+ All Categories
Home > Documents > PROJECT SUMMARY REPORTchem.chem.rochester.edu/~wdjgrp/doe/doe04.pdfExecutive Summary of FY 2001-4...

PROJECT SUMMARY REPORTchem.chem.rochester.edu/~wdjgrp/doe/doe04.pdfExecutive Summary of FY 2001-4...

Date post: 21-Oct-2020
Category:
Upload: others
View: 0 times
Download: 0 times
Share this document with a friend
25
PROJECT SUMMARY REPORT submitted to THE U.S. DEPARTMENT OF ENERGY for TRANSITION METAL ACTIVATION AND FUNCTIONALIZATION OF CARBON-HYDROGEN BONDS William D. Jones, Principal Investigator University of Rochester Department of Chemistry Rochester, New York 14627 Phone: 585-275-5493 Grant No. DE-FG02-86ER13569 Total Project Period: December 1, 2001 - November 30, 2004 Total Award Amount (3 years): $ 395,000 Unexpended Balance from Previous Year: $0
Transcript
  • PROJECT SUMMARY REPORT

    submitted to

    THE U.S. DEPARTMENT OF ENERGY

    for

    TRANSITION METAL ACTIVATION AND FUNCTIONALIZATION OF CARBON-HYDROGEN BONDS

    William D. Jones, Principal Investigator University of Rochester

    Department of Chemistry Rochester, New York 14627

    Phone: 585-275-5493

    Grant No. DE-FG02-86ER13569

    Total Project Period: December 1, 2001 - November 30, 2004 Total Award Amount (3 years): $ 395,000

    Unexpended Balance from Previous Year: $0

  • 2

    Executive Summary of FY 2001-4 Research in the Chemical Sciences Transition Metal Activation and Functionalization of Carbon-Hydrogen Bonds Grant FG02-86ER13569 William D. Jones, Department of Chemistry, University of Rochester, Rochester, NY 14627 Total Grant Period: 12/1/01-11/30/04, $395,000 for three years

    Overview of Research Goals and Accomplishments: This project has as its overall goal improvement in the intelligent use of our energy

    resources, specifically pertroleum derived products, and is aimed at the development of new routes for the manipulation of C-H and C-C bonds. During the 3 year project period, our research has focussed on the following specific goals: (1) discovery of new carbon-carbon bond cleavage reactions, (2) fundamental studies of C-H bond cleavage reactions of trispyrazolyl-boraterhodium complexes, (3) catalytic C-H and C-C bond functionalization, and (4) carbon-fluorine bond activation. This year we have made progress in several of these areas, as described in the following report.

    The specific accomplishments of the current grant period include:(1) we have successfully measured and qunatitatively modeled the processes available to metal alkyl hydride complexes in a trispyrazolylborate-rhodium complex; (2) we have for the first time determined the isotope effects on the specific fundamental steps involved in alkane oxidative addition and reductive elimination; (3) we have preliminary results that measure for the first time the selectivity for a metal fragment binding to methyl vs methylene groups in a linear hydrocarbon; (4) we have cleaved C-C bonds in biphenylene, aryl-nitriles, and aryl-acetylenes, expanding tremendously the breadth of C-C cleavage; (5) we have established 3 different mechanisms for C-F bond cleavage of fluorocarbons using Cp*2ZrH2; (6) we have synthesized and studied several hemi-labile P-N complexes of platinum and nickel, and demonstrated that the labile nitrogen give enhanced reactivity not observed with the well-known P-P chelate complexes; (7) we have discovered new C-H and C-C functionalizations that allow introduction of reactive boronate and olefin functional groups; (8) we have cleaved C-C bonds in allyl nitriles, leading to isomerization of the C-C skeleton. This reaction is critical to the DuPont synthesis of Nylon from butadiene; (9) we have measured C-H activation selectivities in chloroalkanes, showing a preference for activation of the terminal methyl groups and the facile β-chloride elimination reaction.

    A wide variety of chemistry has been examined, resulting in publication of several manuscripts. The work has been communicated at both national and international meetings. DOE funds have been used for the partial support of 4 graduate students (Andrew Vetter, Brad Kraft, Steve Oster, Karlyn Skugrud) and 1 postdoc (Nicole Brunkan) during the current grant period, as well as several undergraduates (Jason Holt, Mike Evans, John Curley, Brian Warsop, Susan Golisz).

    The continued success of this work will lead to the development of new techniques and processes for the manipulation of petroleum-based hydrocarbons. These new processes will be based upon the new methods for making and breaking strong bonds in organic molecules of the type studied here. The work has the potential to have a significant impact in science and in technologies of interest to DOE.

    A detailed report follows, followed by a listing of the 20 DOE supported publications for the current grant cycle and recent special recognitions received by the PI.

  • DOE Report, 2001-2004 3 William D. Jones

    Detailed Progress Report for the Project Period Dec. 1, 2001- Nov. 30, 2004.

    This report summarizes research that has been performed since during the current 3-year

    grant, as well as work that will be completed and published by the end of the grant period.

    1. Tris-pyrazolylborate Rhodium C-H Activation Studies.

    Our rhodium-trispyrazolylborate studies on hydrocarbon activation make use of the

    reactive 16-electron fragment [HB(3,5-dimethylpyrazolyl)3]Rh(CNCH2CMe3), abbreviated

    herein as Tp'RhL. In the prior 3-year project period, we established that the Tp'RhL fragment

    coordinates an alkane to give a σ-complex. A series of labelling studies allowed the

    determination of the relative rates of the processes available to the alkane σ-complex,

    specifically: (1) C-H activation (called oxidative cleavage), (2) migration down the alkane chain,

    or (3) simple dissociation. These experiments involved modelling of the scrambling of

    deuterium in complexes such as Tp'Rh(L)(CH2CH2CH3)D before loss of propane-d1. Figure 1

    shows the relative rates of these processes for methyl, ethyl, n-propyl, and n-butyl derivatives.

    For methane, C-H activation is strongly preferred over dissociation, whereas for ethane, the rates

    of these two processes are closer. End-to-end migration in ethane is intermediate. For propane,

    terminal C-H activation is favored over dissociation to a lessor extent than methane, but

    comparable to ethane. Migration from the end to the middle of propane is slightly slower than

    C-H activation. For the secondary propane complex, migration to the end and dissociation occur

    at about the same rate. Interestingly, migration down a butane chain (secondary to secondary) is

    the fastest process, accounting for the observed kinetic preference for terminal C-H activation.

