Numerical solution of the Hartree self-consistent field method for He

In this article, we shall calculate the ground state energy of a helium atom in one-dimension using an Excel spreadsheet. From eq5, the equations to be solved iteratively are:

The presence of multiple functions and variables in each of these equations requires us to devise a computation strategy, the first of which is to derive a trial wavefunction.

 

Question

Show that  is a reasonable wavefunction describing the helium ion.

Answer

The one-dimensional Schrodinger equation for the helium ion is , which can be written as


where , , and .

The general solutions of eq27 are of the form or . Let’s use , which is a modified form of a Slater-type orbital, where for the helium ion. Since , the normalisation constant N for y is . So, .

 

As the helium atom experiences an effective nuclear charge that is between that of a hydrogen atom () and a helium ion (), it can expressed by the trial wavefunction , which after normalisation is


Next, the pair of equations mentioned at the top of the article have two distinct wavefunctions, each representing one of the two electrons. To simplify our computation, we assume that the two electrons can be expressed by the same spatial wavefunction. This is known as the restricted case. To iteratively solve for the common wavefunction, we need to transform the two equations into a new pair of equations. Rearranging eq5 for helium in one-dimension, we have

where  and .

Substitute the definition of the 2nd derivative in eq29 and rearranging, we have

To prevent from being undefined when , we modify the Coulomb integral by adding a small constant C to the denominator. It is found that gives reasonable computation results. So,

The definition of the 1st derivative is
or:

Eq30 and eq31 are used in an iterative algorithm to find and . The procedure is as follows:

1) Set the boundary condition and . With reference to eq30, use eq28 as the initial guess wavefunction for the Coulomb integral. Let’s also assume an initial guess gradient of .

2) Let with intervals of (350 intervals in total). This implies that at .

3) Use Simpson’s rule to integrate the Coulomb integral for all . An example of the excel table is shown below.


With reference to eq24, the formulae for cells C4, D4 and E4 are =(C2^2)*EXP(-2*1.5*C2)/(ABS(B4-C2)+0.5), =4*(D2^2)*EXP(-2*1.5*D2)/(ABS(B4-D2)+0.5) and =2*(E2^2)*EXP(-2*1.5*E2)/(ABS(B4-E2)+0.5) respectively. The formula for cell MP4 is =SUM(B4:MN4). Finally, the formula for cell MQ is 13.5*7/350/3*MP4, where 13.5 is the square of the normalisation constant of eq28. The formulae for the remaining cells within C4:MQ354 are populated accordingly.

4) The iterative table is formulated as follows:


The second row is for guess values and the error function. Rows 4 to 354 are for the algorithm. Cells MS2 and MV2 are the initial guess values of and b respectively. The formula for the normalisation constant in cell MU2 is =SQRT(((-MV2*2)^3)/2). Cell MS4 is obviously zero. The formula for cell MT4 is =MS2. With reference to eq29 and eq28, the formulae for cells MS5 and MT5 are, =MS4+MT4*(B5-B4) and =MT4+(MQ5-(2/B5)-MT2)*2*MS5*(B5-B4) respectively. The formulae for the remaining cells within MS4:MT354 are populated accordingly.

Next, we plot a graph using values in cells MS4:MS354 as vertical coordinates and B4:B354 as horizontal coordinates. By visually inspecting the graph, the value in cell MT2 is manually adjusted so that when . To optimise the values in cells MS2 and MV2, we include a trendline in the plot, where the formulae for cells MU4 and MU5 is zero and =MU2*B5*EXP(MV2*B5) respectively. The formulae for the remaining cells within MU4:MU354 are populated accordingly.

The error function formula in cell MW2 is {=SQRT(SUM((MS5:MS365-MU5:MU354)^2))}. The curly brackets {} are obtained by first typing the formula within {} in the cell and pressing CTRL+SHIFT+ENTER. Lastly, the Excel Solver application is use to optimise the values in cells MS2 and MV2, with the following settings:

Set objective: MW2
To: Min
By changing variable cells: MS2,MV2
Selecting a solving method: GRG Nonlinear

5) For the next set of iterations, repeat steps 3 and 4, where the guess values of MS2 and MV2 are the optimised values of the previous set of iterations. The value for MU is no longer the formula =SQRT(((-MV2*2)^3)/2) but the optimised normalisation constant of the previous set of iterations. Similarly, the values of and  used in the Coulomb integral in step 3 are no longer  and but the optimised values of MU2 and MV2 of the previous set of iterations.

