Fact-checked by Grok 2 weeks ago

Electron spectroscopy

Electron spectroscopy is a class of surface-sensitive analytical techniques that probe the electronic structure, , and bonding states of materials by measuring the energies of electrons emitted or scattered from a sample upon by photons, electrons, or other particles. These methods rely on the short of low-energy electrons in solids (typically 0.5–3 nm), enabling high-resolution analysis of the topmost atomic layers. Key variants include X-ray photoelectron spectroscopy (XPS), ultraviolet photoelectron spectroscopy (UPS), Auger electron spectroscopy (AES), and electron energy loss spectroscopy (EELS), each exploiting distinct and detection mechanisms to yield complementary data on core levels, valence bands, and vibrational modes. The foundational principle of electron spectroscopy is the , first observed in 1887 and theoretically explained by in 1905, wherein incident radiation ejects electrons whose (KE) relates to the (), (BE), and spectrometer (φ) via the equation BE = hν - KE - φ. In , for instance, soft X-rays (e.g., Al Kα at 1486.6 eV) ionize core electrons, producing photoelectrons whose binding energies identify elements (except H and He) and reveal chemical shifts due to oxidation states or bonding environments, with detection limits around 0.1 atomic%. employs photons (e.g., He I at 21.2 eV) to map valence and conduction band densities of states, aiding studies of molecular orbitals and Fermi levels. , meanwhile, detects characteristic Auger electrons emitted during core-hole relaxation, offering elemental sensitivity and spatial resolution when combined with scanning electron microscopy () for imaging. All techniques typically require (UHV) conditions to minimize surface contamination, though near-ambient pressure () variants extend applicability to reactive environments. Pioneered by in the 1950s–1960s through his development of high-resolution electron spectroscopy for chemical (ESCA, now synonymous with ), the field earned him the 1981 . Applications span , , and , including characterization of thin films, interfaces, polymers, semiconductors, and electrocatalysts—such as determining oxidation states in electrodes or adsorption sites on surfaces. Depth profiling via and angle-resolved measurements further enhance capabilities for multilayer , making electron spectroscopy indispensable for understanding surface phenomena that govern reactivity and performance.

Fundamentals

Definition and Scope

Electron spectroscopy refers to a class of analytical techniques that investigate the , , and bonding characteristics of materials by measuring the and angular distributions of electrons emitted or scattered from a sample's surface. These methods, including variants such as and , exploit electron-matter interactions to provide detailed surface-specific information. The scope of electron spectroscopy is characterized by its high surface , arising from the limited of low-energy electrons, typically in the range of 20-2000 , which have short inelastic mean free paths of about 0.5-3 in solids. This results in probing depths of approximately 1-10 , allowing analysis of the outermost layers without significant contribution from the . Techniques within this field are applicable to diverse sample types, including solids, liquids, and gases, and can operate under conditions or, in advanced setups, near-ambient pressures to study realistic environments. In contrast to optical spectroscopies, which rely on interactions and often probe deeper into materials, or neutron-based methods that emphasize properties, electron spectroscopy delivers element-specific data with sensitivity to chemical shifts and electronic states, enabling precise insights into surface chemistry and adsorption phenomena. Many electron spectroscopies are grounded in the , where incident radiation ejects electrons from atomic orbitals.