    0 5 10 15

    methane

    ethane

    propane

    butane

    migration from 2°-2°migration from 2°-1°2° dissociation

    migration from 1°1° dissociationactivation

    relative rate (kd1=1)

    fixed ratio

    Figure 1. Relative rates of σ-alkane processes.

  • 4

    These conclusions can be expressed in a schematic fashion as shown in Scheme 1. Our

    studies do not permit the determination of either the absolute or relative energies of the primary

    and secondary alkane complexes, so we cannot establish the relative rates of processes between

    these two intermediates. Relative rates are shown assuming the energies of the two alkane

    complexes are equal.

    Scheme 1: Relative rates of processes available in primary (1°) and secondary (2°) alkane

    complexes.

    *Assuming 1° alkane σ-complex is at the same energy as the 2° alkane σ-complex

    (6x)

    (3x)

    (1x)

    MH

    M C C2H5H

    HCH3

    M +

    M CH

    HH

    koxidative cleavage

    kmigration

    kdissociation

    MH

    primary

    primary

    butane σ-complex

    secondarykoxidative cleavage

    kmigration(back)(3x)*

    (11x)*

    (1.6x)*secondarykdissociation

    CH

    HH

    M

    CH

    HCH3

    Mkmigration(2°-2°)

    (0.01x)*

    With the relative rates of all of these processes now known for any linear alkane, we have

    completed studies to determine which C-H bond of an alkane first binds to the Tp'RhL fragment.

    The execution of this experiment is not completely obvious, as reaction of the fragment with any

    linear hydrocarbon only gives a single product, the n-alkyl hydride (eq 1). One cannot determine

    how the alkane initially bound if a single product is observed.

    kprimary

    ksecondary

    MH

    M C CH3H

    HCH3

    M +

    MH

    M CH

    HH

    (1)

    In the current project period, we established that (1) the Tp'RhL fragment coordinates to a

    linear alkane to give a σ-complex and that the coordination is favored at the methylene group

    over the methyl group by a ratio of 1.5:1. (2) a methyl group in pentane coordinates 1.9 times

  • 5

    faster than the more hindered methyl group of isobutane. (3) the oxidative cleavage of a methyl

    C-H bond (primary C-H) occurs 65K12 times faster than the C-H bond in a methylene group

    (secondary C-H), where K12 represents the equilibrium constant between primary and secondary

    alkane complexes. These conclusions were made building upon our earlier studies of the relative

    rates of oxidative cleavage, migration, and dissociation indicated in Scheme 1. The experimental

    evidence for these conclusions is given below.

    The first new result is determination of the rate at which a reactive metal complex with a

    vacant site coordinates a methyl group C-H vs a methylene C-H. This was determined by

    looking at the product distribution in a competition experiment between pentane and decane. In

    this experiment, the metal intermediate would see an equal number of methyl groups in each

    substrate but differing numbers of methylenes (see Scheme 2). Note that in either case, only the

    terminal activation product would be seen since coordination to an internal methylene would

    lead to migration of the metal along the chain to the end, where oxidative cleavage would occur.

    If

    Scheme 2: Pathways for C-H coordination/activation in pentane-decane competitition.

    kd1kd2kd2 kd2

    kd2

    Tp'LRh

    H3kOC

    H

    km12

    km21

    RhLTp'

    C HH

    H

    RhLTp'

    C HH

    H3C

    RhLTp'

    C HH RhLTp'

    C HH

    RhLTp'

    C HHkm22

    km22

    km22

    km22

    km22

    km22

    km22

    km22

    km22

    km22

    km22

    km22Tp'LRh

    CHH

    Tp'LRh

    CHH

    Tp'LRh

    CHH Tp'LRh

    CHH

    CH3 CHH

    H

    Tp'LRh

    km12

    3kOCH

    RhLTp'

    H

    km22

    km22

    kCH3kCH2kCH2kCH2 kCH2kCH2 kCH2

    kCH2kCH2kCH3 kd2kd2

    kd2 kd2kd1

    A

    B C D E FG H

    I J K L M N O P Q RS T

    1 2 3 4 5 6

    78 9 10 11

    1213 14 15 16 17 18 19 20 21 22

    2324

    2526

    27 28 29 3031

    32

    km21

    L = CNCH2CMe3

    kCH2kCH3

    km21

    km12

    km22

    km22

    RhTp'L

    H

    Tp'LRh

    C HH

    HRhLTp'

    C HH

    H3CRhLTp'

    C HH

    pentane/decane

    3kOCH

    kd1 kd2kd2

    RhLTp'

    C HH

    H3CRhLTp'

    C HH

    H

    Tp'L Rh

    H

    km22

    km22 km12

    km21

    3kOCH

    kCH2 kd2 kd1kCH3kCH2

    Tp'LRh(PhN=L)hν

    -PhN=L[Tp'LRh]

    kCH3 @ 7, 11, 23, 32

    kCH2 @ 8, 9, 10, 24, 25, 26, 27, 28, 29, 30, 31

    kRCH

    kRCH

    the reactive intermediate were to bind only to methyl groups C-H bonds, then a 1:1 product ratio

    would be expected. If the reactive intermediate were to bind only to methylene C-H bonds, then

  • 6

    a product ratio favoring decane activation would be expected (corresponding to the increased

    number of methylene groups).

    Scheme 2 shows all possible intermediates and pathways. It is rather complicated (!), but

    since the earlier studies established many of these relative rates, only the rates of primary

    coordination (kCH3) and secondary coordination (kCH2) need to be determined. In fact, we only

    want to know the ratio of these rates so that the ratio can be adjusted in a simulation to match the

    experimentally observed ratio of products. In the experiment, Tp*Rh(CNR)(carbodiimide) was

    irradiated in a 1:1 mixture of pentane/decane. The alkyl hydride products were quenched with

    CCl4 to give the stable chloro derivatives in a 1.1:1 ratio (Scheme 3). Consequently, the results

    are consistent with a slight preference for methylene group coordination, and the simulation

    indicates an actual preference of kCH2/kCH3 = 1.5:1.

    Scheme 3. Competition between pentane and decane.

