Fact-checked by Grok 2 weeks ago

X-ray notation

X-ray notation is a standardized system in and for designating the principal energy levels (shells) and subshells of atoms, as well as the characteristic spectral lines produced by transitions between these levels during emission or processes. It originated from early 20th-century studies and is essential for identifying structure and elemental composition in techniques like (XRF) and photoelectron spectroscopy. The notation primarily focuses on inner-shell s, using labels such as K for the 1s orbital, L for the 2s and 2p orbitals (subdivided into LI, LII, and LIII), and M for the 3s, 3p, and 3d orbitals (with five subshells), reflecting the quantum mechanical organization of atomic orbitals. Two primary notation systems are employed: the historical Siegbahn notation and the modern IUPAC notation. The Siegbahn system, introduced by in the early 1900s, labels spectral lines by the initial vacancy (e.g., K or L) followed by Greek letters indicating the filling and relative intensity, such as for the strongest transition from the L to the K vacancy, and for transitions from the M . This empirical approach, while widely used, lacks explicit reference to subshell quantum states and can be ambiguous for complex spectra. In contrast, the IUPAC notation, recommended since 1991, explicitly denotes transitions by connecting the initial (upper) and final (lower) energy levels with a , using for subshells (e.g., K-LIII for the 1s vacancy filled by a 2p3/2 , equivalent to 1 in Siegbahn notation). This system accommodates satellite lines from multiple ionizations (e.g., labeled as "sat n") and unresolved doublets (e.g., K-L2,3), promoting consistency across emission, absorption, and Auger spectroscopies. The adoption of X-ray notation has facilitated precise of spectrometers and of data, with key standards like the K-LIII line at 1.5405974 serving as benchmarks for measurements. Its application extends to fields such as , where it aids in non-destructive , and , for interpreting high-resolution spectra from cosmic sources.

Principles of notation

Shell and subshell designations

In X-ray notation, the principal electron shells are designated by letters that reflect their increasing levels and from the nucleus, starting with the innermost shell. The K shell corresponds to the principal n=1, the L shell to n=2, the M shell to n=3, the N shell to n=4, the O shell to n=5, and the P shell to n=6. These labels originated from early observations and are used to describe the binding energies of relevant to X-ray transitions. Subshells within these principal shells are further subdivided using numerical indices to distinguish orbitals based on their angular momentum and spin-orbit coupling, particularly important for inner shells where fine structure effects are pronounced. The K shell (n=1) consists of a single subshell with no further division. The L shell (n=2) is divided into three subshells: L1 corresponding to the 2s orbital, L2 to the 2p_{1/2} orbital, and L3 to the 2p_{3/2} orbital. The M shell (n=3) has five subshells: M1 (3s), M2 (3p_{1/2}), M3 (3p_{3/2}), M4 (3d_{3/2}), and M5 (3d_{5/2}). Higher shells like N, O, and P follow similar patterns with increasing numbers of subshells (seven for the N shell), but detailed subshell designations beyond M are less commonly specified in X-ray contexts. This notation is primarily applied to inner shells up to n=4 (N shell) or occasionally n=5 (O shell), as these have binding energies high enough (typically in the keV range) to produce X-ray emissions, whereas outer shells with lower binding energies are more relevant to optical or spectroscopy. For example, K-shell electrons in heavy elements like exhibit binding energies around 69 keV, enabling lines in the hard X-ray regime, while lighter elements like carbon have K-shell binding energies of only about 0.28 keV. An alternative labeling system, known as (e.g., 1s for K, 2s and 2p for L subshells), provides a more detailed description using quantum numbers but is often used alongside X-ray notation for clarity in .

