24390 cm-1 (ε 1660 M-1 cm-1). A corresponding shift in

    the Fe(II/III) redox couple was observed with [Fe(py2[9]aneN2S)Cl]+/2+ E1/2 = 0.66 V,

    [Fe(py2[9]aneN2S)(H2O)]2+/3+ E1/2 = 0.70 V and [Fe(py2[9]aneN2S)(OH)]

    +/2+ E1/2 = 0.37 V

    (NHE) [117].

    The substituted [9]aneN2S ligand, 4,7-bis(2’-methyl-2’-mercaptopropyl)-1-thia-4,7-

    diazacyclononane (bmmpTASN) (Figure 11) as its iron complex, has been investigated

    extensively as a model NO-inactivated iron containing nitrile hydratase (NHases), the first

    non-heme bacterial enzyme characterised by a low spin Fe(III) state [71, 150-154].

    Figure 11. 4,7-bis(2’-methyl-2’-mercaptopropyl)-1-thia-4,7-diazacyclononane (bmmpTASN)

    NHases catalyse the hydration of nitriles to the corresponding amides; they exhibit a

    protein absorption band at 35700 cm-1 in the UV/visible spectrum with a shoulder at 25000

    cm-1 and a less intense band around 14280 cm-1, assigned tentatively as a thiolate- iron charge

    transfer band [155, 156]. The first X-ray structure revealed that NHase consisted of two

    subunits (α and β) with a basic stoichiometry of α1β1M1 (M = Fe(III)) with a N2S3(O) iron

    coordination sphere, the ligands being part of a small peptide, CysXYCysSerCys, the two

    nitrogen donors arising from the amides of the peptide main chain, and the three thiolates

    from the cysteines (Figure 12) [157].

    Figure 12. Active site structure of NO-inactivated iron containing nitrile hydratase; adapted

    from [157]

    S

    N

    N

    SH

    SH

    Fe

    SN

    SN

    O

    O

    O

    O

    OOH

    O

    S

    Cys109

    Henzyme

    enzyme

    Cys112

    Cys114

  • Page 24 of 143

    Acce

    pted

    Man

    uscr

    ipt

    24

    The active form of the enzyme NHaselight reacts with endogenous nitric oxide to yield an

    inactive form NHasedark, The nomenclature employed for this, [Fe-NO]6, and similar types of

    MNO complexes, is derived from the number of d electrons present [158]. In terms of the

    molecular orbital diagram for a six-coordinate complex with linear MNO groups, the electron

    configuration (3a1)2(2e)4(1b2)

    2 would lead to the [MNO]6 nomenclature [158]. The core of the

    NHasedark form exhibits the N2S3Fe-NO coordination with two amido nitrogen donors and

    three sulfur donors from cysteine-[159]. NHasedark is EPR silent and its UV-vis spectrum

    shows the two absorptions at 35700 and 27000 cm-1; however, the broad band at 14280

    cm-1 was absent, suggesting that the presence of this band is directly correlated to the

    enzyme activity. Resonance Raman studies showed that NO was released upon photo-

    irradiation [159].

    The complex [Fe(bmmpTASN)Cl], reacts with NO to give

    [Fe(bmmpTASN)(NO)](BPh4); the X-ray structure of the complex shows that the Fe(III) is

    pseudo-octahedral with the cyclononane macrocycle occupying one face, the two thiolate

    donors cis, and cis to the NO ligand, the arrangement mimicking the coordination of the

    cysteine donors in NHasedark (Figure 13) [71]. The complex was reported as an accurate

    structural model of the enzyme itself. The Fe-N(tertiary) bond distances (average 2.043(7) Ǻ)

    reproduce those seen in the NHasedark enzyme (2.07 Ǻ), as do the Fe-S distances (2.287(3) Ǻ

    model, 2.30 Ǻ enzyme) [71].

    Figure 13. Crystal structure of [Fe(bmmpTASN)(NO)]+ [71]

    The complex displays a υNO stretching frequency in the infrared spectrum at 1856 cm-1,

    whereas the NHasedark displays a band at 1852 cm-1 [71, 155, 156]. In addition, the

    Mössbauer spectra of the model and NHasedark are similar (isomer shift: 0.06 mm s-1, 0.03

    mm s-1, respectively; quadrupole splitting: 1.75 mm s-1, 1.47 mm s-1, respectively) [71, 160].

  • Page 25 of 143

    Acce

    pted

    Man

    uscr

    ipt

    25

    Reaction of [Fe(bmmpTASN)Cl] with Et4NCN results in [Fe(bmmpTASN)(CN)]

    [150, 151] (Figure 14(a)) which displays a υCN infrared stretch at 2083 cm-1; addition of base

    to [Fe(bmmpTASN)Cl] resulted in isolation of [Fe(bmmpTASN)(O)]2 (Figure 14(b)) with an

    infrared frequency at 799 cm-1 typical of a µ-oxo diiron complex [150].