    [Tp'LRh]-PhN=L

    hνTp'LRh(PhN=L)

    L = CNCH2CMe3

    Tp'LRhH

    Tp'LRhH

    +

    -20 °C

    CCl4

    +

    Tp'LRhCl

    Tp'LRhCl

    1

    :

    1.1

    Steric hinderence can interfere with the ability to coordinate to a methyl group C-H bond.

    This can be seen in a competition between pentane and isobutane. In this experiment, two

    methyl groups in a linear alkane compete with three methyl groups in a branched alkane. The

    competitive reactions involved are shown in Scheme 4, with the two different rate constants for

    methyl group binding shown as kCH3 and kCH3'.

  • 7

    Scheme 4. Pathways for C-H coordination/activation in pentane-isobutane competitition.

    3kCH3'

    3kOC1°

    Tp'LRhC

    H

    H

    Tp'LRhC

    H

    HCH3

    Tp'LRhC

    H

    HH

    Tp'LRhH

    3kOC1°

    km22 km22

    km12km212kCH3

    2kCH2

    kCH2

    Tp'LRh

    H

    Tp'LRhC

    H

    HH

    [Tp'LRh]

    -PhN=Lhν

    Tp'LRh(PhN=L)kd1

    kd2

    kd2

    kd2

    kCH3/kCH3' = 1.9

    In the experiment, Tp*Rh(CNR)(carbodiimide) was irradiated in a 1:1 mixture of

    pentane/isobutane. The alkyl hydride products were quenched with CCl4 to give the stable

    chloro derivatives in a 1.35:1 ratio (Scheme 5). Once again, simulation of the experiment while

    varying only the ratio kCH3/kCH3' allows determination of the ratio as 1.9:1. Note that in this

    experiment, we assume that the rate of oxidative cleavage of the C-H bond in the two methyl

    complexes is the same. It is possible to interpret this experiment in terms of an equal rate of

    binding to the two types of methyl groups, but with a 1.9:1 ratio between the rate of the two

    oxidative cleavage rates.

    Scheme 5. Competition between pentane and isobutane.

    [Tp'LRh]-PhN=L

    hνTp'LRh(PhN=L)

    L = CNCH2CMe3

    Tp'LRhH

    Tp'LRhH

    +

    -20 °C

    CCl4+

    Tp'LRhCl

    Tp'LRhCl

    1.35

    :

    1

    In the pentane/decane competition experiment, we learned about the relative rates of

    binding of unhindered CH3 vs CH2. It would be reasonable to expect this rate ratio to apply in

    other competitions involving the binding of similar types of bonds. For example, the methylene

  • 8

    C-H bonds in cyclohexane would be expected to bind to the [Tp*RhL] fragment with the same

    ease as those in pentane, since cyclohexane is an unstrained, 'natural' conformation similar to

    that in a linear alkane. Interestingly, cyclohexane does undergo activation of its C-H bonds by

    the [Tp*RhL] fragment because secondary C-H bonds are the only ones available. In a

    competition between pentane and cyclohexane, the metal fragment should bind to both alkanes

    in a predictable fashion since kCH3/kCH2 is known. From the ratio of products obtained, one

    should be able to determine the relative rates of oxidative cleavage of a secondary C-H bond

    (kOC2°) vs a primary C-H bond (kOC1°), provided the alkane σ-complexes are at the same energy.

    These pathways and the corresponding rate constants are shown in Scheme 6.

    Scheme 6. Pathways for C-H coordination/activation in pentane-cyclohexane competitition.

    kRC2

    2kOC2°

    1312kd2 kCH2

    Tp'L Rh

    H

    RhLTp'

    HH

    kRCH

    kRCH

    [Tp'LRh]-PhN=L

    hνTp'LRh(PhN=L)

    kCH2 kCH3kd1

    kd2kCH2

    3kOCH

    km21

    km12km22

    km22

    Tp'L Rh

    H

    RhLTp'

    C HH

    H

    RhLTp'

    C HH

    H3C

    kd2kd2

    kd1

    3kOCH

    pentane/cyclohexane

    RhLTp'

    C HH

    RhLTp'

    C HH

    H3C

    Tp'LRh

    C HH

    H RhTp'L

    Hkm22

    km22

    km12

    km21

    kCH3kCH2

    L = CNCH2CMe3

    1110987

    654321

    HG FEDCB

    A

    kCH2 @ 8, 9, 10, 12

    kCH3 @ 7, 11, 23, 32

    kOC1°/kOC2° = 65 (with 1° and 2° alkane complexes at same E)

    or otherwise:kOC

    1°/kOC2° = 65 (km12/km21)= 65 (K12)

    I J

    In the competition experiment, Tp*Rh(L)(carbodiimide) was irradiated in a 1:1 mixture of

    pentane/cyclohexane and quenched with CCl4. A 6.5:1 ratio of the n-pentyl and cyclohexyl

    chloride products was seen (Scheme 7). From the simulation, this corresponds to a 65:1 ratio for

    kOC1°/kOC2°, assuming the alkane σ-complexes are at the same energy. If they are at different

    energies, then a 'correction' must be applied corresponding to Keq between the two σ-complexes.

  • 9

    Scheme 7. Competition between pentane and cyclohexane.

    [Tp'LRh]-PhN=L

    hνTp'LRh(PhN=L)

    L = CNCH2CMe3

    Tp'LRhH

    Tp'LRhH

    +

    -20 °C

    +

    Tp'LRhCl

    6.5

    :

    1 Tp'LRhCl

    CCl4

    Finally, we have also looked at activation of cyclopentane vs pentane in a competition

    experiment. Here, due to the strain in the cyclopentane ring, it is not reasonable to assume that

    the rate of methylene coordination is the same as in a linear alkane. Consequently, the scheme

    for this experiment shows different rates for the two types of methylene coordination (Scheme

    8), but the same rate of oxidative cleavage in the methylene-alkane complex.

    Scheme 8. Pathways for C-H coordination/activation in pentane-cyclopentane competitition.

