You are on page 1of 3

Ground state term symbol for oxygen atom

Ground state term symbol. The term symbol for the electronic ground state of oxygen atom is. O2 ground state term symbol. Nitrogen ground state term symbol.

Around 1930, several spectroscopists using high resolution instruments found that lines in the hydrogen atom spectrum actually are not single lines but they are multiplets as shown for an isotopic mixture of hydrogen,\((H^1_{\alpha})\) and deuterium, (\(H^2_{\alpha}\)), in Figure \(\PageIndex{1}\). A multiplet consists of two or more closely spaced
lines. Two lines together form a doublet, three a triplet, etc. Multiplets also are called fine structure.

The term fine structure means the lines are spaced close together, i.e. finely spaced. Such fine structure also was found in spectra of one-electron ions such as \(\ce{He^{+}}\).You should recall that the \(H^1_{\alpha}\) line in the Balmer series at 656.279 nm was understood as resulting from a single transition of an electron from the n = 3 energy
level to the n = 2 level. The observation of fine structure revealed that an orbital energy level diagram does not completely describe the energy levels of atoms. This fine structure also provided key evidence at the time for the existence of electron spin, which was used not only to give a qualitative explanation for the multiplets but also to furnish
highly accurate calculations of the multiplet splittings. Specifying the orbital configuration of an atom does not uniquely identify the electronic state of the atom because the orbital angular momentum, the spin angular momentum, and the total angular momentum are not precisely specified. For example in the hydrogen 2p1 configuration, the electron
can be in any of the three p-orbitals, \(m_l\) = +1, 0, and –1, and have spins with \(m_s\) = +1/2 or –1/2. Thus, there are 3 times 2 different possibilities or states. Also, the orbital and spin angular momentum of the electrons combine in multiple ways to produce angular momentum vectors that are characteristic of the entire atom not just individual
electrons, and these different combinations can have different energies. This coupling of orbital and spin angular momentum occurs because both the electron spin and orbital motion produce magnetic dipole moments. As we have seen previously, the relationship between the angular momentum and the magnetic moment is given by the
gyromagnetic ratio. These magnetic dipoles interact just like two tiny bar magnets attracting and repelling each other.

This interaction is called spin-orbit interaction. The interaction energy is proportional to the scalar product of the magnetic dipole moments, which are proportional to the angular momentum vectors. \[E_{s-o} \propto S \cdot L \label{8.11.1}\] with the following terms added to the Hamiltonian \[\hat {H} _{s-o} \propto \hat {S} \cdot \hat {L}
\label{8.11.2}\] where the constant of proportionality is called the spin-orbit coupling constant. The spin-orbit interaction couples the spin motion and orbital motion of all the electrons together. This coupling means that exact wavefunctions are not eigenfunctions of the spin and orbital angular momentum operators separately.
Rather the total angular momentum \(J = L+S\), the vector sum of the spin and orbital angular momentum, is required to be coupled for a completely accurate description of the system. Trying to describe the coupled system in terms of spin and orbital angular momentum separately is analogous to trying to describe the positions of two coupled bar
magnets independently. It cannot be done; their interaction must be taken into account (Figure \(\PageIndex{1}\)). Figure \(\PageIndex{1}\): An electron that orbits the nucleus induces a "current" and a magnetic field. Similarly, the rotating electron (really intrinsic spin) of the electron induces a different magnet. Both a magnetic (dipole) moments
will interact with each other via spin-orbit coupling to generate fine structure atomic spectra.(CC-SA OpenStax). Higher energy or excited orbital configurations also exist. The hydrogen atom can absorb energy, and the electron can be promoted to a higher energy orbital. The electronic states that result from these excited orbital configurations also
are characterized or labeled by term symbols. The details of how to determine the term symbols for multi-electron atoms and for cases where both the orbital and spin angular momentum differ from zero are given elsewhere, along with rules for determining the relative energies of the terms. We have found that the selection rules for promoting a
single electron moving from one atomic orbital to another via the absorption or emission of light are \[ \Delta l = \pm 1 \label {8.11.6}\] \[ \Delta m_l = 0, \pm 1 \label{8.11.7}\] These selection rules arise from the conservation fo angular momentum during a spectroscopic transition and the fact that a photon has a spin 1. Within the limits of L-S
coupling, these rules can be expressed in terms of atomic term symbols resulting in the resulting Russell-Saunders selection rules: \[ \begin{align} \Delta S &= 0 \label {8.11.8} \\[4pt] \Delta L &= 0, \pm 1 \label {8.11.9} \\[4pt] \Delta J &= 0, \pm 1,\label {8.11.10} \end{align} \] but the \(J =0\) to \(J= 0\) transition is forbidden \[\Delta m_J = 0, \pm 1
\label {8.11.11}\] but the \(m_J = 0\) to \(m_J = 0\) transition is forbidden if \(\Delta J = 0\).

These selection rules result from the general properties of angular momentum such as the conservation of angular momentum and commutation relations. The \(\Delta L =0\) option in Equation \(\ref{8.11.9}\) does not violate the conservation of angular momentum discussed previously, since \(\Delta l = \pm 1\) is still required. The orbital angular
momentum of an electron must change upon absorption, but this does not necessarily affect the overall momentum of the state given by Equation \(\ref{8.11.9}\). The selection rules apply only to atoms that can be described with Russell-Saunders (LS) coupling.

These rules fail as the atomic number increases because the \(S\) and \(L\) quantum numbers become "bad" quantum numbers; this occurs when the jj-coupling coupling approach is more applicable.
For example, the transition between single (\(S=1/2\) and triplet \(S=1\) states (volition of selection rule in Equation \(\ref{8.11.8}\)) are allowed and experimentally observed, in heavy atoms. Example \(\PageIndex{1}\): Sodium Atoms An example of this fine structure is the emission of sodium atoms. Figure above does not show this splitting). 616.07
nm 615.42 nm 589.00 nm 589.59 nm 568.82 nm 568.26 nm How can these transitions be described in terms of transitions between microstates Solution We need to discussed states in terms of not only electron configurations, but in terms of microstates (i.e., term symbols) and the principal quantum number of the valence electron, \(n\): The ground
state has a \((Ne]ns^1\) configuration, which has only one microstate \(^2S_{1/2}\) The excited state with the valence electron in the p-orbitals has an electron configuration of \([Ne]np^1\), which has two microstates: \(^2P_{3/2}\) and \(^2P_{1/2}\) The excited state with the valence electron in the p-orbitals has an electron configuration of \
([Ne]nd^1\), which has two microstates of \(^2D_{5/2}\) and \(^2D_{3/2}\) Observed lines can be explained: \(5S \rightarrow 3P\) gives two lines since the initial configuration has two microstates: 616.07, 615.42 nm \(3P \rightarrow 3S\) gives two lines since the terminal configuration has two microstates: 589.00, 589.59 nm \(4D \rightarrow 3P\)
gives two lines since the terminal configuration has two microstates: 568.82, 568.26 nm Splitting of the Sodium D Line One notable atomic spectral line of sodium vapor is the so-called D-line, which may be observed directly as the sodium flame-test line and also the major light output of low -pressure sodium lamps (these produce pressure sodium
lamps (these produce an unnatural yellow).
The D-line is one of the classified Fraunhofer lines in Sodium vapor in the upper layers of lines.
Sodium vapor in the upper layers of the sun creates a dark line in the emitted spectrum of electromagnetic radiation by absorbing visible light in a band of wavelengths around 589.5 nm. This wavelength corresponds to transitions in atomic sodium in which the valence-electron transitions from a 3s to 3p electronic state. Closer examination of the
visible spectrum of atomic sodium reveals that the D-line actually consists of two lines called the \(D_1\) and \(D_2\) lines at 589 6 nm and 589.0 nm, respectively. The splitting between these lines arises because of spin-orbit coupling. Na has one unpaired electron (\(S = ½\)). If we consider the \(S \rightarrow P\) transition, then for the excited state, \
(P\), we have \(L = 1\).
