Correlation diagram

A correlation diagram is a graphical representation that shows how the energies of the various states of a free -metal ion change when placed in an octahedral ligand field of varying strength. It is useful for studying  transitions, which are electronic transitions that occur between molecular orbitals (MOs) of -metal octahedral complexes that are mostly metal in character.

Consider a transition metal ion, such as , with a electronic configuration. Its correlation diagram (see diagram above) exhibits the following features:

    1. The free metal ion states, represented by atomic term symbols without spin-orbit coupling, are tabulated on the extreme left.
    2. To the right of the term symbols, we assume that the free metal ion is subject to a weak octahedral ligand field, which is just strong enough to lift the degeneracy of microstates associated with each term symbol. The symmetries of the resulting states of distinct energy are denoted by irreducible representations of the point group.
    3. On the extreme right, we list the irreducible representations corresponding to the electronic configurations of the metal ion in a strong octahedral ligand field.
    4. Additionally, lines connect states of the same symmetry.

The atomic terms () of such a free metal ion are constructed using the Russell Saunders coupling techniques. Each term corresponds to a specific energy level of the electronic configuration. Within each term, microstates with the same energy are grouped together. Overall, there are 45 microstates distributed across these five terms. The wavefunction for each microstate is the product of two one-electron -orbital wavefunctions, , where , and its corresponding energy comes from the solution of . Hund’s rules predict that the terms are, in order of increasing energy  and . However, spectroscopic measurements have shown that the actual order is  and .

In the presence of a weak octahedral ligand field, the formation of the MOs of the complex is assumed to involve a small perturbation on the energies corresponding to the atomic terms, such that the symmetries of the MOs are characteristic of an octahedral environment. The energies of these MOs are eigenvalues of the Hamiltonian , where is the unperturbed Hamiltonian and is the perturbed Hamiltonian due to the weak ligand field. The method to determine the symmetries of microstates involves examining how a one-electron wavefunction of the metal ion behaves under symmetry operations and then extending the results to .

A one-electron hydrogenic total wavefunction has the formula , where the spatial function . Here, , , and represent the radial, angular (spherical harmonics) and spin functions, respectively. Any symmetry transformation of affects only the angular part of the total wavefunction because the radial and spin functions remain invariant under all symmetry operations of a point group. This implies that the character table of a regular point group is generated solely via spherical harmonics.

 

Question

Why are the radial and spin functions invariant under all symmetry operations of a point group?

Answer

The radial function is a function of the distance of an electron from the origin. Symmetry operations do not affect . The spin function describes an intrinsic form of angular momentum of electrons known as spin, which is independent of the spatial coordinate system. However, it is possible to generate character tables that depend on both spatial and spin coordinates. Such point groups are known as double groups. The main difference between a regular point group and a double group lies in the basis functions used to generate them: spherical harmonics for regular point groups and total wavefunctions for double groups. In other words, representations of double groups account for the combined symmetry properties of the full wavefunction, not just the spatial part. The character table of a double group can be perceived as an expansion of the character table of a regular group, with additional representations and classes. We will explore how certain properties of double groups are useful in constructing the correlation diagram of the electronic configuration.

 

Let’s analyse the symmetry operations of the point group on . When the rotation operator acts on by an angle around the -axis, only the azimuthal angle is affected and each basis is transformed into . The transformation for all basis functions is summarised as:

The character of the rotation matrix is , which is a geometric series , where , and . This implies that

Using Euler’s formula of ,

When , . Applying L’Hopital’s rule,

Since the basis functions for the electronic configuration are -orbitals, which are even with respect to inversion,

 

Question

What is equivalent to?

Answer

Let . Applying the chain rule, . So, and .

 

An improper rotation combines two symmetry operations: rotation and reflection. A reflection of the angular function about the plane perpendicular to the rotation axis affects only the polar angle and transforms it to . Noting that the explicit expression of the un-normalised spherical harmonics is , we have . As for each basis function,

Therefore,

and

It follows that when ,

Although, the angular momentum quantum number in eq111 through eq115 is associated with a single electron, it can be replaced by the total angular momentum quantum number in those equations for two reasons. Firstly, the number of degenerate states for both and is the same (degenerate states for and  correspond to distinct values of and , respectively). Secondly, a two-electron basis function for an atomic term (e.g. a term) is given by  , where , and . is analogous to the one-electron spherical harmonics .