    5kCH2'

    3kOC1°

    Tp'LRhH

    2kOC2°

    km22 km22

    km12km212kCH3

    2kCH2

    kCH2

    Tp'LRh

    H

    kd1

    kd2

    kd2

    kd2

    kCH2 /kCH2 = 1.7'

    Tp'LRhC

    H

    H

    Tp'LRhC

    H

    HCH3

    Tp'LRhC

    H

    HH

    Tp'LRhC

    H

    H

    [Tp'LRh]

    -PhN=Lhν

    Tp'LRh(PhN=L)

    In the competition experiment, Tp*RhL(carbodiimide) was irradiated in a 1:1 mixture of

    pentane and cyclopentane, and CCl4 added to quench the products. A 4.5:1 ratio of n-pentyl to

    cyclopentyl products were seen (Scheme 9), showing that cyclopentane is slightly more reactive

    than cyclohexane, as anticipated. From the kinetic simulation of the competition, a ratio for

    kCH2'/kCH2 of 1.7 was obtained.

  • 10

    Scheme 9. Competition between pentane and cyclopentane.

    [Tp'LRh]-PhN=L

    hνTp'LRh(PhN=L)

    L = CNCH2CMe3

    Tp'LRhH

    +

    -20 °C

    +Tp'LRhCl

    4.5 : 1Tp'LRhH

    Tp'LRhCl

    CCl4

    As a check that this value represents the relative coordination abilities of cyclopentane vs

    cyclohexane, a competition experiment was run by irradiation of Tp*RhL(carbodiimide) in a 1:1

    mixture of cyclopentane/cyclohexane. A 1.9:1 ratio of cycloalkyl chloride products was

    observed. The excellent agreement between the competition ratio from these two independent

    experiments argues that the assumptions made in the simulations are reasonable.

    One of the more interesting side-lights from these studies comes from the independent

    determination of isotope effects for both the 'oxidative cleavage' and the 'reductive coupling'

    steps of the C–H activation reaction indicated in equation 2. These isotope effects, both kinetic

    isotope effects on a fundamental reaction step, were found to be normal isotope effects. The

    overall effect on alkane reductive elimination, however, is to generate an inverse kinetic isotope.

    The initial equilibrium isotope effect between the alkyl hydride complex and the alkane sigma-

    complex is inverse, not because either of the individual rates are inverse, but because the ratio of

    these isotope effects is inverse. As this was the first known system where these effects had been

    completely sorted out, we published a more didactic article in Accounts of Chemical Research to

    shed light on the analysis of this controversial subject.

    slow+Lfast Ln+1M

    LnMCH2R

    D

    [LnM] + R-CH2D

    LnMCH2R

    DLnM

    CHDR

    H kRCH

    kOCH

    kOCD

    kRCD

    (2)

    2. C-C Bond Cleavage Studies

    Our DOE supported work showed that several types of C-C bonds can be cleaved. We

    have discovered 3 distinct classes of C-C bonds that can be cleaved: (1) strained rings such as

    biphenylene undergo sp2-sp2 C-C cleavage with a number of metal complexes to give a variety

    of products. (2) diphenylacetylene undergoes sp-sp2 cleavage photochemically when attached to

    PtL2 complexes. (3) arylnitriles undergo sp2-sp C-CN cleavage when reacted with NiL2

  • 11

    fragments, and η2-nitrile adducts can be observed as reaction intermediates. These results are

    described in more detail below.

    Earlier DOE supported work showed that the complexes Pt(PEt3)3, Pd(PEt3)3, and [Ni

    (dippe)H]2 cleave the C-C bond of biphenylene to give (PEt3)2Pt(2,2'-biphenyl), (PEt3)2Pd(2,2'-

    biphenyl), and (dippe)Ni(2,2'-biphenyl), respectively. These complexes underwent catalytic

    chemistry to give functionalized products, such as tetraphenylene, biphenylene, fluorenone, or

    phenanthrenes (Scheme 10).

    Scheme 10:

    ML

    L

    + ML2

    L2 = (PEt3)2, dippeM = Ni, Pd, Pt

    H2

    CO

    O

    RC CRR R

    In the case of Ni(dippe), it was found that catalysis required the introduction of small

    amounts of oxygen, just enough to oxidize the phosphine to phosphine oxide. The remaining

    'naked' metal was efficient at catalyzing the insertion of alkynes into biphenylene to give

    phenanthrenes. The requirement of a labile chelate led us to investigate the use of the hemi-

    labile ligand Pri2PCH2CH2NMe2 in these same metal systems during the current grant period.

    The platinum complexes proved to be the easiest to synthesize and study, since the adducts

    are fairly stable. The strategy was to use a source of Pt(0) in the presence of the P-N chelate and

    an alkyne, to isolate the complex, and then to look at reactions of the complex. With

    diphenylacetylene, the preparation of the adduct was straightforward and allowed comparison of

    the P-P chelate with the P-N chelate (eq 3).

    Pt

    PR2

    ER2´

    + PtP

    E

    R2

    R2´- CODCH2Cl2

    E = N, R = iPr, R´= Me: 1E = P, R = R´= iPr: 2 (3)

  • 12

    While the P-P chelated complex showed no reactivity with added diphenylacetylene even

    after heating, the P-N chelated complex reacted at room temperature to give a

    metallacyclopentadiene complex. Further studies of the system showed that the nitrogen of the

    chelate is quite labile, and that once it dissociates the phosphine ligands can then redistribute

    between metals to generate two observable intermediates (Scheme 11). Ultimately, however,

    thermodynamics takes over and one winds up with quantitative production of the metallacycle.

    Heating this sample results in the slow catalytic formation of hexaphenylbenzene.

    Scheme 11:

    + 2 NMe2

    Pt

    (Pri)2 P C

    C

    Ph

    Ph

    NMe2

    Pt

    (Pri)2 P

    Ph

    Ph

    Ph

    Ph

    Pt

    Me2NPt

    PPri2

    PhPh

    PhPh+ PtMe2N

    R2P

    Pt

    PR2

    Me2N

    2 2

    With trimethylsilylphenylacetylene, the preparation of the initial complex is similar but

    now the alkyne complex undergoes insertion into the C-Si bond at room temperature (eq 4).

    Once again, no such reaction is seen for the complex with a bis-phosphine chelate. Just as in the

    case of diphenylacetylene, ligand dissociation and redistribution is observed at intermediate

    times to give two observable intermediates (Scheme 12). Once again, thermodynamics takes

    over and only a single product, the C-Si insertion complex, is observed at longer reaction times.