Core Principles of Electron Emission

Electron emission in electron spectroscopy primarily arises from the interaction of a sample with either photons or an incident electron beam, resulting in inelastic scattering processes that eject s from atomic orbitals. In photon-induced emission, such as in photoelectron spectroscopy, incident photons with sufficient energy excite s, leading to their ejection if the photon energy exceeds the of the . Similarly, electron beam excitation, as employed in techniques like , involves primary s penetrating the surface and undergoing inelastic collisions that transfer energy to core or valence s, causing their emission. These processes generate photoelectrons or Auger s with characteristic kinetic energies that reflect the electronic structure of the material. The serves as the foundational mechanism for photon-driven emission. The of these emitted s is a key parameter for spectroscopic and is typically measured using electron analyzers. Hemispherical analyzers, which employ electrostatic deflection to separate electrons by , offer high and are widely used for precise measurements across a range of energies. Retarding field analyzers, on the other hand, operate by applying a retarding potential to filter electrons based on their ability to overcome an electrostatic barrier, providing simpler implementation for lower- applications. Both types of analyzers detect electrons after they traverse a path, minimizing and ensuring accurate determination. From the measured kinetic energy, the binding energy of the emitted electron can be determined using the energy conservation principle. For photon excitation, the binding energy E_{\text{binding}} is given by E_{\text{binding}} = h\nu - E_{\text{kinetic}} - \phi where h\nu is the energy of the incident photon, E_{\text{kinetic}} is the measured kinetic energy of the photoelectron, and \phi is the work function of the spectrometer or sample surface. This equation accounts for the energy required to overcome the surface potential barrier, enabling the mapping of electron binding energies to specific atomic or molecular orbitals. In electron beam excitation, such as in Auger electron spectroscopy, the kinetic energy of the Auger electron is characteristic of the element and its chemical environment, given by KE \approx BE (initial ionized level) - BE (two final levels) - \phi, and does not depend on the incident beam energy provided it is sufficient to create the initial core hole. The surface specificity of electron spectroscopy stems from the limited of emitted , governed by their (IMFP), which represents the average distance an travels before undergoing an energy-loss event. For with kinetic energies between 50 and 1000 , the IMFP in most solids is approximately 0.5–2 , confining the probed region to the outermost layers. This short path length arises from frequent interactions with the 's and phonons, making the technique ideal for surface and while contributions are negligible. Experimental databases confirm these values vary modestly with composition and energy, with minima near 50–100 . Beyond energy, the angular distribution of emitted electrons provides insights into the emission geometry and orbital symmetry, often exhibiting cosine-like patterns relative to the surface normal due to momentum conservation. Spin polarization, which measures the imbalance in spin-up and spin-down electron populations, adds further detail on magnetic properties and spin-orbit coupling effects during emission. These parameters are quantified using specialized analyzers with angular and spin filters, enhancing the of electronic states.

Theoretical Foundations

Photoelectric Effect

The photoelectric effect, in which electrons are emitted from a material upon absorption of photons, provided the foundational quantum mechanical basis for electron spectroscopy. In 1905, extended Max Planck's 1900 concept of energy quantization to explain the effect, proposing that light consists of discrete energy packets (, later termed photons) with energy h\nu, where h is Planck's constant and \nu is the . Einstein demonstrated that photoemission occurs only when the photon energy exceeds the electron's , with the excess energy appearing as the photoelectron's , resolving classical wave theory's inability to account for the effect's threshold behavior and intensity dependence on rather than . Quantum mechanically, the photoelectric process involves the of a by a bound , leading to its ejection into the if h\nu > E_b, where E_b is the . The for this is governed by , which in the dipole approximation yields the probability per unit time as W = \frac{2\pi}{\hbar} |\langle f | \hat{H}_\text{int} | i \rangle|^2 \rho(E_f), where |i\rangle and |f\rangle are the initial bound and final states, \hat{H}_\text{int} = -\vec{\mu} \cdot \vec{E} is the interaction Hamiltonian with electric operator \vec{\mu}, and \rho(E_f) is the density of final states at energy E_f = h\nu - E_b. This perturbative treatment assumes weak coupling and a of final states, enabling quantitative predictions of emission probabilities central to spectroscopic techniques. The photoionization cross-section \sigma, proportional to the transition matrix element squared, determines the likelihood of ejection and follows dipole selection rules derived from the angular momentum conservation in the \vec{r} \cdot \vec{E} interaction term. For atomic orbitals, these rules dictate changes in orbital angular momentum \Delta l = \pm 1 and \Delta m = 0, \pm 1, forbidding certain transitions like s to s or p to p while allowing s to p or p to s/d. Such rules influence the observable spectral features, with cross-sections typically peaking near threshold and decreasing at higher energies due to the radial overlap integral's behavior. In core-level spectroscopy, photoemission probes tightly bound inner-shell electrons (e.g., 1s or 2p), requiring higher energies (keV range) and yielding sharp lines due to localized orbitals, whereas valence-level targets outer electrons (eV range) with broader features from delocalized bands. Accompanying the primary photoemission, secondary processes like shake-up (discrete excitation of a to an unoccupied state) and shake-off (continuum ionization of another electron) produce satellite peaks shifted by 5–20 eV, arising from the sudden change in potential during core-hole creation under the sudden approximation. These satellites provide insights into many-body interactions, with intensities scaling as the square of the overlap between initial and final many-electron wavefunctions. Photoionization intensities vary with and Z, often exhibiting minima due to destructive interference in the matrix element. A prominent example is the Cooper minimum, observed in transitions like np to εd in atoms such as , where the cross-section drops sharply (e.g., by factors of 10–100) around 20–50 eV as the radial wavefunctions' nodes cause near-zero overlap. This Z-dependent phenomenon, shifting to higher energies for larger Z, aids in identifying orbital contributions and probing electronic structure across the periodic table.

