Fact-checked by Grok 2 weeks ago

Spectroscopic notation

Spectroscopic notation is a standardized system in atomic and molecular physics for labeling the quantum mechanical states of electrons in atoms, ions, and molecules, particularly emphasizing the coupling of their orbital and angular momenta. In the most common form, known as term symbols, it uses the notation ^{2S+1}L_J, where L (denoted by letters S, P, D, F, etc., for L = 0, 1, 2, 3, \ldots) represents the total orbital , S is the total angular momentum , the superscript $2S+1 indicates the multiplicity of the state, and the subscript J denotes the total (ranging from |L - S| to L + S). This notation succinctly captures the energy levels and spectroscopic properties of multi-electron systems under Russell-Saunders coupling (also called LS coupling), which is applicable to lighter atoms where -orbit interactions are relatively weak. The origins of spectroscopic notation trace back to the late 19th and early 20th centuries, evolving from empirical classifications of spectral lines observed in atomic emission spectra. Early spectroscopists, such as Johann Balmer in 1885, identified patterns in lines, leading to formulas that categorized series as "sharp," "principal," "diffuse," and "fundamental," with corresponding letters s, p, d, and f initially denoting these line types rather than quantum numbers. By the , the discovery of electron spin and the need to explain prompted Henry Norris Russell and Frederick Albert Saunders to formalize the LS coupling scheme in their 1925 paper, introducing the modern format to describe multiplet structures in spectra like those of alkaline-earth elements. This development integrated with spectroscopic observations, providing a framework that linked atomic energy levels to observable transitions. In practice, spectroscopic notation is essential for predicting and interpreting atomic spectra, determining selection rules for allowed transitions (e.g., \Delta L = 0, \pm 1, \Delta S = 0, \Delta J = 0, \pm 1), and analyzing configurations in multi- atoms. For example, the of carbon is denoted as 2p^2 \, ^3P_0, indicating two s in the 2p orbital with total L=1, S=1, and J=0. While LS coupling dominates for lighter elements (typically up to Z ≈ 50–60), alternative schemes like jj coupling are used for heavier atoms where relativistic effects strengthen spin-orbit interactions. Today, this notation underpins applications in , laser physics, and , enabling precise modeling of atomic interactions and diagrams.

Basic Principles

Quantum Numbers

In , the spectroscopic notation for atomic states relies on a set of fundamental quantum numbers that describe the quantum mechanical state of electrons in atoms. These numbers arise from the solutions to the for the and its multi-electron extensions, providing essential labels for energy levels, orbital shapes, spatial orientations, and intrinsic s. The four primary quantum numbers—principal, orbital , magnetic, and —uniquely specify each electron's state within an atom, forming the foundation for more complex notations used in . The principal n determines the energy level and average distance of the from the , taking positive values n = 1, 2, 3, \dots. Higher values of n correspond to higher states and larger orbital sizes. The orbital quantum number l specifies the shape of the orbital and ranges from $0ton-1for a givenn, influencing the electron's angular momentum magnitude. The [magnetic quantum number](/page/Magnetic_quantum_number) m_ldescribes the orbital's orientation in space relative to an external [magnetic field](/page/Magnetic_field), with possible values from-lto+lin [integer](/page/Integer) steps. Finally, the [spin quantum number](/page/Spin_quantum_number)m_saccounts for the electron's intrinsic [spin](/page/Spin), which can be either+\frac{1}{2}or-\frac{1}{2}$, representing the two possible spin projections along a quantization axis. These numbers collectively define the possible states available to electrons in an . The concept of these quantum numbers emerged in the early 20th century through the development of atomic models. In 1913, introduced the principal quantum number n in his model of the to quantize electron orbits and explain series, postulating discrete energy levels. extended this in 1916 by incorporating relativistic effects and elliptical orbits, introducing the orbital angular momentum l (initially as a secondary quantum number) to account for in spectra, along with the for Zeeman splitting. The m_s was later proposed by and in 1925 to explain spin-orbit coupling and anomalous Zeeman effects. The Pauli exclusion principle, formulated by Wolfgang Pauli in 1925, states that no two electrons in an atom can occupy the same quantum state, meaning they cannot share identical values for all four quantum numbers n, l, m_l, and m_s. This principle governs the filling of orbitals by ensuring that each orbital (defined by n, l, and m_l) holds at most two electrons with opposite spins, leading to the structured buildup of atomic electron configurations and the periodic table's organization. These quantum numbers provide the basis for constructing term symbols in spectroscopic notation, which combine individual electron states to describe total atomic angular momenta.