    NMe2

    Pt

    (Pri)2 P

    SiMe3

    C6D6, 5 days

    22 °C

    NMe2

    Pt

    (Pri)2 P SiMe3

    (4)

  • 13

    Scheme 12:

    +

    2 2

    SiMe3

    + 2 NMe2

    Pt

    (Pri)2 P C

    C

    SiMe3

    Ph

    PtMe2N

    R2P

    Pt

    PR2

    Me2N

    SiMe3

    NMe2

    Pt

    (Pri)2 P SiMe3

    NMe2

    Pt

    (Pri)2 P SiMe3

    Me2 N

    Pt P(Pri)2

    SiMe3

    Pt

    Me3Si

    While the dippe P-P chelate complex is unreactive thermally, it does react photochemically

    to give a C-Si insertion product (eq 5). Remarkably, however, this complex reverts to the Pt(0)

    alkyne complex thermally! This observation leads to the important conclusion that oxidative

    addition is favored thermodynamically by the presence of the P-N ligand, whereas reductive

    elimination is thermodynamically favored by the P-P ligand. This conclusion implies that the

    study of a chelating, sterically hindered, bis-(dialkylamino)ethane ligand should give a metal

    fragment that will strongly favor C-C cleavage. This hypothesis remains to be tested.

    C6D6

    PtP

    P

    SiMe3Pt

    P

    P

    SiMe3Cy2

    Cy2 Cy2

    Cy2hν

    3 4 (5)

    In addition to the above work with platinum, we have also prepared the analogous nickel

    complex (eq 6). These complexes have proven to be efficient catalysts for the selective coupling

    of biphenylene and alkynes to give phenanthrenes (eq 7).

    Ni + +

    P(iPr)2

    NMe2

    NiP

    N- 2 COD(1)

    (iPr)2

    Me2

    1 (6)

    PhPhPhPh

    +

    T = 70°C

    NiP

    NPh

    Ph(iPr)2

    Me2

    (7)

  • 14

    We have also discovered that the nickel complex [Ni(dippe)H]2 reacts with benzonitrile to

    give first an η2-nitrile complex, which then undergoes C-C cleavage of the carbon-CN bond (eq

    8). Furthermore, the reaction does not go to completion but forms and equilibrium mixture of

    the η2-nitrile and C-CN oxidative addition product. We know of no such example of reversible

    C-C cleavage in the literature. Other examples of aryl C-CN cleavage are under investigation.

    N

    C

    Ph

    Ni

    Pri2 P

    PPri2

    Ni

    Pri2 P

    PPri2

    HNi

    Pri2 P

    PPri2

    H

    + 2 Ph-C N

    2-H2

    KeqNi

    Pri2 P

    PPri2

    Ph

    CN

    (8)

    We have also looked at the effects of electron withdrawing and electron donating groups

    on the aryl cyanide. The results show that both the rate and equilibrium for C-CN cleavage is

    affected, as shown in the plots in Figure 2. For the equilibration, Keq is found to have a ρ value

    of +6.1. This large and positive value for ρ indicates that negative charge is being stabilized on

    the ipso carbon of the substituted aryl group, consistent with the organometallic nature of this

    bond as possessing substantial Niδ+ —Cδ– character. From the rate of approach to equilibrium,

    the rate of the forward reaction k1 can be determined. Similarly, a plot of ln k1 vs. σ, also shown

    in Figure 2b, gives a ρ value of +1.3. While there is somewhat more scatter in this plot, the

    positive slope correlation is unmistakable and is consistent with the localization of charge

    density on the ipso carbon in the transition state for C-CN bond cleavage, although not so much

    as the Ni(II) product.

    -5

    -3

    -1

    1

    3

    5

    -1 -0.5 0 0.5 1

    σp

    ln( K

    eq/ K

    H)

    (a)

    -1

    -0.5

    0

    0.5

    1

    1.5

    -0.75 -0.25 0.25 0.75σp

    ln( k

    1/ kH)

    (b)

    Figure 2. (a) Hammett plot for the equilibrium constants (Keq) from equation 8 vs. σP at 54 °C. (b) Hammett plot for the forward rate constants (k1) vs. σP at 54 °C.

  • 15

    We have also discovered an important new type of C-C bond oxidative addition, cleavage

    of sp-sp2 C-C bonds in aryl acetylenes. This is a new class of C-C bond cleavage, and offers

    many exciting possibilities. We have found that irradiation of either P-P or P-N chelate

    complexes of Pt-(diphenylacetylene) leads to the clean and quantititive formation of the

    oxidative addition product (eq 9). The reaction works for all complexes we have examined to

    date. Furthermore, the oxidative cleavage reaction is reversible, in that heating the Pt(II)

    complexes results in their reversion to the Pt(0)-alkyne complexes. Consequently, we now have

    methodology to break and make C-C———C bonds, and further development of this reaction will be

    the subject of future studies.

    ER2'

    R2P

    Pthν

    C6D6 ER2'

    R2P

    Pt

    E = N, R = iPr, R'= MeE = P, R = R'= iPrE = P, R = R'= Cy

    (9)

    Recently, we have initiated investigations of a new class of C-C cleavage, that of allyl-

    nitriles. At a metal center this cleavage reaction generates both a strong metal-cyanide bond and

    a π-allyl ligand, and hence has been found to be both facile and reversible. Indeed, at nickel(0),

    this reaction forms the basis of DuPont's synthesis of adiponitrile for the production of nylon via

    addition of HCN to butadiene, to the tune of over 400 thousand metric tons per year!

    We have discovered that the reactive hydride [(dippe)NiH]2, which serves as a room

    temperature source of [Ni(dippe)], reacts with allylcyanide to give initially a π-olefin complex

    (dippe = bis-(diisopropylphosphino)ethane). This species can be seen at low temperature by

    NMR spectroscopy, and upon warming to RT competitive C-H and C-CN cleavage takes place.

    C-H activation gives a π-allyl hydride complex that is not observed, because the hydride is

    transferred back to the opposite end of the allyl group to give a very stable crotononitrile

    complex (both cis and trans are formed). C-CN activation, however, leads to a metastable π-

    allyl cyanide complex that can be isolated and was structurally characterized. C-CN cleavage is

    reversible, so that ultimately, all nickel winds up as the crotononitrile complexes (Scheme 13).