Auger Effect and Other Processes

The Auger effect is a fundamental relaxation mechanism in electron spectroscopy, occurring after the creation of a core-level vacancy in an . In this process, an electron from a higher-energy orbital fills the core hole, releasing energy that ejects a second from the same or another orbital, termed the Auger . This leaves the in a doubly ionized state and produces electrons with discrete kinetic energies characteristic of the emitting 's electronic structure. The kinetic energy of the Auger electron is determined by the difference in binding energies of the involved levels, adjusted for the surface work function: E_{\text{kinetic}} = E_B^{(\text{initial})} - E_B^{(\text{Auger})} - E_B^{(\text{final})} - \phi where E_B^{(\text{initial})} is the binding energy of the initial core hole, E_B^{(\text{Auger})} is the binding energy of the filling electron's orbital, E_B^{(\text{final})} is the binding energy of the orbital from which the Auger electron originates, and \phi is the work function (typically 4–5 eV). This energy is independent of the initial excitation method, providing a robust identifier for elements. The Auger process competes with radiative decay via , but for light elements ( Z < 30), the Auger yield dominates because fluorescence yields decrease sharply with decreasing Z, while Auger probabilities increase. Seminal compilations confirm fluorescence yields below 0.1 for Z \approx 10, making Auger emission the primary pathway. Auger transitions exhibit chemical shifts in kinetic energy due to changes in the local electronic environment, such as oxidation state or bonding, which alter relaxation energies. The Auger parameter, defined as \alpha = E_{\text{kinetic}} + E_{\text{binding}} (where E_{\text{binding}} is the binding energy of the initial core level from photoelectron spectroscopy), quantifies these shifts and is particularly useful for distinguishing chemical states, as it is less affected by instrumental factors like photon energy or surface charging. For example, in sulfur compounds, shifts up to 7 eV reflect valence electron involvement. Related multi-electron processes include Coster-Kronig transitions, which are super-Auger decays within subshells of the same principal quantum number (e.g., L₂ to L₃), producing additional low-energy electrons and enhancing ionization cascades. These transitions have high probabilities in mid-Z elements and contribute to spectral complexity. Beyond core-hole relaxation, other electron emission processes observed in spectroscopy arise from electron-matter interactions. Secondary electron emission involves cascades of low-energy electrons (<50 eV) generated by inelastic scattering of primary electrons with valence or conduction band electrons, forming a broad background in spectra. Plasmon excitation refers to collective oscillations of the electron density, leading to discrete energy losses (typically 10–25 eV for bulk plasmons) that appear as peaks or satellites in emission spectra. Bremsstrahlung, or braking radiation, produces a continuum of low-energy photons and electrons from the deceleration of charged particles, contributing to nonspecific backgrounds in both electron and X-ray spectra.