    (a) (b)

    Figure 14. Crystal structures of (a) [Fe(bmmpTASN)(CN)] [151] and (b)

    [Fe(bmmpTASN)(O)]2 [150]

    The EPR spectra of [Fe(bmmpTASN)Cl] displayed a single line at g = 4.28, consistent with a

    high spin Fe(III) (S = 5/2) ground state in a rhombic environment; [Fe(bmmpTASN)(CN)]

    displayed a rhombic signal (g1 = 2.31, g2 = 2.16, g3 = 1.96) consistent with an S = ½, low

    spin, ground state [150]. The [Fe(bmmpTASN)(NO)](BPh4), NHasedark, mimic had an S = 0

    ground state [71, 158]. The authors concluded that the nature of the ligand X in the complex

    [Fe(bmmpTASN)(X)], rather than the presence of the two cis thiolato ligands, determined the

    high- or low-spin behaviour with π-accepting ligands promoting low-spin and the π-donating

    ligands high-spin [150]. The visible spectra of the four complexes were analysed in terms of

    their low spin ([Fe(bmmpTASN)(CN)], [Fe(bmmpTASN)(NO)](BPh4)) or high spin

    ([Fe(bmmpTASN)Cl] and [Fe(bmmpTASN)(O)]2) properties. The low spin complexes

    display a Fe(III)-thiolate charge transfer band at 15200 cm-1 and 15400 cm-1, respectively,

    whereas for the high-spin complexes the band occurs at 16000 cm-1 and 19010 cm-1,

    respectively [150]. The cyclic voltammograms of the four complexes display an irreversible

    oxidation ranging from +0.29 to +0.96 V (Ag/AgCl) assigned to a thiolate to thiyl oxidation

    [150]; in addition, the high spin complexes display a quasi-reversible reduction around -0.66

  • Page 26 of 143

    Acce

    pted

    Man

    uscr

    ipt

    26

    V, assigned as the Fe(III/II) couple. The low spin cyanide complex was, however, more

    difficult to reduce (-0.88 V) [150].

    Reaction of [Fe(bmmpTASN)Cl] with thallium triflate resulted in the isolation of

    [Fe(bmmpTASN)](CF3SO3) [152]. The complex displayed an S = ½ ground state in non-

    coordinating solvents with the room temperature µeff = 1.78 µΒ. The mass spectrum

    displayed m/z = 376.08 as expected for [Fe(bmmpTASN)]+ with no evidence of anion or

    solvent coordination under mass spectrometry conditions. The EPR spectrum in acetonitrile

    (77 K) confirmed binding of solvent with, in addition to a sharp axial signal of

    [Fe(bmmpTASN)]+ (g1 = 2.04, g2 = 2.02 and g3 = 2.01), a rhombic signal (g1 = 2.27, g2 =

    2.18 and g3 = 1.98) attributed to [Fe(bmmpTASN)(CH3CN)]+, the g values typical of low-

    spin iron(III). [Fe(bmmpTASN)(CH3CN)]+ was in equilibrium with [Fe(bmmpTASN)]+

    with a binding constant of Keq = 4.7 at room temperature [152]. [Fe(bmmpTASN)]+ was

    found also to coordinate a variety of solvents resulting in six-coordinate complexes of the

    form [Fe(bmmpTASN)(solvent)]+ (solvent = H2O, DMF, methanol, DMSO, and pyridine) to

    form high-spin six-coordinate complexes, the EPR spectra of which display significant strain

    in the rhombic zero-field splitting term E/D. Addition of triflic acid to

    [Fe(bmmpTASN)(O)]2 resulted in the formation of [(Fe(bmmpTASN))2OH](CF3SO3), the

    complex characterised by X-ray crystallography (Figure 15).

    Figure 15. Crystal structure of [(Fe(bmmpTASN))2OH]+ [152]

    In aqueous solution three distinct species were formed depending on solution pH:

    [Fe(bmmpTASN)(H2O)]+ (pKa = 5.04±0.1), [(Fe(bmmpTASN))2(OH)]

    + (pKa = 6.52 ±0.05)

    and [(Fe(bmmpTASN))2O] [152].

  • Page 27 of 143

    Acce

    pted

    Man

    uscr

    ipt

    27

    The crystal structure of NHasedark, showed that αCys114 had been post-

    translationally modified to a cysteine sulfenic acid (Cys–SOH) [159]. Thus, NHase requires

    oxygen addition to sulfur for modification of the cysteine residues to sulfur oxygenates. In

    order to model this, the oxygen sensitivity of [Fe(bmmpTASN)(CN)] and

    [Fe(bmmpTASN)Cl] has been examined [151]. Oxygen exposure of the low-spin complex

    [Fe(bmmpTASN)(CN)] over a period of several days resulted in the disulfonate complex

    [Fe(bmmp-O6-TASN)(CN)] as an olive-green solid (Figure 16) [151].