  • 16

    Scheme 13. C-C and C-H Bond Activation in allylcyanide.

    k2, k3

    [Ni(dippe)H2]2

    CN+

    PPri2

    Ni

    Pri2 P

    CN

    PPri2

    Ni

    Pri2 P

    H

    CN

    PPri2

    Ni

    Pri2 P CN

    PPri2

    Ni

    Pri2 P

    NC

    k1, C-C activation

    C-H activation

    C-C formation, k-1 ∆H‡ (kcal/mol) ∆S‡ (e.u.) ∆G‡ at 40 °C (kcal/mol)k1 (C-C cleavage) 18.0(0.3) -14.8(0.8) 22.71k-1 (C-C formation) 27.5(0.5) 11.7(1.6) 23.75k2 (C-H cleavage) 13.4(0.2) -29.9(0.6) 22.95k3 (C-H cleavage) 13.2(0.2) -31.2(0.7) 23.14

    By monitoring the distribution of species over time, we have been able to extract the rate

    constants for all of these species by kinetic simulation. In addition, by measuring the

    distribution of species as a function of temperature, we can obtain activation parameters for the

    various steps. The results are quite interesting, in that we find that while C-H activation and C-C

    activation have small temperature dependences, C-C cleavage has a large temperature

    dependence. The result is that by raising the temperature, one can selectively drive the reaction

    in the direction of the less-favorable π-allyl cyanide complex. This is good news, since the

    DuPont catalysis requires that the C-C cleavage dominate the C-H cleavage. The activation

    parameters support the mechanism for C-H anc C-C cleavage shown in Scheme 14.

    Scheme 14:

    Ni

    CN r.d.s.CN

    Ni

    H

    crotononitrilesfast

    Ni

    CN H

    H

    r.d.s.Ni

    CNNi

    CN

    Ni

    NC

    N

    CNi

    NiN

    C

    HH

    Ni

    CN

    We have also investigated the effect of the Lewis acid BPh3 upon the isomerization

    reaction. Addition of BPh3 to a cold solution of (dippe)Ni(η2-allylCN) leads to the immediate

    and quantitative formation of the BPh3 adduct of the π-allyl cyanide adduct. Addition of a

    second equivalent of the Lewis acid leads to the removal of the cyanide ligand as the

  • 17

    [Ph3B-CN-BPh3]- anion, leaving behind the [(dippe)Ni(π-allyl)]+ cation (Scheme 15). With only

    one equivalent of BPh3, there is evidence of linkage isomerism of the initial Ni-C-N-BPh3 adduct

    to the Ni-N-C-BPh3 adduct.

    Scheme 15. Effect of Lewis acid on the C-C and C-H Bond Activation in allylcyanide.

    PR2

    Ni

    R2P

    HNi

    H

    PR2

    R2P CN

    BPh3

    slow

    PiPr2

    Ni

    iPr2 P

    RT PiPr2

    Ni

    iPr2 P

    BPh3NC

    PiPr2

    Ni

    iPr2 P

    BPh3CN

    slow

    [Ph3B-NC] [Ph3B-CN]

    3. C-H and C-C Bond Functionalization Studies

    We have also initiated investigations of the above systems for their ability to serve in

    further functionalization reactions of hydrocarbons. For example, we have found that C-C bonds

    can not only be cleaved, but functionalized using 'standard' organic reagents. A nice example of

    the application of this chemistry appears in our work with palladium catalyzed reactions of

    biphenylene. Using a Pd(0) precursor, we can activate the C-C bond of biphenylene and then

    protonate one of the Pd-C bonds using a weak acid of the appropriate pH. Next, one can perform

    Heck or Suzuki-type couplings using olefins or boronic acids to give functionalized products (eq

    10, 11). Furthermore, other acidic C-H bonds can be added across the activated C-C bond using

    pH control. For example, α-keto C-H bonds or α-nitrile C-H bonds can add across biphenylene

    to give functionalized biaryls (eq 12, 13).

    +R

    (1.3 equiv)

    Pd(PPh3)4 (5 mol%)

    p-cresol (1 equiv)

    C6D6, 120 °C, 6 dR

    + R

    (10)

  • 18

    C6D6, 120 °C

    p-cresol (1 equiv)

    Pd(PPh3)4 (5 mol%)

    (1 equiv)

    +

    R

    B(OH)2

    R

    (11)

    C6D6, 120 °C

    p-cresol (10 mol%)

    Pd(PPh3)4 (5 mol%)+ R

    OR

    O (12)

    C6D6, 120 °C, 1 d

    p-cresol (10 mol%)

    Pd(PPh3)4 (5 mol%)

    (1 equiv)

    +

    Me

    CN

    CN

    Me (13)

    We have conducted an extensive investigation of C-H activation in chloroalkanes using the

    reactive precursor Tp*Rh(CNR)(carbodiimide). Remarkably, the C-Cl bond does not undergo

    oxidative addition. Rather, we find a strong selectivity for exclusive terminal methyl group C-H

    bond activation. Thus, 1-chloro alkane gives the 5-chloropentyl hydride as the only product. 3-

    chloropentane gives the 3-chloropentyl hydride product. In this case, however, two

    diastereomers are formed in a 1:1 ratio, since the metal is chiral and the 3-chloro substituent

    renders the ligand chiral. If a chlorine is present in a β-position, then β-chloride elimination

    occurs to give an olefin and the metal chloride. Therefore, 2-chloropropane gives only propene

    and the hydrido chloride Tp*Rh(CNR)HCl. 2-chloropentane gives a mixture of the C-H

    activation product 4-chloropentyl hydride, pentene, and the hydrido chloride. These reactions

    are summarized in Scheme 16.

    We have also begun to investigate aliphatic nitrile complexes. It appears that there is once

    again a selectivity for terminal methyl groups. With acetonitrile, the adduct Tp*LRh(CH2CN)H

    is formed, and is found to be stable for days at 60 °C! This is the most stable alkyl hydride in

    this series yet, and we may be able to do new functionalization chemistry with this derivative.

  • 19

    Scheme 16. Reactions of Chloroalkanes with [Tp*Rh(CNR)].

