Analytical solution of the energy equation for He using the Hartree self-consistent field method

The analytical solution of the energy equation of He involves finding analytical expressions for all terms in eq8 using trial one-electron wavefunctions. As per the numerical method, guess values are used for the variables in all trial one-electron wavefunctions except one, e.g. . Eq8 is then minimised to obtain the solutions for and . The process is repeated until and all become invariant. For He, eq8 is

where and .

Our computation employs the following assumptions:

  • We base our iteration on the unrestricted case, where electrons and are distinctively expressed by and respectively.
  • All terms on the RHS of eq48 are determined using Slater-type orbitals, where

Substituting in , using the identity  for and integrating by parts for , we have

Similarly, substituting in , we have

Substituting and eq47 in and multiplying the resultant equation by , where is the spherical harmonics for , we have

Due to the orthogonality of the spherical harmonics, the only integral that survives upon expanding the summation is when and . Since, , the above equation becomes:

We proceed by integrating with respect to first. Since ranges from 0 to , we can split the integral into two parts, one from 0 to and the other from to . Supposing , the above equation becomes:

As increases from 0 and approaches , it must be less than and so . Similarly, as increases from to , it must be greater than and so . Therefore,

Using the identity  for the first integral within parentheses, integrating by parts for the second integral within parentheses, and employing the identity for the integral with respect to yields

Substituting eq50, eq51 and eq55 in eq48 gives

Differentiating eq56 with respect to and results in

Eq57 and eq58 are used in an iterative algorithm to find , and  . The procedure is as follows:

  • Substitute an initial guess value of , e.g., , in eq57 and solve for by setting eq57 equal to zero. We then substitute the solution of  and the initial guess value of in eq56 to find .
  • To obtain an improved estimate of , we substitute the solution of from the previous step in eq58, and solve for by setting eq58 equal to zero. We then substitute the solution of and the value of found in the previous step in eq56 to find a better estimate of .
  • Steps 1 and 2, which form an iteration set, are repeated until the values of and are invariant up to six decimal points.

Alternatively, we can set up an iterative table in Excel as follows:

Cell C2 is the initial guess value of . The formula for D2:D7 is eq56, i.e. =(B2^2)/2-2*B2+(C2^2)/2-2*C2+(B2*C2*((B2^2)+3*B2*C2+(C2^2)))/((B2+C2)^3). The formulae for B3, C4, B5, C6 and B7 are =B2, =C3, =B4, =C5 and =B6, respectively. To compute the first iteration set, we employ the Excel Solver application to minimise D2 with respect to B2, with the following settings:

Set objective: D2
To: Min
By changing variable cells: B2
Selecting a solving method: GRG Nonlinear

We then solve for D3 by changing the ‘Set objective’ field and ‘By changing variable cells’ field to D3 and C3 respectively. This procedure is repeated for subsequent iteration sets until the values of and are invariant up to six decimal points.

The results are as follows:

The final theoretical value of has a deviation of about 1.93% versus the experiment data of the ground state of helium.

 

Next article: Antisymmetriser
Previous article: The expansion of \(\frac{1}{r_{ij}}\) (the reciprocal of the inter-electron distance)
Content page of computational chemistry
Content page of advanced chemistry
Main content page

 

Analytical solution of the energy equation for Li using the Hartree-Fock method

The analytical solution of the energy equation for Li involves finding analytical expressions for all terms in eq94 using one-electron trial wavefunctions. Guess values are used for the variables in all the wavefunctions except one, e.g. . Eq94 is then minimised to obtain the solutions for E and . The process is repeated until E and all become invariant.

After removing and due to spin orthogonality from eq94, we have

where


Our computation employs the following assumptions:

  • The ground state of Li is described by a single Slater determinant.
  • Eq113 is minimised iteratively via the restricted open-shell method, where the 1s orbital is doubly occupied with , while the 2s orbital is singly occupied. Therefore, we can substitute eq56 in eq113, where , to give

 

Question

Using the normalised 1s Slater-type orbital and the un-normalised generic 2s Slater-type orbital , derive via the Gram-Schmidt process.

Answer

With reference to eq111, we have

where N’ is the normalisation constant.