Thus, \(J = 3/2\) or \(J=1/2\). Now we want to apply these ideas to understand why multiplet structure is found in the luminescence spectrum of hydrogen and single electron ions. As we have said, the \(H_{\alpha}\) line in the Balmer series at 656.279 nm can be understood via a transition of an electron in a n = 3 atomic orbital to a n = 2 atomic
orbital. When this spectral line was examined using high-resolution instruments, it was found actually to be a doublet, i.e. two lines separated by 0.326 cm-1.
There are 9 degenerate orbitals associated with the n = 3 level, and 4 associated with the n = 2 level. Since an electron can be in any orbital with any one of two spins, we expect the total number of states to be twice the number of orbitals. The number of orbitals is given by \(n^2\) so there should be 8 states associated with n = 2 and 18 states
associated with n = 3. Using the ideas of vector addition of angular momentum, the terms that result from having an electron in any one of these orbitals are given in Table \(\PageIndex{1}\). Orbital Configuration Term Symbols Degeneracy Table \(\PageIndex{1}\): H Atom Terms Originating from n = 1, 2, and 3 1s1 \(^2S_{1/2}\) 2 2s1 \(^2S_{1/2}\)
2 2p1 \(^2P_{1/2}\), \(^2P_{3/2}\) 2, 4 3s1 \(^2S_{1/2}\) 2 3p1 \(^2P_{1/2}\), \(^2P_{3/2}\) 2, 4 3d1 \(^2D_{3/2}\), \(^2D_{5/2}\) 4, 6 Table \(\PageIndex{1}\) shows that there are three terms associated with n = 2, and 5 terms associated with n = 3. In principle, each term can have a different energy. The degeneracy of each term is determined by
the number of projections that the total angular momentum vector has on the z-axis. These projections depend on the \(m_J\) quantum number, which ranges from \(+J\) to \(–J\) in integer steps. J is the total angular momentum quantum number, which is given by the subscript in the term symbol. This relationship between \(m_J\) and J (\(m_J\) varies
from \(+J\) to \(–J\) in integer steps) is true for any angular momentum vector. Exercise \(\PageIndex{3}\) Confirm that the nine term symbols in Table \(\PageIndex{1}\) are correct. Exercise \(\PageIndex{4}\) Confirm that the values for the degeneracy in Table \(\PageIndex{1}\) are correct and that the total number of states add up to 8 for n = 2 and
18 for n = 3. The energies of the terms depend upon spin-orbit coupling and relativistic corrections that need to be included in the Hamiltonian operator in order to provide a more complete description of the hydrogen atom. As a consequence of these effects, all terms with the same \(n\) and (J\) quantum numbers have the same energy, while terms
with different values for \(n\) or \(J\) have different energies. The theoretical term splittings as given by H.E. White, Introduction to Atomic Spectra (McGraw-Hill, New York, 1934) pp. 132-137 and are shown in Figure \(\PageIndex{2}\). Figure \(\PageIndex{2}\): Energy level diagram for the electronic states of hydrogen corresponding to an electron
with the principal quantum number n = 3 and n = 2. While the relative energies of the terms are drawn to scale, the energy of n = 3 relative to n = 2 is not. Transitions to the 2P3/2 and (2P1/2 and 2S1/2) levels correspond to the two lines in the \(H^1_{\alpha}\) band, which have a measured separation of 0.326 cm-1. Figure \(\PageIndex{2}\) shows 5
allowed transitions for the electron in the states associated with n = 3 to the states associated with n = 2. Of these five, two are most intense and are responsible for the doublet structure. These two transitions are indicated by the wide black lines at the bottom of the figure to correspond to the lines observed in the photographic spectrum shown in
Figure \(\PageIndex{2}\). The other transitions contribute to the width of these lines or are not observed. The theoretical value for the doublet splitting is 0.328 cm-1, which is in excellent agreement with the measured value of 0.326 cm-1. The value of 0.328 cm-1 is obtained by taking the difference, 0.364 – 0.036 cm-1,in the term splittings. As we
have just seen, the electronic states, as identified by the term symbols, are essential in understanding the spectra and energy level structure of atoms, but it also is important to associate the term symbols and states with the orbital electron configurations. The orbital configurations help us understand many of the general or coarse features of spectra
and are necessary to produce a physical picture of how the electron density changes because of a spectroscopic transition. Exercise \(\PageIndex{5}\) Use the Russell-Saunders selection rules to determine which transitions contribute to the \(H_{\alpha}\) line in the hydrogen spectrum. Notation in quantum physics In atomic physics, a term symbol is
an abbreviated description of the total spin and orbital angular momentum quantum numbers of the electrons in a multi-electron atom. So while the word symbol suggests otherwise, it represents an actual value of a physical quantity. For a given electron configuration of an atom, its state depends also on its total angular momentum, including spin
and orbital components, which are specified by the term symbol. The usual atomic term symbols assume LS coupling (also known as Russell–Saunders coupling) in which the all-electron total quantum numbers for orbital (L), spin (S) and total (J) angular momenta are good quantum numbers. In the terminology of atomic spectroscopy, L and S
together specify a term; L, S, and J specify a level; and L, S, J and the magnetic quantum number MJ specify a state. The conventional term symbol has the form 2S+1LJ, where J is written optionally in order to specify a level. L is written using spectroscopic notation: for example, it is written "S", "P", "D", or "F" to represent L = 0, 1, 2, or 3
respectively.
For coupling schemes other that LS coupling, such as the jj coupling that applies to some heavy elements, other notations are used to specify the term. Term symbols apply to both neutral and charged atoms, and to their ground and excited states. Term symbols usually specify the total for all electrons in an atom, but are sometimes used to describe
electrons in a given subshell or set of subshells, for example to describe each open subshell in an atom having more than one. The ground state term symbol for neutral atoms is described, in most cases, by Hund's rules. Neutral atoms of the chemical elements have the same term symbol for each column in the s-block and p-block elements, but differ
in d-block and f-block elements where the ground-state electron configuration changes within a column, where exceptions to Hund's rules occur.
Ground state term symbols for the chemical elements are given below.