    Tp'LRh

    + PhNCNR

    L = CNCH2CMe3

    -20 °C

    Cl

    ClTp'LRh

    HCl

    Tp'LRh

    Cl

    H

    Cl

    +

    Tp'LRhClH

    + Tp'Rh(L)HCl

    + Tp'Rh(L)HCl

    ClTp'LRhN

    C

    R

    NPh

    4. C-F Bond Cleavage Studies

    We have now begun studies with the soluble, more reactive Cp*2ZrH2 and found that this

    molecule cleaves a wide variety of aromatic and aliphatic C-F bonds. Systematic studies have

    shown that primary, secondary, and tertiary C-F bonds can all be cleaved with progressively

    greater difficulty (Scheme 17). In addition, di-fluorosubstituted carbons can be made to react

    with even more forcing conditions. Trifluoromethyl groups scarcely react at all even under

    extreme conditions.

    Scheme 17:

    F

    F

    F

    H

    H

    H

    ++

    C6D12, H2 2 d, 25oC

    C6D12, H2

    C6D12, H2

    4 d, 120oC

    1 d, 150oC

    primary C-F:

    secondary C-F:

    tertiary C-F:

    ZrH

    H

    Me5

    Me5

    ZrF

    H

    Me5

    Me5

    Most remarkable, however, even trifluoromethyl C-F bonds can be easily cleaved if they

    are adjacent to a double bond. 3,3,3-trifluoropropene is completely defluorinated in 5 min at

    room temperature to give the zirconium-n-propyl hydride complex (Scheme 18).

    Perfluoropropene undergoes a similar reaction to give the same product. Details of the

  • 20

    Scheme 18:

    F2C CF3

    F

    d12

    Cp*2ZrHF +

    CF3

    F2C CH3

    H2

    CH3CH2CH3 + Cp*2ZrH2

    22°C

    + Cp*2ZrHF

    Cp*2ZrH2

    22°C

    d12

    excess

    + Cp*2ZrHF

    H2

    ZrH

    H

    Me5

    Me5

    ZrH

    CH2CH2CH3

    Me5

    Me5

    ZrH

    CH2CH2CH3

    Me5

    Me5

    mechanism are under further study. Defluorination reactions are also seen with

    nonafluorohexene, perfluorocyclobutene, perfluorocyclopentene, perfluorobenzene,

    trifluorotoluene, and related substrates. Chlorofluorocarbons (CFCs) react rapidly to give first

    fluorocarbons (HFCs), which then are converted to hydrocarbons (HCs) in accord with the above

    established reactivities (Scheme 19). Mechanistic investigations into the aliphatic fluorocarbons

    has revealed evidence for a radical chain mechanism. Further mechanistic work with the

    fluoroolefins is underway suggesting an insertion/β-fluoride elimination pathway.

    Scheme 19. x Cp*2ZrH2

    CFCl2H

    CF2Cl2

    CF2ClH

    CH3F

    CF2H2

    CF2H2

    CH4

    CH4

    CH4

    25oC, 5 min.

    +

    25oC, 5 min.

    25oC, 5 min.

    RT1 day

    120oC

    120oCx = 3, 4

    slow

    slow

    We have now continued our investigation of the C-F cleavage in perfluoroolefins. These

    appear to be a special class of substrate, in that the mechanism of C-F cleavage may be different

    than that seen in our earlier studies with Cp*2ZrH2. Reaction with perfluoropropene gives first

    the selective formation of E-CHF=CFCF3. Further reaction with zirconium hydride leads to

    complete defluorination with no further intermediates being seen (Scheme 20).

  • 21

    Scheme 20. Reaction of Cp*2ZrH2 with perfluoropropene.

    +CF3

    F

    H

    F

    6 Cp*2ZrH2

    - 5 Cp*2ZrHFCF3

    F

    F

    F+

    ZrF

    H

    ZrH

    H-40 C

    ZrH

    The mechanism for the reaction could involve hydridic attack on the olefin with H/F

    metathesis, or an insertion/elimination mechanism. The olefin could approach centrally,

    between the two hydrides, or laterally, with the two hydrides remaining cis to each other. In

    order to investigate these possibilities, we have initiated a collaboration with a theory group in

    Montpellier. Odile Eisenstein and Eric Clot have provided high level calculations investigating

    these systems, and the work is providing guidance for the mechanism of reaction. Initial studies

    were done on the insertion reaction of ethylene with Cp2ZrH2. We then progressed to

    trifluoropropene, which has been shown experimentally to undergo an insertion/β-fluoride

    elimination pathway. The calculations done thus far are able to reproduce the selective internal

    insertion product, and further work is underway to look at the β-fluoride elimination step

    (Scheme 21).

    Scheme 21. Calculations on the reaction of Cp*2ZrH2 with trifluoropropene.

    externalcoordination

    internalcoordination

    internalcoordination

    0.0

    -22.8

    -36.8

    C2

    D2

    C1

    Cp

    Zr

    CpH

    H

    CF3Cp

    Zr

    CpH

    H

    CF3

    Cp

    Zr

    CpH

    H

    CF3

    -22.6

    Cp2ZrH2 + C2H3CF3

    E2

    -31.6

    Cp

    Zr

    CpH

    H

    CF3D1E1

    -32.1

    Cp

    Zr

    CpH

    CF3Cp

    Zr

    CpH

    CF3

    -32.0

    Further calculations have been recently completed with Cp2ZrH2 reacting with

    perfluoropropene. The results show that interaction of the olefin internally with Cp2ZrH2 leads

    to olefin insertion followed by an external β-fluoride elimination. Other pathways were

    investigated but all were found to lie at higher energies (Scheme 22).

  • 22

    Scheme 22. Calculations on the reaction of Cp*2ZrH2 with perfluoropropene.

    ���

    ���

    ����

    ����

    ����

    �����

    �����

    ��

    ����

    �����

    ��

  • 23

    Publications appearing during the current grant cycle acknowledging DOE support, December 1, 2001 - November 30, 2004:

    Manuscripts in print:

    1. “Perspectives: Synthetic Chemistry. The Key to Successful Organic Synthesis is…,” William D. Jones, Science, 2002, 295, 289-290.

    2. “Formation of Tetrafluorobenzyne by β-Fluoride Elimination In Zirconium-Perfluorophenyl Complexes,” Bradley M. Kraft, Rene J. Lachicotte, and William D. Jones, Organometallics, 2002, 21, 727-731.