Substituting  and in the above equation and integrating using the identity , and then normalising and rearranging the result, we have

where


 

Let’s evaluate the remaining terms in eq114, beginning with . Substituting eq116 in and integrating,

Next, we evaluate , which is equal to . Substituting eq116 and eq47 in and integrating (refer to the integration procedure in this article), we have

Finally, we evaluate . Referencing the steps in integrating , we have

Substituting eq119, eq129, eq142 and the atomic number  in eq114, we have the analytical expression of E, which is used in an iterative algorithm to determine the ground state of Li. The procedure involves setting up an iterative table in Excel as follows:

Cell C2 is the initial guess value of . Cell F2 contains the entire formula of the analytical expression of E, with links to B1, C1, D1 and E1. Cell D2 and E2 contain the formulae of N and M respectively, with links to B1 and C1. The formulae for B3, D3 and E3 are ‘=B2’, ‘=D2’ and ‘=E2’ respectively. To compute the first iteration set, we employ the Excel Solver application to minimise F2 with respect to B2, with the following settings:

Set objective: F2
To: Min
By changing variable cells: B2
Selecting a solving method: GRG Nonlinear

We then solve for F3 by changing the ‘Set objective’ field and ‘By changing variable cells’ field to D3 and C3 respectively. This procedure is repeated for subsequent iteration sets until the values of and are invariant up to seven decimal points.

The results are as follows:

The final theoretical value of has a deviation of about 0.80% versus the experiment data of the ground state of Li.

 

Next article: Analytical solution of the Hartree-Fock method for Be, B and C
Previous article: Exchange integral
Content page of computational chemistry
Content page of advanced chemistry
Main content page

Analytical solutions of the Hartree-Fock method for Be, B and C

The analytical Hartree-Fock method for determining the ground state of Li can be extended to Be, B and C. The energy equations to minimise using Excel Solver for Be, B and C are

respectively; where .

The results for Be are

The final theoretical value of has a deviation of about 0.76% versus the experiment data of the ground state of Be.

The results for B are

The final theoretical value of has a deviation of about 0.65% versus the experiment data of the ground state of B.

The results for C are

The final theoretical value of has a deviation of about 0.66% versus the experiment data of the ground state of C.

 

Next article: Koopman’s theorem
Previous article: Analytical solution of the energy equation for Li using the Hartree-Fock method
Content page of computational chemistry
Content page of advanced chemistry
Main content page

Koopmans’ theorem

Koopmans’ theorem states that the energy to remove an electron from an orbital of an atom, whose state is described by a single Slater determinant, is approximately the negative value of the corresponding Hartree-Fock orbital energy. In other words, the theorem states that the negative value of the orbital energy of an atom is approximately equal to the ionisation energy IE for the k-th electron in the atom.

To prove the theorem, we begin with the ionisation process, which is generally expressed as

where is an atom, which is irradiated with a photon of energy , and is the corresponding ion with an electron removed.

In terms of energies,

where is the state of the ion, is the state of the atom and is the kinetic energy of the expelled electron.

As ,

Substituting eq95 in eq143 gives

 

Question

Show that .

Answer

Let . Expanding the equation and eliminating common terms,

Since , we can add it to , giving

Furthermore, and . So,

Relabelling i to j for

 

Substituting eq145 in eq144 yields

Relabelling t with j and with in eq109, we have

Multiplying the k-th equation in eq147 by and integrating over , we have

Since

Substituting eq148 in eq146 results in

Eq149 is known as Koopmans’ theorem.

Lastly, let’s study the relation between the relation between the orbital energy  of an atom and the total energy E of the atom. From eq148,

Substituting eq94 in the above equation gives

 

Question

Show that the alternative to eq150 is .

Answer

Adding  to both sides of eq148 and summing throughout from ,

Since ,

 

Next article: Atomic orbital energies
Previous article: Analytical solutions of the Hartree-Fock method for Be, B and C
Content page of computational chemistry
Content page of advanced chemistry
Main content page

Atomic orbital energies

Atomic orbital energies are eigenvalues of the Hartree-Fock equations. These eigenvalues are expressed by eq148, and are based on a specific state of the atom. As illustrated in the article on the analytical solution of the Hartree-Fock method for the ground state of Li, we can theoretically compute for any occupied orbital of an atom. For example, of the 2p orbital of Li can be determined using the excited electron configuration state of 1s22p1. However, if we want to determine the energies of unoccupied orbitals of an atom in its ground state, we would have to use the Hartree-Fock-Roothaan method, where the basis set includes a linear combination of the 1s, 2s and 2p wavefunctions. In general, for atoms with to are depicted in the diagram below.

For atoms with , their 4s orbital energies lie below their respective 3d orbital energies (see diagram below). An example is the atom  with the ground state electron configuration [Ar]4s1. This implies that the theoretical ground state energy of is closer to the experimental value when we use a 4s Slater-type orbital rather than a 3d Slater-type orbital for computing .

With that in mind, one may conclude that the ground state electron configuration of Sc is [Ar]3d3. However, it is [Ar]3d14s2. This is because the orbital energies and for [Ar]3d3 are different from those for [Ar]3d14s2 (refer to eq148), even though for both configurations. For example,  for the electron configurations [Ar]3d14s2 and [Ar]3d24s2 are (in Hartree units)

and

respectively, where .