Angular Momentum Notation

In spectroscopic notation, the orbital quantum number l for a single is represented by letters derived from early classifications of series in atoms. These include s for l = 0 (sharp series), p for l = 1 (principal series), d for l = 2 (diffuse series), f for l = 3 (fundamental series), g for l = 4, and subsequent letters of the alphabet for higher values. This lettering system originated in the late from observations by spectroscopists such as George Liveing and , who categorized line sharpness in alkali spectra, with expanding the series descriptions around 1890 and formalizing the notation in 1927 to align with quantum mechanical subshells. For multi-electron atoms, the total orbital angular momentum quantum number L employs uppercase letters following the same sequence: S for L = 0, P for L = 1, D for L = 2, F for L = 3, and so forth. The magnitude of L arises from the vector sum of the individual orbital angular momenta of equivalent electrons, expressed as \mathbf{L} = \sum_i \mathbf{l}_i, where the possible values of L range from the maximum sum down to 0 in steps of 1, depending on the electron configuration. The total spin angular momentum quantum number S is denoted by the multiplicity $2S + 1, which serves as a left superscript in spectroscopic descriptions to indicate the number of possible spin states. This convention, part of the broader Russell-Saunders coupling scheme, reflects the degeneracy due to orientation and was standardized in early 20th-century to classify energy levels based on experimental spectra.

Atomic Notation

Electron Configurations and Orbitals

Electron configurations in spectroscopic notation describe the distribution of electrons among atomic orbitals for atoms in their ground or excited states. This notation specifies the principal quantum number n, the azimuthal quantum number \ell (represented by letters s, p, d, f for \ell = 0, 1, 2, 3), and the number of electrons in each subshell as a superscript. For example, the ground state configuration of neon is $1s^2 2s^2 2p^6, indicating two electrons in the 1s orbital, two in 2s, and six in 2p. The arrangement follows the , which states that electrons occupy orbitals starting from the lowest energy levels, ordered by increasing n + \ell, and for equal n + \ell, by increasing n. This building-up process, combined with the (limiting each orbital to two electrons of opposite ), and Hund's rule (maximizing unpaired electrons by filling degenerate orbitals singly with parallel spins before pairing), determines the configuration. Hund's rule minimizes electron-electron repulsion and exchange energy, leading to higher total and orbital for stability./Quantum_Mechanics/10:_Multi-electron_Atoms/Electron_Configuration) Atomic orbitals vary in shape and electron capacity based on \ell: s orbitals (\ell=0) are spherical and hold up to 2 electrons; p orbitals (\ell=1) are dumbbell-shaped along the x, y, or z axes and accommodate 6 electrons; d orbitals (\ell=2) have more complex cloverleaf or double-dumbbell shapes and hold 10 electrons. These shapes arise from the angular part of the wave function and influence electron probability density. For the first 20 elements, configurations follow the Aufbau order, filling 1s, then 2s and 2p, 3s and 3p, and 4s. Anomalies occur due to the stability of half-filled or fully filled subshells, as seen in (Z=24), which adopts [Ar] 4s^1 3d^5 instead of [Ar] 4s^2 3d^4, prioritizing the half-filled 3d subshell for lower energy. (Z=29) similarly shows [Ar] 4s^1 3d^{10} over [Ar] 4s^2 3d^9. The table below lists configurations for through calcium:
ElementAtomic NumberGround State Configuration
$1s^1
He$1s^2
$1s^2 2s^1
Be4$1s^2 2s^2
B5$1s^2 2s^2 2p^1
C6$1s^2 2s^2 2p^2
N7$1s^2 2s^2 2p^3
O8$1s^2 2s^2 2p^4
F9$1s^2 2s^2 2p^5
Ne10$1s^2 2s^2 2p^6
11$1s^2 2s^2 2p^6 3s^1
12$1s^2 2s^2 2p^6 3s^2
$1s^2 2s^2 2p^6 3s^2 3p^1
$1s^2 2s^2 2p^6 3s^2 3p^2
$1s^2 2s^2 2p^6 3s^2 3p^3
16$1s^2 2s^2 2p^6 3s^2 3p^4
Cl17$1s^2 2s^2 2p^6 3s^2 3p^5
Ar18$1s^2 2s^2 2p^6 3s^2 3p^6
K19$1s^2 2s^2 2p^6 3s^2 3p^6 4s^1
Ca20$1s^2 2s^2 2p^6 3s^2 3p^6 4s^2
Excited states are denoted similarly but with electrons promoted to higher-energy orbitals, such as the configuration $1s^2 2s 2p for an excited helium atom, where one electron moves from 2s to 2p. These states have higher energy than ground states and are relevant in spectroscopy for observing transitions.

