177
Coordination chemistry and reactivity of early transition metal complexes with the amine phenolate ligands Thesis for the Degree of “Doctor of PhilosophyBy Stanislav Groysman Submitted to the Senate of Tel Aviv University September 2005

Coordination chemistry and reactivity of early …primage.tau.ac.il/libraries/theses/exeng/free/2021687.pdf5. S. Groysman, S. Segal, M. Shamis, I. Goldberg, M. Kol, Z. Goldschmidt

  • Upload
    others

  • View
    7

  • Download
    0

Embed Size (px)

Citation preview

  • Coordination chemistry and reactivity of early transition

    metal complexes with the amine phenolate ligands

    Thesis for the Degree of “Doctor of Philosophy”

    By

    Stanislav Groysman

    Submitted to the Senate of Tel Aviv University

    September 2005

  • Coordination chemistry and reactivity of early transition

    metal complexes with the amine phenolate ligands

    Thesis for the Degree of “Doctor of Philosophy”

    By

    Stanislav Groysman

    Submitted to the Senate of Tel Aviv University

    September 2005

  • This work was carried out under the supervision of Professor Moshe Kol and

    Professor Israel Goldberg.

  • I am deeply grateful to my teacher and mentor, Professor Moshe Kol (Shiko),

    for guidance, support, brilliant ideas, and never-ending inspiration.

    I wish to thank Professor Israel Goldberg for his guidance in the beautiful field of X-

    ray crystallography.

    I wish to thank my group members, Dr. Edit Tshuva, Dr. Dalia Gut-Regev, Adi Yeori,

    Shimrit Gendler, Sharon Segal, Ekaterina (Katy) Sergeeva, and Sheba D. Bergman,

    for their support. Especially, I wish to express my deep appreciation to Mrs. Dvora

    Reshef for the everyday help, and for the help in olefin polymerization project.

    I wish to thank Professor Zeev Goldschmidt and Dr. Elisheva Genizi from Bar Ilan

    university, and Dr. Michael Shuster from Carmel Olefins Ltd, for our fruitful

    collaboration.

  • This work is dedicated to my parents, Alexander and Galina, and to my wife,

    Marianna.

  • Table of contents

    1 List of publications

    3 Abstract

    15 Introduction

    15 I-1. Alkoxo (phenolate) ligand vs. Cp ligand

    16 I-2. From a simple alkoxo to a bulky phenolate ligand

    17

    I-3. Multidentate ligands: Enhanced stabilization and easier design of

    metal centers

    19 I-4. Multidentate phenolate ligands

    25

    I-5. Synthetic pathways to early transition metal complexes bearing

    phenolate ligands

    26

    I-6. Reactivity of early transition metal complexes supported by

    multidentate phenolate ligands

    27 I-6.1. Olefin polymerization by well-defined Group IV metal complexes

    32

    I-6.2. Stabilization of metal-carbon double bonds and olefin metathesis

    reactivity

    34 I-6.3. Phenolate ligands in bioinorganic chemistry

    37 I-7. Amine phenolate ligands

    38 I-7.1. The amine tris(phenolate) ligands

    39 I-7.2. The amine bis(phenolate) ligands with the additional, sidearm, donor

    41 I-7.3. The amine mono(phenolate) ligands with two sidearm donors

    43 Discussion

    43 D-1. Ligand synthesis

    45 D-2. Complex synthesis

    47 D-3. Coordination chemistry of the amine phenolate metal complexes

  • 54

    D-4. Reactivity of the early transition metal complexes with amine

    phenolate ligands

    54 D-4.1. Olefin polymerization

    60 D-4.2. H-abstraction reactions

    63 D-4.3. Modeling the active site of haloperoxidase

    67 References

    72 Appendix: Symbols and abbreviations

    Publications

    תקציר א

  • List of publications

    1

    List of publications presented in this work

    1. S. Groysman, I. Goldberg, M. Kol, E. Genizi and Z. Goldschmidt Group IV complexes of an Amine Bis(phenolate) Ligand Featuring a THF Sidearm Donor: From Highly Active to Living Polymerization Catalysts of 1-Hexene Inorg. Chim. Acta 2003 345, 137.

    2. S. Groysman, I. Goldberg, M. Kol, E. Genizi and Z. Goldschmidt From THF to Furan: Activity Tuning and Mechanistic Insight via Sidearm Donor Replacement in Group IV Amine Bis(phenolate) Polymerization Catalysts

    Organometallics 2003, 22, 3013. 3. S. Groysman, E. Y. Tshuva, D. Reshef, S. Gendler, I. Goldberg,M. Kol, Z.

    Goldschmidt, M. Shuster and G. Lidor High-Molecular Weight Atactic Polypropylene Prepared by Zirconium Complexes of an Amine Bis(phenolate) Ligand Isr. J. Chem. 2002, 42, 373.

    4. S. Groysman, E. Tshuva, I. Goldberg, M. Kol, Z. Goldschmidt, and M. Shuster Diverse Structure-Activity Trends in the Amine Bis(phenolate) Titanium Polymerization Catalysts Organometallics 2004, 23, 5291.

    5. S. Groysman, S. Segal, M. Shamis, I. Goldberg, M. Kol, Z.

    Goldschmidt and E. Hayut-Salant Tantalum(V) Complexes of an Amine Triphenolate Ligand: a Dramatic Difference in Reactivity Between the Two Labile Positions J. Chem. Soc. Dalton Trans. 2002, 3425.

    6. S. Groysman, S. Segal, I. Goldberg, M. Kol and Z. Goldschmidt Ta(V) Complexes of a Bulky Amine Tris(phenolate) Ligand: Steric Inhibition vs. Chelate Effect

    Inorg. Chem. Commun. 2004, 7, 938. 7. S. Groysman, I. Goldberg, M. Kol and Z. Goldschmidt

    Pentabenzyltantalum: Straightforward Synthesis, X-ray Structure and Application in the Synthesis of [O2N]TaBn3-Type and [O3N]TaBn2-Type complexes Organometallics 2003, 22, 3793.

    8. S. Groysman, I. Goldberg, M. Kol, E. Genizi and Z. Goldschmidt

    Tribenzyl Tantalum(V) Complexes of Amine Bis(phenolate) Ligands: Investigation of α-Abstraction vs. Ligand Backbone β-Abstraction Paths Organometallics 2004, 23, 1880.

    9. S. Groysman, I. Goldberg, M. Kol, E. Genizi and Z. Goldshmidt Exploring Routes to Tantalum(V) Alkylidene Complexes Supported by

  • List of publications

    2

    Amine Tris(Phenolate) Ligands Adv. Synth. Cat 2005, 347, 409.

    10. S. Groysman, I. Goldberg, M. Kol and Z. Goldschmidt Vanadium(III) and Vanadium(V) Amine Tris(phenolate) Complexes Inorg. Chem. 2005, 44, 5073. 11. S. Groysman, E. Sergeeva, I. Goldberg and M. Kol Group IV Complexes of a Tetradentate Amine Mono(phenolate) Ligand: A Second Sidearm Donor Stabilizes Cationic Species Inorg. Chem. 2005, 44, 8188. 12. S. Groysman, E. Sergeeva, I. Goldberg and M. Kol Salophan complexes with Group IV Metals Eur. J. Inorg. Chem. 2005, 2480.

  • Abstract

    3

    Abstract

    Living polymerization of 1-hexene by titanium amine bis(phenolate) complexes

    This work is fully described in articles 1 and 2.

    Ti complexes of the amine bis(phenolate) ligands having bulky t-Bu groups in the

    ortho positions of the phenolate rings, and possessing a sidearm donor, have all shown a

    tendency to catalyze a living polymerization, to a certain degree. Aiming at the most living

    catalyst of this system, we postulated that a strong sidearm oxygen donor obtained via a THF

    substituent, would induce a pronounced living polymerization activity of the derived catalyst.

    TiO

    R

    t-Bu

    t-Bu

    O

    R

    Nt-Bu

    Ot-Bu

    R = CH2Ph, Me

    TiO

    R

    t-Bu

    t-Bu

    O

    R

    Nt-Bu

    Ot-Bu

    R = CH2Ph

    Figure A-1. Dialkyl Ti complexes bearing the amine bis(phenolate) THF and furan ligands.

    Dialkyl Ti-THF complexes (Figure A-1) were activated toward 1-hexene by the

    reaction with B(C6F5)3. This polymerization system showed an unprecedented behavior: The

    polymerization was living for almost a week (at RT), and the resulting polymer had a very

    high Mw (around 106), and a very narrow MWD (PDI = 1.09) (Figure A-2). In addition, this

    catalytic system enabled the copolymerization of two α-olefins, 1-hexene, and 1-octene at

    RT, forming a di-block copolymer of narrow MWD. To further support our hypothesis that

    the pre-condition for the living polymerization in the amine bis(phenolate) catalysts was the

    dative strength of the oxygen sidearm donor, we prepared the dibenzyl amine bis(phenolate)

    Ti complex, having a weak oxygen donor: furan. The structure of the analogous Zr complex

    indicated that the furan oxygen was bound to the metal, and that the bonding was indeed

  • Abstract

    4

    weak. As anticipated, the Ti-furan/B(C6F5)3 system showed a different character of the

    polymerization catalysis, when compared with Ti-THF: the polymerization in this system was

    ca. 10-fold faster, but was not “truly” living, as the MWD was narrow only in the first few

    hours of the reaction. Further investigation of the THF-Ti and furan-Ti catalysts enabled a

    mechanistic insight into the reactivity of the amine bis(phenolate) catalysts, supporting a

    mononuclear nature of the latter, and displaying no polymeryl transfer.

    Figure A-2. “Immortal” polymerization of 1-hexene by THF-Ti: dependence of molecular weight (Mw) on time.

    The numbers indicate the polydispersity index (PDI) values of the polymer samples.

    Highly active zirconium catalysts for 1-hexene and propylene polymerization

    This work is fully described in articles 1, 2 and 3.

    In analogy to the Ti species, the dibenzyl amine bis(phenolate) Zr complexes,

    providing the most convenient form of the pre-catalyst, were prepared by a toluene-

    elimination reaction of tetrabenzylzirconium with the ligand precursor. The activation of the

    THF-Zr complex with a borane co-catalyst gave rise to a highly efficient 1-hexene

    polymerization catalyst, whose activity (21,000 gpol mmolcat-1 h-1) is defined as “very high”.