6) The cycle continues until the values of , N and b are invariant up to five decimal points.

The results are summarized below:

To find  we use eq9, where

is taken from the 12th iteration set in the table above, while the double integral in eq32 is evaluated using form 2 of eq25. We being by computing the inner integral as per step 3, using and . Next, we apply the Simpson rule a second time for the outer integral. The Excel table is as follows:

The formula for cell NA4 is =(4.838127^2)*(B4^2)*EXP(-1.802042*2*B4))*MQ4. With reference to form 2 of eq25, the formula in NC4 is =NA4*NB4*7/350/3, representing the double integral for . The formulae for the remaining cells within NA4:NB354 are populated accordingly. Summing NC4:NC354, we obtain the total value of the double integral of 1.1386946, which when substituted in eq32 together with gives . This theoretical value has a deviation of about 1.91% versus the experiment data of the ground state of helium.

 

Next article: The angle between two vectors in spherical coordinates
Previous article: Slater-type orbitals
Content page of computational chemistry
Content page of advanced chemistry
Main content page

Slater-type orbitals

Slater-type orbitals (STO) are mathematical functions that resemble hydrogenic wavefunctions. Introduced by John Slater in 1930, they are used as trial wavefunctions in computational chemistry to approximate the energies of molecular systems. The general form of an STO is:

where

N is a normalisation constant

n, and are quantum numbers

is the distance between an electron and the nucleus of an atom.

is the effective charge of the nucleus

represents the spherical harmonics

STOs with different n but the same and are not orthogonal to one another. When working with such STOs, the Gram-Schmidt process is used to convert the non-orthogonal orbitals to orthogonal ones.

Question

Show that the normalisation constant of a 1s orbital is .

Answer


Using the identity , we have .

 

Question

Using the identity (see this article for proof), show that the 1s orbitals are orthogonal to each other but not orthogonal to the 2s orbital.

Answer


For the orbitals to be orthogonal to each other, either or needs to be zero, which is forbidden because the effective nuclear charge is not zero. Hence, the orbitals are not orthogonal to each other.

 

Next article: Numerical solution of the Hartree self-consistent method for He
Previous article: Simpson’s rule
Content page of computational chemistry
Content page of advanced chemistry
Main content page

Simpson’s rule

Simpson’s rule is a numerical method that uses quadratic functions to approximate definite integrals.

With reference to the diagram above, is the area under the cubic curve demarcated by two equal intervals, each of magnitude h. The area is estimated using 3 points on the quadratic function , where

Substitute , and in , we have , and . Combining the last two equations gives . Substituting this equation and in eq23 yields

Similarly, if we consider the domain from h to h‘ (see above diagram),

Therefore, using 5 points on , we have

Extending the logic for n points on , we have,


where and .

This implies that there must be an even number of equal intervals for the rule to work, i.e. n must be odd.

 

Question

Evaluate

Answer

Substituting eq24 in the above equation,

So,

where (form 1)

or equivalently (form 2),

These integrals are useful for computing the numerical solution of the Hartree self-consistent field method for helium.

 

Next article: Slater-type orbitals
Previous article: Proof of the Hartree equations
Content page of computational chemistry
Content page of advanced chemistry
Main content page

Lagrange method of undetermined multipliers

The Lagrange method of undetermined multipliers is a technique for finding the maximum or minimum value of a function subject to one or more constraints.

Consider the function . The total differential of the function is

Since x and y are independent variables (see below diagram), the maximum value of the function occurs at the point where and , i.e.

In general, the extremum of a function with i-independent variables is at the point where all .

If we want to find the extremum of subject to a constraint, e.g. , we will realise that one of the variables, e.g. y, in is no longer independent. This is because we can substitute the constraint in the function to eliminate the dependent variable y and then let the 1st derivative of the function with respect to the remaining independent variable x be zero. However, this simple method of substitution becomes complicated when the problem involves more than two variables and more than one constraint.

In 1804, Joseph-Louis Lagrange devised another procedure to handle such problems. For the above example, the procedure involves transforming the constraint into the function , where . The total differential of g is

where and .