Ionization States

In spectroscopic notation, the ionization state of an atom or ion is indicated by appending a Roman numeral to the chemical symbol of the element, where I denotes the neutral atom, II the singly ionized species (one electron removed), III the doubly ionized species, and so on. For example, H I represents neutral hydrogen, while Fe II signifies singly ionized iron, commonly observed in stellar atmospheres. This system provides a standardized way to specify the charge state responsible for particular spectral lines, facilitating the identification of elements in diverse astrophysical environments. The Roman numeral notation originated in the late 19th and early 20th centuries, pioneered by astronomers such as and Alfred Fowler during analyses of solar and stellar spectra. Lockyer distinguished "" lines from high-excitation sources like sparks, associating them with ionized states, while Fowler formalized the numbering in laboratory studies of arc and spark spectra around , linking specific series to ionization levels. This convention was widely adopted by the for classifying spectra from hot stars and plasmas, where higher ionization stages dominate due to elevated temperatures and energies. In observed spectra, ionization states influence the prominence of or lines, as higher temperatures in stellar photospheres or nebular regions promote greater , shifting observable features to lines from more charged . For instance, in hotter O-type stars (surface temperatures exceeding 30,000 K), lines from highly ionized iron like III or IV appear strongly, whereas cooler stars exhibit neutral or lowly ionized lines such as I. For highly ionized resembling hydrogen-like atoms (one remaining), the notation follows the same rule, such as He II for singly ionized or Li III for doubly ionized , whose spectra mimic scaled hydrogen lines due to dominant interactions. In planetary nebulae, common diagnostic lines from specific ions reveal ionization structure and excitation conditions; for example, the [O III] λ5007 forbidden line from traces high-ionization zones near the central star, while [N II] λ6583 from singly ionized highlights cooler, lower-ionization regions at the nebula's edges. These lines, along with others like [O II] and [S II], enable mapping of nebular abundances and dynamics, as their ratios correlate with the parameter and .