Term symbols are also used to describe angular momentum quantum numbers for atomic nuclei and for molecules. For molecular term symbols, Greek letters are used to designate the component of orbital angular momenta along the molecular axis. The use of the word term for an atom's electronic state is based on the Rydberg–Ritz combination
principle, an empirical observation that the wavenumbers of spectral lines can be expressed as the difference of two terms. This was later summarized by the Bohr model, which identified the terms with quantized energy levels, and the spectral wavenumbers of these levels with photon energies. Tables of atomic energy levels identified by their term
symbols are available for atoms and ions in ground and excited states from the National Institute of Standards and Technology (NIST).[1] Term symbols with LS coupling The usual atomic term symbols assume LS coupling (also known as Russell–Saunders coupling), in which the atom's total spin quantum number S and the total orbital angular
momentum quantum number L are "good quantum numbers". (Russell–Saunders coupling is named after Henry Norris Russell and Frederick Albert Saunders, who described it in 1925[2]). The spin-orbit interaction then couples the total spin and orbital moments to give the total electronic angular momentum quantum number J. Atomic states are
then well described by term symbols of the form: 2 S + 1 L J {\displaystyle ^{2S+1}L_{J}} where S is the total spin quantum number for the atom's electrons. The value 2S + 1 written in the term symbol is the spin multiplicity, which is the number of possible values of the spin magnetic quantum number MS for a given spin S. J is the total angular
momentum quantum number for the atom's electrons. J has a value in the range from |L − S| to L + S. L is the total orbital quantum number in spectroscopic notation, in which the symbols for L are: L = 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 ... S P D F G H I K L M N O Q R T U V (continued alphabetically)[note 1] The orbital symbols S, P, D and F
are derived from the characteristics of the spectroscopic lines corresponding to s, p, d, and f orbitals: sharp, principal, diffuse, and fundamental; the rest are named in alphabetical order from G onwards (omitting J, S and P). When used to describe electronic states of an atom, the term symbol is often written following the electron configuration. For
example, 1s22s22p2 3P0 represents the ground state of a neutral carbon atom. The superscript 3 indicates that the spin multiplicity 2S + 1 is 3 (it is a triplet state), so S = 1; the letter "P" is spectroscopic notation for L = 1; and the subscript 0 is the value of J (in this case J = L − S).[1] Small letters refer to individual orbitals or one-electron quantum
numbers, whereas capital letters refer to many-electron states or their quantum numbers. Terminology: terms, levels, and states For a given electron configuration, The combination of an S {\displaystyle S} value and an L {\displaystyle L} value is called a term, and has a statistical weight (i.e., number of possible states) equal to ( 2 S + 1 ) ( 2 L + 1 )
{\displaystyle (2S+1)(2L+1)} ; A combination of S {\displaystyle S} , L {\displaystyle L} and J {\displaystyle J} is called a level. A given level has a statistical weight of 2 J + 1 {\displaystyle 2J+1} , which is the number of possible states associated with this level in the corresponding term; A combination of S {\displaystyle S} , L {\displaystyle L} , J
{\displaystyle J} and M J {\displaystyle M_{J}} determines a single state. The product ( 2 S + 1 ) ( 2 L + 1 ) {\displaystyle (2S+1)(2L+1)} as a number of possible states | S , M S , L , M L ⟩ {\displaystyle |S,M_{S},L,M_{L}\rangle } with given S and L is also a number of basis states in the uncoupled representation, where S {\displaystyle S} , M S
{\displaystyle M_{S}} , L {\displaystyle L} , M L {\displaystyle M_{L}} ( M S {\displaystyle M_{S}} and M L {\displaystyle M_{L}} are z-axis components of total spin and total orbital angular momentum respectively) are good quantum numbers whose corresponding operators mutually commute. With given S {\displaystyle S} and L {\displaystyle
L} , the eigenstates | S , M S , L , M L ⟩ {\displaystyle |S,M_{S},L,M_{L}\rangle } in this representation span function space of dimension ( 2 S + 1 ) ( 2 L + 1 ) {\displaystyle (2S+1)(2L+1)} , as M S = S , S − 1 , … , − S + 1 , − S {\displaystyle M_{S}=S,S-1,\dots ,-S+1,-S} and M L = L , L − 1 , . . . , − L + 1 , − L {\displaystyle M_{L}=L,L-1,...,-L+1,-
L} . In the coupled representation where total angular momentum (spin + orbital) is treated, the associated states (or eigenstates) are | J , M J , S , L ⟩ {\displaystyle |J,M_{J},S,L\rangle } and these states span the function space with dimension of ∑ J = J min = | L − S | J max = L + S ( 2 J + 1 ) {\displaystyle \sum _{J=J_{\min }=|L-S|}^{J_{\max
}=L+S}(2J+1)} as M J = J , J − 1 , … , − J + 1 , − J {\displaystyle M_{J}=J,J-1,\dots ,-J+1,-J} . Obviously, the dimension of function space in both representations must be the same. As an example, for S = 1 , L = 2 {\displaystyle S=1,L=2} , there are (2×1+1)(2×2+1) = 15 different states (= eigenstates in the uncoupled representation) corresponding
to the 3D term, of which (2×3+1) = 7 belong to the 3D3 (J = 3) level. The sum of ( 2 J + 1 ) {\displaystyle (2J+1)} for all levels in the same term equals (2S+1)(2L+1) as the dimensions of both representations must be equal as described above. In this case, J can be 1, 2, or 3, so 3 + 5 + 7 = 15. Term symbol parity The parity of a term symbol is
calculated as P = ( − 1 ) ∑ i ℓ i , {\displaystyle P=(-1)^{\sum _{i}\ell _{i}},} where ℓ i {\displaystyle \ell _{i}} is the orbital quantum number for each electron. P = 1 {\displaystyle P=1} means even parity while P = − 1 {\displaystyle P=-1} is for odd parity. In fact, only electrons in odd orbitals (with ℓ {\displaystyle \ell } odd) contribute to the total
parity: an odd number of electrons in odd orbitals (those with an odd ℓ {\displaystyle \ell } such as in p, f,...) correspond to an odd term symbol, while an even number of electrons in odd orbitals correspond to an even term symbol. The number of electrons in even orbitals is irrelevant as any sum of even numbers is even. For any closed subshell, the
number of electrons is 2 ( 2 ℓ + 1 ) {\displaystyle 2(2\ell +1)} which is even, so the summation of ℓ i {\displaystyle \ell _{i}} in closed subshells is always an even number.
The summation of quantum numbers ∑ i ℓ i {\textstyle \sum _{i}\ell _{i}} over open (unfilled) subshells of odd orbitals ( ℓ {\displaystyle \ell } odd) determines the parity of the term symbol.
If the number of electrons in this reduced summation is odd (even) then the parity is also odd (even). When it is odd, the parity of the term symbol is indicated by a superscript letter "o", otherwise it is omitted: 2Po1⁄2 has odd parity, but 3P0 has even parity. Alternatively, parity may be indicated with a subscript letter "g" or "u", standing for gerade
(German for "even") or ungerade ("odd"): 2P1⁄2,u for odd parity, and 3P0,g for even. Ground state term symbol It is relatively easy to calculate the term symbol for the ground state of an atom using Hund's rules. It corresponds with a state with maximum S and L. Start with the most stable electron configuration.
Full shells and subshells do not contribute to the overall angular momentum, so they are discarded.
If all shells and subshells are full then the term symbol is 1S0.
Distribute the electrons in the available orbitals, following the Pauli exclusion principle. First, fill the orbitals with highest mℓ value with one electron each, and assign a maximal ms to them (i.e. +1⁄2). Once all orbitals in a subshell have one electron, add a second one (following the same order), assigning ms = −1⁄2 to them. The overall S is calculated
by adding the ms values for each electron. According to Hund's first rule, the ground state has all unpaired electron spins parallel with the same value of ms, conventionally chosen as +1⁄2. The overall S is then 1⁄2 times the number of unpaired electrons. The overall L is calculated by adding the m ℓ {\displaystyle m_{\ell }} values for each electron (so
if there are two electrons in the same orbital, add twice that orbital's m ℓ {\displaystyle m_{\ell }} ). Calculate J as if less than half of the subshell is occupied, take the minimum value J = |L − S|; if more than half-filled, take the maximum value J = L + S; if the subshell is half-filled, then L will be 0, so J = S. As an example, in the case of fluorine, the
electronic configuration is 1s22s22p5. Discard the full subshells and keep the 2p5 part. So there are five electrons to place in subshell p ( ℓ = 1 {\displaystyle \ell =1} ). There are three orbitals ( m ℓ = 1 , 0 , − 1 {\displaystyle m_{\ell }=1,0,-1} ) that can hold up to 2 ( 2 ℓ + 1 ) = 6 {\displaystyle 2(2\ell +1)=6} electrons. The first three electrons can
take ms = 1⁄2 (↑) but the Pauli exclusion principle forces the next two to have ms = −1⁄2 (↓) because they go to already occupied orbitals. m ℓ {\displaystyle m_{\ell }} +1 0 −1 m s {\displaystyle m_{s}} ↑↓ ↑↓ ↑ S = 1⁄2 + 1⁄2 + 1⁄2 − 1⁄2 − 1⁄2 = 1⁄2; and L = 1 + 0 − 1 + 1 + 0 = 1, which is "P" in spectroscopic notation. As fluorine 2p subshell is more
than half filled, J = L + S = 3⁄2. Its ground state term symbol is then 2S+1LJ = 2P3⁄2. Atomic term symbols of the chemical elements In the periodic table, because atoms of elements in a column usually have the same outer electron structure, and always have the same electron structure in the "s-block" and "p-block" elements (see block (periodic
table)), all elements may share the same ground state term symbol for the column. Thus, hydrogen and the alkali metals are all 2S1⁄2, the alkaline earth metals are 1S0, the boron column elements are 2P1⁄2, the carbon column elements are 3P0, the pnictogens are 4S3⁄2, the chalcogens are 3P2, the halogens are 2P3⁄2, and the inert gases are 1S0, per
the rule for full shells and subshells stated above.