Since , and . This means that we can multiply throughout by a factor and add the result to (or subtract from) eq10 to obtain the total differential of a new function:

where .

The extremum of this function occurs when

If there is a , where , then it renders the dependent variable term , and we are left with the independent term . Solving the equations , and the constraint simultaneously, we can determine x, y and . This procedure is called the Lagrange method of undetermined multipliers. The factor is the multiplier, which is undetermined because it need not be solved to find the extremum of the function.

To evaluate the extremum of the n-independent variables function subject to m constraints, where , eq11 becomes

where .

If we can find a set of values of that renders all the m dependent variable terms zero, we have m equations of

leaving us with independent variable terms and equations of

This implies that each in eq12 can now vary arbitrarily. The values of all the variables and multipliers corresponding to the extremum are finally obtained by solving eq13, eq14 and all the constraint equations simultaneously. The Lagrange method of undetermined multipliers is used in the derivations of the Boltzmann distribution, the Hartree self-consistent field method, the Hartree-Fock method and the Hartree-Fock-Roothaan method.

 

Next article: Functional variation
Previous article: the hartree self-consistent field method
Content page of computational chemistry
Content page of advanced chemistry
Main content page

The Hartree self-consistent field method

The Hartree self-consistent field method is an iterative procedure that optimises an approximate wavefunction using the variational principle. Its goal is to estimate the eigenvalues of a modified form of the non-relativistic multi-electron Hamiltonian. This method was developed by the English scientist Douglas Hartree in 1927.

In an earlier article, we showed that the three dimensional Hamiltonian operator for an n-electron atom (excluding spin-orbit and other interactions) is:

where , and .

Our objective now is to solve eigenvalue equations for . The eigenvalue equation of the hydrogen atom has an exact solution (i.e. a solution represented by a formula). However, for an atom with more than one electron, the inter-electronic repulsion term poses a problem, which is a three-body problem that has no exact solution. Therefore, approximation methods are needed.

Firstly, we simplify the Hamiltonian by rewriting it in a system devised by Hartree called “atomic units” (abbreviated to a.u.), where ( is the Bohr radius, where ). For example, if the charge of a particle is 3.4 a.u., it is in S.I. units. Eq1 therefore becomes

with the eigenvalues of also in atomic units.

Secondly, we approximate the form of a multi-electron wavefunction. As hydrogenic wavefunctions have exact solutions, it is logical to express the multi-electron wavefunction as products of orthonormal one-electron spatial wavefunctions :

Thirdly, we utilise the variational principle, which states that the calculated energy is always greater than or equal to the ground state of any system. This implies that we can calculate an upper bound to of any system using a trial function . The closer the trial function is to the actual or best wavefunction of the system , the closer E will be to .

Using eq3 and eq2, we have

Simplifying the above equation and minimising E (see this article for proof), we obtain the following set of n ‘one-particle’ equations called the Hartree equations:

where .

Question

How do we interpret eq5?

Answer

Eq5 can be perceived as a set of n Schrodinger equations, each consisting of a modified one-electron non-relativistic Hamiltonian. Solving each of the n equations gives a minimised value of corresponding to the i-th electron.

To further explain the modified inter-electronic repulsion term, we refer to eq2. Focussing on electron 1, we have

We regard electrons 2, 3, …, n as being smeared out into a continuous and spherically symmetrical charge distribution through which electron 1 moves. For example, the interaction between electron 1 and electron 2 is

where is the charge density of (i.e. charge per unit volume) of electron 2.

Since , and , we have and hence . So,

The total potential energy between electron 1 and all other electrons is therefore

 

Since eq5 is the outcome of minimising eq4, subject to the constraints of the orthonormality of one-electron wavefunctions that make up the overall wavefunction of a system, any solution set of from eq5 is associated with a set of , each of which is also obtained from multiplying eq5 by and integrating over all space:

where and .

Summing all of eq6, we have

Since for the Coulomb term (see this article for explanation),

To establish the relation between E and , we rewrite eq4 as

where we have used the fact that is equivalent to (see this article for explanation).