Term Symbols

Term symbols provide a complete spectroscopic notation for the levels of multi-electron atoms, incorporating the effects of spin-orbit coupling and specifying the quantum numbers relevant to transition probabilities and selection rules. The full term symbol is denoted as ^{2S+1}\mathrm{L}_J, where \mathrm{L} represents the total (with letters S for 0, P for 1, D for 2, etc.), S is the total angular momentum quantum number, and J is the total quantum number arising from the vectorial coupling of \mathbf{L} and \mathbf{S}. The multiplicity $2S+1 indicates the degeneracy, while J takes or values from |L - S| to L + S in steps of 1, leading to $2J+1 magnetic sublevels for each term. This notation, formalized in the Russell-Saunders () coupling scheme, is particularly applicable to atoms where electrostatic interactions dominate over spin-orbit effects. To construct term symbols from electron configurations, the possible values of L and S are determined by vector addition of individual orbital (l_i) and spin (s_i = 1/2) angular momenta, subject to the Pauli exclusion principle for equivalent electrons (those in the same n l subshell). For non-equivalent electrons, all combinations are allowed, but for equivalent electrons, antisymmetrization restricts the terms; for example, in the p^2 configuration, the allowed terms are ^1S, ^3P, and ^1D, excluding higher-spin or symmetric states that violate Pauli requirements. The resulting terms are then split by spin-orbit coupling into levels labeled by J, with energies shifted according to the Landé interval rule. The Landé interval rule states that the energy separation between consecutive J levels within a multiplet increases with J: specifically, ΔE(J to J+1) = A(J + 1), where A is the spin-orbit coupling constant. The energy of each level is E_J = \frac{A}{2} [J(J+1) - L(L+1) - S(S+1)]. In the LS (Russell-Saunders) coupling scheme, prevalent for light atoms with low atomic numbers (Z ≲ 40), such as those up to , the individual orbital angular momenta couple first to form \mathbf{L} = \sum \mathbf{l}_i and spins to \mathbf{S} = \sum \mathbf{s}_i, followed by \mathbf{J} = \mathbf{L} + \mathbf{S}. For heavier atoms, where spin-orbit coupling is stronger, the jj coupling scheme is more appropriate, in which each electron's j_i = l_i + s_i couples to total J, though intermediate schemes often apply in practice. A representative example is the of the neutral carbon atom ( $1s^2 2s^2 2p^2), which yields the ^3P with J=0,1,2 levels; the lowest is ^3P_0 at approximately 0 cm^{-1}, followed by ^3P_1 at 16.4 cm^{-1} and ^3P_2 at 43.5 cm^{-1}, illustrating fine-structure splitting consistent with the Landé rule. These term symbols underpin selection rules for electric dipole transitions, which govern allowed spectral lines in atomic spectra. The primary rules in LS coupling are ΔL = 0, ±1 (no 0 ↔ 0), with a change in , ΔS = 0 (no spin flip), and ΔJ = 0, ±1 (with no J=0 \to 0 transitions), ensuring that only certain multiplet components are observable. These rules arise from the symmetry of the operator and conservation of , enabling prediction of transition intensities and facilitating .

Molecular Notation

Molecular Orbitals

In molecular , the notation for orbitals in diatomic and linear polyatomic molecules differs from notation due to the cylindrical along the internuclear , which replaces the spherical of atoms. Instead of labeling orbitals by n and \ell (s, p, d, etc.), molecular orbitals (MOs) are classified by the absolute value of the projection quantum number \lambda, which represents the component of the orbital along the bond . This projection \lambda = |\Lambda|, where \Lambda is the eigenvalue of the L_z in cylindrical coordinates, determines the orbital's around the axis. The symmetry labels derive directly from \lambda: \sigma for \lambda = 0 (non-degenerate orbitals with no angular momentum projection, symmetric under rotation); \pi for \lambda = 1 (doubly degenerate pair of orbitals); \delta for \lambda = 2 (fourfold degenerate, quadrupolar); and higher Greek letters like \phi for \lambda = 3, though rarer in common molecules. These labels are appended with the principal quantum number or atomic orbital origin (e.g., $2s, $2p) to specify energy level, as in \sigma_{2s} or \pi_{2p}. Bonding orbitals lack an asterisk (*), antibonding ones are marked with *, and non-bonding orbitals (typically from atomic orbitals with minimal overlap) use n, such as n\pi. For homonuclear diatomics, additional parity labels g (gerade, even) or u (ungerade, odd) under inversion through the molecular center may be included, but the core symmetry notation focuses on \sigma, \pi, etc. Electron configurations in molecular notation list filled orbitals in order of increasing energy, with superscripts indicating occupancy, similar to atomic configurations but using MO labels. For the hydrogen molecular ion H_2^+, the ground-state configuration is (\sigma_{1s})^1, derived from the single electron in the bonding \sigma orbital formed by 1s atomic orbitals. For neutral H_2, it becomes (\sigma_{1s})^2. In heavier homonuclear diatomics, inner shells are abbreviated: K denotes a filled 1s^2 core on each atom (total four electrons), and KK represents helium-like cores for second-row elements. For N_2, the valence configuration is KK (\sigma_{2s})^2 (\sigma^*_{2s})^2 (\pi_{2p})^4 (\sigma_{2p})^2, yielding a triple bond from the net two electrons in bonding \pi and \sigma orbitals after core and antibonding cancellation. This notation extends to polyatomic linear molecules, where MOs retain \sigma, \pi labels based on axial symmetry, though degeneracy and filling orders vary with molecular geometry.