Historical Development

Early Discoveries

The discovery of electron emission phenomena began in the late 19th century, laying the empirical groundwork for what would later develop into electron spectroscopy. In 1887, observed that ultraviolet light incident on a metal surface facilitated the discharge of sparks across an air gap in his electromagnetic wave experiments, an effect he attributed to the removal of negative electricity from the illuminated electrode, marking the first documented instance of photoelectric emission. This observation, while not fully understood at the time, demonstrated that light could eject charged particles from matter, challenging classical wave theories of light. Building on Hertz's findings, Philipp Lenard conducted detailed experiments in 1902 to investigate the properties of these emitted electrons. Using a vacuum apparatus with a photocathode and collector, Lenard measured the range and energy of photoelectrons emitted from alkali metals under ultraviolet illumination, finding that the electrons penetrated only a short distance into air or other materials before losing energy, with maximum kinetic energies on the order of 0.5–2 electron volts depending on the light frequency. These results highlighted the quantized nature of the emission process and the dependence of electron energy on light frequency rather than intensity, though Lenard initially interpreted them through classical models. A pivotal theoretical advancement came in 1905 when Albert Einstein provided a quantum mechanical explanation for the photoelectric effect in his seminal paper "On a Heuristic Viewpoint Concerning the Production and Transformation of Light." Einstein proposed that light consists of discrete energy packets (quanta, later called ) with energy E = h\nu, where h is and \nu is the frequency, and that electrons are emitted only if this energy exceeds the metal's work function \phi, yielding maximum kinetic energy K_{\max} = h\nu - \phi. This work, which bridged with observable emission, earned Einstein the 1921 Nobel Prize in Physics for his services to theoretical physics, particularly the photoelectric law. Experimental confirmation of Einstein's equation arrived in 1916 through the meticulous work of , who used a photoelectric setup with sodium and lithium electrodes under monochromatic light to measure stopping potentials and thus electron energies. Millikan's data across various frequencies yielded a precise value for h = 6.57 \times 10^{-27} erg-seconds, aligning closely with Einstein's prediction within 0.5% accuracy and validating the linear frequency dependence of K_{\max}. Despite his initial skepticism toward the quantum light concept, Millikan's verification established the equation's empirical foundation. Parallel to photoelectric studies, early 20th-century investigations into thermionic emission provided initial measurements of electron energies from heated surfaces. Owen Richardson, starting in 1901, systematically quantified the temperature dependence of electron emission from hot filaments, reporting in 1911–1916 that the saturation current followed I = A T^2 e^{-\phi / kT}, where \phi is the work function, and deriving electron energy distributions that revealed average kinetic energies scaling with temperature. These efforts, extended by collaborators like John Ambrose Fleming in the 1910s–1920s, enabled the first spectroscopic-like analyses of electron velocities using retarding fields, contributing essential techniques for energy-resolved emission studies.

Key Milestones and Modern Evolution

The , underlying (AES), was discovered independently in the 1920s by and through observations of electron emission following ionization. Practical AES emerged in the 1950s with the use of retarding field analyzers for surface analysis, and advanced in the 1960s with the introduction of cylindrical mirror analyzers, enabling routine elemental mapping. In the 1950s, Kai Siegbahn pioneered high-resolution X-ray photoelectron spectroscopy (XPS), developing electron analyzers that achieved energy resolutions below 0.1 eV, enabling precise determination of atomic binding energies and chemical shifts in solids. His work on (ESCA), which integrated XPS with chemical state identification, laid the foundation for surface-sensitive elemental analysis and earned him the 1981 Nobel Prize in Physics, shared with Nicolaas Bloembergen and Arthur Schawlow for contributions to laser spectroscopy. During the 1960s, David W. Turner advanced ultraviolet photoelectron spectroscopy (UPS) by constructing instruments that measured valence electron binding energies with resolutions around 0.01 eV using helium discharge lamps, providing insights into molecular orbital structures and band alignments in gases and solids. Turner's innovations, including time-of-flight detection for rapid spectral acquisition, facilitated the study of valence bands and vibronic coupling, marking a shift toward gas-phase and molecular applications of photoemission. In the 1970s, the integration of (AES) with (SEM) enabled spatially resolved surface chemical mapping at micrometer scales, as demonstrated by early implementations where AES spectrometers were mounted within SEM chambers to correlate topography with elemental composition. This development, exemplified by MacDonald and coworkers' 1970 demonstration of Auger spectra in SEM, expanded AES from bulk analysis to imaging applications in materials science, such as defect detection in alloys. From the 1990s to the 2000s, synchrotron radiation sources revolutionized photoemission spectroscopy by providing tunable, high-brightness X-ray beams with flux densities exceeding 10^12 photons/s/mm²/mrad²/eV, surpassing laboratory sources by orders of magnitude and enabling angle-resolved studies with sub-meV resolution. Facilities like the and facilitated bulk-sensitive photoemission and time-resolved experiments, driving discoveries in correlated electron systems and surface magnetism during this era. Post-2020 advancements have further enhanced electron spectroscopy's capabilities for real-world conditions and complex data handling. Ambient-pressure XPS (AP-XPS) has matured for operando studies, allowing spectral acquisition at pressures up to 1 mbar to monitor catalytic surfaces under reaction environments, as shown in 2025 investigations of atomic layer deposition chemistry for TiO₂. Spin-resolved angle-resolved photoemission spectroscopy (ARPES) has advanced topological material characterization, achieving spin-polarization mapping with efficiencies improved by Gaussian process regression, revealing helical edge states in insulators like Bi₂Te₃ as of 2024. Machine learning integration for spectral analysis has accelerated interpretation, with models like generative adversarial networks augmenting electron energy loss spectroscopy (EELS) datasets to predict fine structures, reducing analysis time from hours to minutes in 2024 studies. Additionally, aberration-corrected EELS has pushed spatial resolutions to sub-angstrom levels and energy resolutions to a few meV through monochromator enhancements, enabling atomic-scale mapping of phonons and bonds in 2D materials by 2025.