In short, the Hartree-Fock computation using the electron configuration of [Ar]3d14s2 gives a ground state energy that is closest to experimental values. Qualitatively, we rationalise the above with the fact that 3d orbitals are smaller in size compared to the 4s orbitals. Electrons occupying 3d orbitals therefore experience greater repulsions than electrons residing in 4s orbitals, with the order of increasing repulsion being:

where V is the potential energy due to repulsion.

Therefore, to determine the stability of an atom in the ground state, we need to consider the net effect of the relative energies of 4s/3d orbitals and the repulsion of electrons. Calculations for the overall energies of Sc show that

 

Next article: Hartree-Fock-Roothaan method
Previous article: Koopmans’ theorem
Content page of computational chemistry
Content page of advanced chemistry
Main content page

Hartree-Fock-Roothaan method

The Hartree-Fock-Roothaan method, an extension of the Hartree-Fock method, uses spin orbitals that are linear combinations of a set of basis wavefunctions .

Unlike the Hartree-Fock method, in which the parameters of the wavefunctions are varied, the best basis wavefunctions are chosen in the Hartree-Fock-Roothaan method and the coefficients are instead varied in an iterative process to determine the state of the system.

To derive the Hartree-Fock-Roothaan equations, we begin by substituting and in eq88, where we have relabelled and as and respectively, to give

where   .

Next we substitute , , and in eq92 to give

where   and .

Substituting eq152 and eq153 in eq94 gives

The constraint for the Lagrange method is found by substituting  and  in to give:

where .

Therefore, the Lagrangian is:

where  and are the undetermined multipliers.

As shown in the derivation of the canonical Hartree-Fock equations, we can select a set of coefficients and that diagonalises the Hermitian matrix with elements . The Lagrangian becomes

The total differential of the Lagrangian is

where .

Eq155 has the same form as eq12. If we can find a set of values of that renders the dependent variable terms of zero, we are left with the independent variable terms. Consequently, all the coefficients of are equal to zero and they form a set of equations that can be solved simultaneously. To simplify eq155, substitute eq154 in it to give

The next step involves the following:

    1. Renaming the dummy indices of the 5th and 6th terms by swapping i and j, and , and , and the dummy coordinates and .
    2. Noting that and .
    3. Noting that (see this article for explanation).
    4. Expanding each term of the equation and carrying out the partial differentiation.

We have,

Switching the dummy labels , and , , and the dummy coordinates and for the 1st term on the RHS of the above equation,

where the 1st term of the above equation is the complex conjugate of the 2nd term.

Since all coefficients of and are equal to zero, we have

where .

Similarly, the complex conjugate in eq156 gives:

Eq157 and eq158 are known as the Hartree-Fock-Roothaan equations.

Eq157 has non-trivial solutions if the determinant equals to zero. The computation process involves:

  1. Evaluating the integrals and either analytically or numerically using initial guess values of , and with basis wavefunctions where the parameters are fixed.
  2. Substituting the evaluated integrals in the characteristic equation and solving for , which is then used to obtain improved values of for the next iteration.
  3. Repeating steps 1 and 2 with the improved values of from the previous iteration until self-consistency is attained.
NEXT article: Solution of the Hartree-Fock-Roothaan equations for helium
Previous article: Atomic orbital energies
Content page of computational chemistry
Content page of advanced chemistry
Main content page

Antisymmetriser

An antisymmetriser is an operator that makes a wavefunction antisymmetric.

Consider a system of two-electrons, each of which is described by a spin-orbital.

 

Question

What is the difference between an orbital and a spin-orbital?

Answer

An orbital is a one-electron spatial wavefunction , i.e. a function that gives the best estimated energy eigenvalue of an electron when operated on by the Hamiltonian. For a given spatial orbital , we can form two spin-orbitals and , where and are spin functions. We say that the spatial orbital is doubly occupied by two electrons with the same energy. The abbreviated symbol of a spin-orbital is , where  represents the set of all 4 coordinates (3 spatial coordinates and 1 spin coordinate) associated with the electron.

 

The symmetric and antisymmetric forms of the total wavefunction of the system are and respectively. We have

where

      1.  .
      2.   is the identity operator.
      3. , like the exchange operator, swaps the labels of any two identical particles, i.e.

Eq59 implies that the antisymmetriser transforms the symmetric form of the wavefunction into a linear combination of states, each with a distinct permutation of spin-orbital labels. 

We can also expressed the antisymmetriser as , where an even (odd) numbered r represents an even (odd) number of times the labels are permutated. When the antisymmetriser acts on the wavefunction of an n-electron system, , we would expect the linear combination to have terms since there are  ways to permutate the labels. The antisymmetriser is therefore:

where N is the normalisation constant.

