CH2422 MCE

download CH2422 MCE

of 15

Transcript of CH2422 MCE

  • 8/3/2019 CH2422 MCE

    1/15

    CH2422 Quantum Chemistry; Frontier Orbital Theory

    Frontier orbital theory is necessary as it accounts for a better description of reactivity than the vague

    application of curly arrows and often ill -informed electronic arguments, the deciding factor in

    whether a reaction occurs or not is often one of symmetry. For example, there are Diels-Alder

    reactions that appear to involve the reaction of two positive nodes that are explainable with Frontier

    Orbital theory but would otherwise seem unlikely.

    CH2

    OH O

    CH2

    OH

    O

    O OH

    OH

    O

    CH2+

    O-

    OH

    CH2+

    O-

    OH

    Figure 1 - An example Diels-Alder reaction that is allowed

    by frontier orbital theory yet would seem counter-intuitive

    Frontier orbital theory is also useful for explaining a preference of one reaction over another, for

    example the combination of cyclopentadiene rings in solution.

    Figure 2 - The dimerisation of cyclopentadiene showing a

    preference for one reaction, despite both being possible

    Frontier Orbitals of Commonly Used Molecular Fragments

    Determining the Molecular Orbitals of Conjugated Hydrocarbons

    For conjugated systems it is the electrons that are of importance. The movement of electrons

    within these systems can be considered analogous to that of the electron in a box and using this

    model it is possible to calculate the relative signs and orbital coefficients, c, of the p-orbitals at each

    carbon of the conjugated chain.

  • 8/3/2019 CH2422 MCE

    2/15

    Figure 3 - Wavefunctions of conjugated systems

    represented as sin waves for molecular orbitals, n=1(--),2(--),3(--),4(--)

    Each orbital can contain up to 2 electrons, so for an 8 system n=1 to n=3 would be fully occupied,

    with n=4 unoccupied, in this case the highest occupied molecular orbital (the HOMO) is n=3, and the

    lowest unoccupied molecular orbital (the LUMO) is n=4. These are the frontier orbitals, and it is the

    symmetry allowed overlap of a molecular LUMO with a molecular HOMO that dictates whether a

    reaction is forbidden or not.

    To calculate the orbital signs and coefficients across the molecule the appropriate wave-form can besplit into natoms+1, the p-orbitals are found at each split, the respective sign being that of the wave,

    the coefficient determined by the magnitude of the peak or trough at that point.

    The sum of the coefficients squared must be normalised, i.e. c2=1.

    Ethene

    For the simplest example, a 2 electron molecule, it is necessary to consider the first two waves of

    an electron in a box.

    Figure 4 - The wavefunctions used for describing

    the frontier orbitals in ethene

    The 2 electrons occupy the first wavefunction, both have the same sign and magnitude (sin 60 and

    sin 120 are equal), since the magnitudes must be normalised c2

    +c2

    =1 and so c=0.707. This is theHOMO of ethene.

    0 180

    0 360

    0 540

    0 720

    0 180

    0 360

  • 8/3/2019 CH2422 MCE

    3/15

    Since the first waveform is occupied, the LUMO is described by the second function. Here the

    magnitudes are the same, however the signs are opposing. This represents a node in the centre of

    the ethene molecule.

    Figure 5 - The frontier orbitals of ethene

    Butadiene

    For butadiene there are 4 atoms, and so the wavefunctions are split into 5 parts. There are 4

    electrons and so the first two waves are occupied (the second therefore being the HOMO), the third

    is empty and so is the LUMO.

    Figure 7 - The frontier orbitals of butadiene

    0 180

    0 360

    0 540

    Figure 6 - The wavefunctions used for

    describing the frontier orbitals in butadiene

  • 8/3/2019 CH2422 MCE

    4/15

    Allyl Cations and Anions

    Allyl cations and anions are 2 and 4 systems respectively.

    Figure 9 - The frontier orbitals of

    the allyl cation and anion

    0 180

    0 360

    0 540

    Figure 8 - The wavefunctions used for

    describing the frontier orbitals in allyl

    cations and anions

  • 8/3/2019 CH2422 MCE

    5/15

    Pentadienyl Cations and Anions

    Pentadienyl cations and anions are 4 and 6 respectively.

    Figure 11 - The frontier orbitals of the

    pentadienyl cation and anion

    0 180

    0 360

    0 540

    0 720

    Figure 10 - The wavefunctions used for

    describing the frontier orbitals in pentadienyl

    cations and anions

  • 8/3/2019 CH2422 MCE

    6/15

    Hextriene

    Hexatriene is a 6 system.

    Figure 12 - The wavefunctions used for describing

    the frontier orbitals in hexatriene

    Figure 13 - The frontier orbitals of hexatriene

    Using Frontier Orbital Theory to Predict the Product of a Diels-Alder Reaction

    Consider the reaction of a 1-methoxy butadiene with furan-2,5-dione. There are 3 stereogenic

    centres formed, with both the endo and exo product theoretically possible.