Electronic and Vibronic Terms

In molecular spectroscopy, electronic states of diatomic molecules are described using term symbols that account for the projection of angular momenta along the internuclear axis, spin multiplicity, and symmetry properties. The general form of a molecular term symbol is ^{2S+1}\Lambda_{\Omega}, where S is the total spin quantum number, \Lambda is the absolute value of the projection of the orbital angular momentum along the axis (\Lambda = 0 for \Sigma, 1 for \Pi, 2 for \Delta, and so on), and \Omega is the total angular momentum projection including spin-orbit coupling (\Omega = |\Lambda + \Sigma|, with \Sigma the spin projection). For homonuclear diatomic molecules, the symbol includes parity notation g (gerade, even) or u (ungerade, odd) to indicate behavior under spatial inversion through the molecular center, and for \Sigma states, a superscript + or - denotes reflection symmetry through a plane containing the axis. These symbols arise from the coupling of electron spins and orbital motions in the molecular field, differing from atomic spherical symmetry by emphasizing axial projections. For example, the ground state of O_2 is denoted ^{3}\Sigma_g^-, reflecting two unpaired electrons with \Lambda = 0, triplet spin multiplicity, even parity, and antisymmetric reflection; an excited state is ^{1}\Delta_g, a singlet with \Lambda = 2 and even parity. Similarly, the first excited state of O_2 is ^{1}\Sigma_g^+, illustrating how configuration interactions determine allowed terms. Vibronic states combine electronic terms with vibrational levels, denoted by appending the vibrational v (v = , 1, 2, ...) to the . The vibrational for a in an anharmonic potential is given by G(v) = \omega_e \left(v + \frac{1}{2}\right) - \omega_e x_e \left(v + \frac{1}{2}\right)^2, where \omega_e is the and \omega_e x_e the , both in units. For instance, the C^3\Pi_{0u}^+ (v=) state of N_2 represents the lowest vibrational level of the triplet \Pi state with \Omega = , ungerade parity, and positive . Rotational structure is incorporated by adding the total angular momentum quantum number J, with energy levels depending on the coupling scheme. In \Pi and \Delta states (\Lambda > 0), \Lambda-doubling splits each rotational level into a closely spaced pair due to interactions lifting the \pm \Lambda degeneracy, observable in high-resolution spectra. The appropriate coupling scheme for rotational and electronic angular momenta follows Hund's cases. Case (a) applies to light molecules where the spin-orbit interaction exceeds the rotational coupling (high \omega_e / B ratio, with B the rotational constant), preserving \Lambda and \Sigma as good quantum numbers before coupling to J. Case (b) suits heavier molecules or weak spin-orbit coupling, where rotation couples first to orbital motion (forming N), then to spin, effectively uncoupling spin from the axis.

Advanced Applications

Quarkonium States

Quarkonium refers to the bound states of a heavy and its antiquark, such as charmonium (c\bar{c}) and bottomonium (b\bar{b}), where the spectroscopic notation adapts the convention to describe their quantum mechanical properties under (QCD). The notation follows the form n^{2S+1} L_J, where n is the principal radial quantum number starting from 1, L denotes the orbital with letters S (L=0), P (L=1), D (L=2), etc., S is the total (0 for , 1 for triplet), and J is the ranging from |L - S| to L + S. These states possess definite P = (-1)^{L+1}, arising from the intrinsic parities of the quark-antiquark pair combined with the orbital contribution, and charge conjugation parity C = (-1)^{L+S} for neutral systems invariant under quark-antiquark exchange. The notation draws from and but incorporates relativistic and QCD effects for these tightly bound, heavy systems. For instance, the J/ψ meson, a charmonium state, is denoted as $1^3 S_1 (c\bar{c} with n=1, L=0, S=1, J=1), while its singlet partner η_c is $1^1 S_0 (n=1, L=0, S=0, J=0). Higher states include the χ_c family as $1^3 P_J (n=1, L=1, S=1, J=0,1,2) and the radial excitation ψ' as $2^3 S_1 (n=2, L=0, S=1, J=1). These labels facilitate comparison with experimental spectra observed in e^+ e^- collisions and hadron decays. Quarkonia were discovered in the 1970s, with the J/ψ observed independently by SLAC and Brookhaven teams in November 1974 at a mass of approximately 3.1 GeV, marking the "November Revolution" and evidence for the charm quark. The subsequent charmonium and bottomonium spectra confirmed the bound-state interpretation, with notation applied to organize states by their quantum numbers. The masses of quarkonium states follow a non-relativistic Schrödinger equation analogous to the hydrogen atom, but with a QCD-inspired Cornell potential V(r) = -\frac{\alpha}{r} + \sigma r, combining short-range Coulomb-like attraction from one-gluon exchange and linear confinement at large distances. This potential yields level splittings consistent with observed spectra, such as the ~117 MeV gap between J/ψ and η_c due to spin-spin interactions.