Techniques

X-ray Photoelectron Spectroscopy (XPS)

X-ray photoelectron spectroscopy (XPS), also known as electron spectroscopy for chemical analysis (ESCA), is a surface-sensitive technique that utilizes soft X-rays to ionize core-level electrons in atoms, enabling the determination of elemental composition, chemical bonding states, and electronic structure within the top few nanometers of a sample surface. The method relies on the photoelectric effect, where incident photons eject photoelectrons whose kinetic energies reveal the binding energies of the core electrons. Developed primarily by Kai Siegbahn in the mid-20th century, XPS has become a cornerstone for non-destructive surface analysis in materials science and chemistry. In XPS, a monochromatic soft X-ray source, typically aluminum Kα radiation at 1486.6 eV, irradiates the sample, ejecting photoelectrons from core orbitals such as 1s, 2p, or 3d levels. The kinetic energy (KE) of these photoelectrons is measured using an electron energy analyzer, and the binding energy (BE) is calculated via the relation BE = hν - KE - Φ, where hν is the X-ray photon energy and Φ is the spectrometer work function. Spectral features include sharp core-level peaks characteristic of each element, with positions insensitive to the sample's bulk properties but highly sensitive to the local chemical environment. Chemical shifts in peak positions, typically ranging from 1 to 5 eV, arise from variations in electron density due to bonding or oxidation states; for instance, the binding energy of carbon 1s increases by about 2-3 eV when transitioning from elemental carbon to oxidized forms like carboxyl groups. Valence band features, appearing at lower binding energies (0-20 eV), provide insights into the density of states near the , though they are broader and less element-specific. Quantitative analysis in XPS involves integrating the peak areas after background subtraction to determine atomic concentrations. The relative concentration C_x of element x is given by C_x = \frac{I_x / S_x}{\sum (I_i / S_i)}, where I_x is the integrated peak intensity for element x, and S_x is the sensitivity factor accounting for photoionization cross-section, transmission efficiency, and escape depth. Sensitivity factors are empirically derived or theoretically calculated, with values standardized in databases like those from for theoretical estimates or for experimental data, achieving atomic concentration accuracies of ±10%. This enables precise compositional mapping, such as determining oxidation states in alloys or thin films. Depth profiling extends XPS beyond surface analysis by varying the angle of photoelectron emission or using ion sputtering. In angle-resolved XPS (ARXPS), the takeoff angle is adjusted to probe depths from 1 to 10 nm, as the inelastic mean free path of photoelectrons is approximately 1-3 nm in most materials, with 95% of signal originating from within 3λ. Sputtering with inert ions like argon removes layers sequentially, revealing subsurface composition at rates of about 0.1-2 nm per minute, though this can introduce artifacts like preferential etching. These capabilities make XPS ideal for interface studies in semiconductors or catalysts. Despite its strengths, XPS has limitations, particularly with insulating samples where positive charging shifts peak positions, requiring charge neutralization via low-energy electron flood guns. Radiation damage from prolonged X-ray exposure can alter sensitive organics or biomolecules, while the technique's vacuum requirements (typically <10^{-9} Torr) preclude in-situ analysis of liquids or reactive gases without specialized setups. Overall, XPS provides unparalleled chemical specificity for surface characterization.