Combining eq7 and eq8,

E is greater than the ground state of the system, except when the solution set is associated with the set of one-electron wavefunctions that best describes the ground state of the system, in which case E is the best estimate of . This best set of and the corresponding set of are found using the Hartree self-consistent field method, which involves the following steps:

  • Make guesses for all single-electron wavefunctions except for one, e.g. use trial functions for except . The corresponding equation to solve is:

Solving the above equation means finding and , which can only be done numerically because of the presence of unknown functions and variables (see this article for an example). The solutions are then the first approximations of and , which leads to a first improvement to eq3.

  • To further improve the wavefunction of eq3, we solve for and using

  • As per the previous step, the same trial functions for are employed together with the computed . The solutions are then the 1st approximation of and , and hence a further improvement to eq3.
  • Next, and derived from the previous two steps are used along with trial functions for to solve for and . The process is repeated until and are solved. The entire cycle is then repeated, i.e. step 1 is repeated using improved wavefunctions of to solve for an even better approximation of , and then followed by step 2 and step 3. This cyclical process stops when all and become invariant, which implies that the best approximation of eq3 for the particular system has been derived.
  • Finally, E is computed with eq9 using the derived and .

As the numerical method of calculating E is usually quite complex, an analytical method is often used. Such a method 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 E and . The process is repeated until E and all become invariant.

In conclusion, the Hartree self-consistent field method produces relatively accurate results when compared to experiment data for the ground state energy of a small atom like helium. However, it is less accurate for finding E of larger atoms, as it does not take into consideration exchange forces due to spin exchange interactions. A more accurate method, call the Hartree-Fock method is needed for such atoms.

 

Next article: Lagrange method of undetermined multipliers
Content page of computational chemistry
Content page of advanced chemistry
Main content page

 

The variational method

The variational method is a mathematical technique for approximating the energy state of a system, most often the ground state of a multi-electron system.

To illustrate the method, we begin with the expectation value for the ground state energy of a system, E_0:

E_0=\frac{\int\psi_0^*\hat{H}\psi_0d\tau}{\int\psi_0^*\psi_0d\tau}\; \; \; \; \; \; \; \; 293

where \psi_0 is the ground state wavefunction of the system.

 

Question

Derive eq293.

Answer

Eq293 is obtained by multiplying the Schrodinger equation from the left by \psi_0^* and integrating over all space.

 

As mentioned in an earlier article, E_0 is, by definition, the lowest energy value of the system. If we replace \psi_0 (the mathematical description of the ground state energy of the system) with any wave function \psi that describes any energy state of the system,

E_{\psi}=\frac{\int\psi^*\hat{H}\psi d\tau}{\int\psi^*\psi d\tau}\geq E_0\; \; \; \; \; \; \; \; 294

where E_{\psi}, E_0\leftrightarrow \psi,\psi_0.

If \psi is normalised,

E_{\psi}=\int\psi^*\hat{H}\psi d\tau\geq E_0\; \; \; \; \; \; \; \; 295

Eq295 is called the variational principle, i.e. the principle behind the variational method. It is evident that eq295 is a functional: a function E_{\psi} of a function \psi. Therefore, to estimate the ground state energy of a system, we substitute an appropriate trial wave function that depends on one or more arbitrary parameters (variational parameters) into eq295, and minimise E_{\psi} with respect to the parameters.

For example, we can estimate the ground state energy of a hydrogen atom using the trial wavefunction \psi=e^{-\alpha r}, where \alpha is the variational parameter. Substituting \psi=e^{-\alpha r}, \hat{H}=-\frac{\hbar^2}{2m_er^2}\frac{d}{dr}\left ( r^2\frac{d}{dr} \right )-\frac{e^2}{4\pi\varepsilon_0r} in eq295, and using the identity \int_0^{\infty}x^ne^{-a x}dx=\frac{n!}{a^{n+1}} (see this article for proof), we have

E_{\psi}=\frac{\hbar^2\alpha^2}{2m_e}-\frac{e^2\alpha}{4\pi\varepsilon_0}\; \; \; \; \; \; \; \; 296

Clearly, eq296 describes a parabola with a minimum value. Setting \frac{dE_{\psi}}{d\alpha}=0, solving for \alpha, and substituting the expression for \alpha back into eq296, we have

E_{est}=-\frac{m_ee^4}{32\pi^2\varepsilon_0^{\;2}\hbar^2}

where E_{est} is the estimated ground state energy of hydrogen.