Term symbols for the ground states of most chemical elements[3] are given in the collapsed table below.[4] In the d-block and f-block, the term symbols are not always the same for elements in the same column of the periodic table, because open shells of several d or f electrons have several closely spaced terms whose energy ordering is often
perturbed by the addition of an extra complete shell to form the next element in the column. For example, the table shows that the first pair of vertically adjacent atoms with different ground-state term symbols are V and Nb. The 6D1⁄2 ground state of Nb corresponds to an excited state of V 2112 cm−1 above the 4F3⁄2 ground state of V, which in turn
corresponds to an excited state of Nb 1143 cm−1 above the Nb ground state.[1] These energy differences are small compared to the 15158 cm−1 difference between the ground and first excited state of Ca,[1] which is the last element before V with no d electrons. vteTerm symbol of the chemical elements Group → 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
16 17 18 ↓ Period 1 H 2S1/2 He1S0 2 Li2S1/2 Be1S0 B 2P1/2 C 3P0 N 4S3/2 O 3P2 F 2P3/2 Ne1S0 3 Na2S1/2 Mg1S0 Al2P1/2 Si3P0 P 4S3/2 S 3P2 Cl2P3/2 Ar1S0 4 K 2S1/2 Ca1S0 Sc2D3/2 Ti3F2 V 4F3/2 Cr7S3 Mn6S5/2 Fe5D4 Co4F9/2 Ni3F4 Cu2S1/2 Zn1S0 Ga2P1/2 Ge3P0 As4S3/2 Se3P2 Br2P3/2 Kr1S0 5 Rb2S1/2 Sr1S0 Y 2D3/2 Zr3F2 Nb6D1/2
Mo7S3 Tc6S5/2 Ru5F5 Rh4F9/2 Pd1S0 Ag2S1/2 Cd1S0 In2P1/2 Sn3P0 Sb4S3/2 Te3P2 I 2P3/2 Xe1S0 6 Cs2S1/2 Ba1S0 Lu2D3/2 Hf3F2 Ta4F3/2 W 5D0 Re6S5/2 Os5D4 Ir4F9/2 Pt3D3 Au2S1/2 Hg1S0 Tl2P1/2 Pb3P0 Bi4S3/2 Po3P2 At2P3/2 Rn1S0 7 Fr2S1/2 Ra1S0 Lr2P1/2? Rf3F2 Db4F3/2? Sg5D0? Bh6S5/2? Hs5D4? Mt4F9/2? Ds3F4? Rg2D5/2?
Cn1S0? Nh2P1/2? Fl3P0? Mc4S3/2? Lv3P2? Ts2P3/2? Og1S0? La2D3/2 Ce1G4 Pr4I9/2 Nd5I4 Pm6H5/2 Sm7F0 Eu8S7/2 Gd9D2 Tb6H15/2 Dy5I8 Ho4I15/2 Er3H6 Tm2F7/2 Yb1S0 Ac2D3/2 Th3F2 Pa4K11/2 U 5L6 Np6L11/2 Pu7F0 Am8S7/2 Cm9D2 Bk6H15/2 Cf5I8 Es4I15/2 Fm3H6 Md2F7/2 No1S0 s-block f-block d-block p-block Term symbols for an
electron configuration The process to calculate all possible term symbols for a given electron configuration is somewhat longer. First, the total number of possible states N is calculated for a given electron configuration. As before, the filled (sub)shells are discarded, and only the partially filled ones are kept. For a given orbital quantum number ℓ
{\displaystyle \ell } , t is the maximum allowed number of electrons, t = 2 ( 2 ℓ + 1 ) {\displaystyle t=2(2\ell +1)} . If there are e electrons in a given subshell, the number of possible states is N = ( t e ) = t ! e ! ( t − e ) ! . {\displaystyle N={t \choose e}={t! \over {e!\,(t-e)!}}.} As an example, consider the carbon electron structure: 1s22s22p2. After
removing full subshells, there are 2 electrons in a p-level ( ℓ = 1 {\displaystyle \ell =1} ), so there are N = 6 ! 2 ! 4 ! = 15 {\displaystyle N={6! \over {2!\,4!}}=15} different states. Second, all possible states are drawn. ML and MS for each state are calculated, with M = ∑ i = 1 e m i {\displaystyle M=\sum _{i=1}^{e}m_{i}} where mi is either m ℓ
{\displaystyle m_{\ell }} or m s {\displaystyle m_{s}} for the i-th electron, and M represents the resulting ML or MS respectively: m ℓ {\displaystyle m_{\ell }} +1 0 −1 ML MS all up ↑ ↑ 1 1 ↑ ↑ 0 1 ↑ ↑ −1 1 all down ↓ ↓ 1 −1 ↓ ↓ 0 −1 ↓ ↓ −1 −1 one up one down ↑↓ 2 0 ↑ ↓ 1 0 ↑ ↓ 0 0 ↓ ↑ 1 0 ↑↓ 0 0 ↑ ↓ −1 0 ↓ ↑ 0 0 ↓ ↑ −1 0 ↑↓ −2 0
Third, the number of states for each (ML,MS) possible combination is counted: MS +1 0 −1 ML +2 1 +1 1 2 1 0 1 3 1 −1 1 2 1 −2 1 Fourth, smaller tables can be extracted representing each possible term. Each table will have the size (2L+1) by (2S+1), and will contain only "1"s as entries.
The first table extracted corresponds to ML ranging from −2 to +2 (so L = 2), with a single value for MS (implying S = 0). This corresponds to a 1D term. The remaining terms fit inside the middle 3×3 portion of the table above. Then a second table can be extracted, removing the entries for ML and MS both ranging from −1 to +1 (and so S = L = 1, a
3P term).