    Figure 16. [Fe(bmmp-O6-TASN)(CN)]; adapted from [151]

    The complex displayed characteristic peaks in the IR spectrum at 2062 cm-1 assigned

    to a υCN asymmetric stretch as well as bands consistent with the presence of sulfur-

    oxygenates, a result confirmed after 18O substitution [151]. Oxygen exposure of the high-

    spin complex [Fe(bmmpTASN)Cl] results in disulfide formation and decomplexation of the

    metal with subsequent iron-oxo cluster formation [151]. A natural bond order/natural

    localized molecular orbital covalency analysis revealed that the low-spin complex

    [Fe(bmmpTASN)CN] contained Fe–Sthiolate bonds with calculated covalency of 75 and 86%,

    while for the high-spin complex [Fe(bmmpTASN)Cl] the calculated covalencies were 11 and

    40% [151]. The results indicate the degree of covalency of the Fe–S bonds plays a major role

    in determining the reaction pathway associated with oxygen exposure of iron thiolates [151].

    The iron complex of the tetradentate ligand 4-((1-methyl-1H-imidazol-2-yl)methyl)-

    1-thia-4,7-diazacyclononane (mimTASN) [Fe(mimTASN)Cl2](FeCl4) has been

    characterised by X-ray crystallography (Figure 17) [154].

    S

    N

    HN N

    N

    Fe

    NN

    CNO

    S

    OS

    O O SO

    O

  • Page 28 of 143

    Acce

    pted

    Man

    uscr

    ipt

    28

    (a) (b)

    Figure 17. (a) 4-((1-methyl-1H-imidazol-2-yl)methyl)-1-thia-4,7-

    diazacyclononane (mimTASN) and (b) Crystal structure of [Fe(mimTASN)Cl2]2+ [154]

    The structure revealed that the iron(III) exhibited a pseudo-octahedral environment with the

    three nitrogen donors of the ligand coordinated facially. Replacement of the

    tetrachloridoferrate(III) anion with PF6- resulted in [Fe(mimTASN)Cl2](PF6) the structure of

    which revealed meridional coordination of the three nitrogen donors of the ligand [154].

    Cyclic voltammetry of [Fe(mimTASN)Cl2](PF6) in acetonitrile revealed a single Fe(III)/(II)

    reduction (-280 mV (versus Fc+/Fc). In methanol solution the cyclic voltammogram revealed

    a broad cathodic wave due to partial exchange of one chloride for methoxide with half-

    potentials of -170 mV and -440 mV for [Fe(mimTASN)Cl2]+/0 and

    [Fe(mimTASN)(OCH3)Cl2]+/0 with K(chloride exchange) = 7 x 10

    -4 M-1 for Fe(III) and 2 x 10-8

    M-1 for Fe(II). In aqueous solutions, and as a function of pH, three complexes are available

    after chloride exchange: in strongly acidic conditions the aqua complex

    [Fe(mimTASN)Cl(H2O)]2+ (pKa = 3.8 ±0.1), in mildly acidic conditions, the µ-OH complex

    [(Fe(mimTASN)Cl)2(OH)]3+ (pKa = 6.1 ± 0.3) and the µ-oxo complex

    [(Fe(mimTASN)Cl)2(O)]2+ under basic conditions [154].

    In addition to studies of the iron complex of bmmpTASN, as a model NO-inactivated

    iron-containing nitrile hydratase (NHases) [71, 150-154], the ruthenium complexes have also

    been investigated. In a series of papers the reactions of bmmpTASN, in addition to 4-(2′-

    methyl-2′-sulfinatopropyl)-7-(2′-methyl-2′-mercapto-propyl)-1-thia-4,7-diazacyclononane

    (bmmpO2TASN), and 4-(2′-methyl-2′-sulfinatopropyl)-7-(2′-methyl-2′-sulfenato-propyl)-1-

    thia-4,7-diazacyclononane) (bmmpO3TASN), in the form of their respective ruthenium(II)

    complexes, [Ru (bmmpOnTASN) PPh3] and [Ru(bmmpOnTASN) (PPh2CH3)] (n = 1–3), have

    been used to investigate how these complexes mimic the reaction of nitrile hydratase [161-

    166]. Under O2 under limiting conditions, the complex [Ru(bmmpTASN)(PPh3)] reacted to

    yield a mixed sulfenato/sulfinato product [Ru (bmmpO3TASN)(PPh3)] [162]. The complex

    prepared with 16O2 displayed intense bands at 1140 and 1020 cm−1 in the IR spectrum

    attributed to the asymmetric and symmetric S═O stretches, respectively, of the sulfinato

    donor. Reaction with 18O2 resulted in a shift of these bands to 1095 and 982 cm−1 for the