Ultraviolet Photoelectron Spectroscopy (UPS)

Ultraviolet photoelectron spectroscopy (UPS) is a powerful technique for investigating the valence electronic structure of materials and molecules by measuring the kinetic energies of photoelectrons ejected upon absorption of ultraviolet photons. Unlike higher-energy methods, UPS employs photon energies typically in the range of 10–50 eV to selectively probe valence orbitals, offering high energy resolution for mapping electronic states near the or highest occupied molecular orbital (HOMO). The process is governed by the , where the binding energy of an electron is determined from E_B = h\nu - E_{\text{kin}} - \phi, with h\nu as the photon energy, E_{\text{kin}} the kinetic energy, and \phi the work function. Common excitation sources in UPS include helium discharge lamps, such as the He I line at 21.2 eV and He II at 40.8 eV, which provide monochromatic UV radiation suitable for ejecting valence electrons from gases, solids, or surfaces. These sources enable the acquisition of spectra that reflect the valence band density of states (DOS), revealing the distribution of occupied electronic levels and facilitating the identification of orbital symmetries through comparison with theoretical calculations. Orbital assignments are often aided by Koopmans' theorem, which approximates the ionization potential (IP) of an electron from a molecular orbital as the negative of its Hartree-Fock eigenvalue, IP ≈ −ε_HOMO, providing a direct link between experimental spectra and computed orbital energies under the frozen-orbital approximation. Angular-resolved UPS (ARUPS) extends the technique by collecting photoelectrons as a function of emission angle, allowing the determination of band dispersions in crystalline materials. The in-plane momentum component of the photoelectron, conserved parallel to the surface, is calculated using \mathbf{k}_\parallel = \frac{\sqrt{2m E_{\text{kin}}}}{\hbar} \sin\theta, where m is the electron mass, \hbar is the , and \theta is the polar emission angle relative to the surface normal; this enables mapping of E(\mathbf{k}) relations for valence bands. ARUPS is particularly valuable for studying momentum-dependent electronic properties in semiconductors and metals, revealing details such as effective masses and Fermi surface topologies. UPS finds widespread applications in analyzing gas-phase molecules, where it elucidates molecular orbital energies and vibronic coupling, and on solid surfaces, where it probes adsorption-induced band shifts and interface electronics. A related variant, He* Penning ionization, uses metastable helium atoms (excited to 19.8 or 23.6 eV) to induce electron emission from surfaces, offering complementary sensitivity to outer valence states and orientation effects in adsorbates. Energy resolution in conventional UPS typically ranges from 10–50 meV, limited by source linewidth and analyzer performance, while time-resolved UPS variants, employing femtosecond laser pump-probe setups, capture ultrafast dynamics such as charge transfer or relaxation processes on picosecond timescales.

Auger Electron Spectroscopy (AES)