    0 180

    0 360

    0 540

    0 720

  • 8/3/2019 CH2422 MCE

    7/15

    The trajectory of this reaction must be considered, with the diene approaching from out of the plane

    there are two possible conformations for attack. The exo product involves less stearic interaction

    between the furan and the diene, however the endo product is preferred. This is due to secondary

    orbital interactions, if the furan is considered to be electronically analogous to hexatriene (an

    assumption that in this case works) then the atoms either side of the reacting double bond possess

    orbitals of the same sign as those in the middle of butadiene, this interaction is energetically

    favourable and so, despite stearic effects, the endo product is favoured.

    Diels-Alder Reactions - Normal And Inverse Electron Demand

    Typically in a Diels-Alder the reaction will occur between the HOMO of the diene and the LUMO of

    the dienophile, this is said to follow normal electron demand. There are cases in which the opposite

    is true, in which the LUMO of the diene interacts with the HOMO of the dienophile, these are

    considered to follow inverse electron demand. Inverse electron demand Diels-Alder reactions,

    DAINV, occur between electron-rich dienophiles and electron-poor dienes, an example of which is the

    reaction between acrolein and methyl vinyl ether (Figure 14). DAINV reactions often involve

    heteroatoms and are a useful route to heterocyclic molecules. The mechanism of a DAINV reaction is

    believed to be slightly asynchronous, but otherwise the same as that for a typical normal electrondemand reaction. The time-delay between scission and formation of bonds is believed to be

  • 8/3/2019 CH2422 MCE

    8/15

    minimal, however, there is still dispute within the literature as to how significant a difference there

    is.

    Figure 14 - The reaction between acrolein and methyl vinyl ether

  • 8/3/2019 CH2422 MCE

    9/15

    Fleming Notation and the Orbitals Involved in Diels-Alder Reactions

    Fleming describes three types of substituent that can form part of a diene or dienophile for use in a

    Diels-Alder reaction. These are allyl anions, ally cations and conjugated carbon systems; X, Z and C

    respectively.

    For dienophiles, considering only the LUMO;

  • 8/3/2019 CH2422 MCE

    10/15

    For 1-substituted dienes, considering only the HOMO;

  • 8/3/2019 CH2422 MCE

    11/15

    For 2-substituted dienes;

    Electrocyclic Reactions

    Consider the formation of a 6-membered cycle from a substituted hexatriene molecule (Figure 15).

    Figure 15 - The electrocyclic reaction of 1,6-dimethyl hexatriene

    Considering the HOMO of this molecule (assuming there is no activation of the hexatriene derivative

    first) it is possible to determine the stereochemistry of the product by considering the directions in

    which the carbons must rotate to achieve successful orbital overlap. Figure 16 shows that in order to

    overlap the two carbon atoms must rotate in opposing direction, this reaction is therefore

    disrotatory.

  • 8/3/2019 CH2422 MCE

    12/15

    Figure 16 - The HOMO of 1,6-dimethyl hexatriene and the resulting cis product

    When considering the photochemical reaction, in which an electron is first promoted by excitement

    with light, it is the LUMO that must be considered, as seen in Figure 17. Here the reaction is

    conrotatory and leads to the antiproduct.

    Figure 17 - The LUMO of 1,6-dimethyl hexatriene and the resulting anti product

    In some cases the distinction between thermal and photochemical electrocyclic reactions may

    dictate the reaction scheme chosen. For example, the formation of a 5-membered heterocycle

    shown below (Figure 18) can be controlled to give a cis or trans product based on the first reaction

    step. By deprotonating the cis product will form as the symmetry of the resulting HOMO encourages

    disrotatory motion of the bond-forming carbon atoms. If the trans product is desired it could be

    proposed that the anionic intermediate is somehow isolated and photochemically excited, as the

    LUMO would have a conrotatory bond forming step. This is not possible, however, and so a different

    approach must be adopted, the formation of a cation in the first step alters the HOMO, allowing

    conrotatory bond formation and the resulting trans product.

  • 8/3/2019 CH2422 MCE

    13/15

    Figure 18 - The formation of two different products through exploitation of frontier orbital theory

    Sigmatropic Reactions

    A sigmatropic reaction is one in which a -bond is broken. The simplest involves the movement of a

    hydrogen from one carbon to another in a conjugated system. A 1,5 and a 1,7 sigmatropic reaction

    are shown below (Figure 19), by considering the HOMO of the -bond with the LUMO of the

    conjugated system it is possible to rationalise the face to which the hydrogen will move. Thesereactions are then categorised according to the movement of the hydrogen, with reactions in which

    the hydrogen is retained on the same face described as suprafacial and those in which it switches

    faces are described as antarofacial.

  • 8/3/2019 CH2422 MCE

    14/15

    Figure 19 - 1,5 and 1,7 sigmatropic reactions

    The Cope rearrangement (Figure 20) is a 3,3-sigmatropic reaction in which bonds are broken and

    formed between carbon atoms.

    Figure 20 - The Cope rearrangement

    The Cope rearrangement is reversible, and so controlling the outcome of a reaction in which it is a

    necessary step requires intelligent reaction design. The most common method is to make the

    product as stable as possible relative to a starting material or intermediate. In Figure 21 two

    methods of achieving this are shown, the first being the tautomerisation of an alcohol to a much

    more stable product, the second being the incorporation of a cyclopropane moiety in the starting

    material, the resulting ring strain discourages the return rearrangement.

  • 8/3/2019 CH2422 MCE

    15/15

    Figure 21 - Controlling the Cope rearrangement