Transition line notations

In X-ray notation, spectral lines arising from electron transitions between atomic shells are designated using a combination of series identifiers and Greek letters, with the series named after the lower-energy shell that receives the transitioning . The K-series encompasses transitions to the K shell (n=1), the L-series to the L shell (n=2, comprising 2s and 2p subshells), and the M-series to the M shell (n=3, including 3s, 3p, and subshells), following the principal shell designations. Within each series, Greek letters such as α, β, and γ denote specific transitions ordered by increasing (decreasing ), with α representing the lowest-energy transition from the adjacent higher , β from the next higher , and so on. For instance, in the K-series, the Kα lines result from electrons transitioning from the L to the , while Kβ lines involve transitions from the M to the , producing higher-energy photons due to the greater energy difference. Subscripts further distinguish from subshell splittings, typically with 1 indicating the transition involving the j=3/2 subshell and 2 the j=1/2 subshell. The Kα line, often the most intense in the K-series, exemplifies this notation as the aggregate of Kα₁ (from L₃ or 2p_{3/2} to K or 1s) and Kα₂ (from L₂ or 2p_{1/2} to 1s), where Kα₁ carries slightly higher energy than Kα₂ owing to the subshell differences. In the L-series, Lα denotes transitions from the M shell to the L₃ subshell, such as Lα₁ (M₅ or 3d_{5/2} to L₃) and Lα₂ (M₄ or 3d_{3/2} to L₃), while Lβ involves higher-energy transitions from N-shell subshells to L subshells. The M-series follows analogously, with Mα for N to M transitions, though these lines are generally weaker and occur at lower energies for heavier elements.
SeriesGreek LetterTypical TransitionExample (Mo, energies in eV)
Kα₁L₃ → K17,479
Kα₂L₂ → K17,374
Kβ₁M₃ → K19,608
Lα₁M₅ → L₃2,293
Lα₂M₄ → L₃2,290
Lβ₁N → L₃2,395

Relation to quantum mechanics

Correspondence to quantum numbers

The X-ray notation for atomic shells directly corresponds to the principal quantum number n and azimuthal quantum number l, with subshell splittings arising from the total angular momentum quantum number j due to spin-orbit interactions. The K shell is designated as the n=1 level with l=0, corresponding to the 1s orbital. For the L shell (n=2), the notation distinguishes subshells based on l and j: L1 represents the 2s orbital (l=0, j=1/2); L2 the 2p_{1/2} orbital (l=1, j=1/2); and L3 the 2p_{3/2} orbital (l=1, j=3/2). This splitting of the p subshell into j=1/2 and j=3/2 components is a direct consequence of spin-orbit coupling, where the interaction between the electron's spin and orbital angular momentum lifts the degeneracy of states with the same l but different j. The M shell (n=3) extends this mapping further: M1 is the 3s orbital (l=0, j=1/2); M2 and M3 are the 3p_{1/2} (l=1, j=1/2) and 3p_{3/2} (l=1, j=3/2) orbitals, respectively, again split by spin-orbit coupling; while M4 and M5 correspond to the (l=2, j=3/2) and (l=2, j=5/2) orbitals. Spin-orbit effects become more pronounced in these inner orbitals of heavier elements, where relativistic corrections significantly influence energy levels and transition probabilities. This notation primarily addresses inner shells up to n=4 (N shell), as higher principal quantum numbers (n>4) are less relevant for X-ray processes due to their lower binding energies and weaker relativistic influences. The focus on inner orbitals highlights the role of relativistic effects, including spin-orbit coupling, in determining the observed in X-ray spectra.

