Добавил:
Опубликованный материал нарушает ваши авторские права? Сообщите нам.
Вуз: Предмет: Файл:
Computational Chemistry - introduction to the theory and applications of molecular and quantum mechanics.pdf
Скачиваний:
309
Добавлен:
08.01.2014
Размер:
18.42 Mб
Скачать

122 Computational Chemistry

Diagonalization of the four-basis-function matrix of Eq. (4.65) gives

The energy levels and MOs from these results are shown in Fig. 4.17. Note that all these matrix diagonalizations yield orthonormal eigenvectors: and as required the fact that the Fock matrices are symmetric (see the discussion of matrix diagonalization in section 4.3.3).

4.3.5 The simple Hückel method – applications

Applications of the SHM are discussed in great detail in several books [21]; here we will deal only with those applications which are needed to appreciate the utility of the method and to smooth the way for the discussion of certain topics (like bond orders and atomic charges) in later chapters. We will discuss: the nodal properties of the MOs; stability as indicated by energy levels and aromaticity (the 4n + 2 rule); resonance energies; and bond orders and atomic charges.

The nodal properties of the MOs

A node of an MO is a plane at which, as we proceed along the sequence of basis functions, the sign of the wavefunction changes (Figs 4.15–4.17). For a given molecule,

the number of nodes in the

orbitals increases with the energy. In the two-orbital

system (Fig. 4.15),

has zero nodes and

has one node. In the three-orbital system

(Fig. 4.16),

and

have zero, one and two nodes, respectively. In the cyclic

four-orbital system (Fig. 4.17),

has zero nodes,

and

which are degenerate

(of the same energy) each have one node (one nodal plane), and

has two nodes.

In a given molecule, the energy of the MOs increases with the number of nodes. The nodal properties of the SHM orbitals form the basis of one of the simplest ways of

Introduction to Quantum Mechanics 123

understanding the predictions of the WoodwardHoffmann orbital symmetry rules [38]. For example, the thermal conrotatary and disrotatary ring closure/opening of polyenes can be rationalized very simply in terms of the symmetry of the highest occupied MO of the open-chain species. That the highest MO should dominate the course of this kind of reaction is indicated by more detailed considerations (including extended Hückel calculations) [38]. Figure 4.18 shows the situation for the ring closure of a 1,3-butadiene to a cyclobutene. The phase (+ or –) of the HOMO at the end carbons (the atoms that bond) is opposite on each face, because this orbital has one node in the middle of the chain. You can see this by sketching the MO as the four AOs contributing to it, or even – remembering the node – drawing just the end AOs. For the electrons in to bond, the end groups must rotate in the same sense (conrotation) to bring orbital lobes of the same phase together. Remember that plus and minus phase has nothing to do with electric charge, but is a consequence of the wave nature of electrons (section 4.2.6): two electron waves can reinforce one another and form a bonding pair if they are “vibrating in phase”; an out-of-phase interaction represents an antibonding situation. Rotation in opposite senses (disrotation) would bring opposite-phase lobes together, an antibonding situation. The mechanism of the reverse reaction is simply the forward mechanism in reverse, so the fact that the thermodynamically favored process is the ring-opening of a cyclobutene simply means that the cyclobutene shown would open to the butadiene shown on heating. Photochemical processes can also be

124 Computational Chemistry

accommodated by the WoodwardHoffmann orbital symmetry rules if we realize that absorption of a photon creates an electronically excited molecule in which the previous lowest unoccupied MO (LUMO) is now the HOMO. For more about orbital symmetry and chemical reactions see, e.g. the book by Woodward and Hoffmann [38].