To evaluate N, we have

Due to the orthonormal property of , we find, after expanding the product of the linear combinations of and its complex conjugate, that the integrals result in terms of unity. Hence,

Substitute eq61 in eq60,

 

Next article: Properties of the Antisymmetriser
Previous article: Analytical solution of the energy equation for He using the Hartree self-consistent field method
Content page of computational chemistry
Content page of advanced chemistry
Main content page

The expansion of \(\frac{1}{r_{ij}}\) (the reciprocal of the inter-electron distance)

The expansion of enables the Hartree equations to be solved analytically. Consider two electrons and  as depicted in the diagram below:

According to the law of cosines , which is equal to or

where .

Due to inter-electronic repulsion,  if is acute and hence , which implies that we can express eq35 as a binomial series, where  . Substituting  in the series, rearranging and displaying the terms of up to yields

Since the coefficients of the powers of are the Legendre polynomials , and that , we have

As mentioned in the previous article, is dependent on , , and (see diagram below). This implies that a function of like is a function of four angles. If we rotate the z-axis such that it lies on  , the vectors are expressed in a new coordinate system, where the new polar angle and azimuthal angle are   and  respectively. Consequently, we have , which is a simpler function to work with. We can also simplify the function in the original coordinate system by holding   and  as constants, resulting in a function of the variables and , or .

With reference to the new coordinate system, and its complex conjugate are eigenfunctions of with eigenvalue . For the original coordinate system, , and its complex conjugate are eigenfunctions of with the same eigenvalue .

 

Question

Show that the eigenvalue of   is in any three-dimensional coordinate system.

Answer

From eq108 and eq109, the quantum orbital angular momentum ladder operators for the original and new coordinate systems are and  respectively. The corresponding eigenfunctions of  and are and  respectively. Following the steps of determining the eigenvalues of and , we have and respectively.

 

Furthermore, the general solution of the eigenfunction of  is , where are the associated Legendre polynomials and are the coefficients of the basis polynomials in the linear combination. Therefore,

Question

Why is ?

Answer

For a particular value of , the linear combination is a sum of basis functions of a particular quantum angular momentum value . For example, when , we have a linear combination of the hydrogenic -orbital wavefunctions. Therefore, the eigenvalues of each term of the sum is always degenerate, which ensures that is an eigenfunction of .

 

To solve for , we multiply both sides of eq37 by the eigenfunction  and integrate over all space:

Due to the orthogonal property of associated Legendre polynomials, only the  term in the RHS expansion of the above equation survives,

From eq390, . Substituting this equation in the above equation and rearranging gives

To evaluate the integral in eq38, we expand the eigenfunction as a linear combination of basis functions in the new coordinate system:

 

Question

Why can we express in the original coordinate system as a linear combination of basis functions in the new coordinate system?

Answer

As mentioned earlier in the article,   and share the same set of eigenvalues. For a particular value of , is an eigenfunction of with a specific eigenvalue of , which is the same eigenvalue associated with for the same value of . In other words, , which implies that is invariant with respect to rotation of the coordinate system.

 

To solve for , multiply both sides of eq39 by , integrate over all space and repeat the steps in evaluating . We have

Next, eq39 must hold for the case of , with and . Expanding eq39,

From eq363a, , which is equal to zero if  and is equal to one if (because ), i.e.

So the only term that survives in eq41 is  and eq41 becomes

Substitute eq40, where and noting that , in eq43,

Substitute eq44 in eq38

Combining eq45, eq37 and eq36

Substitute in eq46, we have

 

 

Next article: Analytical solution of the energy equation for He using the Hartree self-consistent field method
Previous article: The angle between two vectors in spherical coordinates
Content page of computational chemistry
Content page of advanced chemistry
Main content page

Functional variation

Functional variation refers to the change in a functional’s output as a result of a small change in the functional’s input on a specified domain.

 

Question

What is a functional?

Answer

A functional is a function whose value depends on a function instead of an independent variable. For example, the expectation value of the Hamiltonian is a functional, where .

 

For a small variation in ,

where is a small arbitrary change in .

Since is small, , and

where .

If , we have or

which means that a small change in the functional’s input yields no change in the functional’s output. This implies a minimum energy according to the variational principle.

 

Next article: Proof of HArtree equations
Previous article: Lagrange method of undetermined multipliers
Content page of computational chemistry
Content page of advanced chemistry
Main content page

The angle between two vectors in spherical coordinates

The angle between two vectors in spherical coordinates is depicted in the diagram below:

We have

where and are the unit vectors of and , respectively.

Substituting eq182b (the unit vector in spherical coordinates) in eq33, noting that , and rearranging, gives

 

Next article: The expansion of \(\frac{1}{r_{ij}}\) (the reciprocal of the inter-electron distance)
Previous article: Numerical solution of the Hartree self-consistent field method for He
Content page of computational chemistry
Content page of advanced chemistry
Main content page
Mono Quiz