In other words, we vary the variational parameter to arrive at a good approximation of E_{\psi}.

To determine the ground state energy of a multi-electron system, the variational method is used together with the Hartree self-consistent method, which will be explained in subsequent articles.

 

Question

Provide a proof of the variational principle.

Answer

Since any well-behaved wavefunction can be expressed as a linear combination of a complete orthonormal set of basis wavefunctions, we substitute \psi=\sum_{n=0}^{\infty}c_n\phi_n in eq295.

E_{\psi}=\sum_{m=0}^{\infty}\sum_{n=0}^{\infty}\int c_m^*\phi_m^*\hat{H}c_n\phi_nd\tau=\sum_{m=0}^{\infty}\sum_{n=0}^{\infty}c_m^*c_nE_n\int\phi_m^*\phi_nd\tau=\sum_{m=0}^{\infty}\vert c_m\vert^2E_m

Since \sum_{m=0}^{\infty}\vert c_m\vert^2=1, we have E_0=\sum_{m=0}^{\infty}\vert c_m\vert^2E_0. So,

E_{\psi}-E_0=\sum_{m=0}^{\infty}\vert c_m\vert^2E_m -\sum_{m=0}^{\infty}\vert c_m\vert^2E_0=\sum_{m=0}^{\infty}\vert c_m\vert^2(E_m-E_0)

E_0 is, by definition, the lowest energy value of the system. So, \sum_{m=0}^{\infty}\vert c_m\vert^2(E_m-E_0) \geq 0 and is zero if and only if E_m=E_0. Therefore, E_{\psi}\geq E_0.

 

 

Next article: Overlap integral
Previous article: Hund’s rules
Content page of quantum mechanics
Content page of advanced chemistry
Main content page

Hund’s rules

Hund’s rules, created by Friedrich Hund in 1927, consist of three principles used to identify the lowest energy term symbol of a given atomic configuration. They are:

  • The term symbol with the highest multiplicity represents state(s) with the lowest energy.

This is due to the spin correlation between spin angular momenta of electrons, where the expectation value of the distance between a pair of electrons in the triplet state is greater than the expectation value of the distance between a pair of electrons in the singlet state. Consequently, there is less electron-electron repulsion for the triplet state (higher multiplicity) than for the singlet state.

  • For a particular multiplicity, the term symbol with the highest L is lowest in energy.

This results from the Coulomb interaction between orbital angular momentum of a pair of electrons, where electrons orbiting in the same direction meet less often than when they are orbiting in opposite directions. This implies that \vert\boldsymbol{\mathit{r}}_i-\boldsymbol{\mathit{r}}_j\vert_{ave}(same)>\vert\boldsymbol{\mathit{r}}_i-\boldsymbol{\mathit{r}}_j\vert_{ave}(opposite). Since the individual angular momentum of each electron orbiting in the same direction adds vectorially to give a higher total angular momentum (and hence a higher value of the quantum number L) than electrons orbiting in opposite directions, the term symbol with the highest L is lowest in energy.

  • Generally, for a particular term describing an atom with a less than half-filled outermost subshell, the term symbol with the lowest J is lowest in energy. If the subshell is more than half-filled, the term symbol with the highest J is lowest in energy.

This is because the value of A in eq289, which is empirically evaluated, generally changes from positive to negative when the subshell is more than half-filled.

 

Question

What about a half-filled subshell?

Answer

Examples of electron configuration of half-filled subshells are p3 and d5, with terms 4S and 6S respectively and corresponding levels of 4S3/2 and 6S5/2 respectively. Since the level of a half-filled subshell is described by only one term symbol, which always has the highest multiplicity of all microstates of the associated electron configuration, it is always lowest in energy. In other words, no rules are needed to determine the lowest energy term symbol of electron configurations like p3 and d5.

 

Since term symbols are based on Russell-Saunders coupling, where we assume that spin-spin coupling > orbit-orbit coupling > spin-orbit coupling, the three rules are applied in the above order (from rule 1 to rule 3) when we determine the lowest energy term symbol of a given configuration of an atom. The ground states of carbon and nitrogen are therefore 3P0 and 4S3/2 respectively.