    The resulting polymer was of relatively high Mw, having a PDI of around 2, indicating a

    “single-site” species. Similarly, the furan-Zr catalyst was highly active in 1-hexene

    polymerization. In contrast to the THF-Zr, furan-Zr led to a low-Mw polymer.

    1.12

    1.061.06

    1.051.07

    1.061.11

    1.09

    0100000200000300000400000500000600000700000800000900000

    0 2 4 6 8Polym. Time (days)

    Mol

    ecul

    ar W

    eigh

    t

  • Abstract

    5

    Next, we investigated the reactivity of the Zr-based catalysts in the polymerization of

    the industrially challenging monomer, propylene. Two types of the pre-catalyst were

    employed: The first having two benzyl groups at the labile positions, and the second having

    two chloro groups. Both pre-catalysts, upon activation with MAO, led to a very efficient

    polymerization of propylene (Scheme A-1). The optimized activity of this system was as high

    as 13,000 gpol mmolcat-1 h-1. The resulting polypropylene was uncommon, combining

    complete atacticity and relatively high molecular weight, thereby displaying elastomeric

    properties.

    Menn

    Atactic polypropylene

    Mw = 8,000 - 400,000250 - 1,100 equiv of MAO Me

    Zr pre-catalyst

    Scheme A-1. Propylene polymerization by the amine bis(phenolate) Zr catalyst.

    Influence of the phenolate substituents on the reactivity of titanium catalysts

    This work is fully described in article 4.

    Aiming at developing of new activity modes, and discovering new structure activity

    relationships, we prepared a series of Ti-based catalysts, possessing ortho (para) substituents

    of varying steric and electronic nature (Scheme A-2). As the reactivity of these systems may

    depend on the nature of the sidearm donor, two such series were prepared: One possessing a

    NMe2 donor, and the other possessing an OMe donor. For the NMe2 series, the gradual

    reduction of steric hindrance led to a considerable increase of the catalyst activity: The

    catalyst having ortho Me groups had an activity of 1,400 vs. 30 for the catalyst having t-Bu

    groups. An additional increase in activity (up to 8,000) was obtained using the chloro-

    substituted ligands. Surprisingly, the reduction of the steric bulk, and use of electron-

    withdrawing groups had no effect on the activity of catalysts featuring the OMe sidearm

  • Abstract

    6

    donor. The 13C NMR characterization of the polymer samples shed light on the origin of

    these diverse trends for the two series: All catalysts of the NMe2 series produced a relatively

    regioregular poly(1-hexene), whereas the reduction of the steric bulk in the OMe series led to

    a substantial increase in the degree of regioerrors. The latter are known to decrease

    polymerization rate significantly.

    TiO

    PhH2C

    t-Bu

    t-Bu

    D

    CH2Ph

    Nt-Bu

    Ot-Bu

    t-Bu vs. Me

    D = NMe2, OMe

    TiO

    PhH2C

    Me

    Me

    D

    CH2Ph

    NMe

    OMeTi

    O

    PhH2C

    Me

    D

    CH2Ph

    N

    OMe

    TiO

    PhH2C

    Cl

    Cl

    D

    CH2Ph

    NCl

    OCl

    Me

    Me Me vs. Cl

    D = NMe2, OMeD = NMe2, OMeD = NMe2

    Me vs. H

    Scheme A-2. Dibenzyl amine bis(phenolate) Ti(IV) complexes, displaying varying steric and electronic

    influence at the metal center

    The most efficient catalyst of these series led to an unusual polymer: Poly(1-hexene)

    of an extremely high molecular weight (Mw = 4,300,000). In contrast to the “regular” poly(1-

    hexene), which is a sticky oily substance, this polymer is in a rubbery state, showing the

    properties of an elastomer.

    Coordination chemistry of Ta(V) complexes with amine tris(phenolate) ligands

    This work is fully described in articles 5 and 6

    Wishing to extend the chemistry of the amine phenolate ligands beyond group IV

    metals, we turned to investigate their chemistry with the group V metal, tantalum (V). The

  • Abstract

    7

    amine tris(phenolate) ligand, bearing Me groups at the ortho, para positions of the phenolate

    rings, was chosen as a first ligand in this study. Two metal precursors, Ta(OEt)5 or

    Ta(NMe2)5, reacted successfully with a ligand precursor, leading to a well-defined rigid

    octahedral complexes of Cs-symmetry (Scheme A-3). Subsequently, these complexes were

    treated with excess of Me3SiCl. Surprisingly, only one of the “labile” monodentate ligands

    was exchanged. The solid-state structure determination revealed that the remaining OEt

    (NMe2) group is trans to the central amine donor. The differentiation between the “labile”

    positions was so strong, that the remaining OEt group did not undergo a metathesis reaction

    at all, even when the reaction was carried out for weeks, or when stronger reagents (HCl, for

    example) were employed. The desired dichloro complex could be obtained only via the

    replacement of the more basic dimethylamido group, after prolonged reaction times.

    TaO

    X

    Me

    Me

    O

    X

    NMe

    OMe

    Me

    Me

    HO

    N

    OHMe

    Me

    Me

    Me

    HO

    Me

    Me

    TaX5

    X = OEt, NMe2

    TaO

    Cl

    Me

    Me

    O

    X

    NMe

    OMe

    Me

    Me

    xs. Me3SiCl

    xs. Me3SiClTa

    O

    Cl

    Me

    Me

    O

    Cl

    NMe

    OMe

    Me

    Me

    X = OEt, NMe2

    4 days, X = NMe2

    TaCl5

    Scheme A-3. Coordination chemistry of Ta(V) in the amine tris(phenolate) environment

    In contrast to the non-bulky ligand, the reaction of the sterically-crowded amine

    tris(phenolate) ligand with Ta(NMe2)5 proceeded slowly, via a relatively stable dinuclear

    intermediate, whose nature was established by means of X-ray diffraction. When formed, the

    bis(dimethylamido) complex of the sterically-crowded ligand did not lead to a clean

  • Abstract

    8

    metathesis reaction with any chlorinating agent, while the analogous di(ethoxo) complex

    allowed a clean substitution of a single group.

    Synthesis, structure, and reactivity of pentabenzyltantalum

    This work is fully described in article 7.

    Pentabenzyltantalum was prepared by Schrock in 1976 by a two-step reaction

    sequence: a reaction between tantalum pentachloride and dibenzyl zinc to produce

    TaCl2(CH2Ph)3, which was further reacted with dibenzylmagnesium. We were able to

    develop a more direct procedure, by preparation of pentabenzyltantalum via the single-step

    reaction of TaCl5 with PhCH2MgCl. As the X-ray structure of this fundamental compound

    was never reported, we solved its crystal structure as well (Figure A-3). Subsequently, we

    evaluated the potential of Ta(CH2Ph)5 as a precursor to Ta(V) organometallic complexes with

    the amine phenolate ligands. Two representative ligands of the amine bis(phenolate) and the

    amine tris(phenolate) types were reacted with Ta(CH2Ph)5. Both ligands reacted smoothly,

    leading to the expected tribenzyl, and dibenzyl complexes, respectively, in high yields.

    Figure A-3. Molecular structure of pentabenzyltantalum.

  • Abstract

    9

    Investigation of α-abstraction vs. ligand backbone β-abstraction paths in the tribenzyl

    tantalum(V) complexes of amine bis(phenolate) ligands

    This work is fully described in article 8.

    A variety of tribenzyl amine bis(phenolate) Ta(V) complexes could be prepared

    smoothly using the pentabenzyltantalum route. All these complexes displayed a similar

    propensity: They are hexacoordinate, and the fourth, sidearm donor is a “dormant” donor in

    these species. We postulated that such a “dormant” donor may cause various organometallic

    transformations in the parent tribenzyl species, and thus we carried out a thorough SAR study

    of the H-abstraction reactions in this system as a function of various structural modifications.

    TaO

    PhH2C

    R

    R

    D

    CHPh

    NR

    ORTaO

    PhH2C

    R

    R

    D

    CH2Ph

    NR

    OR

    α-Hβ-H

    C

    Ta

    N

    PhH2C O

    O CH2Ph

    Ph HH

    DHH

    R

    R

    α

    β

    Scheme A-4. Possible H-abstraction routes for the tribenzyl complexes

    When subjected to thermolysis, these species underwent two well-defined reactions:

    Abstraction of an �-proton (from one of the benzyl groups), forming the metal-carbon double

    bond (benzylidene), or abstraction of a β-proton (from the benzyl position in the amine

    bis(phenolate) ligand), forming a new metal-carbon single bond (Scheme A-4). The choice of

    the abstraction pathway was found to depend on two structural parameters: The presence of

    the “dormant” donor, and the steric bulk at the ortho positions of the phenolate rings. The

    presence of the sidearm donor, accompanied by small (H, Cl) ortho substituents led to

    smooth �-H abstraction. When the sidearm donor was present, and the ortho substituents were

  • Abstract

    10

    bulkier, the main reaction was facile β−Η abstraction. In one particular case, i.e. t-Bu

    phenolate substituents, the β−Η activation took place even at RT. In these cases, the sidearm

    donor was bound to the metal following the H-abstraction and toluene elimination. When the

    donor was absent (an [ONO]-type ligand), the reaction took the β−Η abstraction pathway

    exclusively, and proceeded much slower. All the H-abstraction reactions were of first order in

    the reactant, indicating an intramolecular mechanism.

    Exploring routes to tantalum(V) alkylidene complexes supported by amine tris(phenolate)

    ligands.

    This work is fully described in article 9.

    At the next step, an alkylidene functionality supported by amine tris(phenolate)

    ligands was pursued. On the way to the desired alkylidene species we employed two

    fundamental organometallic species, tris(neopentyl) mono(neopentylidene) tantalum(V)

    (Ta(CH2t-Bu)3(CHt-Bu)), and pentabenzyltantalum, as starting materials. The reaction of the

    amine tris(phenolate) ligand precursor with Ta(CH2t-Bu)3(CHt-Bu) proceeded via the O-H

    addition to the metal-carbon double bond, forming a phenoxo-tetraneopentyl complex, one of

    the most sterically congested organometallic species. Further reactivity of this complex was

    less well-defined, as it decomposed into unidentified species, not containing an alkylidene

    function.