Conversion from spectroscopic notation

The conversion from , which uses the principal n, azimuthal l (with letters s for l=0, p for l=1, etc.), and total j for , to X-ray notation involves mapping the core shells and subshells based on their quantum mechanical designations. This process is essential for interpreting spectra, where the notation simplifies labeling of inner-shell transitions while retaining the underlying quantum structure. To perform the conversion step by step, first identify the principal n from the shell designation in X-ray notation, where K corresponds to n=1, L to n=2, M to n=3, N to n=4, O to n=5, P to n=6, and Q to n=7. Next, determine l from the subshell type: s indicates l=0, p indicates l=1, d indicates l=2, and f indicates l=3. Finally, account for j-splitting due to spin-orbit coupling in subshells with l \geq 1, where X-ray subshell labels (numbered I, II, III, etc., often written as 1, 2, 3) are assigned in order of increasing j within each shell, with ties resolved by increasing l. For example, the 2p subshell splits into $2p_{1/2} (j=1/2) and $2p_{3/2} (j=3/2), mapped to and , respectively. The following table illustrates key mappings for the K and L shells:
Spectroscopic notationX-ray notationnlj
1s101/2
2sL_I201/2
2p_{1/2}L_{II}211/2
2p_{3/2}L_{III}213/2
In multi-electron atoms, X-ray notation focuses on core levels (typically up to the M or N shell), ignoring electrons in higher n shells that do not participate in transitions. For instance, the core configuration of iron (Z=26) in is K: 1s², L: 2s² 2p⁶, which directly translates to X-ray notation as (fully occupied 1s), L_I (2s²), L_{II} (2p_{1/2}²), and L_{III} (2p_{3/2}⁴), with the 3d⁶ 4s² omitted. A common pitfall arises when using non-relativistic , which omits j-splitting; for light elements (low ), where spin-orbit is weak and splitting energies are small (e.g., <1 eV in the L shell for Z<20), approximations treat subshells like L_{II} and L_{III} as a single 2p without resolving the j levels.

Applications

In X-ray spectroscopy

In X-ray spectroscopy, X-ray notation serves as the standard system for labeling peaks in emission spectra generated by characteristic X-rays, which occur when an atom is ionized in an inner electron shell and a higher-energy electron from an outer shell de-excites to fill the vacancy, releasing a photon with discrete energy. This notation identifies the originating shell (e.g., K for the 1s orbital, L for 2s/2p) and the transition type (e.g., α for the strongest lines from L to K, β for transitions from M to K), enabling unambiguous assignment of spectral features to specific elements based on their unique transition energies. The system, rooted in early spectroscopic conventions, directly corresponds to the quantum mechanical description of electron shells while providing a practical shorthand for experimental data interpretation. These notations are integral to key analytical techniques, including energy-dispersive X-ray (EDX) spectroscopy and wavelength-dispersive X-ray (WDX) spectroscopy, where they facilitate elemental identification and quantification in diverse materials. In EDX, often integrated with scanning electron microscopy, the full spectrum of emitted X-rays is collected and analyzed by energy, with peaks labeled such as Cu Kα at 8.04 keV or Fe Lβ around 0.7–0.8 keV to pinpoint the presence and relative abundance of elements from beryllium to uranium. WDX, by contrast, employs diffraction gratings or analyzing crystals to isolate individual wavelengths with superior energy resolution (down to 5–10 eV), allowing precise measurement of weaker lines like Kβ or Lγ that might overlap in EDX spectra, thus enhancing accuracy for trace element detection in complex samples. The observed intensities of these labeled lines follow quantum selection rules for electric dipole transitions, which require a change in the orbital angular momentum quantum number of Δl = ±1 between initial and final states, along with Δj = 0, ±1 for the total angular momentum. This rule favors α lines, such as Kα₁ (from 2p_{3/2} to 1s) and Kα₂ (from 2p_{1/2} to 1s), over β lines like Kβ₁ (from 3p to 1s), because the former involve more probable l-changing transitions and higher orbital degeneracies in the p states, typically resulting in Kα:Kβ intensity ratios of about 8:1 to 10:1 depending on the element. These relative strengths provide diagnostic information on transition probabilities without needing full quantum calculations, aiding in spectrum deconvolution and validation of experimental setups. In contemporary applications, particularly scanning electron microscopy with energy-dispersive spectroscopy (SEM/EDS), X-ray notation ensures consistent reporting of compositional data across global laboratories, where spectra from alloys, semiconductors, or biological tissues are annotated with lines like Al Kα or Si Lα to map elemental distributions at micron scales with detection limits around 0.1–1 wt%. This standardization supports industries from materials science to forensics by enabling direct comparison of results from diverse instruments, while the notation's link to transition line designations simplifies database integration for automated analysis.