Auger electron spectroscopy (AES) is a surface-sensitive technique that analyzes the energy of Auger electrons emitted from a sample following excitation by a focused electron beam, enabling elemental composition mapping at the nanoscale. The method relies on the Auger effect, where an incident electron ejects a core-level electron, creating a vacancy that is subsequently filled by a higher-energy electron, with the energy difference transferred to another electron emitted as an Auger electron. These Auger electrons, with kinetic energies typically in the range of 20–2000 eV, escape only from the top 1–10 nm of the surface due to their limited mean free path in solids, providing high surface specificity. The excitation source in AES is a high-energy electron beam, usually operating at 1–30 keV, which ionizes core levels such as K, L, or M shells in atoms. This leads to Auger transitions predominantly of the core-valence-valence (CVV) type, involving two valence electrons, or core-core-valence (CCV) for deeper levels, with the kinetic energy of the emitted Auger electron given by E_k = E_1 - E_2 - E_3 - \phi, where E_1 is the binding energy of the initial core vacancy, E_2 and E_3 are the binding energies of the participating electrons, and \phi is the work function (as derived in the section). Spectra are acquired by measuring the energy distribution of these emitted electrons using analyzers like cylindrical mirror analyzers or hemispherical analyzers. In direct mode, the spectrum plots electron intensity N(E) versus kinetic energy E, but it often suffers from high background noise; to mitigate this, derivative mode is commonly employed, plotting dN(E)/dE versus E via lock-in amplification with a modulated beam, which enhances peak visibility and reduces noise. Quantification in AES is achieved by measuring the peak-to-peak heights in the derivative spectra for characteristic , normalized by sensitivity factors specific to each element, which account for ionization cross-sections, escape depths, and detector efficiencies. Sensitivity factors vary by about an order of magnitude across the , with transition metals exhibiting particularly high sensitivity due to favorable transition probabilities and peak intensities. For example, the sensitivity factor for is approximately 1.0 (reference), while for it is around 0.2, allowing detection limits of 0.1–1 atomic percent for most elements. This relative quantification is semi-quantitative without standards but provides reliable compositional ratios for surface layers. AES excels in spatial resolution, achieving ≈5–10 nm in hybrid scanning electron microscopy-AES (SEM-AES) systems, where the electron beam is raster-scanned across the surface to generate elemental maps and line scans via Auger peak intensities at each pixel. This enables imaging of nanoscale features, such as nanoparticle distributions or thin film interfaces, with lateral resolution limited primarily by beam diameter and interaction volume. However, artifacts can compromise data quality: beam-induced damage from high-energy electrons may cause atomic displacement, bond breaking, or preferential sputtering, particularly in organics or beam-sensitive materials, often mitigated by lowering beam energy or current. Additionally, overlap with low-energy secondary electrons contributes to spectral background, though derivative mode helps distinguish true Auger signals from this continuum.

Electron Energy Loss Spectroscopy (EELS)

Electron energy loss spectroscopy (EELS) is a transmission-based analytical technique integrated with transmission electron microscopy (TEM) that examines electronic and vibrational excitations in materials by recording the energy losses of high-energy electrons passing through ultrathin samples. The method employs an incident electron beam with energies typically ranging from 100 to 300 keV, directed at specimens thinner than approximately 100 nm to ensure primarily single inelastic scattering events and high signal-to-background ratios. Energy losses, spanning 0.1 eV to over 1000 eV, arise from interactions such as valence electron excitations and core-shell ionizations, providing detailed insights into the sample's composition, electronic structure, and local chemistry. The EELS spectrum features distinct components that reveal specific material properties. The zero-loss peak corresponds to elastically scattered electrons that traverse the sample without energy transfer. In the low-loss region (typically <50 eV), plasmon losses dominate, including bulk plasmons from volume collective oscillations of valence electrons and surface plasmons at interfaces, which yield information on electron density and dielectric response. Higher-energy core-loss edges display fine structure termed (ELNES), which probes the unoccupied density of states and bonding coordination around core-excited atoms, enabling differentiation of chemical environments such as oxidation states in transition metals. Quantitative analysis of EELS data extracts key physical parameters. The bulk plasmon energy \omega_p follows the Drude model relation \omega_p = \sqrt{\frac{n e^2}{\epsilon_0 m}}, where n denotes the valence electron density, e the elementary charge, \epsilon_0 the vacuum permittivity, and m the electron mass; measured peak positions thus allow estimation of n for metals and semiconductors. Sample thickness t, crucial for accurate interpretation, is determined relative to the inelastic mean free path \lambda via \frac{t}{\lambda} = \ln\left(\frac{I_\text{total}}{I_0}\right), with I_\text{total} the integrated spectrum intensity and I_0 the zero-loss peak area, assuming Poisson statistics for scattering. When coupled with scanning TEM (STEM), EELS achieves spatial resolutions below 1 nm, particularly with aberration-corrected optics that minimize probe delocalization, facilitating nanoscale elemental mapping and spectrum imaging of nanostructures. This high resolution supports applications in mapping chemical variations across interfaces or defects. Compared to emission spectroscopies, EELS excels in analyzing insulators due to the transmission geometry avoiding surface charging and enables momentum-transfer q-resolved studies by varying collection angles, revealing excitation dispersion relations inaccessible in zero-momentum techniques.