    The pentabenzyltantalum route led to the octahedral dibenzyl complexes with a

    variety of amine tris(phenolate) ligand precursors. Unlike the tribenzyl amine bis(phenolate)

    complexes, the dibenzyl complexes were found to be remarkably stable toward thermolysis,

    decomposing only at 120 ºC after prolonged reaction times. Instead of forming the

    mononuclear terminal alkylidene functionality, the reaction unexpectedly led to the rare

    dinuclear µ-benzylidene complex, in which both Ta(V) centers were of octahedral geometry

    (Figure A-4).

  • Abstract

    11

    Figure A-4. Molecular structure of µ-benzylidene di-Ta(V) complex.

    Vanadium(III) and Vanadium(V) amine tris(phenolate) complexes

    This work is fully described in article 10.

    Following the investigation of the chemistry of the heavier group V member,

    tantalum, we turned to the lighter group V member, vanadium. The chemistry of this metal

    was studied especially in light of its bioinorganic relevance, as vanadium participates in the

    active site of several enzymes. Three amine tris(phenolate) ligands featuring different steric

    and electronic influence were employed in this study (Scheme A-5). V(III) complexes of

    these ligands were obtained by reaction between the ligand precursors and VCl3(THF)3, in the

    presence of Et3N. All the resulting complexes were of almost perfect TBP geometry, in which

    the three phenolate oxygens lied in the equatorial positions, one axial positions was occupied

    by the central nitrogen, and the other by an additional ligand: THF. The V(III) complexes of

    the “electron-rich” ligands, having ortho, para Me and t-Bu phenolate substituents, were

    found to undergo smooth and clean oxidation to V(V) oxo species, when exposed to air.

    Alternatively, the V(V) complexes could be prepared directly, starting from VO(OPr)3. The

    crystallographic and multinuclear NMR studies of these complexes supported their structure

  • Abstract

    12

    to be TBP, similar to the V(III) species, with the sole difference being replacement of the

    axial neutral THF ligand by the dianionic oxo ligand. This structure is noteworthy: It mimics,

    to a great extent, the structure of the (resting) active site in V-dependent haloperoxidase.

    Scheme A-5. Coordination chemistry of V(III) and V(V) amine tris(phenolate) complexes

    The reaction of the “electron-poor” ortho, para dichloro substituted ligand with

    VO(OPr)3 resulted in the formation of V(V) species having a different structure. The X-ray

    solution disclosed an octahedral amine tris(phenolate) complex, carrying an aqua ligand, in

    addition to the oxo ligand. In addition to the octahedral species, the products mixture

    contained a small fraction of a TBP complex. Thus, it appears that the amine tris(phenolate)

    complexes are able to switch between the TBP and octahedral geometries. In addition, the

    V(V) amine tris(phenolate) species have shown an oxygen-transfer reactivity, being able to

    catalyze, albeit slowly, oxygen transfer from a peroxide to styrene and stilbene.

    Group IV complexes of a tetradentate amine mono(phenolate) ligand: a second extra donor

    stabilizes cationic species

    This work is fully described in (submitted) article 11.

    VO

    OO

    NR

    R

    R

    RR

    R

    O

    VO

    L

    Cl

    Cl

    ClO

    O

    NCl

    O

    Cl

    Cl

    N

    OHHO

    OH

    R

    R

    R

    R

    R

    RVCl3(THF)3

    VO

    OO

    NR

    R

    R

    RR

    R

    O

    R =

    R = Me, t-Bu, Cl

    Me, t-BuairVO(OPr)3

    R = Cl

    VO(OPr)3R = Me, t-Bu

    - L, ∆

    + L

  • Abstract

    13

    The amine mono(phenolate) ligand precursor was prepared by a single-step Mannich

    condensation between a secondary amine, a substituted phenol, and formaldehyde. Upon

    reaction with suitable Group IV metal precursors, this ligand precursor led to the highly

    flexible tris(alkoxo) complexes with Ti(IV) and Zr(IV), and the tribenzyl complex with

    Zr(IV) metal. At the next step, the removal of one of the alkoxo (alkyl) groups by means of

    Lewis or Bronsted acid (B(C6F5)3 or [PhNMe2H][B(C6F5)4]) led to a fast and clean formation

    of cationic complexes, in which the “extra” donor was bound to the metal (Scheme A-6). The

    cationic alkoxo complexes display a rigid C1-symmetry, while the analogous alkyl species are

    presumably of Cs-symmetry. The species demonstrate a remarkable stability: The cationic

    alkoxo species are stable both in solution and in solid state for weeks, while the cationic alkyl

    species are stable for days.

    MO

    O OR

    OR

    OR

    Nt-Bu

    t-BuO

    Me

    Me

    OH

    N

    OMe

    MeO

    t-Bu

    t-Bu

    M(OR)4 MO

    O O

    OR

    Nt-Bu

    t-Bu

    Me Me

    OR

    +

    B(C6F5)4 -

    [PhNMe2H]

    [B(C6F5)4]

    Scheme A-6. Synthesis and structure of the neutral tris(alkoxo) and cationic bis(alkoxo) Group IV metal

    complexes with the amine mono( phenolate) ligand. M = Ti(IV), Zr(IV), R = Ot-Bu, Oi-Pr.

    Salophan complexes of group IV metals

    This work is fully described in article 12.

    Sequential diamine bis(phenolate) Salophan ligands are similar to Salan ligands, with

    the only difference between them being the nature of the link between the two nitrogen

    donors: Flexible ethylene (diamine) bridge in Salan, and rigid ortho-phenylene (diamine)

    bridge in Salophan. Early transition metal complexes of Salan ligands are in the focus of

  • Abstract

    14

    intensive investigation by several research groups; in contrast, no early transition metal

    complexes of Salophan ligands have been reported thus far, with the exception of a single Mo

    dioxo complex. We synthesized three new Salophan ligand precursors, featuring ortho Me,

    ortho, para di-chloro, and ortho, para di-t-Bu phenolate substituents, in addition to the

    previously reported prototypical Salophan ligand, bearing only hydrogen substituents at the

    phenolate rings. The ligand pecursors were prepared by a simple sequence of condensation

    and reduction. All the ligand precursors were reacted with Ti(Oi-Pr)4 and Zr(Ot-Bu)4

    (Scheme A-8). We found that the unsubstituted Salophan led to a complex product mixture

    for both metal precursors, whereas the ortho-Me substituted ligand led to a clean complex

    only with zirconium. Other, namely ortho, para di-chloro, and ortho, para di-t-Bu substituted

    ligands, gave a well-defined coordination chemistry with both metals. NMR analysis

    indicated that all the Salophan ligands act as dianionic ligands, forming hexa-coordinate

    complexes of C2-symmetry. X-ray structure analysis of these species revealed that the

    wrapping mode of the ligands was fac-fac, and the orientation of the labile groups was cis

    (Scheme A-7).

    Scheme A-7. Wrapping mode of the Salophan ligands around Group IV metals

    N N

    OHR1

    R2

    HOR1

    R2

    H H

    O

    N

    MN

    OOR

    R1

    R1

    R2

    R2RO

    H

    H

    M(OR)4M=Zr: R=Ot-Bu, R1=Me, R2=H

    M=Zr: R=Ot-Bu, R1=R2=Cl

    M=Ti: R=Oi-Pr, R1=R2=Cl

    M=Ti: R=Oi-Pr, R1=R2=t-Bu

    M=Ti: R=Oi-Pr, R1=R2=t-Bu

  • Introduction

    15

    Introduction

    This thesis is devoted to the study of early transition metal complexes with amine

    phenolate ligands, and presents their coordination chemistry and reactivity. In the

    introduction I will describe various ligand systems with emphasis on oxygen-donor ligands,

    their characteristic chemistry, and reactivity in the fields relevant to this work.

    I-1. Alkoxo (phenolate) ligand vs. Cp ligand

    In recent years we are witnessing a huge interest in the development of new ligand

    systems that may be used as an alternative to the cyclopentadienyl (Cp) ligand in the

    chemistry of early transition metals. Ligands based on the alkoxo- and amido donors have

    attracted the most considerable attention, as these “hard” multi-electron donors successfully

    stabilize high oxidation states of the early transition metals, that are defined as “hard acids”.1

    The alkoxo (phenolate) ligands are especially attractive: The ligand precursors are

    sufficiently acidic, bind the “oxophilic” early transition metals firmly, and may be viewed as

    isoelectronic with the cyclopentadienyl ligand.2,3

    Several important differences between cyclopentadienyl ligands and alkoxo ligands

    should be taken into account. The alkoxo ligand is less electron-donating than the Cp ligand,

    thus the alkoxo complexes are less electron-saturated.2-4 A simple alkoxo or phenoxo ligand

    is less bulky than the cyclopentadienyl ligand. As a result, the use of a simple alkoxo

    (phenoxo) ligand normally leads to complex product mixtures.2-4 An additional obstacle in

    the chemistry of the alkoxo (phenoxo) ligands is their ability to bridge metal atoms.4 An

    aggregation of metal complexes causes ligands redistribution, and provides low-energy

    pathways for the decomposition of coordinatively unsaturated metal complexes.

  • Introduction

    16

    I-2. From a simple alkoxo to a bulky phenolate ligand.

    Two general strategies have been developed to overcome the obstacles encountered in

    the use of simple alkoxo or phenolate ligands: Employing bulky monodentate ligands, or

    designing multidentate ligands. The use of bulky ligands allows accommodation of only a

    limited number of ligands around the metal center, which could lead to the desired structure

    of a metal complex, and to prevent aggregation. Throughout the last two decades, the motif of

    bulky monodentate alkoxo, siloxo, and phenoxo ligands has been carefully investigated.2,3,5,6

    The steric bulk of a ligand may be approximated using the “cone angle” concept.1,7 For the

    Cp ligand, the cone angle was estimated to be 136º.7 Wolczanski and coworkers have

    introduced a number of bulky alkoxo and siloxo ligands, with cone angles close to this value

    (Figure I-1).2 The substantial steric bulk of the ligand allowed a clean metathesis reaction,

    leading to the alkoxo/siloxo complexes of desired structures. As a result of the coordinative

    unsaturation at the metal center, some of these complexes exhibited an unusual reactivity,

    including carbon monoxide cleavage, oxidative addition of various substrates, and formation

    of metal-metal bonded complexes.2

    Figure I-1. Cone angles of the Cp, Ot-Bu, and “tritox” ligands

    Bulky phenolates were subject of even more thorough investigation, especially by the

    groups of Rothwell and Schrock.3,5,6 The phenolate may be viewed as a more "user-friendly"

    ligand than simple alkoxide, as it is more acidic and its properties are more easily controlled

    M

    O

    Ct-Bu

    t-But-Bu

    M

    M

    O

    CMe MeMe

    136 o < 90 o 125 o

  • Introduction

    17

    by the utilization of the appropriate substituents. As phenolate is a planar rather than a cone-

    shaped ligand, the “cone angle” term is less appropriate for this ligand. The steric bulk of the

    phenolate is normally transferred via its ortho substituents that point toward the metal; the

    bulky ortho groups prevent aggregation of the metal centers and lead to the well-defined

    constitution of a metal complex. For example, for 2,6 di-t-Bu phenolate ligands, complexes

    containing only two such ligands (in axial positions), in addition to three alkyl ligands, were

    typically formed for tantalum (V) (Figure I-2).5

    CH3

    Ta

    CH3H3C

    O O

    t-Bu

    t-Bu

    t-Bu

    t-Bu

    Figure I-2. A well-defined mononuclear complex of Ta(V) carrying a 2,6 di-t-Bu phenolate ligand.