Spectral Series Notation

In the late 1880s and early 1890s, spectroscopists observed regular patterns of spectral lines in the emission and absorption spectra of alkali metals, such as sodium and potassium, which were grouped into distinct Rydberg series. These series were first systematically analyzed by Johannes Rydberg, building on earlier work by George Liveing and James Dewar, who identified sharp, principal, and diffuse groupings based on line appearance and wavelength regularity. A fourth series, the fundamental, was later recognized in 1907 by Arno Bergmann. In alkali spectra, the principal series involves transitions from np states (upper) to the ground ns state (lower), the sharp series from ns' (n'>n) to np, the diffuse series from nd to np, and the fundamental series from nf to nd, where n denotes the principal quantum number. The series were labeled using the initial letters S (sharp, l=0 for upper s states), P (principal, l=1 for upper p states), D (diffuse, l=2 for upper d states), and F (fundamental, l=3 for upper f states), reflecting the azimuthal quantum number l of the excited electron's orbital in the upper level. These empirical notations, introduced around 1890, predate and were based on the visual sharpness or diffuseness of lines on photographic plates, with sharper lines corresponding to s-like transitions and more diffuse ones to higher l. Rydberg formalized the wavenumber ν of lines in these series using the \nu = R \left( \frac{1}{n_1^2} - \frac{1}{n_2^2} \right), where R is the , n_1 is fixed for each series (e.g., n_1=2 for Balmer-like in ), and n_2 > n_1 varies, linking the patterns to discrete energy levels. This formula accurately predicted line positions across and spectra, with series converging to limits as n_2 → ∞. Prominent examples include the sodium D lines at 589.0 nm and 589.6 nm, the first members of the principal series arising from 3p → 3s transitions, which appear as a bright yellow in flame tests and dominate the visible sodium . In , the Balmer series in the visible region corresponds to P-like transitions (Δl=±1, upper l=1 relative to n=2 lower levels), analogous to the alkali principal series but simplified due to the single-electron nature. Quantum mechanically, these series arise from electric dipole-allowed transitions obeying the selection rule Δl = ±1 between states described by term symbols ^{2S+1}L, where L = S, P, D, F,... denotes the total (l for single in alkalis). For instance, principal series lines connect ^2P upper terms to ^2S lower terms, sharp series ^2S to ^2P, diffuse ^2D to ^2P, and fundamental ^2F to ^2D, with the emerging from the Coulomb potential's energy levels E_n = -\frac{h c R}{n^2}, where h is Planck's constant and c the . This interpretation, formalized in the 1920s by Bohr and others, unifies the empirical series with quantization, explaining line intensities and series limits via and conservation (ΔS=0).