Consequently, the characters for the reducible representation corresponding to term , where , are

7 1 -1 -1 -1 7 -1 1 -1 -1

can then be decomposed using eq27a into . The three irreducible representations have the same spin multiplicity as because the perturbed Hamiltonian only causes a small change in the microstates’ energies and does not affect their common spin multiplicity. Using the same logic, the reducible representations of  and decompose to the corresponding terms shown under the weak field column in the correlation diagram above.

 

Question

For , the constituent irreducible representations have the same spin multiplicity as . What happens if the constituent representations of a reducible representation corresponding to an electronic configuration are associated with both singlet and triplet microstates?

Answer

As mentioned earlier, is derived using spherical harmonics. When considering the total wavefunction of a microstate, which includes the spin wavefunction, we must search for clues to assign the correct spin multiplicities to the constituent irreducible representations of . Fortunately, for the weak field case, the atomic terms serve as a reference for assignment. When the free metal ion is subject to a strong ligand field, there are scenarios where the constituent irreducible representations of a reducible representation, for instance , can correspond to both singlet and triplet microstates (see below for details). In such cases, we form a new representation using total wavefunctions as basis functions. The resultant reducible representation, which belongs to a double group, is enlarged: .

 

In the presence of a strong octahedral ligand field, the field’s contribution to the energy of an MO now dominates over the energy due interelectronic repulsion in the ion. The ligand field splitting parameter is large and the only possible configurations are , and . The wavefunction of each of the configurations is assumed to be the product of two one-electron -orbital wavefunctions. This implies that the symmetries of the three configurations can be derived using the concept of direct product representations. With reference to the character table of the point group, we have

9 0 1 1 1 9 1 0 1 1
6 0 0 0 -2 6 0 0 -2 0
4 1 0 0 4 4 0 1 4 0

Even though the three representations are direct product representations, their characters are fundamentally linked to eq111 through eq115, which are functions of the total angular momentum of a state. Since the number of microstates are the same under a weak or strong field, both angular and spin momenta of the microstates are preserved. In other words, we would expect the three reducible representations to decompose into irreducible representations that correspond one-to-one with the eleven states in the weak field scenario.

Using eq27a, the reducible representations in the above table decompose to

 

Question

What is the difference between orbital symmetries and configuration symmetries?

Answer

The degenerate and orbitals are one-electron wavefunctions that transform together according to the irreducible representation of the point group. For the case of the configuration, there are six possible ways to fill these orbitals with two electrons (see diagram below). Each of these six microstates is described by a two-electron wavefunction that is a product of two one-electron -orbital wavefunctions. These wavefunctions transform according to a reducible representation , which can be decomposed into the symmetries  and .

 

The next step is to determine the spin multiplicities associated with the irreducible representations of each configuration. One approach is to use the descent of symmetry method, as spin multiplicities are intrinsic properties of electrons that remain independent of the system’s symmetry. For this method to be effective, the chosen subgroup must facilitate the mapping of irreducible representations from the parent group to different representations within the subgroup. One subgroup that satisfies this mapping criterion for is the point group.

 

Question

What does of the point group correlate to in the point group? Show that of the point group is equivalent to .

Answer

Using the correlation table above, of the point group correlates to of the point group. Let and be the characters of and , respectively. is equivalent to as they are related by a similarity transformation. So,

Since and , . Consequently, the relationship between and can be described by the diagram below.

 

The reducible representations  of of and of share the same set of basis functions. Even though these functions are spherical harmonics, their total wavefunctions must be associated with certain spin states. For instance, the basis functions of the block of  describe either a singlet or triplet spin state, as they are a degenerate pair. The pair also transforms according to the block of . To proceed with the determination of the spin multiplicity of wavefunctions that transform according to , we need to analyse the microstates in the environment.

 

Microstate analysis

According to the character table, the and  orbitals are degenerate and belong to . The perturbation field is different from the field, resulting in the degeneracy of these orbitals being lifted in the environment ( and transform according to and , respectively). Nevertheless, the number of microstates remain at six (see above diagram). Any two-electron wavefunctions that transform according to or  must be singlets to satisfy Pauli’s exclusion principle, while those that transform according to can be expressed either as singlet or triplet states.

 

Configuration

Direct product

Spin state

Symmetry of microstate

Singlet
Singlet, triplet
Singlet

Since the wavefunction that transforms according to of the block represents a singlet, the degenerate pair of wavefunctions corresponding to the block of must also describe a singlet state. This implies that the wavefunction that transforms according to of the block also represents a singlet. As there are two triplet microstates out of the six microstates in the environment, the remaining representation must be associated with triplet wavefunctions. Consequently, the states that arise from the configuration in the environment are grouped into  and , which can be expressed as the double group representation .