The remaining table is a 1×1 table, with L = S = 0, i.e., a 1S term. S = 0, L = 2, J = 21D2 MS 0 ML +2 1 +1 1 0 1 −1 1 −2 1 S=1, L=1, J=2,1,03P2, 3P1, 3P0 MS +1 0 −1 ML +1 1 1 1 0 1 1 1 −1 1 1 1 S=0, L=0, J=01S0 MS 0 ML 0 1 Fifth, applying Hund's rules, the ground state can be identified (or the lowest state for the configuration of
interest). Hund's rules should not be used to predict the order of states other than the lowest for a given configuration. (See examples at Hund's rules § Excited states.) If only two equivalent electrons are involved, there is an "Even Rule" which states that, for two equivalent electrons, the only states that are allowed are those for which the sum (L +
S) is even. Case of three equivalent electrons For three equivalent electrons (with the same orbital quantum number ℓ {\displaystyle \ell } ), there is also a general formula (denoted by X ( L , S , ℓ ) {\displaystyle X(L,S,\ell )} below) to count the number of any allowed terms with total orbital quantum number L and total spin quantum number S. X ( L ,
S , ℓ ) = { L − ⌊ L 3 ⌋ , if S = 1 / 2 and 0 ≤ L < ℓ ℓ − ⌊ L 3 ⌋ , if S = 1 / 2 and ℓ ≤ L ≤ 3 ℓ − 1 ⌊ L 3 ⌋ − ⌊ L − ℓ 2 ⌋ + ⌊ L − ℓ + 1 2 ⌋ , if S = 3 / 2 and 0 ≤ L < ℓ ⌊ L 3 ⌋ − ⌊ L − ℓ 2 ⌋ , if S = 3 / 2 and ℓ ≤ L ≤ 3 ℓ − 3 0 , other cases {\displaystyle X(L,S,\ell )={\begin{cases}L-\left\lfloor {\tfrac {L}{3}}\right\rfloor ,&{\text{if }}S=1/2{\text{ and }}0\leq
L<\ell \\\ell -\left\lfloor {\tfrac {L}{3}}\right\rfloor ,&{\text{if }}S=1/2{\text{ and }}\ell \leq L\leq 3\ell -1\\\left\lfloor {\tfrac {L}{3}}\right\rfloor -\left\lfloor {\tfrac {L-\ell }{2}}\right\rfloor +\left\lfloor {\tfrac {L-\ell +1}{2}}\right\rfloor ,&{\text{if }}S=3/2{\text{ and }}0\leq L<\ell \\\left\lfloor {\tfrac {L}{3}}\right\rfloor -\left\lfloor {\tfrac {L-\ell }
{2}}\right\rfloor ,&{\text{if }}S=3/2{\text{ and }}\ell \leq L\leq 3\ell -3\\0,&{\text{other cases}}\end{cases}}} where the floor function ⌊ x ⌋ {\displaystyle \lfloor x\rfloor } denotes the greatest integer not exceeding x. The detailed proof can be found in Renjun Xu's original paper.[5]For a general electronic configuration of ℓ k {\displaystyle \ell
^{k}} , namely k equivalent electrons occupying one subshell, the general treatment, and computer code can also be found in this paper.[5] Alternative method using group theory For configurations with at most two electrons (or holes) per subshell, an alternative and much quicker method of arriving at the same result can be obtained from group
theory. The configuration 2p2 has the symmetry of the following direct product in the full rotation group: Γ(1) × Γ(1) = Γ(0) + [Γ(1)] + Γ(2), which, using the familiar labels Γ(0) = S, Γ(1) = P and Γ(2) = D, can be written as P × P = S + [P] + D. The square brackets enclose the anti-symmetric square. Hence the 2p2 configuration has components with
the following symmetries: S + D (from the symmetric square and hence having symmetric spatial wavefunctions); P (from the anti-symmetric square and hence having an anti-symmetric spatial wavefunction).
The Pauli principle and the requirement for electrons to be described by anti-symmetric wavefunctions imply that only the following combinations of spatial and spin symmetry are allowed: 1S + 1D (spatially symmetric, spin anti-symmetric) 3P (spatially anti-symmetric, spin symmetric). Then one can move to step five in the procedure above, applying
Hund's rules. The group theory method can be carried out for other such configurations, like 3d2, using the general formula Γ(j) × Γ(j) = Γ(2j) + Γ(2j−2) + ⋯ + Γ(0) + [Γ(2j−1) + ⋯ + Γ(1)]. The symmetric square will give rise to singlets (such as 1S, 1D, & 1G), while the anti-symmetric square gives rise to triplets (such as 3P & 3F). More generally,
one can use Γ(j) × Γ(k) = Γ(j+k) + Γ(j+k−1) + ⋯ + Γ(|j−k|) where, since the product is not a square, it is not split into symmetric and anti-symmetric parts. Where two electrons come from inequivalent orbitals, both a singlet and a triplet are allowed in each case.[6] Summary of various coupling schemes and corresponding term symbols Basic
concepts for all coupling schemes: ℓ {\displaystyle {\boldsymbol {\ell }}} : individual orbital angular momentum vector for an electron, s {\displaystyle \mathbf {s} } : individual spin vector for an electron, j {\displaystyle \mathbf {j} } : individual total angular momentum vector for an electron, j = ℓ + s {\displaystyle \mathbf {j} ={\boldsymbol {\ell
}}+\mathbf {s} } . L {\displaystyle \mathbf {L} } : Total orbital angular momentum vector for all electrons in an atom ( L = ∑ i ℓ i {\displaystyle \mathbf {L} =\sum _{i}{\boldsymbol {\ell }}_{i}} ). S {\displaystyle \mathbf {S} } : total spin vector for all electrons ( S = ∑ i s i {\displaystyle \mathbf {S} =\sum _{i}\mathbf {s} _{i}} ). J {\displaystyle
\mathbf {J} } : total angular momentum vector for all electrons. The way the angular momenta are combined to form J {\displaystyle \mathbf {J} } depends on the coupling scheme: J = L + S {\displaystyle \mathbf {J} =\mathbf {L} +\mathbf {S} } for LS coupling, J = ∑ i j i {\displaystyle \mathbf {J} =\sum _{i}\mathbf {j} _{i}} for jj coupling, etc. A
quantum number corresponding to the magnitude of a vector is a letter without an arrow, or without boldface (example: ℓ is the orbital angular momentum quantum number for ℓ {\displaystyle {\boldsymbol {\ell }}} and ℓ ^ 2 | ℓ , m ℓ , … ⟩ = ℏ 2 ℓ ( ℓ + 1 ) | ℓ , m ℓ , … ⟩ {\displaystyle {{\hat {\ell }}^{2}}\left|\ell ,m_{\ell },\ldots \right\rangle ={\hbar
^{2}}\ell \left(\ell +1\right)\left|\ell ,m_{\ell },\ldots \right\rangle } ) The parameter called multiplicity represents the number of possible values of the total angular momentum quantum number J for certain conditions. For a single electron, the term symbol is not written as S is always 1/2, and L is obvious from the orbital type. For two electron
groups A and B with their own terms, each term may represent S, L and J which are quantum numbers corresponding to the S {\displaystyle \mathbf {S} } , L {\displaystyle \mathbf {L} } and J {\displaystyle \mathbf {J} } vectors for each group. "Coupling" of terms A and B to form a new term C means finding quantum numbers for new vectors S = S
A + S B {\displaystyle \mathbf {S} =\mathbf {S} _{A}+\mathbf {S} _{B}} , L = L A + L B {\displaystyle \mathbf {L} =\mathbf {L} _{A}+\mathbf {L} _{B}} and J = L + S {\displaystyle \mathbf {J} =\mathbf {L} +\mathbf {S} } . This example is for LS coupling and which vectors are summed in a coupling is depending on which scheme of coupling
is taken. Of course, the angular momentum addition rule is that X = X A + X B , X A + X B − 1 , … , | X A − X B | {\displaystyle X=X_{A}+X_{B},X_{A}+X_{B}-1,\dots ,|X_{A}-X_{B}|} where X can be s, ℓ, j, S, L, J or any other angular momentum-magnitude-related quantum number. LS coupling (Russell–Saunders coupling) Coupling scheme: L
{\displaystyle \mathbf {L} } and S {\displaystyle \mathbf {S} } are calculated first then J = L + S {\displaystyle \mathbf {J} =\mathbf {L} +\mathbf {S} } is obtained.