Hund’s rules are often stated as a single rule at introductory courses, with the rule being: electrons occupy the orbitals of a subshell singly and in parallel before pairing up. This statement is another way to state the first rule, which is the most dominant. Using the example of the n=2 shell of the ground state of carbon, the two electrons in the 2p orbitals are parallel because of the first rule.

 

Question

Why are the two electrons in the 2s orbital anti-parallel?

Answer

This is also a consequence of the first rule, where the spatial wavefunction associated with a triplet state is \psi=\frac{1}{\sqrt{2}}[\psi_a(r_1)\psi_b(r_2)-\psi_b(r_1)\psi_a(r_2)]. If r_1=r_2,  and \psi=0 and \int\vert\psi\vert^2d\tau=0, which implies that there is zero probability of finding the two parallel electrons at the same point.

 

 

Next article: The variational method
Previous article: Constructing atomic terms (coupled representation)
Content page of quantum mechanics
Content page of advanced chemistry
Main content page

Constructing atomic terms (coupled representation)

The construction of term symbols in the spin-orbit coupled representation (i.e. levels) is an extension of the procedure for deriving term symbols in the spin-orbit uncoupled representation (i,e, terms).

For the p2 configuration of the carbon atom, we have shown in the previous article that the terms are 1D, 3P and 1S with degeneracies (2L+1)(2S+1) of 5, 9 and 1 respectively. The total degeneracy of 15 corresponds to 15 microstates in the uncoupled representation, which are the basis states for the coupled representation. Since the total number of microstates describing a system is independent of the chosen representation, we have 15 coupled states after linearly combining the uncoupled basis states. The procedure to construct levels is to use the Clebsch-Gordan series, which describes the allow values of the total angular momentum number J for a given value of L and a given value of S, while ensuring that the total number of microstates (degeneracy) of each term (e.g. 3P) is equal to the total number of microstates of the associated levels (e.g. 3P2, 3P1, 3P0), i.e. (2L+1)(2S+1)=\sum_{allowed\: J}(2J+1)_{allowed\: J}.

 

Term

Degeneracy \boldsymbol{\mathit{J}}=\boldsymbol{\mathit{L}}+\boldsymbol{\mathit{S}} Level

Degeneracy

1D

5 2 = 2 + 0 1D2

5

3P

9

2 = 1 + 1 3P2 5
3P1

3

3P0

1
1S 1 0 = 0 + 0 1S0

1

 

For the p3 configuration of nitrogen, the terms are 4S, 2D, 2P and the levels are:

Term

Degeneracy \boldsymbol{\mathit{J}}=\boldsymbol{\mathit{L}}+\boldsymbol{\mathit{S}} Level Degeneracy
4S 4 3/2 = 0 + 3/2 4S3/2

4

2D

10 5/2 = 2 + 1/2 2D5/2 6
2D3/2

4

2P

6 3/2 = 1 + 1/2 2P3/2 4
2P1/2

2

You’ll notice that the degeneracies for both carbon and nitrogen are not fully lifted (i.e. each level may contain more than one state) when spin-orbit interactions are considered. Other effects, like an external magnetic field, are required to completed lift the degeneracy of an atom.

 

Next article: Hund’s rules
Previous article: Constructing atomic terms (uncoupled representation)
Content page of quantum mechanics
Content page of advanced chemistry
Main content page

Constructing atomic terms (uncoupled representation)

Consider the coupling of two electrons. From eq205, we have

\hat{J}=\hat{L}\otimes I+I\otimes\hat{S}

where \hat{L}=\hat{l}^{(1)}\otimes I+I\otimes\hat{l}^{(2)} and \hat{S}=\hat{s}^{(1)}\otimes I+I\otimes\hat{s}^{(2)}.

This implies that the state corresponding to \hat{L} is a coupled representation of \hat{l}^{(1)} and \hat{l}^{(2)} (orbit-orbit coupling), and that the state corresponding to \hat{S} is a coupled representation of \hat{s}^{(1)} and \hat{s}^{(2)} (spin-spin coupling). The overall state can be either a spin-orbit coupled representation \vert J,M_J,L,S\rangle, or a spin-orbit uncoupled representation \vert L,M_L,S,M_S\rangle. As mentioned in a previous article, if we neglect spin-orbit coupling, we can construct atomic terms using the uncoupled representation.