 

Question

How do we verify that is correct?

Answer

The and irreducible representations are associated with one and two linearly independent basis functions, respectively. A singlet spin state and a triplet spin state are associated with one and three spin wavefunctions, respectively. Therefore, we can form microstates for .

 

 

Repeating the above logic, the subgroup that satisfies the criterion for mapping of is the point group. The fifteen microstates of transform according to the reducible representation , which correlates to of .

 

Microstate analysis

The degeneracy of the and orbitals in of  are lifted in , with belonging to and transforming according to . We can use six configurations to describe the fifteen microstates (see table below, where the symmetries of and are denoted by and , respectively.

 

Configuration

Direct product

Spin state

Symmetry of microstate

Singlet
Singlet, triplet
Singlet, triplet
Singlet
Singlet, triplet
Singlet

Basis functions that transform according to one of the blocks of must describe either singlet or triplet states. Since there are three triplet microstate symmetries in the environment, one of which is , the only possibility is for the wavefunction that transforms according to the block to represent a triplet, while the wavefunctions belonging to the remaining blocks describe singlets. In other words, the states that arise from the configuration in the environment are grouped into  and . Similarly, we can verify that we have microstates for .

The remaining configuration is associated with twenty-four microstates. As the two -electrons are in different orbitals, their spins may be either paired or unpaired. Without using the descent of symmetry method, it is obvious that the only way to assign the twenty-four microstates is to form the double group representation .

Finally, we can complete the correlation diagram by connecting the states between the weak and strong fields using the non-crossing rule.

 

Previous article: Subgroup (descent of symmetry)
Content page of group theory
Content page of advanced chemistry
Main content page

Subgroup (descent of symmetry)

A subgroup is a collection of elements within a larger group that forms its own distinct group. For example, is a subgroup of (see multiplication table below).

The irreducible representations of the subgroup and its parent group, generated from the same set of basis functions, exhibit identical characters associated with symmetry operations that are common to both groups (see character tables below). This occurs because

A basis function undergoing a specific symmetry operation is associated with the same transformation matrix and, consequently, the same trace in both the subgroup and its parent group.

We say that the representations of a subgroup and its parent group are correlated. For example, and of the point group are correlated to and of the point group, respectively. The correlation of representations of two point groups can be used to determine the nature of certain properties of a basis function. This is especially relevant when the properties, for instance spin, are independent of the spatial coordinate system and therefore invariant under any symmetry operations. Further explanation can be found in the next article.

Let’s consider a more complicated case: and its subgroup . When comparing the character tables (see below) of both groups, it becomes evident that the symmetry operations and of are not preserved in . However, apart from the identity and inversion symmetry operations, the remaining symmetry operations of are partially preserved (e.g. five of the fifteen symmetry operations). To simplify the correlation of irreducible representations between the two groups, we focus on the symmetry operations that are unambiguously preserved. In the descent of symmetry from to , the preserved operations include , , , , and . Consequently, , , and of are correlated to , , and of , respectively.

of does not appear to correlate with any single irreducible representation of . However, there must exist a basis function that transforms according to both of and a representation of . The logic behind this begins with the fact that if a chemical species is invariant under the symmetry operations of a group , it is also invariant under the symmetry operations of a subgroup of . Therefore, a basis function that transforms according to an irreducible representation of must also transform to a representation of the subgroup of . Since a basis function associated with of  does not transform according to any irreducible representation of , it must transform according to a reducible representation of .

As mentioned above, a basis function undergoing a specific symmetry operation is associated with the same transformation matrix and, consequently, the same trace in both the subgroup and its parent group. This implies that of  correlates to the  reducible representation that is a direct sum of the irreducible representations . Similarly, we have

 

Question

How does the reducible representation of correlate to ?

Answer

The corresponding transformation matrix of the representation of must be a direct sum of matrices of .

 

In summary, a descent in symmetry occurs when a subgroup inherits symmetries from its parent group, preserving certain properties while reducing the overall complexity. It is useful in constructing weak-strong ligand field correlation diagrams.

 

Next article: Correlation diagram
Previous article: ligand field theory
Content page of group theory
Content page of advanced chemistry
Main content page

Non-crossing rule

The non-crossing rule in quantum chemistry states that distinct eigenvalues of two eigenstates with the same symmetry cannot become equal as a function of an external perturbation parameter.

Consider a system with two orthogonal eigenstates and of the unperturbed Hamiltonian , with distinct eigenvalues of and , respectively. In the presence of a weak ligand field, the unperturbed Schrodinger equation , becomes

where , is the perturbed Hamiltonian, is the eigenvalue corresponding to the perturbed eigenstate , which is assumed to be a linear combination of the unperturbed eigenstates.