From a practical point of view, it means L, S and J are obtained by using an addition rule of the angular momenta of given electron groups that are to be coupled. Electronic configuration + Term symbol: n ℓ N ( ( 2 S + 1 ) L J ) {\displaystyle n{\ell ^{N}}{{(}^{(2S+1)}}{L_{J}})} . ( ( 2 S + 1 ) L J ) {\displaystyle {{(}^{(2S+1)}}{{L}_{J}})} is a
term which is from coupling of electrons in n ℓ N {\displaystyle n{\ell ^{N}}} group. n , ℓ {\displaystyle n,\ell } are principle quantum number, orbital quantum number and n ℓ N {\displaystyle n{\ell ^{N}}} means there are N (equivalent) electrons in n ℓ {\displaystyle n\ell } subshell. For L > S {\displaystyle L>S} , ( 2 S + 1 ) {\displaystyle (2S+1)}
is equal to multiplicity, a number of possible values in J (final total angular momentum quantum number) from given S and L. For S > L {\displaystyle S>L} , multiplicity is ( 2 L + 1 ) {\displaystyle (2L+1)} but ( 2 S + 1 ) {\displaystyle (2S+1)} is still written in the term symbol.
Strictly speaking, ( ( 2 S + 1 ) L J ) {\displaystyle {{(}^{(2S+1)}}{L_{J}})} is called level and ( 2 S + 1 ) L {\displaystyle {^{\left(2S+1\right)}{L}}} is called term. Sometimes right superscript o is attached to the term symbol, meaning the parity P = ( − 1 ) ∑ i ℓ i {\displaystyle P={{\left(-1\right)}^{{\underset {i}{\mathop {\sum } }}\,{\ell
_{i}}}}} of the group is odd ( P = − 1 {\displaystyle P=-1} ). Example: 3d7 4F7/2: 4F7/2 is level of 3d7 group in which are equivalent 7 electrons are in 3d subshell. 3d7(4F)4s4p(3P0) 6F09/2:[7] Terms are assigned for each group (with different principal quantum number n) and rightmost level 6Fo9/2 is from coupling of terms of these groups so
6Fo9/2 represents final total spin quantum number S, total orbital angular momentum quantum number L and total angular momentum quantum number J in this atomic energy level. The symbols 4F and 3Po refer to seven and two electrons respectively so capital letters are used. 4f7(8S0)5d (7Do)6p 8F13/2: There is a space between 5d and (7Do).
It means (8S0) and 5d are coupled to get (7Do). Final level 8Fo13/2 is from coupling of (7Do) and 6p. 4f(2F0) 5d2(1G) 6s(2G) 1P01: There is only one term 2Fo which is isolated in the left of the leftmost space. It means (2Fo) is coupled lastly; (1G) and 6s are coupled to get (2G) then (2G) and (2Fo) are coupled to get final term 1Po1.
jj Coupling Coupling scheme: J = ∑ i j i {\displaystyle \mathbf {J} =\sum _{i}\mathbf {j} _{i}} . Electronic configuration + Term symbol: ( n 1 ℓ 1 j 1 N 1 n 2 ℓ 2 j 2 N 2 … ) J {\displaystyle {{\left({n_{1}}{\ell _{1}}_{j_{1}}^{N_{1}}{n_{2}}{\ell _{2}}_{j_{2}}^{N_{2}}\ldots \right)}_{J}}} Example: ( 6p 1 2 2 6p 3 2 ) o 3 / 2 {\displaystyle
{{\left({\text{6p}}_{\frac {1}{2}}^{2}{\text{6p}}_{\frac {3}{2}}^{}\right)}^{o}}_{3/2}} : There are two groups. One is 6p 1 / 2 2 {\displaystyle {\text{6p}}_{1/2}^{2}} and the other is 6p 3 2 {\displaystyle {\text{6p}}_{\frac {3}{2}}^{}} . In 6p 1 / 2 2 {\displaystyle {\text{6p}}_{1/2}^{2}} , there are 2 electrons having j = 1 / 2
{\displaystyle j=1/2} in 6p subshell while there is an electron having j = 3 / 2 {\displaystyle j=3/2} in the same subshell in 6p 3 2 {\displaystyle {\text{6p}}_{\frac {3}{2}}^{}} . Coupling of these two groups results in 1 {\displaystyle {1}} (coupling of j of three electrons). 4d 5 / 2 3 4d 3 / 2 2 ( 9 2 , 2 ) 11 / 2 {\displaystyle {\text{4d}}_{5/2}^{3}
{\text{4d}}_{3/2}^{2}~\ {{\left({\frac {9}{2}},2\right)}_{11/2}}} : 9 / 2 {\displaystyle 9/2} in () is J 1 {\displaystyle J_{1}} for 1st group 4d 5 / 2 3 {\displaystyle {\text{4d}}_{5/2}^{3}} and 2 in () is J2 for 2nd group 4d 3 / 2 2 {\displaystyle {\text{4d}}_{3/2}^{2}} .
Subscript 11/2 of term symbol is final J of J = J 1 + J 2 {\displaystyle \mathbf {J} =\mathbf {J} _{1}+\mathbf {J} _{2}} . J1L2 coupling Coupling scheme: K = J 1 + L 2 {\displaystyle \mathbf {K} =\mathbf {J} _{1}+\mathbf {L} _{2}} and J = K + S 2 {\displaystyle \mathbf {J} =\mathbf {K} +\mathbf {S} _{2}} . Electronic configuration + Term
symbol: n 1 ℓ 1 N 1 ( t e r m 1 ) n 2 ℓ 2 N 2 ( t e r m 2 ) ( 2 S 2 + 1 ) [ K ] J {\displaystyle {n_{1}}{\ell _{1}}^{N_{1}}\left(\mathrm {term} _{1}\right){n_{2}}{\ell _{2}}^{N_{2}}\left(\mathrm {term} _{2}\right)~\ {^{\left(2{S_{2}}+1\right)}{{\left[K\right]}_{J}}}} . For K > S 2 , ( 2 S 2 + 1 ) {\displaystyle K>S_{2},(2S_{2}+1)} is equal to
multiplicity, a number of possible values in J (final total angular momentum quantum number) from given S2 and K. For S 2 > K {\displaystyle S_{2}>K} , multiplicity is ( 2 K + 1 ) {\displaystyle (2K+1)} but ( 2 S 2 + 1 ) {\displaystyle (2S_{2}+1)} is still written in the term symbol. Example: 3p5(2Po1/2)5g 2[9/2]o5: J 1 = 1 2 , l 2 = 4 , s 2 = 1 / 2
{\displaystyle {J_{1}}={\frac {1}{2}},{l_{2}}=4,~{s_{2}}=1/2} .
9 / 2 {\displaystyle 9/2} is K, which comes from coupling of J1 and ℓ2.
