Slater determinant

A Slater determinant represents a multi-electron wavefunction that satisfies the Pauli exclusion principle. It was named after the American physicist John Slater for his contribution to quantum mechanics.

The normalised anti-symmetric wavefunction for an n-electron system with spin-orbitals can be expressed as , where and is the antisymmetriser. From eq59 and 62,

Comparing the above equation with the definition of a determinant, i.e. , we have or

which is called an n-electron Slater determinant.

Since   (see this article for proof), the Slater determinant can also be written as


 

Next article: Unitary transformation
Previous article: Properties of the antisymmetriser
Content page of computational chemistry
Content page of advanced chemistry
Main content page

Unitary transformation

A unitary transformation of a set of vectors to another set of vectors preserves the lengths of the vectors and the angles between the vectors.

In other words, a unitary transformation is a rotation of axes in the Hilbert space. This implies that if the transformation involves matrices of eigenvectors, the eigenvalues of the eigenvectors are preserved. Consider 2 complete sets of orthonormal bases:

where is the identity matrix.

The relation between the two basis sets are

where are elements of the transformation matrix U.

We say that is transformed to by . Similarly, we have .

The reverse transformation of eq68 is:

Question

Show that .

Answer

Similarly,

 

 

Next article: Slater determinant of unitarily transformed spin orbitals
Previous article: Slater determinant
Content page of computational chemistry
Content page of advanced chemistry
Main content page

Slater determinant of unitarily transformed spin orbitals

The Slater determinant of unitarily transformed spin orbitals is related to the Slater determinant of the original spin orbitals by the product of  and the determinant of the unitary matrix .

A unitary matrix is defined by eq70, where . Let , where

Substituting and in eq66 yields


and

respectively.

Using the identity of for square matrices (see this article for proof) results in

Substituting eq71a and eq71b in eq72 gives

From eq70, . Since (see this article for proof), we have . Therefore, (or ) and

The average value of a physical property of an electron, which is described by , is , which is the same when the electron is described by . In other words, any expectation value of a physical property of an electron is invariant to a unitary transformation of the electron’s wavefunction that is expressed as a Slater determinant. Such a consequence is used to derive the Hartree-Fock equations.

Lastly, if , and are the matrix elements of , and , we have

where we have replaced the dummy variable within the parentheses with .

The reverse transformations are

With reference to eq71,

Next article: Slater-Condon rule for a one-electron operator
Previous article: Unitary transformation
Content page of computational chemistry
Content page of advanced chemistry
Main content page

Slater-Condon rule for a one-electron operator

The Slater-Condon rule for a one-electron operator is an expression of the expectation value of the one-electron operator involving Slater determinants.

Let’s consider the expectation value of the one-electron Hamiltonian operator: , where  and is given by eq66.

 

Question

Show that the product of  and is Hermitian.

Answer

or .

With reference to eq64 and the fact that , we have . Using eq65, we have , i.e. the product of and is Hermitian. This implies that , where is also Hermitian.

 

Substituting eq59 in and using

i) eq65
ii) the Hermitian property of
iii) the Hermitian property of twice

yields

Substitute eq62 and eq63 in the above equation,

Substituting and in eq85 and simplifying gives

where .

Given that any expectation value of a physical property of an electron is invariant to a unitary transformation of the electron’s wavefunction expressed as a Slater determinant, we substitute eq74 and eq75 in eq86 to give

where we have changed the dummy variable from to .

Substitute eq82 in eq87,

Eq88 is used in the derivation of the Hartree-Fock equations.

 

Next article: Slater-Condon rule for a two-electron operator
Previous article: Slater determinant of unitarily transformed spin orbitals
Content page of computational chemistry
Content page of advanced chemistry
Main content page

Slater-Condon rule for a two-electron operator

The Slater-Condon rule for a two-electron operator is an expression of the expectation value of the two-electron operator involving Slater determinants.

Let’s consider the expectation value of the two-electron Hamiltonian operator: , where  and  is given by eq66.