Substituting in eq281j,

Multiplying eq281k from the left by and in turn, and integrating over all space, we have the following two equations:

where .

In matrix form, the pair of equations can be expressed as

where is the identity matrix.

Eq281l is a linear homogeneous equation, which has non-trivial solutions if

Expanding the determinant,

whose roots are

The unperturbed Hamiltonian is invariant to all symmetry operations of a point group and transforms according to the fully symmetric representation of the group. In the context of a metal ion subject to a ligand field, the perturbed Hamiltonian is a summation of potentials between metal valence electrons and ligand valence electrons:

is also totally symmetric, as the distances between metal and ligand valence electrons are invariant to all symmetry operations of a point group. If we assume that both and have a complete set of orthogonal eigenfunctions that are associated with real eigenvalues, and  are Hermitian. This implies that  is also Hermitian. Using eq35, eq281n becomes

The term  is positive and varies with the ligand field strength. If and  have the same symmetry, then is generally non-zero according to the vanishing integral principle. It follows that and are always distinct regardless of ligand field strength. This is known as the non-crossing rule.

On the other hand, if and  have different symmetry, then is necessarily zero. Consequently, the eigenvalues of the two states may be the same at a certain ligand field strength where .

For example, the diagram below depicts some states of the configuration of a metal ion in an octahedral ligand field. The  states, which have the same symmetry, do not cross as the ligand field strength varies, while and , which have different symmetries, do.

 

Next article: Relativistic one-electron Hamiltonian
Previous article: Time-dependent Perturbation theory
Content page of quantum mechanics
Content page of advanced chemistry
Main content page

Ligand field theory

Ligand field theory describes the electronic and magnetic properties of coordination complexes resulting from the interaction of peripheral donor atoms with -orbitals of the central metal atom.

Molecular orbital (MO) diagrams of -metal complexes (see above diagram for an example, and read the previous article on how it is constructed) provide the theoretical foundation for the ligand field theory. In the context of an octahedral complex, the energy separation between  and , known as the ligand field splitting parameter , is a crucial aspect of this theory. This energy difference is responsible for distinguishing the electronic and magnetic properties among different complexes. The magnitude of for a complex with a specific metal is dependent on the type of ligands. Ligands (e.g. and ) that interact strongly with the metal orbitals are called strong-field ligands, while ligands with weak interactions (e.g. and )  are known as weak-field ligands.

 

Question

Why does the interaction of a strong-field ligand with the metal lead to a larger ?

Answer

A ligand like is a -acceptor ligand. It has low-lying vacant orbitals that can overlap with the orbitals of the metal ion. This -back-donation from metal orbitals to ligand orbitals lowers the energy of the MOs and therefore increases the energy separation between and . The stronger the -acceptor ability of the ligand, the greater the splitting of the orbitals.

 

Consider the following two cases:

    1. Strong-field: is significantly large, making it energetically advantageous to fill the orbitals before the orbitals.
    2. Weak-field: is small, which makes it energetically preferable to fill the  orbitals before the orbitals are entirely occupied.

The diagram below shows the ground state configuration of -complexes, where , for both the strong-field and weak-field cases. For instance, it is energetically favourable for a -complex to adopt the  configuration if is large. When is small, the more stable configuration is , where the electrons are in separate orbitals with parallel spins. Complexes with configurations of 3 or more unpaired spins are classified as high-spin complexes, while configurations of less than 3 unpaired spins are known as low-spin complexes.

The above MO diagram is useful for studying charge transfer transitions of an octahedral complex. Charge transfer transitions occur between MOs that are mostly metal in character and those that are mostly ligand in character. These transitions depend on the type of ligand: they occur only when the metal is bound to ligands that are -donors or -acceptors. For instance, the transition from to is known as a ligand to metal charge transfer (LMCT). If we’re interested in studying  transitions, which are electronic transitions that occur between MOs that are mostly metal in character, a correlation diagram comes in handy.

 

Next article: Subgroup (descent in symmetry)
Previous article: Group theory and its applications in inorganic chemistry
Content page of group theory
Content page of advanced chemistry
Main content page

Group theory and its applications in inorganic chemistry

Group theory plays a pivotal role in understanding molecular symmetry and electronic properties in inorganic chemistry, particularly when applied to transition metal compounds.