    I-3. Multidentate ligands: Enhanced stabilization and easier design of metal centers.

    Exploitation of the chelate effect provides a second possible strategy for the

    stabilization of metal-alkoxide complexes. In general, multidentate ligands bind the metal

    firmly due to the chelate effect,1 and, not less importantly, may enable precise control of the

    structure and reactivity at the metal center. The following non-alkoxo ligand system, relevant

    to the studies presented in this work, illustrates this trend. The triamidoamine [N3N] ligand

    system, investigated initially by Verkade,8 was introduced to the early transition metals realm

    by Schrock.9 These ligands were shown to bind to metals in a tetradentate, almost uniformly

    C3v-symmetrical manner, creating a 3-fold-symmetric axial “pocket” (Figure I-3). In that

    pocket, 3 orbitals are available for bonding with additional ligand(s): two π orbitals and one σ

    orbital. In addition to the electronic consideration, this pocket was influenced sterically by

    bulky substituents at the amido nitrogens. All this led to extremely rich chemistry of the

  • Introduction

    18

    metal-ligand multiple bonds. The formation of metal-ligand multiple bonds in this position

    was frequently unavoidable: For instance, the reaction of various [N3N]TaCl2 complexes with

    any Grignard reagent bigger than methyl led uniformly to the formation of the tantalum

    alkylidene (M=CR2) function, and tungsten alkyl complexes [N3N]WR readily lost

    dihydrogen at RT to form metal alkylidyne (M≡CR) complexes. In addition, the unique

    ability of this ligand system to stabilize multiple bonds allowed to fulfill one of the “Holy

    Grails” of modern inorganic chemistry: catalytic dinitrogen activation to ammonia via a 14-

    step cycle including the states of imide and nitride, 10 of the intermediate species being

    isolated and characterized.10

    N

    NN

    R R

    NR

    z

    dz2 dxz

    N

    N NTa

    H R'R R

    NR

    N

    NNW

    R'R R

    NR

    N

    NNMo

    NR R

    NR

    N

    NN

    Mo

    PR R

    NR

    Figure I-3. Orbitals available for the axial ligand(s) in a 3-fold symmetrical pocket (only one orbital of the

    degenerate set (dxz, dyz) is shown); examples of multiply bonded ligands in the [N3N] metal complexes

  • Introduction

    19

    I-4. Multidenatate phenolate ligands.

    Several multidentate phenolate-based ligand systems became widely known and

    extensively used in recent years.4 Simple methylene- or ethylene-bridged bis(phenolate)

    ligands have been investigated by several groups. These ligands, providing only two donors,

    forming an 8-membered chelate ring (or a 9-membered chelate ring), and being relatively

    flexible, do not impose a rigid geometry at the metal. Thus, the resulting early transition

    metal complexes may be tetra-, penta-, or hexacoordinate, depending on the reaction

    conditions (Scheme I-1).11 The use of a mononuclear metal precursor for these ligands

    generally leads to a formation of a mononuclear product. In contrast, for dinuclear metal

    precursors (Mo2(NMe2)6 and W2(NMe2)6), the ligand can bind either to a single metal or to

    bridge between the metal centers (Scheme I-2).12

    OHR1

    R2

    OHR1

    R2

    TiCl4

    OR1

    R2

    OR1

    R2

    Ti

    Cl Cl

    + THF- THF

    OR1

    R2

    OR1

    R2

    TiCl

    ClTHF

    ZrCl4THF2

    OR1

    R2

    OR1

    R2

    ZrCl Cl

    THFTHF

    Scheme I-1. Coordination behavior of methylene-bridged bis(phenolate) ligand with Group IV metals.

  • Introduction

    20

    OR1

    R2

    O R1

    R2

    Mo2(NMe2)6

    OR1

    R2

    OR1

    R2

    Mo Mo

    OR1

    R2

    OR1

    R2

    R2N

    NR2 +Mo

    MoO

    R1

    R2

    OR1

    R2

    NR2

    R2N

    OH

    R1

    R2

    OH

    R1

    R2

    Scheme I-2. Coordination behavior of the bis(phenolate) ligand with Mo(VI) dinuclear metal centers.

    Another example of a relatively simple chelate ligand system is the α, α’-bridged

    bis(naphtholate) (BINOL) ligand.13 This ligand leads mostly to tetra- or penta-coordinate

    complexes; in the case of bulky ortho substituents, the formation of bis(homoleptic)

    complexes is diminished to a large extent (Scheme I-3). The most unique feature of this

    ligand is its intrinsic chirality; thus, BINOL provides a chiral backbone in the resulting metal

    species. For the Group IV metals possessing identical monodentate ligands (X), the resulting

    tetrahedral complexes are C2-symmetrical, and the labile positions are homotopic.14

    OHOH

    OO

    MXX

    MX4

    R

    R

    R

    R

    Scheme I-3. A BINOL ligand precursor, and the resulting tetrahedral M(IV) complex.

    Sequential methylene-bridged tris(phenolate) ligands were investigated by Kawaguchi

    and coworkers.15 As in the respective bis(phenolate) ligands, the phenolate rings in the

    tris(phenolate) ligands are connected at the ortho positions through methylene linkers. These

  • Introduction

    21

    ligands demonstrated even more flexible behavior than the corresponding bis(phenolates):

    Several conformations for the coordination of this type of ligands to the metal center were

    reported. For the early transition metals (Ti(IV), Ta(V), Nb(V)), these ligands normally form

    dinuclear complexes. The nearby metal centers in these complexes could be bridged by

    additional monodentate (hydrido) or bidentate (DME) ligands, or by the phenolate oxygens

    themselves (Figure I-4).

    Figure I-4. A dinuclear di-Ti complex with the “open-chain” tris(phenolate) ligand.

    The ligands presented in the previous paragraphs are all sequential (“open-chain”)

    phenolate-only ligands. A very important family of multidentate phenolate ligands is the

    family of cyclic poly(phenolate) ligands, termed calixarenes (“bowl-shaped”) after Gutsche.16

    This general term includes all cyclic poly(phenolates), connected by methylene bridge; most

    of the inorganic and organometallic chemistry was performed with the ligand containing four

    phenolate units (calix[4]arene).17 This ligand may adopt various conformations, with the

    “cone” conformation being the most common for the early transition metal complexes

    (Scheme I-4). The basic calix[4]arene ligand is tetraanionic, which generally results in the

    formation of dinuclear complexes for the Group IV metals, and may also lead to the dinuclear

    species with Group V metals (Scheme I-4). For Group V metals, formation of a mononuclear

    comlex may be achieved by using a Cp group as an additional, fifth ligand. In contrast, this

    O

    O

    Ot-Bu

    Ti

    O

    O

    O

    t-Bu

    Ti

    Cl

    Cl

  • Introduction

    22

    tetraanionic tetradentate ligand normally forms a mononuclear well-defined species with

    Group VI (M(VI)) metal centers. A variety of di- and trianionic calix[4]arene derivatives

    could be synthesized by a selective Williamson alkylation at the phenolate oxygens. For the

    corresponding dianionic ligands, mononuclear Group IV complexes are obtained.

    Scheme I-4. Top: coordination chemistry of the tetraanionic calix[4]arene toward early transition metals.

    Bottom: A mononuclear Zr(IV) complexes of a dianionic calix[4]arene.

    The chemistry of chalcogen-bridged bis(phenolate) and bis(naphtolate) ligands

    ([OEO], E = S, Te) with Group IV metals has been studied mainly by Kakugo and coworkers,

    and by Okuda and coworkers.18 These ligands are dianionic and tridentate, binding to the

    OO O O

    HH H H

    t-Bu t-Bu t-Bu t-Bu

    Ti(NMe2)4

    OO O O

    t-Bu t-Bu t-Bu t-Bu

    OOOO

    t-But-But-But-Bu

    TiTi OOOO

    t-But-But-But-Bu

    WO

    W(O)Cl4calix[4]arene

    CpTaCl4

    OOOO

    t-But-But-But-Bu

    Ta

    OO O O

    HMe H Me

    t-Bu t-Bu t-Bu t-Bu

    n-BuLi

    ZrCl4O

    O O OMe Me

    t-Bu t-Bu t-Bu t-Bu

    ZrCl Cl

    OO O O

    Me Me

    t-Bu t-Bu t-Bu t-Bu

    Zr

    PhH2C CH2Ph

    PhCH2MgCl

  • Introduction

    23

    metal through the phenolate oxygens and the central neutral donor. Due to the 5-membered

    chelate rings, these ligands tend to adopt a more rigid coordination at the metal. The

    coordination mode of these ligands is facial (fac) in octahedral geometry, which leads to the

    cis disposition of the phenolate oxygens. For potentially bridging labile ligands (i.e., OR, Cl),

    the resulting metal derivatives are dimeric, possessing nearly octahedral environment

    (Scheme I-5). In contrast, the organometallic complexes of these ligands are monomeric and

    pentacoordinate. Further studies on Mo(VI), W(VI), and Sm(III) metal complexes with this

    type of ligand have all indicated a facial binding of the [OEO] ligand to the metal center.19

    S

    O Ti

    O

    R

    RCl

    Cl Cl

    S

    OTi

    O

    R

    R

    Cl

    OH OH

    RRTiCl4

    MeLi

    O O

    RR

    Ti

    S

    S

    Me Me

    S

    O Ti

    O

    R

    RTHF

    Cl Cl

    THF

    Scheme I-5. Coordination chemistry of the [OSO] ligands at Ti(IV) metal centers.