References

  1. [1]
    Atomic Spectroscopy - Different Coupling Schemes | NIST
    Oct 3, 2016 · LS Coupling (Russell-Saunders Coupling). Some of the examples given below indicate notations bearing on the order of coupling of the electrons.Missing: definition | Show results with:definition
  2. [2]
    Atomic Term Symbols - Chemistry LibreTexts
    Jan 29, 2023 · The sequence of "s, p, d, f, g, h, i, k..." is known as the spectroscopic notation for atomic orbitals. By introducing spin as a nature of ...
  3. [3]
    Spectroscopic Notation
    Spectroscopic notation is a standard way to write angular momentum quantum numbers, using the form N^{2s+1}L_j, where N, s, L, and j are quantum numbers.Missing: definition | Show results with:definition
  4. [4]
    A Primer on Quantum Numbers and Spectroscopic Notation
    Quantum numbers are n (principal), l (azimuthal), m (magnetic), and s (spin). Spectroscopic notation describes the state of electrons in an atom or ion.Quantum Numbers · Spectroscopic Notation · Terms, Configurations, and...
  5. [5]
    Quantum Numbers and Electron Configurations
    These quantum numbers describe the size, shape, and orientation in space of the orbitals on an atom.Missing: m_s | Show results with:m_s
  6. [6]
  7. [7]
    [PDF] Atomic Spectroscopy - National Institute of Standards and Technology
    Mar 3, 1999 · The notation for Russell-Saunders basis states has been used only for concreteness; the eigenvectors may be expressed in any coupling scheme, ...Missing: original | Show results with:original
  8. [8]
    How Sommerfeld extended Bohr's model of the atom (1913–1916)
    Jan 30, 2014 · Sommerfeld's extension of Bohr's atomic model was motivated by the quest for a theory of the Zeeman and Stark effects.
  9. [9]
    [PDF] Spectroscopic Notation - Macmillan Learning
    Spectroscopic notation uses letters (spdfgh...) for angular momentum, shells (KLMNO...) for principal quantum number, and n, j, and s for quantum numbers. For ...Missing: definition | Show results with:definition
  10. [10]
    Spectroscopy - Orbital, Spin, Momentum - Britannica
    Oct 15, 2025 · The total orbital angular momentum is the sum of the orbital angular momenta from each of the electrons; it has magnitude Square root of√L(L + 1) ...
  11. [11]
    [PDF] 3 Atomic Electron Configurations and Periodicity
    Electron Configuration Notation (spectroscopic notation) shows: • Values for n and l for each orbital that is occupied by at least one electron. • Number of ...
  12. [12]
    4.6 Electronic configuration | The atom - Siyavula
    Using spectroscopic notation, the electron configuration of fluorine is 1s22s22p5 and the electron configuration of argon is 1s22s22p63s23p6.
  13. [13]
    [PDF] Principles of Chemical Science - MIT OpenCourseWare
    Sep 17, 2014 · Electron configurations can be simplified by considering core versus ... Mo and Ag are counterpart anomalies to Cr and Cu in 4th period.Missing: first 20
  14. [14]
    3.6 The Shape of Atomic Orbitals – Chemistry LibreTexts
    The s sublevel is composed of a single spherical orbital. The p sublevel is composed of 3 dumbbell shaped orbitals oriented along the x, y, and z axes.Missing: spd | Show results with:spd
  15. [15]
    [PDF] Electron Configurations
    Electron Configurations: Excited States. Which of the electron configurations to the right corresponds to phosphorous? Answer: c. The rest are excited states of.
  16. [16]
    NIST: Atomic Spectra Database - Spectral Lines Help File
    To indicate the spectrum of a given element, enter either a Roman numeral or an Arabic numeral after the element name. The Roman numerals must be separated ...
  17. [17]
    [PDF] Chapter 7 ATOMIC SPECTRA - Astronomy Research Centre (ARC)
    Oct 7, 2012 · Today, we use the symbols C I , or Fe I, or Zr I , etc., to denote the spectrum of the neutral atom, regardless of the source, and C II , C III ...
  18. [18]
    Alfred Fowler. 1868-1940 - jstor
    Lockyer had proposed a notation for such g example, had its lines represented by Sil,. SilV, corresponding to successive groups laboratory and represented ...
  19. [19]
    The Strongest Lines of Singly Ionized Atoms*
    In 1914 Fowler[2] definitely associated enhanced lines with ionized atoms of the elements. It is now known that each chemical element is characterized by Z ...