In atomic structure analysis

In theoretical atomic physics, X-ray notation plays a crucial role in by specifying core orbitals within multi-electron wavefunctions, particularly for modeling screening effects that influence inner-shell binding energies. The notation designates subshells such as K, L_I, L_II, and L_III based on principal quantum number n, orbital angular momentum \ell, and total angular momentum j, enabling precise construction of for relativistic self-consistent field methods like the approach. For instance, in computations of K-shell ionization potentials, screening by outer electrons is accounted for through the DHFS potential with corrections like Latter's tail, which adjusts the effective nuclear charge and yields binding energies in good agreement with experimental data for elements across the . Binding energy tables in atomic structure analysis often rely on empirical relations derived using X-ray notation to correlate subshell energies with atomic number Z. Moseley's law provides a foundational example, approximating the energy of K\alpha transitions—originating from L to K shell—as E \approx (Z-1)^2 \times constant, where the screening constant accounts for the incomplete shielding by the K-shell electron itself. This relation, E_{K\alpha} = 10.2 \, \text{eV} \times (Z-1)^2 in simplified form, facilitates the tabulation of core-level energies for predictive modeling, with the (Z-1) term reflecting effective nuclear charge in inner-shell processes. Relativistic effects become prominent in inner shells, where adjustments via the Dirac equation in Hartree-Fock frameworks reveal j-dependent energy shifts that X-ray notation explicitly highlights. For L-shell subshells, the relativistic splitting distinguishes L_{II} (j=1/2) from L_{III} (j=3/2), with the former exhibiting lower energies due to direct relativistic stabilization, while indirect effects dominate in higher-Z atoms. These shifts, computed using , scale with Z^4 for spin-orbit coupling and are essential for accurate wavefunctions in heavy elements. In plasma physics applications, X-ray notation simplifies the modeling of highly ionized atoms by denoting core configurations in stripped ions, aiding simulations of ionization equilibria and radiative processes under extreme conditions. For example, notations like describe hollow-ion states in dense hot plasmas, where core vacancies drive enhanced X-ray emission, and relativistic use these labels to parameterize electron configurations for spectral diagnostics in astrophysical environments.

Historical development

Origins in early X-ray research

The discovery of X-rays by in 1895 marked the beginning of intensive investigations into their properties and interactions with matter, laying the groundwork for the development of X-ray notation. Röntgen's observation of these penetrating rays, produced by electron impacts on a , prompted subsequent researchers to explore their spectral characteristics and absorption behaviors in various elements. This foundational work set the stage for systematic studies of X-ray emission and absorption spectra, which would eventually lead to empirical labeling systems for atomic energy levels. Between 1906 and 1910, Charles Glover Barkla conducted pivotal experiments on X-ray absorption and fluorescence, identifying distinct absorption edges that he designated as K and L series. Using filtered X-rays from tubes incident on elements, Barkla observed that each element emitted characteristic fluorescent radiations upon excitation, with the K series being significantly more penetrating (approximately 300 times harder) than the L series. These findings, derived from absorption measurements through varying thicknesses of materials, revealed that the hardness of these radiations increased with atomic weight and required increasingly energetic incident beams for excitation. Barkla's empirical classification of these series as K (for the most penetrating) and L (for the next) provided the initial framework for notating X-ray spectral lines based on their energy and penetration properties, without reliance on atomic theory. In 1913, Henry Gwyn Jeffreys Moseley advanced this empirical approach through precise measurements of X-ray emission spectra, focusing on the Kα lines across elements from calcium to zinc. Employing a crystal spectrometer to resolve wavelengths, Moseley demonstrated that the frequency ν of these Kα lines followed a linear relationship with the square root of the atomic number Z, expressed as √ν ∝ Z - b (where b is a screening constant). This periodicity directly tied spectral lines to atomic number rather than atomic weight, resolving inconsistencies in the periodic table and establishing the K, L, and later M shells as discrete energy levels associated with inner electrons. Moseley's work, conducted just before Niels Bohr's quantum model, solidified the shell notation as an empirical tool for ordering elements and predicting missing ones (such as Z=43, 61, 72, and 75). In the mid-1910s, particularly from 1916 onward, Manne Siegbahn refined the notation by resolving fine structures within the series through high-resolution crystal diffraction spectroscopy, with the detailed notation system formalized in the 1920s. Observing multiple sharp discontinuities in absorption spectra, Siegbahn identified three distinct edges within the L absorption, labeling them LI, LII, and LIII based on their decreasing energy (LI highest, LIII lowest). These subshell designations arose from the relative intensities and positions of lines in fluorescent spectra, providing a more granular empirical classification that distinguished transitions between closely spaced energy levels. Siegbahn's contributions, building on Barkla's series, enhanced the precision of X-ray notation in the pre-quantum era, linking observed spectral features to hierarchical atomic structures without theoretical underpinnings.