Substituting in eq85,

Substituting in the above equation and simplifying, we have

where and .

and are known as the Coulomb integral and the exchange integral, respectively.

 

Question

Show that .

Answer

Since , we can add terms of , where  to , resulting in , i.e.

 

Given that any expectation value of a physical property of an electron is invariant to a unitary transformation of the electron’s wavefunction expressed as a Slater determinant, we substitute eq74, eq75, eq76 and eq77 in of eq89 to give

where we have changed the dummy variables from and  to and respectively.

Substituting eq82 in the above equation and simplifying


Similarly, substituting eq83 in the above equation and simplifying

Substituting eq74, eq75, eq76 and eq77 in of eq89 and repeating the above logic, we have

Substitute eq90 and eq91 back in eq89

Eq92 is used in the derivation of the Hartree-Fock equations.

 

Next article: Hartree-Fock method
Previous article: Slater-Condon rule for a one-electron operator
Content page of computational chemistry
Content page of advanced chemistry
Main content page

Hartree-Fock Method

The Hartree-Fock method is an iterative procedure that optimises an approximate asymmetric wavefunction using the variational principle, in the attempt to estimate the eigenvalues of a modified form of the non-relativistic multi-electron Hamiltonian.

One of the deficiencies of the Hartree self-consistent field method is that it does not consider exchange forces due to electron spin interactions. This is because the derivation of the Hartree equations uses eq3, which does not satisfy the Pauli exclusion principle. In 1930, Vladimir Fock developed an improved procedure called the Hartree-Fock method. It replaces the product of orbitals in eq3 with a Slater determinant of spin-orbitals to represent the wavefunction of an atom.

Similar to the Hartree self-consistent field method, the computation of the energy of an atom via the Hartree-Fock method involves simultaneously solving a set of equations called the Hartree-Fock equations:

where .

The derivation of the Hartree-Fock equations is shown in the next article.

 

Next article: Proof of the Hartree-Fock equations
Previous article: Slater-Condon rule for a two-electron operator
Content page of computational chemistry
Content page of advanced chemistry
Main content page

Proof of the Hartree equations

To proof the Hartree equations, we begin by rearranging eq4:


Noting that is normalised and simplifying the 1st and 2nd terms on RHS of eq15, we have

respectively, where , , and .

Substituting eq16 and eq17 back in eq15, noting that (see this article for explanation)

where E is a functional of 2n independent variables, with n variables of and n variables of .

To find the minimum value of E subject to n constraints of , we apply the Lagrange method of undetermined multipliers to form the new functional , where

Question

Show that .

Answer

Taking the complex conjugate of F throughout, noting that E and hence F is real (recall that the functional F is obtained by subtracting the function , where from E),

Comparing the above equation with the original F,

 

Substituting eq18 in and , and noting that the 2nd order terms are approximately zero,




The minimum energy corresponding to F is when a small change in the functional’s input (change in and ) yields no change in the functional’s output, i.e. when , or

 

Question

Show that is Hermitian.

Answer

Since is Hermitian and is real,

 

Using the Hermitian property of the operator

 It can be easily shown, by expanding the Coulomb terms in the above equation, that 1st and 2nd Coulomb terms are the same, and that the 3rd and 4th Coulomb terms are the same. Therefore,

and can now be chosen arbitrarily. If we vary only the k-th function , eq20a becomes

Since is chosen arbitrarily, the only way to satisfy the above equation is for

Similarly, if we vary only the m-th function , where , we have

Repeating this logic to all other functions, we have a set of n simultaneous equations:

and another set of n simultaneous equations in the complex conjugate form:

Eq21 is known as the Hartree equations, and eq22, the complex conjugate form of the Hartree equations. The Hartree equations are sometimes written by changing the dummy variables and to and respectively. For example, eq21 can be expressed as

 

Next article: Simpson’s rule
Previous article: Functional variation
Content page of computational chemistry
Content page of advanced chemistry
Main content page

Canonical Hartree-Fock equations

The canonical Hartree-Fock equations are a set of eigenvalue equations, which when solved iteratively gives the eigenvalues of a modified form of the non-relativistic multi-electron Hamiltonian.