Sigma interactions

Consider an octahedral molecule (see diagram above), which belongs to the point group. Let’s assume that valence atomic orbitals (AO) of the transition metal participate in bonding with the valence orbitals of the ligands , specifically -orbitals with symmetry. The wavefunctions of these ligand orbitals shall be denoted by , where . To generate a molecular orbital (MO) diagram for the complex, we need to perform the following steps:

    1. Find the symmetry of the AOs of .
    2. Determine the symmetry of the symmetry-adapted linear combinations (SALC) of .
    3. Work out which AOs of have non-zero vanishing integrals with valence orbitals of .

For step 1, the central atom lies on all the planes and axes of symmetry of the point group and is invariant to all symmetry operations of the point group. Therefore, the valence AOs of must transform according to respective bases of the character table of the point group.

Valence AOs Irreducible representations of

To accomplish step 2, we need to

    1. Select a basis set to generate a reducible representation of the  point group.
    2. Calculate the traces of the matrices of .
    3. Decompose into irreducible representations of the point group.
    4. Generate a set of orthogonal SALCs.

We shall employ the -vectors in the diagram as a basis set. The direction of each vector indicated in the diagram is regarded as the positive lobe of the ligand -orbital. Instead of carrying out the laborious task of letting act on the basis set to produce matrices for the reducible representation, we can determine the traces by inspection. This is because each -vector is transformed by into another -vector but not into a linear combination of -vectors. Furthermore, a diagonal element of a matrix of is equal to 1 when leaves a -vector invariant. For example, a operation along the -axis leaves and invariant, giving a trace of 2. The result is

6 0 0 2 2 0 0 0 4 2

Using eq27a, is decomposes to .

This implies that the matrices of are block-diagonal matrices of the same form. Each block-diagonal matrix is composed of the direct sum of the three irreducible representations , and of the point group. Since the number of basis functions of an irreducible representation corresponds to the dimension of the representation, there are a total of six such functions for (one for , two for , and three for ).

Instead of generating the basis functions, known as SALCs, using the projection operator, we shall derive them by logic. The SALC that transforms according to must totally symmetric. This is only possible if because the operation simply permutates the order of in , i.e. . There are three SALCs , and that transform according to . Since non-zero vanishing integrals occur only when they overlap with AOs of belonging to , these linear combinations must behave the same as the three -orbitals of under symmetry operations.  By inspection, we find that transforms like of , transforms like  of  and transforms like of . It follows that the two SALCs and  that transform according to have to behave the same as and of under symmetry operations (see diagram above). Therefore, because has positive lobes along the -axis and negative lobes along the -axis. The last SALC is and not because the latter is not orthogonal to the other SALCs. Therefore, the normalised set of SALCs are:

To construct the MO diagram, we refer to eq157 of the Hartree-Fock-Roothaan method. The total wavefunction is , where is the antisymmetriser and and  represents the AOs of and in the SALCs. In general, a solution of eq157 for the complex produces the following result:

The lowest six MOs of , and are bonding orbitals. They are occupied by 12 electrons, which are supplied by the six electron-donor ligands. The highest six MOs of ,  and are antibonding orbitals. This leaves three MOs of symmetry as non-bonding MOs. The valence -electrons of occupied the five MOs in green. The relative order of some of the MOs may vary depending on the types of metal and ligand.

 

Question

Do the remaining -orbitals (other than those of symmetry) and the -orbitals of the ligands participate in bonding with the valence AOs of ?

Answer

Yes, they may participate in bonding. When selected as a basis set, the six -orbtials of the ligands generate a reducible representation with matrices that have the same traces as those of . The SALCs also matches the six SALCs derived above. It follows that the MO diagram, when both -orbitals with symmetry and -orbitals participate in bonding with , will be a superposition of two very similar MO diagrams.

The remaining -orbitals form -bonds with the orbitals of . Such interactions will be discussed below.

 

Pi interactions

We shall utilise the remaining vectors in the diagram as a basis set. The direction of each vector indicated in the diagram is regarded as the positive lobe of the ligand -orbital. Employing the same logic as we did for interactions, we have

12 0 0 0 -4 0 0 0 0 2

Using eq27a, is decomposes to .

The orthonormal SALCs are

Irreducible representations

Orthonormal ligand SALCs Metal atom AOs
Non-bonding Non-bonding
Non-bonding Non-bonding

The MO diagram, which describes both sigma and pi interactions, has the following general form:

The and MOs are non-bonding, while the MOs are now bonding. The relative order of some of the MOs may vary depending on the type of complex and whether the ligands are -acceptors or -donors. For instance, the energies of the MOs are usually lower than those of the  MOs for -donors ligands.  Examples of ligands that can engage in both sigma and pi interactions include and , while examples of -acceptors and -donors ligands are and , respectively. These MO diagrams provide the theoretical foundation for the ligand field theory.