  • Page 29 of 143

    Acce

    pted

    Man

    uscr

    ipt

    29

    labelled [Ru(bmmpO3TASN)(PPh3)] complex. In addition, positive ion ESI-MS of

    [Ru(bmmpO3TASN)(PPh3)] prepared with 16O2, displayed a parent peak at m/z 731.1138

    which shifted to m/z 737.1267 for the complex prepared with 18O2 [162]. The crystal

    structure of [Ru(bmmpO3TASN)(PPh3)] (Figure 18) [162] showed the two oxygen donors of

    the sulfinato donor are directed along the S1−Ru−S3 bond axis, while the sulfenato oxygen is

    oriented toward a nitrogen donor along the P1−Ru−N1 axis. The triphenylphosphine ligand

    appears to impede access to the remaining potential thiaether oxygenation site, possibly

    retarding the rate of further oxygenation under limited O2 [162].

    Figure 18. Crystal structure of

    [Ru(bmmpO3TASN)(PPh3)] [162]

    A subsequent report explored the same chemistry of [Ru(bmmpTASN)(PPh3)] but

    showed that it was possible to isolate a number of sulfur oxygenated derivatives based on

    reaction time [165]. Thus, addition of five equivalents of O2 yielded the thiolato/sulfinato

    complex [Ru(bmmpO2TASN)(PPh3)] (Figure 19) within 5 minutes whereas a reaction time

    of 12 hours gave the sulfenato/sulfinato derivative [Ru(bmmpO3TASN)(PPh3)]. The bis-

    sulfinato complex [Ru(bmmpO4TASN)(PPh3)] was formed with either longer reaction times

    or additional O2; in all cases the oxidation state of the metal ion remained Ru(II) [165].

  • Page 30 of 143

    Acce

    pted

    Man

    uscr

    ipt

    30

    Figure 19. Crystal structure of [Ru(bmmpO2TASN)(PPh3)] [165]

    Stoichiometric hydrolysis of acetonitrile to acetamide by these complexes was investigated

    [165]. The hydrolysis reaction would require either dissociation of the phosphine leading to

    an open coordination site for substrate binding or direct participation of the S-oxygenate

    moieties. The elongation of the Ru−P bond distance as a function of the state of sulfur

    oxygenation suggested a transient five-coordinate intermediate after dissociation of the

    triphenylphosphine ligand. Solutions of [Ru(bmmpO2TASN)(PPh3)] and

    [Ru(bmmpO3TASN)(PPh3)] in a mixed solvent system of acetonitrile, methanol, and buffer

    (PIPES, pH = 7.0) gave small quantities of acetamide after 5 days; [Ru(bmmpTASN)

    (PPh3)] gave no product under the same reaction conditions [165].

    The precatalyst complexes [Ru(Ln)(PPh3)] (n = 1–3; Ln = bmmpTASN,

    bmmpO2TASN or bmmpO3TASN) have been used to study the rate of benzonitrile hydration,

    the study inspired by the metalloenzyme nitrile hydratase [71, 150-154, 161]. Previous

    studies had suggested that dissociation of the triphenylphosphine ligand was facile such that

    the open coordination site could be occupied by substrate, in this case the benzonitrile [165].

    The kinetic data were consistent with a mechanism involving initial activation by complete

    PPh3 dissociation to give the aqua complex [Ru(Ln)(OH2)] , which is in equilibrium with the

    nitrile complex [Ru(Ln)(NCR)]. Subsequent hydration of this nitrile complex through an

    activated water molecule resulted in [Ru(Ln)(NH2C(O)R)], the product complex. For the

    different ligands the hydration rate constants were reported to be 0.37 ± 0.01, 0.82 ± 0.07,

    and 1.59 ± 0.12 M–1 h–1 for L1 to L3, respectively [161]. Substitution of the amide by water

    completes the catalytic cycle [161].

    In the course of these studies the ruthenium(II) dimer [Ru2(bmmpTASN)2] was

    isolated and characterised (Figure 20) [166]. The authors reported that repeated attempts to

    reproduce the synthesis of this complex were

    unsuccessful [166].

  • Page 31 of 143

    Acce

    pted

    Man

    uscr

    ipt

    31

    Figure 20. Crystal structure of [Ru2(bmmpTASN)2] [166]

    Reaction of [(RuCl2(cym))2] and [9]aneN2S in methanol resulted in the mixed-

    sandwich Ru(III) complex [Ru(cym)([9]aneN2S)](BPh4)Cl2·MeCN (Figure 21) [133]. The

    structure shows the cyclononane coordinated facially through the two nitrogen and the

    thiaether donors, the macrocyclic ligand displaying λλλ (or δδδ) configuration. The p-

    cymene is η6 coordinated to the ruthenium(III) with Ru-C bonds in the range 2.196(7)–

    2.233(8) Å. The Ru(III)-N bond lengths were 2.120(7) Ǻ and 2.105(7) Ǻ with the Ru-S bond

    length being 2.324(2) Ǻ [133].