Stability as indicated by energy levels and aromaticity

The MO energy levels obtained from an SHM calculation must be filled with electrons according to the species under consideration. For example, the neutral ethene molecule has two electrons, so the diagrams of Fig. 4.19(a) (cf. Fig. 4.15) with one, two and three electrons, would refer to the cation, the neutral and the anion. We might expect the neutral, with its bonding orbital full and its antibonding orbital empty, to be resistant to oxidation (which would require removing electronic charge from the low-energy and to reduction (which would require adding electronic charge to the

high-energy

The propenyl (allyl) system has two, three or four electrons, depending on whether we are considering the cation, radical or anion (Fig. 4.19(b); cf. Fig. 4.16). The cation

Introduction to Quantum Mechanics 125

might be expected to be resistant to oxidation, which requires removing an electron from a low-lying orbital and to be moderately readily reduced, as this involves adding an electron to the nonbondingorbital a process that should not be strongly favorable or unfavorable. The radical should be easier to oxidize than the cation, for this

126 Computational Chemistry

requires removing an electron from a nonbonding, rather than a lower-lying bonding, orbital, and the ease of reduction of the radical should be roughly comparable to that of the cation, as both can accommodate an electron in a nonbonding orbital. The anion should be oxidized with an ease comparable to that of the radical (removal of an

electron from the nonbonding

but be harder to reduce (addition of an electron to

the antibonding

 

The cyclobutadiene system

(Fig. 4.19(c); cf. Fig. 4.17) can be envisaged with,

amongst others, two (the dication), four (the neutral molecule) and six (the dianion) electrons. The predictions one might make for the behavior of these three species toward redox reactions are comparable to those just outlined for the propenyl cation, radical and anion, respectively (note the analogous occupancy of bonding, nonbonding and antibonding orbitals). The neutral cyclobutadiene molecule is, however, predicted by the SHM to have an unusual electronic arrangement for a diene:

in filling the

orbitals, from the lowest-energy one up, one puts electrons of the

same spin into the degenerate

and

in accordance with Hund’s rule of max-

imum multiplicity. Thus the SHM predicts that cyclobutadiene will be a diradical, with two unpaired electrons of like spin. Actually, more advanced calculations [39] indicate, and experiment confirms, that cyclobutadiene is a singlet molecule with two single and two double C/C bonds. A square cyclobutadiene diradical with four 1.5 C/C bonds would distort to a rectangular, closed-shell (i.e. no unpaired electrons) molecule with two single and two double bonds (Fig. 4.20). This could have been predicted by augmenting the SHM result with a knowledge of the phenomenon known as the Jahn–Teller effect [40]: cyclic systems (and certain others) with an odd number of electrons in degenerate (equal-energy) MOs will distort to remove the degeneracy.

Introduction to Quantum Mechanics 127

What general pattern of MOs emerges from the SHM? Acyclic systems (ethene, the propenyl system, 1,3-butadiene, etc.), have MOs distributed singly and evenly on each side of the nonbonding level; the odd-AO systems also have one nonbonding MO (Fig. 4.21). Cyclic systems (the cyclopropenyl system, cyclobutadiene, the cyclopentadienyl system, benzene, etc.) have a lowest MO and pairs of degenerate MOs, ending with one highest or a pair of highest MOs, depending on whether the number of MOs is even or odd. The total number of MOs is always equal to the number of basis functions, which in the SHM is, for organic polyenes, the number of p orbitals (Fig. 4.21). The pattern for cyclic systems can be predicted qualitatively simply by sketching the polygon, with one vertex down, inside a circle (Fig. 4.22). If the circle is of radius the energies can even be calculated by trigonometry [41 ]. It follows from this

128 Computational Chemistry

pattern that cyclic species (not necessarily neutral) with 2, 6, 10, electrons have filled MOs and might be expected to show particular stability, analogously to the filled AOs of the unreactive noble gases (Fig. 4.23). The archetype of such molecules is, of course, benzene, and the stability is associated with the general collection of properties called aromaticity [17]. These results, which were first perceived by Hückel [ 19] (19311937), are summarized in a rule called the 4n + 2 rule or Hückel’s rule, although the 4n + 2 formulation was evidently actually due to Doering and Knox (1954) [42]. This says that cyclic arrays of-hybridized atoms with electrons are characteristic of aromatic molecules; the canonical aromatic molecule benzene with six electrons corresponds to n = 1. For neutral molecules with formally fully conjugated perimeters this amounts to saying that those with an odd number of C/C double bonds are aromatic and those with an even number are antiaromatic (see resonance energies, below).