 

Next article: Ligand field theory
Previous article: Group theory and its applications in organic chemistry
Content page of group theory
Content page of advanced chemistry
Main content page

The Hückel method

The Hückel method is a semi-empirical method that makes additional assumptions to the -electron approximation in order to determine the energies of  molecular orbitals in planar molecules that are conjugated.

 

Question

What is a semi-empirical method?

Answer

A semi-empirical method combines theoretical approximations with experimental data to derive useful parameters like the energy of a molecular orbital (MO).

 

Like the -electron approximation, the Hückel method assumes that valence electrons in planar conjugated organic compounds are relatively unreactive for reactions that do not involve in breaking these bonds. Furthermore, they are assumed not to interact with the valence electrons, which participate in many reactions. Consequently, the valence electrons are ignored in deriving the energies of the MOs.

To solve for the energies of -electron MOs, we refer to eq157 of the Hartree-Fock-Roothaan method. The total wavefunction is , where is the antisymmetriser and . The index refers to the number of carbon atoms forming the conjugated framework of the planar molecule, while is a real and normalised orbital of the -th carbon atom. Non-trivial solutions are given by the secular determinant:

is known as the Coulomb integral if .

is known as the resonance integral if .

is the normalisation expression of if .

is known as the overlap integral if .

is the eigenvalue associated with .

Unlike ab initio methods, the exact form of the effective Hamiltonian in is not important (read on for explanation). The additional assumptions used in the Hückel method are:

In other words, all Coulomb integrals (interaction energy) have the same value of , which is reasonable if all the atoms forming the conjugated framework are the same. It follows that  for adjacent atoms is similarly a good approximation. From ab initio calculations, both and are negative. As non-adjacent carbons atoms are well separated in space, the conjecture that for non-adjacent carbons atoms and  is also reasonable. The integral is a result of the normalisation of orbitals. However, for adjacent carbon atoms is a poor approximation because orbitals of adjacent carbon atoms are not orthogonal to one another. In reality, the overlap of  orbitals of adjacent carbon atoms leads to a stabilisation (lower energy) of bonding MOs and a destabilisation (increase in energy) of antibonding MOs. Nevertheless, the relative order of the MOs along the energy axis remains unchanged.

Let’s look at a simple example: ethene. The wavefunction is and the secular determinant is

Since the orbitals are real and is Hermitian, and . Furthermore, and we have

This implies that  and the ground state electron configuration of the electrons in ethene is

 

Consequently, the Hückel method provides a quick way of predicting the number and order of a conjugated molecule’s valence MOs that likely to participate in a reaction, without explicitly specifying the Hamiltonian. The parameters and are adjusted to give the best fit to experimental data, which is why the Hückel method is considered a semi-empirical method. The total ground state electronic energy of ethene is .

 

Question

Show that i) when and when , and that ii) the normalisation constant of is .

Answer

i) Multiplying the eigenvalue equation by , integrating over all space and using the Hückel assumptions, we get

For and , we have and , respectively.

Therefore, the occupied MO wavefunction and unoccupied MO are and respectively. Since the two electrons occupying lead to a stable molecule, the MO is known as the bonding MO. The higher energy MO, which obviously results in a relatively less state when occupied, is known as the antibonding MO.

ii)

Since , we have

 

Let’s consider another molecule: 1,3-Butadiene. The secular determinant is:

The Hückel assumptions give

Using the determinant identity  , we multiply both sides of the above equation by to give

where .

Expanding the determinant, we have . The roots to the polynomial are and the energies of the MOs are

If we regard two  electrons in 1,3-Butadiene as localised between carbons atoms  and , and another two between carbon atoms and , we can compare the total ground state electronic energy of 1,3-Butadiene with that of two molecules of ethene.  The difference of implies an energy stabilisation for 1,3-Butadiene that is due to the delocalisation of electrons over the length of the molecule, rather than being localised.

The general form of the wavefunction for the four MOs is  . To determine each individual wavefunction, we apply the Hückel assumptions to eq157, giving

From eq162 and eq165, we have

Substituting eq166 in eq163 gives . When , we have .

Substituting in eq166 and comparing with eq167, . It follows that . Therefore, , which becomes after normalisation. Using the same logic, we have

For more complicated molecules, such as benzene, the Hückel method is combined with group theory to derive the wavefunctions and solve for their associated eigenvalues.