Question

Show that

Answer

Substituting eq75, eq76 and eq102 in and changing the dummy variables and to and respectively, we have

where .

With reference to eq84, is the matrix element of an matrix, which is the product of three matrices. In matrix notation,

Since is Hermitian, we can always select a pair of matrices and such that  produces a diagonal matrix. Therefore, and

 

To derive the canonical equations, we perform a unitary transformation of in eq96, i.e. by substituting eq88, eq92 and eq106 in eq96 to give

where

If we apply the functional variation method as per the previous article, we have

Since , we can remove n terms of , where , from , resulting in .

Therefore, eq107 becomes

To satisfy the above equation, we select a set of such that n terms in eq108 are zero, ensuring all other variable functions and are independent. Using the same logic described in the steps taken from eq20 to eq21, we have the canonical Hartree-Fock equations:

where eq110 is the complex conjugate of eq109.

 

Next article: Gram-Schmidt process
Previous article: Proof of the Hartree-Fock equations
Content page of computational chemistry
Content page of advanced chemistry
Main content page

Gram-Schmidt Process

The Gram-Schmidt process is a mathematical technique for orthogonalising a set of nonorthogonal vectors.

Consider two linearly independent nonorthogonal vectors and of the Hermitian operator :

where is the degenerate eigenvalue corresponding to and .

Let . With reference to the diagram above, , where we have used the fact that the magnitude of the unit vector is 1 in the second equality. Furthermore, ; that is, the vector is a multiple of the unit vector , with the multiple being . Therefore,


is a component of , which is orthogonal to . The eigenvalue of is unchanged versus that of because

In the presence of a third linearly independent vector that is nonorthogonal to and (where ), the vector  that is orthogonal to is (c.f. eq111). To determine the component of that is orthogonal to both and , let  be the component of  that is orthogonal to :

We can immediately see that is orthogonal to both and because it is sum of two vectors and , each of which is orthogonal to (and hence the dot product  is zero). Substituting in the above equation, noting that and are scalars and that , we have

Therefore, for a set of three linearly independent nonorthogonal vectors , the transformed set of vectors, which are orthogonal to one another, is , where

For a set of  linearly independent nonorthogonal vectors , the transformed set of vectors, which are orthogonal to one another, is with the k-th transformed vector as

The corresponding orthonormal vectors are , , … , .

An example of the application of the Gram-Schmidt process is the orthogonalisation of nonorthogonal Slater-type orbitals in the Hartree-Fock method.

 

Next article: Spin-orbital
Previous article: Canonical Hartree-Fock equations
Content page of computational chemistry
Content page of advanced chemistry
Main content page

Spin-orbital

A spin-orbital expresses the state of an electron as a complete wavefunction, which comprises of a spatial component and a spin component.

It is defined as

where is a composite coordinate consisting of three spatial coordinates and one spin coordinate , while and are spin wavefunctions describing the two possible spin states of an electron.

In the derivation of the canonical Hartree-Fock equations, , where . When we substitute the spin-orbitals in their explicit forms into eq109, we have


Multiplying throughout by  and integrating with respect to the spin coordinates, while noting that because of spin orthogonality, we have

 

Question

Why are the spin components of the total wavefunction orthogonal?

Answer

From the article on Hermitian operators, we know that two eigenfunctions of a Hermitian operator that correspond to different eigenvalues are orthogonal. Since the spin components of the total wavefunction and correspond to different eigenvalues of the spin Hermitian operator , they are orthogonal.

 

Hence the spin orbitals reduce to spatial orbitals. This implies that we only need to work with spatial orbitals, such as Slater-type orbitals, when we use the Hartree-Fock method to estimate the ground state energies of atoms.

 

Next article: Exchange integral
Previous article: Gram-Schmidt process
Content page of computational chemistry
Content page of advanced chemistry
Main content page
Mono Quiz