Modern standardization

The modern standardization of X-ray notation was primarily driven by the International Union of Pure and Applied Chemistry (IUPAC), which sought to create a systematic framework aligned with quantum mechanical principles for describing inner-shell s and transitions in . In 1991, IUPAC published recommendations establishing the IUPAC notation as the official system for X-ray emission lines and absorption edges, replacing the ad hoc Siegbahn notation with labels based on principal quantum numbers (K for n=1, L for n=2, etc.) and subshell designations using (e.g., LI for 2s, LII for 2p1/2, LIII for 2p3/2). This approach ensures precise correspondence to electron configurations and total angular momentum J, facilitating consistent reporting in and molecular studies. To address relativistic effects in heavy elements (Z > 50), the notation was refined to explicitly include j-values (total angular momentum of the subshell), distinguishing levels like the 2p1/2 (II) and 2p3/2 (III) where spin-orbit splitting becomes significant; these refinements draw from Dirac-Fock methods for calculating accurate energy levels and transition probabilities. For example, in elements like (Z=79), the L subshell transitions incorporate these j-dependent labels to account for relativistic corrections exceeding 10% in binding energies. As of 2025, the IUPAC notation remains the global standard, seamlessly integrated into computational resources such as the NIST Transition Energies Database, which employs labels like K-LIII for key transitions (e.g., 6403.8 eV for iron Kα1) and supports compatibility with software like Dirac and experimental tools in synchrotron-based . This ensures reliable data exchange and validation across international research efforts.