 

Previous article: The -electron approximation
Content page of computational chemistry
Content page of advanced chemistry
Main content page

Group theory and its applications in organic chemistry

Group theory plays a pivotal role in understanding molecular symmetry and electronic properties in organic chemistry, particularly when applied to the Hückel method.

Consider the benzene molecule. As the Hückel method is based on the -electron approximation, which treats electrons separately from electrons, the six carbon orbitals of benzene are used as a basis to generate a representation of the point group, which the molecule belongs to.

By selecting these six orbitals as a basis, their corresponding wavefunctions form a set of linearly independent wavefunctions that satisfy the effective Hamiltonian of the -electron approximation of benzene. This implies that any linear combination of these six wavefunctions is associated with an eigenvalue of . Since molecular orbitals (MOs) are formed by the combination of atomic orbitals, these linear combinations represent MOs. To generate a set of orthogonal MOs of , we let the symmetry operations act on the basis set to produce the following results:

The transformation matrices form a reducible representation of the  point group with the following traces:

Using eq27a, is decomposed to .

In other words, the matrices of can undergo a similarity transformation to yield block-diagonal matrices of the same form. Each block-diagonal matrix is composed of the direct sum of the four irreducible representations , , and . Since the number of orthogonal basis functions of an irreducible representation corresponds to the dimension of the representation, there are a total of six orthogonal basis functions for (one each for and , and two each for and ). These six basis functions can be generated using the projection operator.

Applying the projection operator on any -orbital wavefunction, for instance , for the 1-dimensional irreducible representations of and , we have

As there are two basis functions each for the 2-dimensional irreducible representations and , we apply the projection operator on two -orbital wavefunctions, for instance and , to give

Some of these symmetry-adapted linear combinations (SALCs) are not orthogonal to one another. As linear combinations of basis functions that transform according to an irreducible representation of a point group also belong to the same irreducible representation, we can take linear combinations of and  (and and ) to produce four SALCs that are orthogonal to and . The six orthogonal SALCs, after normalisation, are

 

Question

How do we normalise , where is a constant?

Answer

Since , we have .

 

The Hückel method assumes that

where

is known as the Coulomb integral if .

is known as the resonance integral if .

is the normalisation expression of if .

is known as the overlap integral if .

To solve for the energies of -electron MOs, we refer to eq157 of the Hartree-Fock-Roothaan method. The total wavefunction is , where is the antisymmetriser and . Non-trivial solutions are given by the secular determinant:

is known as the Coulomb integral if .

is known as the resonance integral if .

is the normalisation expression of if .

is known as the overlap integral if .

is the eigenvalue associated with .

 

Question

Show that .

Answer

Expanding the integral and using the Hückel assumptions, we have . Similarly,

 

Eq109 becomes

Using the determinant identity  , we multiply both sides of the above equation by to give

where .

Using the determinant identity   if is diagonal, we have . Therefore, , , , , and , giving the following MO diagram:

The total ground state electronic energy of benzene is . If we regard two electrons in benzene as localised between carbons atoms 1 and 2, another two between 3 and 4, and another two between 5 and 6, we can compare the total ground state electronic energy of benzene with that of three molecules of ethene. The difference of  implies an energy stabilisation for benzene that is due to the delocalisation of electrons over the benzene ring, rather than being localised.

 

Next article: Group theory and its applications in inorganic chemistry
Previous article: Group theory and infra-red spectroscopy
Content page of group theory
Content page of advanced chemistry
Main content page

The π-electron approximation

The -electron approximation is a method for determining the energies of planar conjugated organic compounds by treating the electrons separately from the electrons.

In this approximation,  valence electrons in conjugated organic compounds are assumed to form the planar molecular framework. They are considered relatively unreactive for reactions that do not involve in breaking these bonds. valence electrons, on the other hand, participate in many reactions. They are perceived as moving in some fixed electrostatic potential due to the valence electrons.

If we further assume that the total valence wavefunction of a planar conjugated organic compound has the separable form , we can express the Hamiltonian as

where

It follows that the total energy of the valence electrons is

where and .

As we have assumed that valence electrons are involved in reactions, we need only to solve for . This is done using eq157 of the Hartree-Fock-Roothaan method, with , where is the antisymmetriser and . The basis functions are the  orbitals of the carbon atoms of the planar molecule.

Evaluating for larger molecules using the aforementioned ab initio procedure may be challenging. To streamline the computation, we can introduce additional assumptions, which leads us to the Hückel method.

 

Question

What does ab initio mean?