    Possibly, the most appreciated veteran in the field of multidentate phenolate ligands is

    the “sequential” tetradentate di(imine) bis(phenolate) ligand, also known as SALEN ligand

    (Scheme I-6).20 Complexes of nearly all early transition metals with this ligand have been

    prepared and characterized. This ligand, being dianionic and tetradentate, leads to well-

    defined penta- or hexacoordinate mononuclear complexes, generally avoiding the formation

  • Introduction

    24

    of the bis(homoleptic) or bridging species. In octahedral geometry, this rigid ligand normally

    imposes a “mer-mer” planar disposition of its four donor atoms, leaving the two remaining

    positions mutually trans (Scheme I-6).

    NN

    HOOHRR

    NN

    OO

    RR

    M

    X

    X

    MX4

    Scheme I-6. SALEN ligand precursor, and the resulting metal (M(IV)) complex.

    Reducing the C=N double bonds in the SALEN ligand leads to the di(amine)

    bis(phenolate) [ONNO]-type SALAN ligand precursor (Scheme I-7). In analogy to the Group

    IV SALEN metal complexes, SALAN metal complexes are also well-defined mononuclear

    species of octahedral geometry. However, these saturated ligands demonstrate completely

    different binding mode: The fac-fac coordination of the [ONN] fragments, resulting in cis

    disposition of the monodentate ligands. Thus, the resulting metal species are C2-symmetrical

    and chiral-at-metal.21 A similar binding mode has been reported for the di(thio)

    bis(phenolate) [OSSO] ligands.22

    NN

    OHR

    R

    HO

    R

    RMX4

    O

    N

    M

    N

    OX

    R

    R

    R

    RX

    Scheme I-7. Wrapping mode of the SALAN ligands around Group IV metal centers

  • Introduction

    25

    “Breaking” the SALEN ligand into two parts leads to bidentate monoanionic

    phenoxy-imine ligands. Group IV metal complexes bind two such ligands, in addition to two

    monodentate ligands (chloro or alkyl groups). Similar to the di(amine) bis(phenolate) ligand

    complexes, and in contrast to the SALEN metal complexes, the phenoxy-imine ligands are

    oriented such that the overall symmetry is C2, the two phenolate oxygens mutually trans, and

    the monodentate ligands are cis to one another (Scheme I-8).23

    Scheme I-8. Phenoxy-imine ligands, and their Group IV complexes. The resulting catalysts are termed “FI”

    catalysts after phenoxy-imine ligands.

    The recent findings in the field of multidentate phenolate-based ligands were recently

    reviewed.4 It should be noted, however, that this field remains relatively unexplored, in

    comparison with the polydentate amido ligands,24 and major effort is invested today in the

    design and exploration of new chelate phenolate-based ligands.

    I-5. Synthetic pathways to early transition metal complexes bearing phenolate ligands

    The salt metathesis route between the ligand salt (Li, Na, etc.) and the metal chloride

    provides the most commonly used method for the preparation of early transition metal

    complexes with various sorts of ligands (Scheme I-9). However, the chemistry of the

    phenolate-based ligands with early transition metals is rather unique, as the phenol function

    in the ligand precursor is substantially more acidic than most other functions (such as simple

    MX4

    R'N

    OH2 MN

    O

    R'N

    O

    R'X X

    R

    R

    R

  • Introduction

    26

    alkohol, amine, or alkane). This may be employed to help avoiding the rather notorious salt

    metathesis route normally required for the amido or cyclopentadienyl ligands. Thus,

    homoleptic metal alkoxides or amides may be employed as the metal precursors in alcohol or

    amine elimination reaction, respectively (Scheme I-10). One of the goals of the current work

    was to show that the homoleptic metal alkyl precursors, such as tetrabenzyltitanium or

    pentabenzyltantalum, could be used as starting materials of choice in the chemistry of

    phenolate ligands.

    ArOH + BuLiMClx ArOMClx-1 + LiClArOLi

    Scheme I-9. The salt metathesis route to early transition metal complexes with the phenolate ligands

    ArOH + M(OR)x ArOM(OR)x-1 + ROH

    ArOH + M(NR2)x ArOM(NR2)x-1 + R2NH

    ArOH + MRx ArOMRx-1 + RH

    Scheme I-10. The route of the alcohol, amine, and alkane elimination reactions, respectively, to early transition

    complexes with phenolate ligands

    I-6. Reactivity of early transition metal complexes supported by multidentate phenolate

    ligands

    In the last decade, phenolate-based chelating ligands became widely abundant in the

    frontier fields of inorganic chemistry and catalysis. Among other fields, early transition metal

    complexes of multidentate phenolates show an extraordinary reactivity in the fields of

    enantioselective transformations,25 olefin polymerization, lactide and caprolactone

  • Introduction

    27

    polymerization,26 olefin metathesis, and bioinorganic chemistry. In the following sub-

    paragraphs, a description of the fields relevant to this work is presented.

    I-6.1. Olefin polymerization by well-defined Group IV metal complexes

    One of the research fields most significantly influenced by the chelating phenolate-

    based ligands is the field of α-olefin polymerization catalysts.27 The field of α-olefin

    polymerization emerged in 1954, when Ziegler discovered that a combination of a simple

    early transition metal complex (TiCl4), and a main group compound (AlEt3) catalyzes a

    polymerization of ethylene to polyethylene under mild conditions.28 Soon after, Natta and

    coworkers reported the stereoregular (isotactic) polymerization of propylene by the same

    catalyst, producing a crystalline polymer of high melting point (Scheme I-11).29 This catalytic

    system was heterogeneous, and therefore ill-defined. As a result, this multi-site system

    produced polymers mixtures of broad molecular weight distribution and different properties,

    and impeded a thorough understanding of the polymerization mechanism. Soon, it became

    understood that the preparation of the well-defined polymers demands the preparation of

    well-defined (“single-site”) catalysts.

    Me

    TiCl4

    AlR3Me

    n

    x +

    Mey

    isotactic polypropylene: crystalline, high-melting

    atactic polypropylene:amorphous, oily

    Scheme I-11. Propylene polymerization by the heterogeneous system TiCl4/AlR3

    The first homogeneous models were constructed from the Group IV bent

    metallocenes (CpMX2) and alkyl aluminum compounds as activators (“co-catalysts”)

  • Introduction

    28

    (Scheme I-12). In contrast to the heterogeneous systems, these catalytic systems showed low

    activity in the polymerization of ethylene, and negligible activity toward propylene and

    higher alpha-olefins.30 Two major discoveries revolutionized this field. First, the reactivity of

    the metallocene-based catalysts was significantly improved by the invention of a new co-

    catalyst: Methylalumoxane (MAO), a partially hydrolyzed Me3Al having an oligomeric

    form.31 Second, the invention of chiral (C2-symmetric) ansa-metallocenes has led the way to

    the desired stereoregular (isotactic) polypropylene.32 Consequently, Cs-symmetric

    metallocenes having enantiotopic sites led to the syndiotactic polypropylene,33 that was

    inaccessible via the heterogeneous group IV Ziegler-Natta catalysts.

    Scheme I-12. Top: Propylene polymerization by C2v-symmetric metallocenes. Middle: Propylene

    polymerization by C2-symmetric ansa-metallocenes. Bottom: Propylene polymerization by Cs-symmetric ansa-

    metallocenes.

    ZrClCl

    AlR3 Men

    Me

    n

    Low activity, low Mw, atactic polypropylene

    ZrCl Cl

    ZrCl Cl

    Me

    n

    Me

    n

    High activity, high Mw, isotactic polypropylene

    High activity, high Mw, syndiotactic polypropylene

    MAO

    Al(Me)-O n(MAO)

    Men

    Me n/2Me

  • Introduction

    29

    One should not underestimate the contribution of metallocene catalysts to the field of

    α-olefin polymerization, as these catalysts shed light on polymerization mechanism, and led

    to novel polymer types.34 However, the versatility of a single ligand family in determining of

    catalyst structure (metal geometry, its electron and steric saturation) is limited by definition.

    In addition, the synthesis of various functionalized Cp ligands, and their metal complexes

    may be quite cumbersome, and almost no metallocenes found their way to the polymer

    industry for various reasons. Thus, in the middle of the 1990’s, a search for new ligand

    systems for polymerization catalysts has begun.

    At first, the amido-based ligands attracted most of the attention, as the derived

    catalysts were shown to lead to “living polymerization”.24a Several criteria have been

    proposed for living polymerization, among them are negligible termination (or chain transfer)

    and fast initiation relative to propagation.35 Living polymerization can be of great importance,

    as it leads to well-defined polymers of high-Mw and narrow molecular weights distribution. In

    addition, living polymerization allows the preparation of block-copolymers. Prior to these

    reports, no living polymerization catalysts, operating at ambient conditions, had been

    reported in the field of Ziegler-Natta polymerization. In 1996, McConville and coworkers

    have reported that a Ti-based catalyst, carrying a diamido ligand, polymerizes 1-hexene in a

    living fashion (Scheme I-13).36 The living character of polymerization was indicated by a

    linear rise in the molecular weight of the resulting polymer vs. time and vs. monomer

    consumed, and by a narrow molecular weight distribution (MWD, or PDI). Since then, the

    amido-based polymerization catalysts were thoroughly studied, especially in relation to living

    polymerization.24a,37, 38

  • Introduction

    30

    Scheme I-13. Living polymerization of 1-hexene by diamido-based Ti catalyst

    The great promise of the phenolate-based polymerization was demonstrated in the

    work of Schaverien and coworkers in 1995 describing highly isotactic polymerization of 1-

    hexene using BINOL Ti(IV) and Zr(IV) dibenzyl complexes (see Scheme I-3).14 It should be

    noted, however, that the major interest in the phenolate-based catalysts has arisen only

    recently, with the development of four families of polymerization catalysts.39 The first family

    is the family of Ti catalysts based on the [OSO] ligands (Scheme I-5) that have led to the

    active polymerization of ethylene and styrene.18 The second family is the family of “FI”

    catalysts (Scheme I-6).23 All the complexes of this family are C2-symmetric, and the steric

    and electronic parameters of this system may be easily controlled by the phenolate

    substituents, and the imine substituents. As a result of fine structural modifications, some of

    these catalysts exhibited unusual reactivities. For example, the FI-Zr catalysts, possessing a

    bulky group at the ortho position of the phenolate rings were reported to be the most active

    ethylene polymerization catalysts ever reported. The sterically-hindered and the electron-

    deficient FI-Ti catalysts, possessing bulky groups at the ortho positions, and the meta-

    difluorophenyl groups bound to the nitrogen donors, led to the living and syndiotactic

    polymerization of propylene (Scheme I-14).