    Figure 21. Crystal structure of

    [Ru(cym)([9]aneN2S)]3+ [133]

    2.3.6 Cobalt(III), Rhodium(III),

    The synthesis of [Co([9]aneN2S)2](ClO4)3 was reported in a manuscript which first

    mentioned the synthesis of the [9]aneN2S ligand itself but concentrated on the chemistry of

    cobalt(III) complexes of two quinquidentate ligands, 4,7-bis(2-aminoethyl)-1-thia-4,7-

    diazacyclononane (dats) and 1,4-bis(2-aminoethyl)-1,4,7-triazacyclononane (datn) (Figure

    22), derived from [9]aneN2S and [9]aneN3, respectively [68]. The 13C NMR of

    [Co([9]aneN2S)2](ClO4)3 was reported to display eight resonances and it was predicted that in

    solution the complex existed as a mixture of cis- and trans-isomers; attempts to separate the

    isomers with chromatographic methods were reportedly unsuccessful [68]. Crystallisation of

    the complex from aqueous ethanol did result in isolation and characterisation by X-ray

    crystallography of the trans-S isomer of [Co([9]aneN2S)2](ClO4)3 [70] The synthesis of the

    complex was reported in a later paper, and again the authors reported the presence of

  • Page 32 of 143

    Acce

    pted

    Man

    uscr

    ipt

    32

    resonances in the 13C NMR indicative of the presence of cis- and trans-isomers, and the lack

    of success in their separation [114]. The two geometric isomers of [Co([9]aneN2S)2]3+ were

    ultimately separated by fractional crystallization and cation-exchange chromatography on SP-

    Sephadex [109]. Attempts to separate the enantiomers of the dissymmetric cis-complex were

    unsuccessful [109]. The isomers underwent a facile base-catalysed isomerisation to a

    cis/trans (4/1) mixture at room temperature with the rate being significant down to pH 5, the

    successful chromatographic separation requiring acidic conditions [109]. The X-ray crystal

    structures of both the cis and trans isomers of [Co([9]aneN2S)2]3+ were subsequently reported

    [167].

    The base hydrolysis of [Co(datn)Cl](ClO4)2 and [Co(dats)Cl](ClO4)2 proceeded for

    both complexes with two consecutive steps, the first reaction in each case being chloride

    hydrolysis (kOH(dats) = 3.6 x 104 M-1 s-1; kOH(datn) = 2.85 x 10

    3 M-1 s-1), the second reaction for

    the dats complex was proposed to be base-catalyzed dissociation of the thiaether resulting in

    the cis-[Co(dats)(OH)2]+ complex whilst for the datn complex the reaction was proposed to

    be base catalysed terminal ring opening (Co-N cleavage) [68].

    (a)

    (b)

    Figure 22. (a) 1,4,7-tris(2-aminoethyl)-1,4,7-

    triazacyclononane (datn) and (b) 1,4-bis(2-aminoethyl)-1-thia-4,7-diazacyclononane (dats)

    [Rh([9]aneN2S)Cl3].H2O was prepared by reaction of RhCl3·H2O and [9]aneN2S in

    refluxing ethanol [133]. The X-ray crystal structure showed that the six coordinate metal ion

    was coordinated by the cyclononane and three chlorido ligands in a facial arrangement

    [133]. The three five-membered chelate rings of [9]aneN2S had the λλλ (or δδδ)

    conformation. The Rh-N bond lengths (2.036(3) Ǻ and 2.040(3) Ǻ) are shorter than in

    [Rh([9]aneN3)2]Br3 (2.061 Ǻ) [168] and the Rh-S bond (2.246(1) Å) was shorter than in the

    [Rh([9]aneS3)2]3+ complex (2.331(2)–2.348(2) Å) [169, 170] suggesting Rh→S back

    donation. Of the three Rh-Cl bonds, one (2.396(1) Ǻ compared with 2.368(1) Ǻ and 2.358(1)

    Ǻ) shows a slight structural influence from the trans thiaether donor [133].