If the two electrons are in the same open subshell, they are called equivalent electrons. An example of such a system is the carbon atom. Even though carbon has six electrons, four of them are in closed shells with zero angular momentum and are therefore ignored when determining atomic terms.

The first step is to tabulate all possible microstates of p2 (arrangements of p2 electrons) that do not violate the Pauli exclusion principle:

Groups

\boldsymbol{\mathit{m}}_l \boldsymbol{\mathit{M}}_L \boldsymbol{\mathit{M}}_S

+1

0

-1

All up

u

u 1

1

u

u 0 1

u

u -1

1

All down d d 1

-1

d d 0

-1

d d -1 -1

One up, one down

ud

2 0

ud 0

0

ud -2

0

u d 1

0

u d 0

0

u d -1

0

d

u 1

0

d

u 0

0

d u -1

0

where u and d represent m_s=+\frac{1}{2} and m_s=-\frac{1}{2} respectively.

The above table is re-organised as:

\boldsymbol{\mathit{M}}_S

+1

0

-1

\boldsymbol{\mathit{M}}_L

+2

1

+1

1 2

1

0

1 3

1

-1

1 2

1

-2

1

where each green number represents the number of microstates corresponding to each (M_L,M_S) combination.

The only microstate with M_L=+2 and M_S=0 when expressed in the uncoupled representation of \vert L,M_L,S,M_S\rangle is \vert 2,2,0,0\rangle because L=M_{L,max}. It must belong to the term 1D, since 1D is when L=2 and S=0, with degenerate states of \vert 2,2,0,0\rangle, \vert 2,1,0,0\rangle, \vert 2,0,0,0\rangle, \vert 2,-1,0,0\rangle  and \vert 2,-2,0,0\rangle. For accounting purposes, we refresh the above table by removing these 5 states of 1D, giving:

\boldsymbol{\mathit{M}}_S

+1

0

-1

\boldsymbol{\mathit{M}}_L

+2

+1

1 1

1

0

1 2

1

-1

1 1

1

-2

Similarly, the only state with M_L=+1,M_S=+1 must be one of the 9 degenerate states of 3P. Again, we refresh the above table by removing these 9 states of 3P, leaving behind one state (M_L=0,M_S=0), which corresponds to the term 1S. Therefore, the atomic terms for the p2 configuration of carbon are 1D, 3P and 1S.

 

Question

Why are the 9 degenerate states of 3P, combinations of M_L=+1,0,-1 and M_S=+1,0,-1?

Answer

Recall that states with same L and same S have the same energy and are grouped into a term. For the term 3P, L=1 and S=1 with degeneracy 9 (3P2 has 5, 3P1 has 3 and 3P0 has 1). For S=1, we have three spin-coupled basis vectors associated with the quantum numbers M_S=+1,0,-1 (formed by the coupling of s_1 and s_2). For L=1, we have another three orbit-coupled basis vectors associated with the quantum numbers M_L=+1,0,-1 (formed by the coupling of l_1 and l_2). The total uncoupled microstates \vert L,M_L,S,M_S\rangle are the number of ways to form Kronecker products of basis vectors from the two vector spaces and hence combinations of M_L=+1,0,-1M_L=+1,0,-1 and M_S=+1,0,-1.

 

For the configuration 1s22s22p3, e.g. nitrogen, we have

Groups \boldsymbol{\mathit{m}}_l \boldsymbol{\mathit{M}}_L \boldsymbol{\mathit{M}}_S
+1 0 -1
All up u u u 0 3/2
All down d d d 0 -3/2
One up two down ud d 2 -1/2
ud d 1 -1/2
d ud 1 -1/2
ud d -1 -1/2
d ud -1 -1/2
d ud -2 -1/2
u d d 0 -1/2
d u d 0 -1/2
d d u 0 -1/2
Two up one down ud u 2 1/2
ud u 1 1/2
u ud 1 1/2
ud u -1 1/2
u ud -1 1/2
u ud -2 1/2
u u d 0 1/2
u d u 0 1/2
d u u 0 1/2

We can re-organise the above table as:

\boldsymbol{\mathit{M}}_S
+3/2 1/2 -1/2 -3/2
\boldsymbol{\mathit{M}}_L +2 1 1
+1 2 2
0 1 3 3 1
-1 2 2
-2 1 1