Subscript 5 in term symbol is J which is from coupling of K and s2. 4f13(2Fo7/2)5d2(1D) [7/2]o7/2: J 1 = 7 2 , L 2 = 2 , S 2 = 0 {\displaystyle {J_{1}}={\frac {7}{2}},{L_{2}}=2,~{S_{2}}=0} . 7 / 2 {\displaystyle 7/2} is K, which comes from coupling of J1 and L2. Subscript 7 / 2 {\displaystyle 7/2} in the term symbol is J which is from coupling of
K and S2. LS1 coupling Coupling scheme: K ℓ = L + S 1 {\displaystyle \mathbf {K} \ell =\mathbf {L} +\mathbf {S_{1}} } , J = K + S 2 {\displaystyle \mathbf {J} =\mathbf {K} +\mathbf {S_{2}} } . Electronic configuration + Term symbol: n 1 ℓ 1 N 1 ( t e r m 1 ) n 2 ℓ 2 N 2 ( t e r m 2 ) L ( 2 S 2 + 1 ) [ K ] J {\displaystyle {n_{1}}{\ell
_{1}}^{N_{1}}\left(\mathrm {term} _{1}\right){n_{2}}{\ell _{2}}^{N_{2}}\left(\mathrm {term} _{2}\right)\ ~L~\ {^{\left(2{S_{2}}+1\right)}{{\left[K\right]}_{J}}}} . For K > S 2 , ( 2 S 2 + 1 ) {\displaystyle K>S_{2},(2S_{2}+1)} is equal to multiplicity, a number of possible values in J (final total angular momentum quantum number) from
given S2 and K. For S 2 > K {\displaystyle S_{2}>K} , multiplicity is ( 2 K + 1 ) {\displaystyle (2K+1)} but ( 2 S 2 + 1 ) {\displaystyle (2S_{2}+1)} is still written in the term symbol. Example: 3d7(4P)4s4p(3Po) Do 3[5/2]o7/2: L 1 = 1 , L 2 = 1 , S 1 = 3 2 , S 2 = 1 {\displaystyle {L_{1}}=1,~{L_{2}}=1,~{S_{1}}={\frac {3}{2}},~{S_{2}}=1}
. L = 2 , K = 5 / 2 , J = 7 / 2 {\displaystyle L=2,K=5/2,J=7/2} . Most famous coupling schemes are introduced here but these schemes can be mixed to express the energy state of an atom. This summary is based on [1]. Racah notation and Paschen notation These are notations for describing states of singly excited atoms, especially noble gas atoms.
Racah notation is basically a combination of LS or Russell–Saunders coupling and J1L2 coupling. LS coupling is for a parent ion and J1L2 coupling is for a coupling of the parent ion and the excited electron. The parent ion is an unexcited part of the atom. For example, in Ar atom excited from a ground state ...3p6 to an excited state ...3p54p in
electronic configuration, 3p5 is for the parent ion while 4p is for the excited electron.[8] In Racah notation, states of excited atoms are denoted as ( ( 2 S 1 + 1 ) L 1 J 1 ) n ℓ [ K ] J o {\displaystyle \left(^{\left(2{{S}_{1}}+1\right)}{{L}_{1}}_{{J}_{1}}\right)n\ell \left[K\right]_{J}^{o}} . Quantities with a subscript 1 are for the parent ion, n and ℓ are
principal and orbital quantum numbers for the excited electron, K and J are quantum numbers for K = J 1 + ℓ {\displaystyle \mathbf {K} =\mathbf {J} _{1}+{\boldsymbol {\ell }}} and J = K + s {\displaystyle \mathbf {J} =\mathbf {K} +\mathbf {s} } where ℓ {\displaystyle {\boldsymbol {\ell }}} and s {\displaystyle \mathbf {s} } are orbital angular
momentum and spin for the excited electron respectively. “o” represents a parity of excited atom. For an inert (noble) gas atom, usual excited states are Np5nℓ where N = 2, 3, 4, 5, 6 for Ne, Ar, Kr, Xe, Rn, respectively in order. Since the parent ion can only be 2P1/2 or 2P3/2, the notation can be shortened to n ℓ [ K ] J o {\displaystyle n\ell
\left[K\right]_{J}^{o}} or n ℓ ′ [ K ] J o {\displaystyle n\ell '\left[K\right]_{J}^{o}} , where nℓ means the parent ion is in 2P3/2 while nℓ′ is for the parent ion in 2P1/2 state. Paschen notation is a somewhat odd notation; it is an old notation made to attempt to fit an emission spectrum of neon to a hydrogen-like theory. It has a rather simple structure to
indicate energy levels of an excited atom. The energy levels are denoted as n′ℓ#. ℓ is just an orbital quantum number of the excited electron. n′ℓ is written in a way that 1s for (n = N + 1, ℓ = 0), 2p for (n = N + 1, ℓ = 1), 2s for (n = N + 2, ℓ = 0), 3p for (n = N + 2, ℓ = 1), 3s for (n = N + 3, ℓ = 0), etc. Rules of writing n′ℓ from the lowest electronic
configuration of the excited electron are: (1) ℓ is written first, (2) n′ is consecutively written from 1 and the relation of ℓ = n′ − 1, n′ − 2, ... , 0 (like a relation between n and ℓ) is kept. n′ℓ is an attempt to describe electronic configuration of the excited electron in a way of describing electronic configuration of hydrogen atom. # is an additional number
denoted to each energy level of given n′ℓ (there can be multiple energy levels of given electronic configuration, denoted by the term symbol). # denotes each level in order, for example, # = 10 is for a lower energy level than # = 9 level and # = 1 is for the highest level in a given n′ℓ. An example of Paschen notation is below. Electronic configuration
of Neon n′ℓ Electronic configuration of Argon n′ℓ 1s22s22p6 Ground state [Ne]3s23p6 Ground state 1s22s22p53s1 1s [Ne]3s23p54s1 1s 1s22s22p53p1 2p [Ne]3s23p54p1 2p 1s22s22p54s1 2s [Ne]3s23p55s1 2s 1s22s22p54p1 3p [Ne]3s23p55p1 3p 1s22s22p55s1 3s [Ne]3s23p56s1 3s See also Quantum number Principal quantum number Azimuthal
quantum number Spin quantum number Magnetic quantum number Angular quantum numbers Angular momentum coupling Molecular term symbol Hole formalism Notes ^ There is no official convention for naming orbital angular momentum values greater than 20 (symbol Z) but they are rarely needed. Some authors use Greek letters (α, β, γ, ...)
after Z. References ^ a b c d NIST Atomic Spectrum Database For example, to display the levels for a neutral carbon atom, enter "C I" or "C 0" in the "Spectrum" box and click "Retrieve data". ^ Russel, H.
N.; Saunders, F. A. (1925) [January 1925]. "New Regularities in the Spectra of the Alkaline Earths". SAO/NASA Astrophysics Data System (ADS). Astrophysical Journal. adsabs.harvard.edu/. 61: 38. Bibcode:1925ApJ....61...38R. doi:10.1086/142872. Retrieved December 13, 2020 – via harvard.edu. ^ "NIST Atomic Spectra Database Ionization Energies
Form". NIST Physical Measurement Laboratory. National Institute of Standards and Technology (NIST). October 2018. Retrieved 28 January 2019. This form provides access to NIST critically evaluated data on ground states and ionization energies of atoms and atomic ions. ^ For the sources for these term symbols in the case of the heaviest
elements, see Template:Infobox element/symbol-to-electron-configuration/term-symbol. ^ a b Xu, Renjun; Zhenwen, Dai (2006). "Alternative mathematical technique to determine LS spectral terms". Journal of Physics B: Atomic, Molecular and Optical Physics. 39 (16): 3221–3239. arXiv:physics/0510267. Bibcode:2006JPhB...39.3221X.
doi:10.1088/0953-4075/39/16/007. S2CID 2422425. ^ McDaniel, Darl H. (1977). "Spin factoring as an aid in the determination of spectroscopic terms". Journal of Chemical Education. 54 (3): 147. Bibcode:1977JChEd..54..147M. doi:10.1021/ed054p147. ^ "Atomic Spectroscopy - Different Coupling Scheme 9. Notations for Different Coupling Schemes".
Nist. National Institute of Standards and Technology (NIST). 1 November 2017. Retrieved 31 January 2019. ^ "APPENDIX 1 - Coupling Schemes and Notation" (PDF). University of Toronto: Advanced Physics Laboratory - Course Homepage. Retrieved 5 Nov 2017. Retrieved from " Atomic term symbols contain two pieces of information. They tell you
the total orbital angular momentum of the atom (\(L\)), and they tell you the multiplicity (\(M\)). \(L\) is denoted by a simple code, similar to the code used to delineate the types of atomic orbitals:Note that while the notation is similar, L does NOT say anything about what types of orbitals the electrons are in. A state that has the term symbol P does
NOT necessarily have an open p‐shell. The multiplicity is indicated by appending a number to the upper left of the symbol. A \(L=2\), \(M=3\) state would be represented by \(^3D\). The secret to writing the term symbols for an atom is to discover what combinations of \(L\) and \(M\) are possible for that atom with that specific electronic configuration.