References

  1. [1]
    [PDF] 10.3.4.8 X-ray spectroscopy - iupac
    X-ray spectroscopy uses a new IUPAC notation based on energy levels, and characteristic X-ray emission consists of discrete spectral lines.
  2. [2]
    None
    ### Summary of X-ray Notation from "Theory of XRF" (https://www.chem.purdue.edu/xray/docs/Theory%20of%XRF.pdf)
  3. [3]
    [PDF] Table 1-1. Electron binding energies, in electron volts, for the ...
    Electron binding energies, in electron volts, for the elements in their natural forms. Element. K 1s. L1 2s. L2 2p1/2. L3 2p3/2. M1 3s. M2 3p1/2. M3 3p3/2. M4 ...Missing: heavy | Show results with:heavy
  4. [4]
    [PDF] X-Ray Data Booklet
    Electron binding energies, in electron volts, for the elements in their natural forms. Element. K 1s. L1 2s. L2 2p1/2. L3 2p3/2. M1 3s. M2 3p1/2. M3 3p3/2. M4 ...
  5. [5]
    [PDF] X-Ray Fluorescence (XRF): Understanding Characteristic X-Rays
    It is somewhat confusing, because three distinct naming systems are used for the atomic shells. Some authors refer to the energy levels (K, L1, L2, M1), some to ...
  6. [6]
    83.07.03: From The Inside Out: The Atomic Basis of Radiation
    The binding energy of the K shell of tungsten is approximately 69 keV, whereas the K shell energy of carbon is only 0.28 keV.
  7. [7]
  8. [8]
    [PDF] X-ray wavelengths and x-ray atomic energy levels
    Inconsistencies in accepted values (in x units) of x-ray reference lines have recently been demonstrated, although all.Missing: authoritative | Show results with:authoritative
  9. [9]
    [PDF] Part VIII. Nomenclature system for X-ray spectroscopy - iupac
    The aim of this document is therefore to present a new notation for X-ray emission lines and absorption edges to be called the IUPAC Notation. Surveys among X- ...
  10. [10]
    a short introduction to atomic structure
    = 1, j = 3/2 is designated as 2p3/2. There is also another notation that is commonly used in X-ray spectroscopy. Shells with n = 1, 2, 3, 4, 5, ...
  11. [11]
    [PDF] X-ray Studies of Quantum Materials
    Sep 3, 2017 · 2s L1. (2,0,1/2). 2p1/2. L2. (2,1,1/2). 2p3/2. L3. (2,1,3/2). 3s M1. (3,0,1/2). 3p1/2. M2. (3,1,1/2). 3p3/2. M3. (3,1,3/2). 3d3/2. M4. (3,2,3/2).
  12. [12]
  13. [13]
    [PDF] An introduction to X-ray photoelectron spectroscopy
    Nov 20, 2019 · The energy split tends to increase with increasing Z (atomic number) The widths may not be exactly the same – usually the higher BE peak is ...
  14. [14]
    [PDF] Energy Dispersive Spectroscopy - MyScope
    In spectroscopy, the most commonly used naming convention for Characteristic X-ray lines is the Siegbahn notation. The rst component of the name is the ...
  15. [15]
    1.12: An Introduction to Energy Dispersive X-ray Spectroscopy
    Aug 28, 2022 · Energy-dispersive X-ray spectroscopy (EDX or EDS) is an analytical technique used to probe the composition of a solid materials.
  16. [16]
    High-precision measurement of the X-ray Cu K-alpha spectrum | NIST
    May 12, 2017 · In this measurement, the region from 8000 to 8100 eV has been covered with a highly precise angular scale, and well-defined system efficiency.
  17. [17]
    Wavelength-Dispersive X-Ray Spectroscopy (WDS) - SERC (Carleton)
    A wavelength-dispersive spectrometer uses the characteristic X-rays generated by individual elements to enable quantitative analyses (down to trace element ...
  18. [18]
    X-Ray Emission - an overview | ScienceDirect Topics
    The selection rules applicable to optical dipole transitions also apply to X-ray transitions. The rules are ΔL=±1, Δj=0,±1. Intensity rules are also the same ...
  19. [19]
    Energy-Dispersive X-Ray Spectroscopy - an overview - ScienceDirect
    Energy-dispersive X-ray (EDX) spectroscopy utilizes the characteristic spectrum of X-rays that is emitted by a sample, following initial excitation by the high ...
  20. [20]
  21. [21]
  22. [22]
    (IUCr) Chapter 16. X-ray spectroscopy
    16.1.​​ Especially Barkla had been able, by this rather simple method, to find some fundamental properties of the X-rays when they were scattered by different ...
  23. [23]
    The dawn of X‐ray spectroscopy
    ### Summary of Early X-ray Spectroscopy Episodes (Pre-1913 Bohr Model)
  24. [24]
    Henry Moseley, X-ray spectroscopy and the periodic table - Journals
    Aug 17, 2020 · This paper provides an introduction to Moseley and his experiments and then traces attempts to 'discover' missing elements by X-ray spectroscopy.
  25. [25]
    [PDF] High-Frequency Spectra of the Elements. 703 - MIT
    By H. G. J. MOSELEY, M.A.*. THE. THE first part of this paper dealt with a method of photographing X-ray spectra, and included the spectra of a dozen elements ...Missing: alpha | Show results with:alpha
  26. [26]
    Manne Siegbahn wins Nobel Prize in Physics for x-ray spectroscopy
    Siegbahn initially used the same kind of spectrometer as Moseley when he started this line of research in 1914 but then made significant improvements when he ...Missing: origins notation development
  27. [27]
    [PDF] Nomenclature of Inorganic Chemistry | IUPAC
    ... Red Book I is the different organization of material, adopted to improve clarity. Thus, Chapters IR-5 (Compositional Nomenclature, and Overview of. Names of ...
  28. [28]