The only state with M_L=0,M_S=+3/2 must be one of the 4 degenerate states of 4S. For accounting purposes, we re-tabulate the above table by removing these 4 states of 4S, giving:

\boldsymbol{\mathit{M}}_S
+3/2 1/2 -1/2 -3/2
\boldsymbol{\mathit{M}}_L +2 1 1
+1 2 2
0 2 2
-1 2 2
-2 1 1

A good guess for the next term is 2D with a degeneracy of 10, because we are left with states of M_S=\pm1/2, spanning M_L from +2 to -2. The remaining states are:

\boldsymbol{\mathit{M}}_S
+3/2 1/2 -1/2 -3/2
\boldsymbol{\mathit{M}}_L +2
+1 1 1
0 1 1
-1 1 1
-2

which obviously belong to the term 2P.

Therefore, the terms for the configuration of nitrogen are 4S, 2D, 2P.

For a system with non-equivalent electrons (electrons in different subshells) in open shells, the way to construct atomic terms is similar to a system with equivalent electrons, except that we do not have to apply the Pauli exclusion principle in the construction process. The easiest method of construction is to use the Clebsch-Gordan series by coupling the l_i values to find out the possible L values, and then coupling the s_i values to give the possible S values, and finally combining the possible L and S values to produce the atomic terms. For example, the terms for the 2p13p1 system are 3D, 1D, 3P, 1P, 3S and 1S.

 

Next article: Constructing atomic terms (coupled representation)
Previous article: Term symbols
Content page of quantum mechanics
Content page of advanced chemistry
Main content page

Term symbols (quantum mechanics)

Term symbols are notations representing energy levels of a particular electron configuration of an atom.

Based on Russell Saunders coupling, a term symbol can be expressed either as:

^{2S+1}L\; \; \; \; \; \; \; \; 291

or

^{2S+1}L_J\; \; \; \; \; \; \; \; 292

where 2S+1 is called the multiplicity of the term symbol, S is the total spin angular momentum quantum number, L is the total orbital angular momentum quantum number and J is the total angular momentum quantum number.

Each value of L is further assigned a letter:

L 0 1 2 3 4 5 6 7 8
Letter S P D F G H I K L

Energy levels denoted by eq291 ignore spin-orbit coupling and are called terms, while those denoted by eq292 take into account spin-orbit coupling and are known as levels. Eq291 implies that for a particular electron configuration of an atom, states with the same set of angular momentum quantum numbers L and S have the same energy. We have shown in an earlier article how this is possible. When spin-orbit coupling effects are considered, degenerate states corresponding to a term are split into multiple levels. From eq248 and eq289, energy levels of a particular electron configuration of an atom are dependent on the quantum numbers J, L and S.

 

Question

Write the term symbols for i) the ground state of the hydrogen atom, and ii) the excited state of the hydrogen atom with the electron in the 2p orbital.

Answer

For the ground state of hydrogen, where S=s=\frac{1}{2} and L=0\equiv S, it is denoted by ^2S_{\frac{1}{2}}. For the excited state, the electron can be in any of the three degenerate 2p orbitals with either an up spin or a down spin. We would therefore expect two term symbols. The first symbol is ^2P_{\frac{3}{2}}, while the second symbol is obtained using the Clebsch-Gordan series of eq219 and is ^2P_{\frac{1}{2}}.

 

Question

Why is a term symbol not described by the principal quantum number ?

Answer

Term symbols focus on the variation of energy resulting from the arrangement and interaction of electrons in subshells. These arrangements, known as microstates, depend on whether the electrons are equivalent (belonging to the same subshell) or non-equivalent (belonging to different subshells). For example, the electrons in the configuration are equivalent, while those in are not. The term symbols for and are and respectively. Although the term symbols for and  are the same, the energies of the terms for , which are determined spectroscopically, differ from those for . Therefore, while the energy associated with does influence the overall energy of the electron within the atom, it doesn’t directly contribute to the variation in energies of electrons within the subshells.

 

Next article: Constructing atomic terms (uncoupled representation)
Previous article: russell saunders coupling
Content page of quantum mechanics
Content page of advanced chemistry
Main content page
Mono Quiz