Hückel’s rule has been abundantly verified [17] notwithstanding the fact that the SHM, when applied without regard to considerations like the JahnTeller effect (see above) incorrectly predicts 4n species like cyclobutadiene to be triplet diradicals. The Hückel rule also applies to ions; for example, the cyclopropenyl with system two electrons, the cyclopropenyl cation, corresponds to n = 0, and is strongly aromatic.

Introduction to Quantum Mechanics 129

Other aromatic species are the cyclopentadienyl anion (six electrons, Hückel predicted the enhanced acidity of cyclopentadiene) and the cycloheptatrienyl cation. Only reasonably planar species can be expected to provide the AO overlap need for cyclic electron delocalization and aromaticity, and care is needed in applying the rule.

Resonance energies

The SHM permits the calculation of a kind of stabilizing energy, or, more accurately, an energy that reflects the stability of molecules. This energy is calculated by comparing the total electronic energy of the molecule in question with that of a reference compound, as shown below for the propenyl systems, cyclobutadiene, and the cyclobutadiene dication.

The propenyl cation, Fig. 4.19(b); cf. Fig. 4.16. If we take the total electronic energy of a molecule to be simply the number of electrons in a MO times the energy level of the orbital, summed over occupied orbitals (a gross approximation, as it ignores interelectronic repulsion), then for the propenyl cation

We want to compare this energy with that of two electrons in a normal molecule with no special features (the propenyl cation has the special feature of an empty p orbital adjacent to the formal C/C double bond), and we choose neutral ethene for our reference energy (Fig. 4.15)

The stabilization energy is then

Since is negative, the electronic energy of the propenyl cation is calculated to be below that of ethene: providing an extra, empty p orbital for the electron pair causes the energy to drop. Actually, resonance energy is usually presented as a positive quantity, e.g. We can interpret this as below a reference system. To avoid a negative quantity in SHM calculations like these, we can use instead of

The propenyl radical, Fig. 4.16. The total electronic energy by the SHM is

For the reference energy we use one ethene molecule and one nonbonding p electron (like the electron in a methyl radical):

The stabilization energy is then

130 Computational Chemistry

The propenyl anion. An analogous calculation (cf. Fig. 4.16, with four electrons for the anion) gives

Thus the SHM predicts that all three propenyl species will be lower in energy than if the electrons were localized in the formal double bond and (for the radical and anion) in one p orbital. Because this lower energy is associated with the ability of the electrons to spread or be delocalized over the whole system, what we have called E(stab) is often denoted as the delocalization energy, and designated Note that (or is always some multiple of (or is zero). Since electron delocalization can be indicated by the familiar resonance symbolism the Hückel delocalization energy is often equated with resonance energy, and designated The accord between calculated delocalization and the ability to draw resonance structures is not perfect, as indicated by the next example.

Cyclobutadiene (Fig. 4.17). The total electronic energy is

Using two ethene molecules as our reference system:

and so for E(stab) we get

Cyclobutadiene is predicted by this calculation to have no resonance energy, although we can readily draw two “resonance structures” exactly analogous to the Kekulé struc-

tures of benzene. The SHM predicts a resonance energy of

for benzene. Equating

with the commonly-quoted resonance energy of

for

benzene gives a value of

for

but this should be taken with more than

a grain of salt, for outside a closely related series of molecules,

has little or no quan-

titative meaning [43]. However,

in contrast to the failure of simple resonance theory

in predicting aromatic stabilization (and other chemical phenomena) [44], the SHM is quite successful.

The cyclobutadiene dication (cf. Fig. 4.17). The total electronic energy is

Using one ethene molecule as the reference:

and so

Introduction to Quantum Mechanics 131

Thus the stabilization energy calculation agrees with the deduction from the disposition of filled MOs (i.e. with the 4n + 2 rule) that the cyclobutadiene dication should be stabilized by electron delocalization, which is in some agreement with experiment [45].