An atom that only has closed shells will always be \(1S\).Each term symbol represents a discrete energy level. We can place these levels in the correct order by using these simple rules:These rules reliably predict the ground state. They have only erratic agreement with experiment when ordering the other levels. Example \(\PageIndex{9}\): Hydrogen
Ground State What are the term symbols for the microstates possible for \(1s^1\) electronic configuration of hydrogen? Solution Since there is only one electron, this is a simple problem. \(L=0\) and \(M=1\), so the only possible term symbol is \(^2S\). With only one electron, \(S = ½\), so \(J = 0 + ½ = ½\). Only one microstate exists for this
configuration and it has a term symbol of \(^2S_{½}\). Example \(\PageIndex{10}\): Boron What are the term symbols for the microstates possible for \(1s^2 2s^2 2p^1\) electronic configuration of boron? Solution There still only one open shell electron, so \(L=1\), \(M=1\) and \(S = ½\). We get a term symbol of the type \(^2P\), which gets split into
separate symbols because \(J = 3/2\) and \(1/2\).
Two possible microstates exist for this system with term symbols of \(^2P_{3/2}\) and \(^2P_{1/2}\). Example \(\PageIndex{11}\): Beryllium Excited State What are the term symbols for the microstates possible for the \(1s^2 2s^1 2p^1\) excited-state electronic configuration of Beryllium? Solution Now we have two electrons to worry about. Since \
(l_1 = 0\) and \(l_2 = 1\), the only possible combination is \(L=1\). The possible combinations of S are: \(S=1,0\). This means that \(M=3\) or \(M=1\). The term symbols will be of the form \(^1P\) and \(^3P\). For the \(^1 P\) state, \(L=1\) and \(S=0\), so \(J=1\). For the second state, \(L=1\) and \(S=1\), so \(J=2,1,0\). There are four microstates for this
configuration with term symbols of \(^1 P_1\) and \(^3 P_2\), \(^3P_1\), and \(^3P_0\). Example \(\PageIndex{12}\): Zirconium What are the term symbols for the microstates possible for the \([Kr] 5s^2 4d^2\) ground-state electronic configuration of zirconium? Solution This is a much harder problem. We will need to use a special technique to
disentangle all of the possible combinations of \(L\) and \(M\). Let’s start be listing the relevant quantum numbers for the two open-shell electrons: \(l_1 = 2\) \(l_2 = 2\) \(m_{l1} = 2, 1, 0, ‐ 1, ‐ 2\) \(m_{l2} = 2, 1, 0, ‐ 1, ‐ 2\) \(m_{s1} = ½, ‐½\) \(m_{s1} = ½, ‐½\) Let’s combine these numbers to generate the atomic quantum numbers: \[L = 4,3,2,1,0 \]
\[M_l = 4,3,2,1,0, ‐ 1, ‐ 2, ‐ 3, ‐ 4\] \[M_s = 1,0\] We know that there will at least one each of S, P, D, F and G. It isn’t immediately clear which of these will be singlets and which will be triplets. To figure this out, we need to systematically examine the possible microstates.
It turns out that there are 45 possible ways to put distribute two electrons between 5 d orbitals. That’s a lot! The easiest way to list the states is to organize them into a chart: \(M_s= -1\) \(M_s= 0\) \(M_s= +1\) Attacking the chart one row at a time. Ask yourself, how many ways can I arrange the two electrons to give me \(M_l = 4\)? It turns out there
is only one possible combination that does this: This state is has \(M_s = 0\).
This means that there is only 1 microstate that corresponds to \(M_l =4\) and \(M_s = 0\), and none that correspond to \(M_l =4\) and \(M_s = \pm ‐1\).
We add this microstate to the chart like this: \(M_l\) \(M_s= -1\) \(M_s= 0\) \(M_s= +1\) 4 0 1 0 Now, how many ways are there to get \(M_l = 3\)?
\(M_l\) \(M_s= -1\) \(M_s= 0\) \(M_s= +1\) 4 0 1 0 3 1 2 1 For \(M_l =2\), we find the following states: \(M_l\) \(M_s= -1\) \(M_s= 0\) \(M_s= +1\) 4 0 1 0 3 1 2 1 2 1 3 1 You should be able to draw the microstates on your own by now. You should find 8 states, four of which are singlet and four of which are triplets. \(M_l\) \(M_s= -1\) \(M_s= 0\) \(M_s=
+1\) 4 0 1 0 3 1 2 1 2 1 3 1 1 2 4 2 There are only nine possible ways to arrange the electrons to get \(M_l =0\) \(M_l\) \(M_s= -1\) \(M_s= 0\) \(M_s= +1\) 4 0 1 0 3 1 2 1 2 1 3 1 1 2 4 2 0 2 5 2 The rest of the chart will be symmetric to the first half, so we do not need to do any more work: \(M_l\) \(M_s= -1\) \(M_s= 0\) \(M_s= +1\) 4 0 1 0 3 1 2 1 2 1 3 1
1 2 4 2 0 2 5 2 -1 2 4 2 -2 1 3 1 -3 1 2 1 -4 0 1 0 Now that we have a listing of all of the microstates, we need to figure out how to divide them up between the term symbols.
It turns out that each term symbol can have, at most, one microstate from each box on the chart. The term symbols always end up claiming a “box” of microstates, centered on the middle of the chart. This is easier shown than said. Attacking the chart from the top, we can see that the \(M_l = 4\) \(M_s = 0\) state clearly belongs to a \(^1G\) symbol.
The \(M_l = ‐4\) \(M_s =0\) box also clearly belongs to this symbol. If I connect these states with a “box,” I get this: \(M_l\) \(M_s= -1\) \(M_s= 0\) \(M_s= +1\) 4 0 1 0 3 1 2 1 2 1 3 1 1 2 4 2 0 2 5 2 -1 2 4 2 -2 1 3 1 -3 1 2 1 -4 0 1 0 The strickthough configurations all belong to the \(^1 G\) state.
Let's subtract them from the chart to indicate that they are not available for other term symbols. \(M_l\) \(M_s= -1\) \(M_s= 0\) \(M_s= +1\) 3 1 2 1 2 1 3 1 1 2 4 2 0 2 5 2 -1 2 4 2 -2 1 3 1 -3 1 2 1 The next row indicates a \(L=3\) state.
Because there are three \(M_s\) values available, this is a triplet. The term symbol will be \(^3F\), which reduces the chart down to \(M_l\) \(M_s= -1\) \(M_s= 0\) \(M_s= +1\) 1 2 4 2 0 2 5 2 -1 2 4 2 The next state will be \(^1D\). Leaving us with \(M_l\) \(M_s= -1\) \(M_s= 0\) \(M_s= +1\) 1 1 1 1 0 1 2 1 -1 1 1 1 Next is a \(^3P\) state. The chart is getting
pretty small now wtih The last remaining microstate comprises the \(^1S\) term symbol. The total listing is \(^1G\), \(^3F\), \(^1D\), \(^3P\), \(^1S\).
Assigning \(J\) values, we get \(^1G_4\), \(^3F_4\), \(^3F_3\), \(^3F_2\), \(^1D_2\), \(^3P_2\), \(^3P_1\), \(^3P_0\), \(^1S_0\) If you can do this problem, you can do almost any atomic term symbol.

You might also like