    3. “Cleavage of the Carbon-Carbon Bond in Biphenylene using Transition Metals,” Christophe Perthuisot, Brian L. Edelbach, Deanna L. Zubris, Nira Simhai, Carl N. Iverson, Christian Müller, Tetsuya Satoh, and William D. Jones, J. Mol. Catal. A, Chemical, 2002, 189, 157-168.

    4. “Chelating P,N versus P,P Ligands: Differing Reactivity of Donor Stabilized Pt-(η2-PhC———CPh) Complexes Towards Diphenylacetylene,” Christian Müller, Rene J. Lachicotte, and William D. Jones, Organometallics, 2002, 21, 1118-1123.

    5. “Thermal and Photolytical Silicon-Carbon Bond Activation in Donor Stabilized Pt(0)-Alkyne Complexes,” Christian Müller, Rene J. Lachicotte, and William D. Jones, Organometallics, 2002, 21, 1190-1196.

    6. “Catalytic C-C Bond Activation in Biphenylene and Cyclotrimerization of Alkynes: Reactivity of P,N versus P,P Substituted Nickel Complexes,” Christian Müller, Rene J. Lachicotte, and William D. Jones, Organometallics, 2002, 21, 1975-1981.

    7. “Mechanism of Vinylic and Allylic Carbon-Fluorine Bond Activation of Non-Perfluorinated Olefins using Cp*2ZrH2,” Bradley M. Kraft, Rene J. Lachicotte, and William D. Jones, J. Am. Chem. Soc., 2002, 124, 8681-8689.

    8. “Cleavage of Carbon-Carbon Bonds in Aromatic Nitriles using Nickel(0),” Juventino J. Garcia, Nicole M. Brunkan, and William D. Jones, J. Am. Chem. Soc., 2002, 124, 9547-9555.

    9. “Carbon-Fluorine Bond Activation of Perfluorinated Arenes with Cp*2ZrH2,” Bradley M. Kraft and William D. Jones, J. Organomet. Chem., 2002, 658, 132-140.

    10. “η2- Coordination and C–H Activation of Electron-poor Arenes,” Carl N. Iverson, Rene J. Lachicotte, Christian Müller and William D. Jones, Organometallics, 2002, 21, 5320-5333.

    11. “Isotope Effects in C-H Bond Activation Reactions by Transition Metals,” William D. Jones, Acc. Chem. Res. 2003, 36, 140-146.

    12. “Preparation, Structure, and Dynamics of a Nickel π-Allyl Cyanide Complex,” Nicole M. Brunkan and William D. Jones, J. Organomet. Chem. 2003, 683, 77-82.

    13. “Activation of C-F Bonds using Cp*2ZrH2: A Diversity of Mechanisms,” William D. Jones, J. Chem. Soc., Dalton Trans. 2003, 3991-3995.

    14. “Synthesis, Characterization, and Reactivity of a Rhenium Complex with a Corannulene Based Ligand,” Robert M. Chin, Benjamin Baird, Michael Jarosh, Shane Rassman, Brian Barry, and William D. Jones, Organometallics 2003, 22, 4829-4832.

  • 24

    15. “Structural Properties and Inversion Mechanisms of [Rh(dippe)(µ−SR)]2 Complexes,” Steven S. Oster and William D. Jones, Inorg. Chim. Acta, 2004, 357, 1836-1846.

    16. “Carbon-hydrogen bond activation of chloroalkanes by a rhodium trispyrazolylborate complex,” Andy J. Vetter and William D. Jones, Polyhedron, 2004, 23, 413-417.

    17. “Kinetics, Thermodynamics and Effect of BPh3 on Competitive C-C and C-H Bond Activation Reactions in the Interconversion of Allyl Cyanide by [Ni(dippe)],” Nicole M. Brunkan, Donna M. Brestensky, and William D. Jones, J. Am. Chem. Soc., 2004, 126, 3627-3636.

    18. “Defluorination of Perfluoropropene using Cp*2ZrH2 and Cp*2ZrHF: A Mechanistic Investigation from a Joint Experimental-Theoretical Perspective,” Eric Clot, Claire Megret, Bradley M. Kraft, Odile Eisenstein, and William D. Jones, J. Am. Chem. Soc., 2004, 126, 5647-5653.

    Manuscripts Submitted or in press:

    19. “Alkane Complexes as Intermediates in C-H Bond Activation Reactions,” William D. Jones, Andrew J. Vetter, Douglas D. Wick, Todd O. Northcutt, ACS Symposium Series, in press.

    20. “Cleavage of Carbon-Carbon bonds in Alkyl Nitriles Using Nickel(0),”.Juventino J. García*, Alma Arévalo, Nicole M. Brunkan, and William D. Jones, Organometallics, submitted.

    Manuscripts in Preparation:

    21. “Alkane Coordination Selectivity in Hydrocarbon Activation by [Tp'Rh(CNneopentyl)]: The Role of Alkane Complexes,” Andrew J. Vetter, and William D. Jones, J. Am. Chem. Soc. 2004, to be submitted.

  • 25

    Recent Special Recognitions Received by the PI:

    ACS Award in Organometallic Chemistry, 2003

    Associate Editor, J. Am. Chem. Soc., January 2003-

    Chair of Organometallic Subdivision, Inorganic Division of the American Chemical Society, 2001.

    Charles F. Houghton Professor of Chemistry, 2000-present

    July 2000- 2003: Chairman, Department of Chemistry

    Organometallic Gordon Conference, Chairman, Newport, RI, 2000.

    Opening speaker at Inorganic Reaction Mechanisms Gordon Conference, Ventura, 2003.

    Opening speaker at Organometallic Gordon Conference, Newport, 2003.

    Speaker at Inorganic Gordon Conference, Newport, 2003.

    Speaker at Isotope Effects Gordon Conference, Ventura, 2004.

    Additional Comments: Our renewal budget for 2001-2004 was less than our budget for the prior grant period. On

    the basis of our success and productivity, as delineated by the DOE request for the specific information included in this progress report, it is becoming increasingly difficult to accomplish such a large body of work with so little funding. It is imperative that the grant be increased so that 2.5 people can work on the projects. Our current stipend for a good student is $22K/year, and the current budget is supporting only about 80% of these costs. We usually obtain several undergraduate workers for 'free' (i.e., paid by the University), but supply and instrument usage costs still need to be paid.

    We apologize for the length of this report.


Recommended