    3.3.7 Nickel(II/III), Platinum(II), Palladium(II)

    N

    S

    N

    H2N NH2

    N

    N

    N

    H2N

    H2N NH2

  • Page 33 of 143

    Acce

    pted

    Man

    uscr

    ipt

    33

    The complex [Ni([9]aneN2S)2](NO3)2 was prepared from reaction of nickel(II) nitrate and

    the ligand in an ethanol water mixture; purple crystals were obtained on standing [125]. The

    X-ray crystal structure of [Ni([9]aneN2S)2]2+ showed that the thiaether donors were trans,

    [125] whereas in the analogue [Ni(daes)2]2+ the sulfur donors were cis [171]. The Ni-N

    distances were 2.122(2) Ǻ and 2.108(2) Ǻ, close to the strain-free distance, while the Ni-S

    distance (2.418(1) Ǻ) was shorter than that seen for the [Ni(daes)2]2+ analogue (2.455 Ǻ)

    [171]. The coordination geometry around the metal ion in [Ni([9]aneN2S)2]2+ was distorted

    from octahedral with the angles around the metal ion significantly different from 90° [125].

    Potentiometric studies revealed that the Ni(II) binds the two ligands strongly with logK1 =

    10.45(2) and logK2 = 9.60(2) [125].

    Two analogues of the [9]aneN2S ligand, 15-thia-1,5,8,11-

    tetraazabicyclo[10.5.2]nonadecane (L1) and 1,11-dithia-4,8,14,18-tetraaza[5.2.2.5]eicosane

    (L2) have been reported (Figure

    23) [77, 172-174].

    (a)

    (b)

    Figure 23. (a) 15-thia-1,5,8,11-tetraazabicyclo[10.5.2]nonadecane (L1) and (b) 1,11-dithia-

    4,8,14,18-tetraaza[5.2.2.5]eicosane (L2) (syn- and anti-isomers)

    Both ligands were synthesised from a ligand base of 1,4,8,11-tetraazacyclononane (cyclam);

    1,11-dithia-4,8,14,18-tetraaza[5.2.2.5]eicosane was synthesised by reaction of 1,4,8,11-

    tetraazacyclononane with chloroacetyl chloride and subsequent reaction with sodium sulfide

    and reduction of the product with borane; anti- and syn-isomers of the ligand were isolated

    [172-174]. The ligand 15-thia-1,5,8,11-tetraazabicyclo[10.5.2]nonadecane was prepared

    after complexation of 4,7-bis(3-aminopropyl)-1-thia-4,7-diazacyclononane (Figure 24) with

    Cu(II), template addition of glyoxal, reduction with sodium borohydride and subsequent

    isolation of the free ligand [77].

    S

    N N

    HNNH

    S

    S

    N N

    NN

    S

    S

    N N

    NN

    anti isomer

    syn isomer

  • Page 34 of 143

    Acce

    pted

    Man

    uscr

    ipt

    34

    Figure 24. 4,7-bis(3-aminopropyl)-1-thia-4,7-diazacyclononane

    Upon complexation the cyclam backbone of 15-thia-1,5,8,11-

    tetraazabicyclo[10.5.2]nonadecane offers the potential for a number of stereochemical

    conformations, including five possible trans isomers (trans-I – trans-V) [175-180] as well as

    a cis isomer; incorporation of the [9]aneN2S fragment adds both anti- and syn-possibilities.

    The cis-[Ni( trans-I,syn-L2)]2+ and the trans-[Ni( trans-IV,anti-L2)]2+ complexes

    have been structurally characterised (Figure 25) [172].

    (a) (b)

    Figure 25. Crystal structures of (a) cis-[Ni( trans-I, syn-L2)]2+ and (b) trans-[Ni( trans-IV,

    anti-L2)]2+ complexes [172]

    The equatorial plane for the cis-[Ni( trans-I, syn-L2)]2+ ion is defined by two N-donors and

    two S-donors, the other two N-donors occupying the axial positions. The equatorial Ni–N

    distances are 2.164(3) and 2.179(3) Ǻ with the Ni–S distances being 2.4265(11) Ǻ and

    2.4512(11) Ǻ [172]. As is common in Ni(II) complexes with the cis-coordinated cyclam

    ligand the geometry around the nickel(II) centre is distorted. In the trans-[Ni( trans-IV, anti-

    L2)]2+ ion, the nickel sits at an inversion centre and the N-donors lie in the equatorial plane;

    the average Ni–N distance was 2.125(4) Ǻ with Ni-S distances of 2.5321(11) Ǻ - the S-

    S

    N

    N

    NH2

    H2N

  • Page 35 of 143

    Acce

    pted

    Man

    uscr

    ipt

    35

    donors occupy the axial positions. The cation exhibits D4h symmetry and the trans-IV

    configuration, the latter imposed by the anti-orientation of the [9]aneN2S rings. This trans-IV

    configuration has the highest strain energy and was the least favoured among the various

    possible stereochemical arrangements (trans-I – trans-V) adopted by the cyclam ring [172].