Answer

Ab initio means “from the beginning” in Latin. An ab initio method involves deriving some parameters from first principles and not from experimental data. The Hartree self-consistent field method, the Hartree-Fock method and the Hartree-Fock-Roothaan method are all ab initio methods.

 

 

Next article: The Hückel method
Previous article: Solution of the Hartree-Fock-Roothaan equations for helium
Content page of computational chemistry
Content page of advanced chemistry
Main content page

Rayleigh-Jeans law

The Rayleigh-Jeans law is a flawed attempt in physics to describe the spectral radiance of electromagnetic radiation as a function of wavelength from a blackbody at a given temperature.

Consider a cube of side , filled with electromagnetic radiation in thermal equilibrium at temperature . If a tiny aperture exists in one of the walls, the radiation that it emits would exhibit the properties of an ideal blackbody. The walls, regarded as perfect conductors, provide boundary conditions for the electromagnetic waves inside the cube.

In a perfect conductor, an idealised model for real metals, electric charges can move without any resistance. An electromagnetic wave propagating in the -direction, with its electric field vector parallel to the wall, causes electrons in the conductor to move in response to the field. The movement of these electrons creates an opposing electric field that cancels out the external electric field. Given that the electric field is zero at the walls, the boundary conditions can only be satisfied by standing waves. These are waves that oscillate in place and have nodes at the cube’s walls. For example, standing waves oscillating in the -direction have wavelengths of , where is an integer (see below diagram).

The amplitude of each of the above standing waves can be expressed as , or in terms of the wave vector pointing in the -direction:

where .

Similarly, the wavelengths of the oscillation modes of the electromagnetic field in the -direction and -direction are and  respectively. Correspondingly, the magnitudes of the wave vectors are given by and  respectively. It follows that the wave vector  of the oscillation modes of the electromagnetic field in an arbitrary direction in the cube lies in a -space, which contain a cubical array of points that are apart (see Figure I below).

Since the magnitude of the wave vector is , boundary conditions are satisfied by each ordered triple . In other words, every mode of oscillation of an arbitrary wave oscillating in the cube is defined by a point in the -space. We can also plot the array of modes in an -space, where the points are now unit length apart from one another (see Figure II above). Substituting in , we have

where is the radius of an octant (one-eighth of the volume of a sphere) in the -space.

For a certain value of , ordered triples of that satisfy eq1 lie on the surface of the octant. Consider the volume of the shell of the octant enclosed by the surfaces at  and . Substituting eq1 and  in , we have

Since the density of the number of modes in the -space is one per unit volume, the number of modes within the shell is the product of and :

As there are two independent polarisations for each mode of an electromagnetic wave (see above diagram),

Lord Rayleigh used the equipartition theorem to assume that each mode has an average energy of . Multiplying the above equation by the average energy per mode, we have

where is the energy density (average energy per mode between and per unit volume of the cube).

Eq2 is known as the Rayleigh-Jeans law, which fails to describe the spectrum of blackbody radiation at high frequencies because it implies that radiated energy would increase without limit as increases, leading to what is called an ultraviolet catastrophe (see diagram below).

 

Next article: Planck radiation law
Previous article: Reduction of the two-particle problem to one-particle problems
Content page of quantum mechanics
Content page of advanced chemistry
Main content page

Planck radiation law

The Planck radiation law explains how a blackbody emits electromagnetic radiation at a specific temperature, based on the assumption that the energy of each oscillator in the body can only have discrete values.

In June 1900, Lord Rayleigh published the Rayleigh-Jeans law, which is now known as a flawed attempt in physics to describe the spectral radiance of electromagnetic radiation as a function of wavelength from a blackbody at a given temperature. The mistake that he made was to use the equipartition theorem to assume that each oscillation mode within a blackbody has an average energy of . In December of the same year, the German physicist Max Planck presented the Planck radiation law, which assumed that the energy of an oscillator of frequency came in discrete bundles:

where and  is a proportionality constant called the Planck constant.

According to the Boltzmann distribution, the probability of a mode with frequency  associated with the state is

The average energy of the mode of frequency is

Let .

Substituting the Taylor series of and in the above equation,

Substituting eq4 in eq2,

which is the mathematical expression of the Planck distribution law.

 

Question

Show that in the classical limit, the average energy of a mode in eq4 is consistent with the equipartition theorem.

Answer

In the classical limit,  and we can expand as the Taylor series . Substituting the series in eq4 and ignoring the higher powers of the series because , we have .

 

Previous article: Rayleigh-Jeans law
Content page of quantum mechanics
Content page of advanced chemistry
Main content page