    TiN

    N

    Me

    Me

    i-Pri-Pr

    i-Pr i-Pr

    B(C6F5)3Bu

    nBu

    nPDI < 1.1

    Living poly(1-hexene)

  • Introduction

    31

    Me

    n

    Living polymerization yielding

    syndiotactic polypropylene

    MAOMe n/2Me

    TiN

    O

    t-Bu

    RN

    O

    t-Bu

    RCl Cl

    FF

    R =

    Scheme I-14. Propylene polymerization by FI-Ti catalyst

    The third family consists of the amine bis(phenolate) metal complexes discussed in

    the present work. Finally, the fourth family is the family of the di(amine) bis(phenolate)

    ligands.21,40 The labile positions, at which the polymerization reaction takes place, are

    homotopic in this catalyst, due to the rigid C2-symmetric structure of the complex. Thus, in

    analogy to the C2-symmetric metallocenes, an isotactic polymerization of various α-olefins

    was observed.21a, 21b, 40 In addition, a living and isotactic polymerization of an α-olefin (1-

    hexene) was achieved by the Zr catalyst of this family (Scheme I-15).21a The precondition for

    the living polymerization in this system is the presence of bulky groups in ortho positions,

    which are presumed to retard termination.

    Scheme I-15. 1-hexene polymerization by di(amine) bis(phenolate) Group IV complex

    nBu

    nBu

    Living and isotactic poly(1-hexene)

    B(C6F5)3

    O

    N

    ZrN

    OCH2Ph

    t-Bu

    t-Bu

    t-Bu

    t-BuPhH2C

    Me

    Me

  • Introduction

    32

    Given these findings, the potential of the phenolate-based multidentate ligands has

    been widely recognized by multiple research groups. Thus, at the present time, novel

    phenolate-based olefin polymerization catalysts are designed, and their reactivity is

    investigated by many research groups worldwide.

    I-6.2. Stabilization of metal-carbon double bonds and olefin metathesis reactivity

    The first metal-carbon double bond of the ”alkylidene” type was reported by Schrock

    in 1974.41 In the course of investigation of Ta(V) homoleptic alkyl complexes, Schrock

    discovered that the compound formed by the reaction between Ta(CH2t-Bu)3Cl2 and two

    equivalents of t-BuCH2Li was actually Ta(=CHt-Bu)(CH2t-Bu)3, instead of Ta(CH2t-Bu)5

    (Scheme I-16). Following the discovery that the alkylidene function mediates olefin

    metathesis,42 metal-carbon multiple bonds became a subject of principal interest in

    organometallic chemistry.43

    Ta(CH2t-Bu)3Cl2 Tat-BuH2C

    t-BuH2C

    CH

    CH2t-Bu

    CH2t-Bu

    t-Bu H

    2 NpLi

    - 2 LiCl

    - NpHTa

    CH2t-Bu

    C(H)t-Bu

    t-BuH2C

    t-BuH2C

    Scheme I-16. The α-hydrogen abstraction reaction, leading to the formation of an alkylidene complex

    The major route that leads to the alkylidene function is the α-hydrogen abstraction

    reaction, taking place in a dialkyl precursor (Scheme I-16).43 If the alkyl group carries β-

    hydrogens, β-H elimination may compete with the α-hydrogen abstraction reaction. The α-

    hydrogen abstraction reaction normally goes via a monometallic mechanism, being therefore

    of first order in the reactant.44

  • Introduction

    33

    TaO

    O

    THF

    R R

    R

    R

    O

    R

    RCt-Bu

    H

    Mo

    N

    CHRO

    O

    i-Pr i-PrR

    R RR

    Figure I-5. Alkylidene complexes, supported by monodentate phenolate ligands.

    Monodentate phenolate-based ligands were extensively used to support the alkylidene

    functionality, and to study its structure and reactivity (Figure I-5).3,5 However, only recently

    multidentate phenolate ligands have been introduced into this field. Following the highly

    efficient Mo(VI) metathesis catalysts, supported by two monodentate alkoxo or phenolate

    ligands, Schrock and coworkers prepared Mo(VI) alkylidene complexes, supported by a

    single bis(phenolate) or BINOL ligands.45 The resulting species are mainly tetracoordinate,

    containing an imido function in addition to the alkylidene and the phenolate oxygens

    (Scheme I-17). Most significantly, these species are chiral, as a result of the ligand backbone,

    and may be prepared in an enantiomerically pure form. These species have been found to be

    highly effective stereoselective catalysts for various asymmetric olefin metathesis reactions.

    For example, enantiomerically-pure small rings could be prepared from achiral precursors in

    high yields, and with excellent enatiomeric excess (Scheme I-17).46

  • Introduction

    34

    Mo

    N

    CHROO

    i-Pr i-Prt-Bu

    t-Bu

    Me

    Me Me

    MeO O

    Me H

    Me

    Me2 mol %

    99% ee, 93% yield

    Scheme I-17. An asymmetric RCM reaction catalyzed by the Schrock-Hoveyda catalyst.

    I-6.3. Phenolate ligands in bioinorganic chemistry

    The field of bioinorganic chemistry is rapidly growing in the last years. In general,

    this field is concerned with the active sites of metallo-enzymes, and with simple model

    compounds, that mimic the structure and reactivity of the active site of the enzyme.47 The

    metallo-enzymes that contain oxophilic early transition metals (mostly V, and Mo to a lesser

    extent) may exhibit an oxygen-rich environment at the active site. For V, two enzyme

    families are known today. The enzyme of the first family, V-dependent nitrogenase, reduces

    dinitrogen into ammonia; the active site of this enzyme contains V ligated mainly by sulfur

    donors.48,49 The second family of V-dependent enzymes is V-dependent haloperoxidases.48,49

    These enzymes catalyze the oxidation of halide to hypohalous acid (using peroxide as the O

    atom donor), which, in turn, halogenates organic substrates (Scheme I-18, top). In addition,

    these enzymes catalyze the asymmetric O atom transfer to sulfides, forming optically active

    sulfoxides (Scheme I-18, bottom).

  • Introduction

    35

    Scheme I-18. Reactivity of V-dependent haloperoxidase. Top: Halogenation of organic substrates (X = Cl, Br).

    Bottom: Oxidation of sulfides to sulfoxides.

    Considerable structural data has been gained about the V-dependent

    haloperoxidases.50 The (resting) active site of all the enzymes of this family contains a V(V)

    metal center in a trigonal bipyramidal geometry, coordinated to four oxygen atoms, and to

    one nitrogen atom (Figure I-6). During the catalytic cycle, the metal retains its oxidation

    state. The proposed catalytic cycle involves a peroxo (side-on) intermediate, in which the

    activation of the O atom takes place. At the final step, before the release of the oxidant, the

    metal center is of hexacoordinate octahedral geometry.51 In the hexacoordinate geometry, V

    binds five oxygen ligands, and one nitrogen ligand (imine) in one of the axial positions.

    X- + H2O2 + H+ "V" HOX + H2O

    HOX + RH RX + H2O

    RS

    R+ H2O2 R

    SR

    + H2O

    O"V"

  • Introduction

    36

    Figure I-6. Proposed catalytic cycle for V-dependent haloperoxidase

    Appropriate structural models for the active site of V-dependent haloperoxidase are

    V(V) complexes, having NO4 donors set (where N is a neutral donor) preferably in a trigonal

    bipyramidal geometry.48,49 It is worth noting that trigonal bipyramidal geometry is rare for

    pentacoordinate V complexes; usually, a distorted square pyramidal geometry is observed.48c

    In addition, the metal center should be able to switch between the penta- and hexacoordinate

    geometries. The number of compounds strictly matching these conditions is very limited.

    However, many complexes that come close to these conditions might be viewed as “good”

    structural models. Most of these models are based on the bi- or tridentate imine-phenolate

    ligands (Figure I-7).48,49 Some of these complexes have been found to be functional models

    as well, being able to conduct oxidation of organic and inorganic substrates, such as olefins

    (to epoxides) and sulfides (to sulfoxides).51 This field is of high interest to the inorganic

    V

    OHN(His404)

    O

    OOH

    N(His496)

    V

    OHN(His404)

    O

    OOH

    N(His496)

    H -OOH

    Cl-

    -2 H2OVO

    O

    O

    N(His496)O

    Cl-

    N(His404)

    H+, H2O

    HOCl

    VO

    O

    O

    N(His496)OH

    N(His404)

    H

    ClO H

    H2O2

  • Introduction

    37

    community nowadays, as it may open the way to efficient oxidation catalysts, operating

    under ambient conditions, and using environment-friendly oxidants, such as hydrogen

    peroxide or molecular oxygen.

    Figure I-7. Selected structural models of the active site of V-dependent haloperoxidase

    I-7. Amine phenolate ligands

    The present work describes the structural chemistry and reactivity of early transition

    metal complexes of the divergent tetradentate amine-phenolate ligands (with a minor

    exception of sequential tetradentate Salophan ligands). This ligand family includes three sub-

    groups: trianionic amine tris(phenolate) ligands; dianionic amine bis(phenolate) ligands,

    having an additional donor on a sidearm; and monoanionic amine mono(phenolate) ligands,

    carrying two such donors (Figure I-8). In contrast to many other ligand precursors, the

    ligands of this family can be easily prepared from commercially available starting materials

    in a single-step Mannich condensation. All these ligands may be viewed as “super-chelating”

    ligands: they provide four donor atoms, and, as divergent ligands, do not require a specific

    wrapping of the ligand precursor for the binding to the metal. In addition, these ligands may

    enable a precise control of the geometry at the metal center, by restricting the number of

    possible isomers in both penta- and hexacoordinate geometries.