    The geometry imposed by the syn- and the anti-L2 is reflected in the spectroscopic

    data for the Ni(II) complexes [172]. The cis-[Ni(syn, trans-I, L2)]2+ ion shows absorption

    bands at 10020 cm-1 (ε = 26 M-1 cm-1 ) and 18590 cm-1 (ε = 16 M-1 cm-1 ) whilst the trans-

    [Ni(anti, trans-IV, L2)] 2+ ion has absorption bands at 10040 cm-1 (ε = 10 M-1 cm-1 ) and

    19160 cm-1 (ε = 11 M-1 cm-1 ) with 10Dq values of 10485 cm-1 for the cis-[Ni( trans-I, syn-

    L)] 2+ ion , and 16690 cm-1 for trans-[Ni( trans-IV anti-L2)]2+ ion, a value considered to be

    abnormally high [172]. The absorption spectra of Ni(II) complexes and the problems

    associated with assignment of the 10Dq value and the origins of the transitions is addressed

    later in this review [171, 172, 181, 182].

    Further studies were undertaken using cyclic voltammetry and EPR [172]. The

    trans-[Ni( trans-IV , anti-L2)]2+ ion showed a reversible wave (E1/2 = 0.91 V) and an axial

    EPR spectrum with g┴ = 2.18 and g║ = 2.01, consistent with D4h symmetry. The cis-

    [Ni( trans-I, syn-L2)]2+ complex exhibited more complicated and scan rate dependent

    electrochemical behaviour. At 1000 mV s-1, the redox cycle for the Ni2+/3+ couple showed a

    quasi-reversible wave with Ep,c = 1.35 V, Ep,a =1.45 V; lowering the scan rate to

  • Page 36 of 143

    Acce

    pted

    Man

    uscr

    ipt

    36

    complexes. The five-membered chelate rings adopt the gauche configuration with the six-

    membered rings in the chair form [172] .

    Figure 26. Crystal structure of [Ni(Ll)(ClO4)]+ [172]

    The UV-visible spectrum for [Ni(II)(L1)]2+ corresponded to that expected for a six-

    coordinate complex. The 3A2g → 3T1g(F) transition observed at 19840 cm-1 is at high energy

    for a high-spin Ni(II) system suggesting that the bicyclic ligand system exerts a strong

    ligand field on the metal centre. Cyclic voltammetry of the Ni(II) complex showed a

    reversible wave in aqueous solution for the Ni3+/2+ couple (0.765 V vs SCE) and two

    reversible waves in CH3CN corresponding to the Ni2+/+ and Ni3+/ 2+ couples (-1.675 V

    and 0.775 V vs Fc+/Fc, respectively). It was proposed that the additional six-membered

    chelate ring imparted an extra stabilisation on the Ni(III) complex attributed to (i) the

    cryptate effect from the additional chelate ring, and (ii) the formation of the 14-

    membered N4 ring which results in a stronger in-plane interaction, thus raising the

    energy of the eg orbitals and making it easier to remove an electron from the metal

    centre. The [Ni(III)(Ll)(H2O)]3+ complex ion exhibited an EPR spectrum characteristic of a

    low-spin d7 ion in a distorted octahedral environment with g┴ = 2.169 and g║ = 2.025

    (frozen solution 77 K) [172].

    The Ni(II) complex with py2[9]aneN2S crystallised with all five of the ligand donors

    bound; a water molecule occupied the sixth coordination site, and the pyridyl groups were

    situated cis (Figure 27) [183].

  • Page 37 of 143

    Acce

    pted

    Man

    uscr

    ipt

    37

    Figure 27. Crystal structure of [Ni(py2[9]aneN2S)(H2O)]2+ [183]

    The olive-green complex [Ni(bmmpTASN)] was prepared from reaction of

    NiCl2·6H2O with H2(bmmpTASN) in ethanol in the presence of KOH [151]. The complex

    reacted rapidly with hydrogen peroxide in acetonitrile, producing [Ni(bmmp-O6-TASN)].

    The X-ray crystal structures of both complexes have been determined [151]. For

    [Ni(bmmpTASN)] the metal ion is in a square-planar environment with an N2S2 donor set.

    The Ni–S(thiolate) and Ni–N bond distances (2.1775(7) Ǻ and 2.1925(7) Ǻ; 1.957(2) Ǻ and

    1.982(2) Ǻ, respectively) are typical for similar square-planar Ni(II) complexes. The axial

    thiaether sulfur of the [9]aneN2S ligand is essentially non-bonding (2.824 Ǻ). The structure

    of [Ni(bmmp-O6-TASN)] has the metal ion in a pseudo-octahedral coordination

    environment, the donor set consisting of the two amine donors (2.103(4) Ǻ and 2.139(5) Ǻ) ,

    the thiaether sulfur (2.3847(14) Ǻ) and the η1 (1.991(4) Ǻ) and η2 sulfonate donors (2.134(4)

    Ǻ and 2.283(4) Ǻ) [151].