More sophisticated calculations indicate that cyclic 4n systems like cyclobutadiene (where planar; cyclooctatetraene, for example, is buckled by steric factors and is simply an ordinary polyene) are actually destabilized by electronic effects: their resonance energy is not just zero, as predicted by the SHM, but less than zero. Such systems are antiaromatic [17,46].

Bond orders

The meaning of this term is easy to grasp in a qualitative, intuitive way: an ideal single bond has a bond order of one, and ideal double and triple bonds have bond orders of two and three, respectively. Invoking Lewis electron-dot structures, one might say that the order of a bond is the number of electron pairs being shared between the two bonded atoms. Calculated quantum mechanical bond orders should be more widely applicable than those from the Lewis picture, because electron pairs are not localized between atoms in a clean pairwise manner; thus a weak bond, like a hydrogen bond or a long single bond, might be expected to have a bond order of less than one. However, there is no unique definition of bond order in computational chemistry, because there seems to be no single, correct method to assign electrons to particular atoms or pairs of atoms [47]. Various quantum mechanical definitions of bond order can be devised [48], based on basis-set coefficients. Intuitively, these coefficients for a pair of atoms should be relevant to calculating a bond order, since the bigger the contribution two atoms make to the wavefunction (whose square is a measure of the electron density; section 4.2.6), the bigger should be the electron density between them. In the SHM the order of a bond between two atoms and is defined as

Here the 1 denotes the single bond of the ubiquitous spectator bond framework, which is taken as always contributing a bond order of unity. The other term is the bond order; its value is obtained by summing over all the occupied MOs the number of electrons n in each of these MOs times the product of the c’s of the two atoms for each MO. This is illustrated in the following examples.

Ethene. The occupied orbital is which has 2 electrons), and the coefficients of and for this orbital are 0.707, 0.707 (Eq. (4.66)). Thus

which is reasonable for a double bond. The order of the

bond is 1 and that of the

bond is 1.

 

The ethene radical anion. The occupied orbitals are

which has 2 electrons, and

which has 1 electron; the coefficients of and for

are 0.707, 0.707 and for

132 Computational Chemistry

0.707, 0.707 (Eq. (4.66)). Thus

The bond order of 0.500 (1,500 – bond order) accords with two electrons in the bonding MO and one electron in the antibonding orbital.

Atomic charges

In an intuitive way, the charge on an atom might be thought to be a measure of the extent to which the atom repels or attracts a charged probe near it, and to be measurable from the energy it takes to bring a probe charge from infinity up to near the atom. However, this would tell us the charge at a point outside the atom, for example a point on the van der Waals surface of the molecule, and the repulsive or attractive forces on the probe charge would be due to the molecule as a whole. Although atomic charges are generally considered to be experimentally unmeasurable, chemists find the concept very useful (thus calculated charges are used to parameterize molecular mechanics force fields chapter 3), and much effort has gone into designing various definitions of atomic charge [47,48]. Intuitively, the charge on an atom should be related to the basis set coefficients of the atom, since the more the atom contributes to a multicenter wavefunction (one with contributions from basis functions on several atoms), the more it might be expected to lose electronic charge by delocalization into the rest of the molecule (cf. the discussion of bond order above). In the SHM the charge on an atom

is defined as (cf. Eq. (4.70))

The summation term is the charge density, and is a measure of the electronic charge on the molecule due to the electrons. For example, having no electrons (an empty p orbital, formally a cationic carbon) would mean a electron charge density of zero; subtracting this from unity gives a charge on the atom of +1. Again, having two electrons in a p orbital would mean a electron charge density of 2 on the atom; subtracting this from unity gives a charge on the atom of 1 (a filled p orbital, formally an anionic carbon). The application of Eq. (4.71) will be illustrated using methylenecyclopropene (Fig. 4.24).

Methylenecyclopropene (Fig. 4.24).

Соседние файлы в предмете Химия