  20. [20]
    [PDF] STELLAR SPECTRA A. Basic Line Formation
    Feb 6, 2007 · For T = 5 000 K element E is predominantly neutral, for T = 10 000 K it is once ionized (E+), for higher temperature stages E2+ and E3+ appear ...
  21. [21]
    [PDF] A Practical Guide to the Partition Function of Atoms and Ions
    For example, the one-electron systems: H I, He II, and Li III with 1s configuration, two-electron systems: He I, Li II, Be III, B IV, and C V with. 1s2 ...
  22. [22]
    Ratios of forbidden [OIII] λλ4959,5007 and [NII] λλ6548,6583 lines in ...
    May 12, 2025 · Emission lines from ionized gas regions in galaxies provide important diagnostic tools for understanding their physical properties.
  23. [23]
    USING STRONG LINES TO ESTIMATE ABUNDANCES IN ...
    Because the two ions have quite different ionization potentials, the [N ii]/[O iii] ratio depends strongly on the ionization parameter. Thus, if this diagnostic ...
  24. [24]
    Termstruktur und Zeemaneffekt der Multipletts | Zeitschrift für Physik ...
    Landé, A. Termstruktur und Zeemaneffekt der Multipletts. Z. Physik 15, 189–205 (1923). https://doi.org/10.1007/BF01330473
  25. [25]
    Atomic Data for Carbon (C ) - Physical Measurement Laboratory
    C I Ground State 1s22s22p2 3P0. Ionization energy 90820.45 cm-1 (11.26030 eV) Ref. J66 C II Ground State 1s22s22p 2P°1/2. Ionization energy 196664.7 cm-1 ...Missing: term symbol
  26. [26]
    [PDF] Structure and spectra of diatomic molecules
    The symmetry labels σ, π, δ, . . . correspond to orbitals with values of 0 ... symmetry do not exist because no σ molecular orbital can. PCV ...
  27. [27]
    [PDF] Molecular Electronic Structure
    Spectroscopic notation for diatomic molecules. 32. D. States of other ... The states are designated σ, π, δ, etc. corresponding to ml = 0, 소1, 소2 ...
  28. [28]
    Molecular Spectra Vol I : Herzberg,Gerhard. - Internet Archive
    Nov 15, 2006 · Molecular Spectra Vol I. by: Herzberg,Gerhard. Publication date: 1950/00/00. Topics: NATURAL SCIENCES, Physics, Mechanics of gases.
  29. [29]
    [PDF] MOLECULAR CONFIGURATION AND STATES - eGyanKosh
    KK π π π σ σ σ. = where KK represent the filled K shell. The configuration can be re-written in familiar notation as: 1*. 4. 2. 2*. 2. 2*. 2. )1()1()3()2()2()1 ...
  30. [30]
    [PDF] The Spectrum of Molecular Nitrogen - Standard Reference Data
    This is a critical review and compilation of the observed and predicted spectroscopic data on the molecule N, and its ions N, N₂+, N₂²+, and the molecule N3.
  31. [31]
  32. [32]
    [PDF] 15. Quark Model - Particle Data Group
    May 31, 2024 · The C-parity can be generalized to the G-parity G = (−1)I+`+s for mesons made of quarks and their own antiquarks (isospin I z = 0), and for the ...
  33. [33]
    [PDF] Quarkonium physics: NRQCD factorization formula for J - Inspire HEP
    spectroscopic notation 2S+1LJ is used to represent a physical quarkonium state, where J is the total angular momentum quantum number. A physical quarkonium ...
  34. [34]
    [PDF] 15. Quark Model - Particle Data Group
    Dec 1, 2017 · The charge conjugation, or C-parity C = (−1)ℓ+s, is defined only for the q¯q states made of quarks and their own antiquarks.
  35. [35]
    [PDF] Charmonium Spectroscopy - INFN Ferrara
    The charmonium mass spectrum can be computed also within the framework of ... Spectroscopic Notation. n2S+1L. J. Page 31. Diego Bettoni. Charmonium. 31. The J ...
  36. [36]
    [PDF] Charmonium Spectroscopy - Duke Physics
    Aug 24, 2006 · Emission of soft gluons, followed by hadronization. Transition to other charmonium state with light hadron emission (ψ'→J/ψππ). Non perturbative ...
  37. [37]
    Getting the numbers right - the lonely struggle of Rydberg | Feature
    Jun 30, 2004 · Electron jumps between these various shells and the atomic 'ground state' were related to the sharp, principal, diffuse and fundamental line ...Missing: notation | Show results with:notation
  38. [38]
    [PDF] The X-ray spectra and the structure of the atoms - Nobel Prize
    For the alkali spectra, in particular, Rydberg found three different series, which he termed the principal series, the sharp, and the diffuse secondary series.