    The complexes [Pd([9]aneN2S)2](PF6)2 and [Pt([9]aneN2S)2]Br2.H2O were prepared

    by reaction of the ligand with Pd(II) acetate and K2PtCl4, respectively [184]. The X-ray

    crystal structures of both complexes have been reported. Under different solvent (3:1

    nitromethane and dichloromethane) and temperature conditions, reaction of PdCl2 and

    [9]aneN2S resulted in [Pd([9]aneN2S)2]Cl2.H2O at room temperature and cis-

    [Pd([9]aneN2S)Cl2] at 80oC. Attempts to produce X-ray quality crystals from cis-

    [Pd([9]aneN2S)Cl2] resulted in the dark red [Pd3([9]aneN2S)4Cl2]Cl4.2H2O complex [185].

    The syntheses of the complexes [Pt([9]aneN2S)2](PF6)2 and [Pt([9]aneN2S)2]Br2 from

    reaction with K2PtCl4 and ligand in aqueous solution have also been reported, the bromide

    anion in the latter arising apparently from the use of the hydrobromide salt of the

    cyclononane ligand [184, 185].

    The structures of the [Pd([9]aneN2S)2](PF6)2 and [Pd([9]aneN2S)2]Cl2·H2O

    complexes differ with respect to both the conformation of the cyclononane ligand and as a

    result the coordination of the thioether group [184, 185]. The [333] type conformation in the

    latter means that the thiaether donor is directed away from the metal ion, whereas the [234]

    conformation in the former leads to a weak interaction with the metal ion. In the case of the

    platinum complexes [Pt([9]aneN2S)2](PF6) and [Pt([9]aneN2S)]Br2 both adopt a [333]

  • Page 38 of 143

    Acce

    pted

    Man

    uscr

    ipt

    38

    conformation where the thiaether is non-coordinating [184, 185]. Based on these

    observations, the fluxional behaviour observed previously in solution for the

    [Pd([9]aneN3)2]2+ analogue was speculated to be a result of an isomerisation process

    between the [234] and [333] conformation of the ligand [185, 186] .

    For [Pd([9]aneN2S)2](PF6)2 the metal atom exhibits a distorted octahedral geometry

    made up of the two nitrogen and one thiaether donor of each facially coordinated

    cyclononane macrocycle [184]. The four nitrogen donors (two from each ligand) form a

    square plane around the metal ion with Pd-N distances of 2.054(3) Ǻ and 2.064(3) Å [184].

    The interaction with the axial thiaether donors is weak, with a Pd-distance of 3.034(1) Å.

    The [234] conformation of the ligand is similar to that reported for the gold(III) complex

    [Au([9]aneN2S)Cl2][AuCl 4] where again the thiaether donor exhibits a long apical interaction

    (2.973(3) Ǻ) [112]. The structure of the platinum(II) complex [Pt([9]aneN2S)2]Br2 complex

    is different [184]. Again, the metal ion is coordinated to four of the nitrogen atoms of both

    cyclononane ligands in a square planar geometry but the thiaethers are folded out and away

    from the central atom, the ligand having a [333] type conformation. The Pt-N bond lengths

    (2.04(2) Ǻ and 2.04(1) Ǻ) are shorter than those for the Pd(II) analogue [184]. Hydrogen

    bonding interactions, average length 3.31-3.41 Å, involving the bromide anions and the

    nitrogen atoms and the water molecule are evident in the structure [184]. The cyclic

    voltammogram of [Pd([9]aneN2S)2](PF6)2 in acetonitrile exhibited a quasi-reversible single

    electron oxidation, E1/2 = +0.30 V (Fc+/Fc) assigned to a Pd(II)/Pd(III) redox pair [184].

    The complex [Pd3([9]aneN2S)4Cl2]+ resulted from attempts to obtain crystals of cis-

    [Pd([9]aneN2S)Cl2] [185]. The structure consists of a central [Pd([9]aneN2S)2]2+ cation

    bridged through its thiaether donor atoms to two [Pd([9]aneN2S)Cl] moieties; there are

    additional Cl- ions acting as counter-ions. The central Pd(II) is coordinated to four nitrogen

    donors of two [9]aneN2S ligands in a square-planar configuration, the thiaether donors from

    each cyclononane ligand of the central [Pd([9]aneN2S)2]2+ cation displaying long apical

    interactions (3.008(2) Ǻ) with one halide from the [Pd([9]aneN2S)Cl] moieties [185].

    Reaction of py2[9]aneN2S with [Pd(CH3CN)4](BF4)2 resulted in

    [Pd(py2[9]aneN2S)](BF4)2 [187]; a separate study reported the same complex as the

    hexafluorophosphate salt [183]. The X-ray structure of the tetrafluoroborate salt showed that

    the