    NO

    Vt-Bu

    Me

    Me

    NO t-Bu

    Me

    Me

    O

    NN

    OOMe

    t-Bu t-Bu

    MeV

    O

    O N

    O O

    V

    OH2

    OH2O

    Br

  • Introduction

    38

    Figure I-8. The family of the amine phenolate ligands

    I-7.1. The amine tris(phenolate) ligands

    The amine tris(phenolate)s are old organic compounds but very young ligands.52 Even

    though they are ideally suitable for stabilization of high-oxidation state early transition

    metals, and despite their structural similarity with the well-known tripodal ligands, such as

    triethanol amine,8 or the triamido amine,8,9 their potential has not been evaluated until the end

    of 1990’s. The first transition metal (Fe(III)) complex of an amine tris(phenolate) ligand was

    reported only in 1998.53a Soon after, the amine tris(phenolate) complexes of main group

    elements (Ga(III), In(III), Si(IV), P(V)) were reported.53b,54 The amine tris(phenolate) ligands

    were introduced into the chemistry of early transition metals by our group in 2001.55 A year

    later, the amine tris(phenolate) Ti complexes were shown to lead to the active lactide

    polymerization catalysts.56 Since then, these ligands draw an ever-increasing attention as a

    rigid and versatile platform for the early transition metals.4

    According to their coordination chemistry with the early transition (Ti(IV)), and the

    main group metals, these ligands are indeed “tripodal”, or “atrane” ligands,8 as they lead to a

    well-defined mononuclear TBP complexes, in which the phenolate oxygens occupy the

    equatorial positions, and the amine occupies one of the axial positions, while the second axial

    position is occupied by a fifth (monodentate) ligand (Figure I-9).55,56 In this respect, a

    N

    OH HOR R

    R

    HO

    N

    OH HO

    D

    R R

    N

    OHD

    D

    Amine tris(phenolate):Tetradentate trianionicno sidearm donors

    Amine bis(phenolate):Tetradentate dianionicone sidearm donor

    Amine mono(phenolate):Tetradentate monoanionictwo sidearm donors

    R

  • Introduction

    39

    comparison between the reactivity of these tripodal oxygen-donor ligands and the reactivity

    of tridodal nitrogen-donor ligands (triamidoamine) is revealing. Aiming at an easier

    preparation of well-defined mononuclear complexes, the amine tris(phenolate) ligands should

    be more useful than the triethanolamine ligands,8 according to the considerations presented

    earlier, i.e., they are more acidic, and allow a precise steric control via their ortho

    substituents. In addition, these ligands may possibly switch between hexacoordinate and

    penta-coordinate metal geometries, thus enabling their participation in various catalytic

    applications.

    Figure I-9. A tripodal amine tris(phenolate) Ti(IV) complex

    I-7.2. The amine bis(phenolate) ligands with the additional “sidearm” donor

    The dianionic tetradentate amine bis(phenolate) ligands, carrying a sidearm donor,

    were first reported in 1988 for Mo(VI).57 Our group has begun to explore the chemistry of

    these ligands with Group (IV) metals in 1999, aiming at the preparation of α-olefin

    polymerization catalysts. The complexes were prepared using per(alkoxo) (Ti(Oi-Pr)4) and

    per(benzyl) (Zr(CH2Ph)4 and Hf(CH2Ph)4) metal precursors (Scheme I-19).58,59 For Ti, the

    preparation of dialkyl complexes was accomplished by the reaction of the dialkoxo complex

    with Me3SiCl, and further alkylation with the corresponding Grignard reagent.58b,58c In all

    cases, the amine bis(phenolate) ligands have led to well-defined mononuclear complexes; in

    most of the cases the sidearm donor was bound to the metal, accomplishing an octahedral

    geometry at the metal center. The resulting complexes exhibited an almost uniform structure

    TiO

    OO

    NR

    R

    R

    RR

    R

    X

  • Introduction

    40

    at the metal center, in which the phenolate oxygens were mutually trans, and the labile

    positions were cis (as required for an α-olefin polymerization pre-catalyst), being consistent

    therefore with an overall Cs-symmetrical structure.

    MO

    X

    t-Bu

    t-Bu

    D

    X

    Nt-Bu

    Ot-Bu

    N

    OH HOt-But-Bu

    t-Bu t-Bu

    D

    MX4

    M = Ti, Zr, Hf

    X = Oi-Pr, CH2Ph

    D = NMe2, OMe, py, SMe, NEt2

    - 2 HX

    Scheme I-19. Synthesis and structure of the amine bis(phenolate) Group IV metal complexes

    In combination with the co-catalyst (B(C6F5)3), the amine bis(phenolate) Group IV

    metal complexes have led to active 1-hexene polymerization catalysts. Keeping in mind a

    uniform structure of the pre-catalyst, a thorough structure-activity relationship study was

    carried out. In general, the amine bis(phenolate) ligand precursor possesses several “degrees

    of freedom”: The nature of the phenolate substituents, the type of the sidearm donor, and the

    bridge between the central amine and the sidearm donor. At the first step of this study, most

    of the ligands under investigation had bulky (t-Bu) substituents in the ortho positions, in

    order to minimize the formation of the bis(homoleptic) complexes for the large metal ions (Zr

    and Hf). The additional degree of freedom is the nature of the metal ion. Overall, the Zr

    amine bis(phenolate) complexes exhibited the highest activity (standing among the highest

    activities ever reported for 1-hexene polymerization),59a-c Hf complexes were somewhat less

    active,59c and Ti complexes possessed much lower activity, displaying, however, a living

    character of the polymerization in several cases.58b, 58d The presence of the sidearm donor was

    found to be the most crucial parameter for high activity in Zr or for living polymerization in

    the Ti series. As for the nature of sidearm donor, the non-bulky hard donors (NMe2 and OMe)

  • Introduction

    41

    led to the most active Zr catalysts, presenting activities of 21,000 and 50,000 gpol mmolcat-1 h-

    1, respectively (Scheme I-20).59c For Ti, the OMe sidearm donor led to an unprecedented

    living polymerization of 1-hexene for 31 h.58d Based on these results, the amine

    bis(phenolate)-supported metal complexes seem to be one of the most promising non-

    metallocene olefin polymerization catalysts today. Thus, further studies, aiming at the

    discovery of highly active and living polymerization/block copolymerization catalysts, and

    new activity modes, are being carried extensively.

    Scheme I-20. Reactivity of Ti and Zr amine bis(phenolate) pre-catalysts in 1-hexene polymerization

    I-7.3. The amine mono(phenolate) ligands with two sidearm donors

    Viewing the tetradentate amine tris(phenolate) ligands and the amine bis(phenolate)

    ligands as a single family, the amine mono(phenolate) ligands, carrying two sidearm donors

    may be regarded as the “missing” relative in this family. These ligands possess an

    “unnecessary”, at first sight, sidearm donor for Group IV M(IV) metals in octahedral

    geometry. However, such a donor may probably be of high importance in cationic species, in

    which one of the labile ligands is removed, and may possibly lead to dicationic

    polymerization catalysts. This monoanionic ligand may successfully stabilize M(III)

    complexes of octahedral geometry, possessing two additional labile groups. Furthermore, this

    MO

    PhH2C

    t-Bu

    t-Bu

    D

    CH2Ph

    Nt-Bu

    Ot-Bu

    Bu

    B(C6F5)3

    Bun

    for M = Zr: highly active polymerization catalysts;

    for M = Ti: living polymerization catalysts

    D = NMe2, OMen

  • Introduction

    42

    “dormant” donor may trigger the α-elimination reaction to form the metal-ligand multiple

    bonds in the high-oxidation state early transition metals. Until now, no early transition metal

    complexes of such ligands have been reported.60

  • Discussion

    43

    Discussion

    This work encompasses coordination chemistry and reactivity of early transition metal

    complexes with a variety of amine phenolate ligands. In this section various parameters that

    bring this work together are discussed, starting from ligand and complex synthesis, and

    concluding with the reactivity at the pre-designed metal sites. In addition, this section

    highlights novelties brought to the realm of inorganic/organometallic chemistry and catalysis

    by complexes of the amine phenolate ligands.

    D-1. Ligand synthesis

    The amine bis(phenolate) ligands and the amine tris(phenolate) ligands presented

    herein have been prepared by the group of Prof. Z. Goldschmidt of Bar Ilan university. The

    symmetrical amine bis(phenolate) ligands have been prepared by a single-step Mannich

    condensation between the corresponding phenol, primary amine and formaldehyde (Scheme

    D-1). The ligands were obtained as crystalline solids after recrystallization. The amine

    tris(phenolate) ligands were synthesized by a straightforward reaction between

    hexamethylene tetraamine and the corresponding phenol. Overall, this route is very practical,

    leading to a large variety of ligand precursors in high yields. The amine mono(phenolate)

    ligand was prepared in our laboratory following a similar preparation.

  • Discussion

    44

    R

    R

    OH+ CH2O D NH2+

    MeOHR

    RN

    OH HOR

    R

    D

    N N

    N

    N

    R

    R

    OH+

    R

    RN

    OH HOR

    RMeOH

    HO

    R

    R

    R

    R

    OH+ CH2O D

    NH+MeOH

    R

    RN

    OHD

    D

    2

    Scheme D-1. Synthesis of various amine phenolate ligands via Mannich condensation

    Sequential (Salophan) ligands were synthesized by a different route, including a

    condensation between the substituted 2-hydroxy benzaldehydes and ortho-phenylene

    diamine, and subsequent reduction with sodium borohydride (Scheme D-2). For the majority

    of the ligands, a useful one-pot synthesis was developed, followed by a simple work-up,

    leading to the pure Salophan ligand precursors without a need for chromatography or

    recrystallization.

    R

    R

    OH

    H

    OH2N

    H2N+

    N

    N

    OH

    OH

    R

    R

    R

    R

    HN

    HN

    OH

    OH

    R

    R

    R

    R

    NaBH4MeOH

    MeOH

    Scheme D-2. Synthesis of Salophan ligands

  • Discussion

    45

    D-2. Complex synthesis

    As described in the Introduction section, the phenolate ligand precursors are

    sufficiently acidic, thus not requiring the notorious salt metathesis route for preparation of the

    metal complexes. A variety of metal precursors, purely inorganic or organometallic, were

    available for this purpose. The most convenient precursors are the commercially available

    homoleptic alkoxide complexes, such as Ti(Oi-Pr)4, and Ta(OEt)5, or the related VO(OPr)3

    (Scheme D-3). As the amine phenolate ligands are “super-chelate” ligands, and the alkoxide

    function is more basic than the phenolate function, the reactions normally proceeded to full

    conversion, and the yields were quantitative. The reactivity of the homoleptic amides, such as

    Zr(NMe2)4, or Ta(NMe2)5, towards the amine phenolate ligand precursors resembled that of

    the metal alkoxides to a large extent.

    Ta(OEt)5 + LigH3 LigTa(OEt)2 + 3 HOEt

    V(=O)(OPr)3 + LigH3 LigV(=O) + 3 HOPr

    Scheme D-3. Samples for alcohol-elimination reactions between the amine tris(phenolate) ligand precursors and

    metal-alkoxide precursors.

    The major achievement of this work from a synthetic point of view is the pursuit of

    the “alkane elimination” reactions between the homoleptic organometallic metal precursors

    and the amine phenolate ligand precursors. These routes are advantageous for the preparation

    of an organometallic complex. Some homoleptic organometallic species e.g., Zr(CH2Ph)4,

    and Hf(CH2Ph)4 are known for their stability, and thus have been widely used as metal

    precursors in “toluene elimination” reactions with acidic ligand precursors. However, the

    remaining member of the Group IV triad, Ti(CH2Ph)4, is much less stable, and therefore was

    seldom used as a starting material. Previously, an alternative route was developed in our

  • Discussion

    46

    group, including the reaction of the ligand precursor with Ti(Oi-Pr)4, followed by

    chlorination with TMSCl, and finally an alkylation with PhCH2MgCl (Scheme D-4).58 In the

    present research, we demonstrated that Ti(CH2Ph)4 is stable enough, and highly useful for

    preparation of titanium benzyl complexes with amine phenolate ligands.

    Ti(Oi-Pr)4 + LigH2 LigTi(Oi-Pr)2

    Ti(CH2Ph)4 + LigH2

    Me3SiCl LigTiCl2PhCH2MgCl LigTi(CH2Ph)2

    LigTi(CH2Ph)2

    Scheme D-4. Top: A previous synthesis of amine bis(phenolate) Ti(IV) dibenzyl complexes, consisting of the

    three steps; Bottom: A single-step synthesis of amine bis(phenolate) Ti(IV) dibenzyl complexes

    Until recently, the chemistry of “toluene elimination” reactions had been confined to

    the tetrabenzyl Group IV metal complexes only. In the course of our study of Ta(V)

    chemistry with the amine phenolate ligands, we attempted the use of the Group V (Ta(V))

    pentabenzyl complex as a starting material in that reaction. First, we developed a simplified

    preparative route for pentabenzyltantalum, and solved its crystal structure, which had not

    been reported previously. Thereafter, we reacted it with a variety of amine bis- and amine

    tris(phenolate) ligands (Scheme D-5). This work proved that pentabenzyltantalum is the

    “precursor of choice” in this chemistry, as it led to the desired organometallic species in high

    yields in a single step.

  • Discussion

    47

    Scheme D-5. Structure, and toluene-elimination reactions of pentabenzyltantalum with the amine phenolate

    ligand precursors

    An additional synthetic route that we employed in this work relies on the reaction

    between a metal chloride precursor and the ligand precursor in the presence of a mild base:

    Et3N. This route was found to be particularly useful when the utilization of metal alkoxide or

    metal alkyl precursors was not possible, as in the case of V(III) complexes. In this route,

    triethylamine serves to absorb HCl, forming the EtN·HCl salt that is insoluble in common

    organic solvents such as ether or THF. Thus, the desired products are obtained by a rather

    simple filtration-recrystallization sequence.

    D-3. Coordination chemistry of the amine phenolate early transition metal complexes

    One of the most pronounced advantages of the amine phenolate ligands is their well-

    defined coordination chemistry. Two isomers are feasible in octahedral geometry for the

    amine bis(phenolate) complexes of the LigMX2 type (Figure D-1). In the first isomer, the

    phenolate oxygens are trans to each other, leading to an approximate Cs-symmetry of the

    metal complexes. In the second isomer, the phenolate oxygens are cis to each other, and the

    resulting symmetry is C1. Throughout this research, we have synthesized dozens of such

    complexes, bearing alkoxo, amido, benzyl or methyl monodentate ligands. All of these

    LigH2LigTa(CH2Ph)3- 2 CH3Ph

    LigH3LigTa(CH2Ph)2

    - 3 CH3Ph

  • Discussion

    48

    complexes exhibited a trans disposition of the phenolate oxygens. However, Mountford and

    coworkers have recently demonstrated the viability of the C1-symmetric zirconium

    complexes as well, in which the oxygens are in a cis disposition.61

    MO

    R

    R1

    R1

    D

    R

    NR1

    OR1

    M = Ti(IV), Zr(IV), Hf(IV)

    R = Oi-Pr, NMe2, Me, CH2Ph, Cl

    R1 = t-Bu, Me, Cl, Br

    D = NMe2, OMe, OEt, THF, furan, py, SMe

    MO

    R

    R1

    O

    R

    NR1

    D

    R1

    R1M = Zr(IV)

    R = Cl

    R1 = t-Bu

    D = py

    Cs-symmetry C1-symmetry

    N

    OH HOR1

    R1

    R1

    R1

    D

    MO

    R

    R1

    R1

    R

    R

    NR1

    OR1

    M = Ta(V)

    R = CH2Ph

    R1 = t-Bu, Me, Cl, Br

    D = NMe2, OMe, Me

    Cs-symmetry

    D

    Figure D-1. Possible wrapping modes of the tetradentate amine bis(phenolate) ligands around a Group IV

    (M(IV)), and Group V (Ta(V)) metals in octahedral geometry

    The second structural parameter under investigation in the chemistry of the amine

    bis(phenolate) ligands was binding of the fourth, sidearm, donor. Previously, it was shown

    that, because of steric pressure, the sidearm donor might remain uncoordinated to the Group

  • Discussion

    49

    IV metal, forming a pentacoordinate complex.59b Thus, the binding of a sidearm donor should

    not be taken for granted a priori. In this work, we employed several types of sidearm donors:

    THF, open-chain ether (-OMe), non-bulky amine (-NMe2), and furan. All these, including the

    weak aromatic donor (furan), were found to bind to the Group IV metal center, thus leading

    to octahedral geometry at the metal.

    For a Group V metal complex (Ta(V)) not bearing multiply bonded ligands, the

    tetradentate dianionic amine bis(phenolate) ligand is expected to bind in a tridentate fashion

    (i.e. the sidearm donor remaining unbound) yielding octahedral geometry at the metal center

    (Figure D-1). This was supported by numerous X-ray structure determinations. However, this

    “dormant” donor may coordinate, or even trigger the removal of one of the benzyl groups in a

    LigTa(CH2Ph)3-type complex, which will be demonstrated in the following section.

    The coordination chemistry of the amine tris(phenolate) complexes was found to be

    well-defined as well. Octahedral LigMX2 complexes of M(V) metal centers may exist as a

    single geometrical isomer (for identical X’s, see Figure D-2). However, these ligands may

    probably lead to pentacoordinate metal centers, in the case of metal centers in lower than

    M(V) oxidation state (V(III), Ti(IV)), or for the metal centers in M(V) oxidation state bearing

    a multiply bound ligand (oxo (=O), alkylidene (=CHR)).

  • Discussion

    50

    MO

    X

    R1

    R1

    O

    X

    NR1

    OR1

    R1

    R1

    MO

    OO

    NR1

    R1

    R1

    R1 R1

    R1

    L

    M = Ta(V)

    X = OEt, NMe2, Cl, CH2Ph, µ-CHPh

    R1 = Me, t-Bu, Cl

    M = V(III),

    L = THF

    R1 = Me, t-Bu, Cl

    MO

    OO

    NR1

    R1

    R1

    R1 R1

    R1

    X

    MO

    OO

    NR1

    R1

    R1

    R1 R1

    R1

    R

    M = Ti(IV)

    L = Oi-Pr

    R1 = Me, t-Bu, Cl

    M = V(V)R = OR1 = Me, t-Bu

    N

    OH HOR1

    R1

    R1

    R1

    HO

    R1

    R1

    C3-symmetry C3-symmetry

    Cs-symmetryC3-symmetry

    Figure D-2. Coordination modes for early transition metals in the amine tris(phenolate) environment

    All the Ta(V) complexes with the amine tris(phenolate) ligands were found to be

    octahedral. As the amine tris(phenolate) is a tripodal ligand, possessing a central donor, and

    three equivalent “arms”, its chemistry may be correlated with the chemistry of other tripodal

    ligands, and in particular triamidoamine. The coordination chemistry of the amine

    tris(phenolate) ligands with Ta(V) stands in sharp contrast to the behavior of the

    triamidoamine ligands: The triamidoamine ligands did not lead to octahedral complexes

    when the metal center was hexa-coordinate, giving instead a C3-symmetrical pocket.9 For one

    specific monodentate ligand (benzyl), the amine tris(phenolate) Ta(V) complexes were C3v-

  • Discussion

    51

    symmetric on the NMR timescale down to 203 K; however, this may probably be explained

    by a dynamic process recently found in the hexacoordinate amine tris(phenolate) Ti(IV)

    complexes.62 Furthermore, the amine tris(phenolate) ligands do not show the tendency found

    for the triamidoamine ligands, as they do not lead to TBP Ta(V) complexes with the axial

    position occupied by a multiply-bonded ligand. Even when the remaining (“monodentate”)

    ligand was nominally “double-bonded” (alkylidene), the complex was of octahedral

    geometry, and the alkylidene ligand was found in a rare bridging (µ-alkylidene) mode. This

    difference in coordination chemistries between two ligand systems may stem from the

    presence of an additional π orbital on the phenolate oxygen, that destabilizes the orbitals

    configuration in the “axial pocket” found for the triamidoamine ligands (Figure I-3).

    In contrast, vanadium complexes of the amine tris(phenolate) ligands normally

    featured a penta-coordinate TBP geometry. For V(III), the remaining axial position was

    occupied by a neutral ligand (THF). For V(V), THF was replaced by a doubly-bonded ligand,

    the oxo group. This structure provides the first evidence for viability of the “triamidoamine-

    type” binding of our tripodal ligand to the metal: i.e., the structure in which the phenolate

    oxygens occupy equivalent equatorial positions, and the multiply bonded ligand is found in

    the axial position.

    We found that an exchange between these two geometries could take place if an

    electron-deficient amine tris(phenolate) ligand was bound to V(V) (Scheme D-6). This

    complex was found to support both geometries at RT, coordinating an additional neutral

    molecule at the sixth coordination site. As expected, the equilibrium between the

    pentacoordinate and the hexacoordinate geometries was temperature-dependent, with the

    TBP form prevailing at elevated temperatures, and the octahedral form prevailing at RT. The

    preference for a hexacoordinate complex in this case is attributed to the overall electron-

    deficiency at