Applications

Surface and Interface Analysis

Electron spectroscopy techniques, such as (XPS) and (AES), are essential for characterizing the elemental composition and chemical states at material surfaces and interfaces, owing to their high surface sensitivity arising from the short inelastic mean free paths of emitted electrons, typically 0.5–3 nm. , in particular, enables the identification of elements present on surfaces to depths of about 5–10 nm with a sensitivity of 0.1–1 at.%, excluding hydrogen and helium, by measuring the binding energies of photoelectrons from core levels. Chemical shifts in these binding energies reveal oxidation states; for instance, the Ti 2p peaks in titanium oxides distinguish metallic Ti (Ti⁰ at ~454 eV), TiO (Ti²⁺ at ~455 eV), Ti₂O₃ (Ti³⁺ at ~457 eV), and TiO₂ (Ti⁴⁺ at ~459 eV), allowing precise determination of surface oxidation. A key application involves quantifying oxide layer thicknesses on surfaces through photoelectron attenuation. In XPS, the intensity of substrate signals decreases exponentially with overlying oxide depth d, following I = I_0 \exp(-d / \lambda \sin \theta), where \lambda is the inelastic mean free path and \theta is the takeoff angle; this method, often using the Strohmeier equation for relative peak intensities, has measured oxide thicknesses as low as 1.26 ± 0.068 nm on silicon spheres. Such analyses are critical for assessing surface passivation and reactivity in materials like metals and semiconductors. In adsorption and catalysis studies, AES and XPS quantify adsorbate coverage and interactions on catalytic surfaces. AES determines coverage \theta of submonolayer adsorbates by monitoring the attenuation of substrate Auger signals, approximated as \theta = 1 - \exp(-d / \lambda), where d is the adsorbate layer thickness and \lambda is the mean free path of electrons in the adsorbate; this approach informs models of adsorbate-substrate bonding. In catalysis, these techniques track adsorbate-induced changes, such as oxygen adsorption on metal catalysts, revealing site-specific reactivity and coverage effects on reaction kinetics. Interface phenomena, including band alignment and diffusion profiles, are probed using XPS to evaluate heterostructure performance. The Kraut method in XPS calculates valence band offsets (VBO) from core-level binding energies and valence band maxima of thin and thick films; for example, a VBO of 3.49 ± 0.08 eV was determined at the Ga₂O₃/Si(111) interface, confirming type I alignment suitable for optoelectronic devices. Depth profiling via angle-resolved XPS or sputtering reveals diffusion profiles, such as intermixing at oxide-semiconductor boundaries, which influences charge transport. In semiconductor processing, electron spectroscopy characterizes critical interfaces like SiO₂/Si, essential for metal-oxide-semiconductor devices. High-resolution XPS of Si 2p core levels identifies intermediate oxidation states (e.g., Si¹⁺ to Si³⁺) at the SiO₂/Si(100) interface, arising from strained Si–O–Si bonds, and quantifies valence band offsets to assess electrical performance in sub-5 nm oxides. AES complements this by detecting contaminants or diffusion barriers, such as titanium interactions forming silicides at SiO₂/Si surfaces during metallization. For corrosion monitoring, XPS and AES analyze passive films and corrosion products on metals. These techniques identify the composition of oxide layers, such as chromate on steel or tarnish on silver, and track their evolution under environmental exposure; for instance, XPS has revealed chloride-induced pitting in stainless steels by detecting localized enrichment of iron oxides. In high-temperature oxidation, AES profiles multilayer scales, identifying sulfur segregation that promotes spalling. Challenges in these analyses include surface contamination from residual gases, which can alter compositions during measurement, and the necessity for ultrahigh vacuum (UHV) conditions around 10⁻¹⁰ Torr to minimize adsorption rates on reactive surfaces like silicon or titanium. Prolonged UHV exposure can even induce carbonization or oxygen accumulation, complicating interpretations, though near-ambient pressure XPS mitigates this for in-situ corrosion studies.

Electronic Structure and Material Characterization

Electron spectroscopy techniques, particularly ultraviolet photoelectron spectroscopy (UPS) and angle-resolved photoelectron spectroscopy (ARPES), enable detailed mapping of valence band structures in semiconductors, revealing Fermi surfaces and band gaps essential for understanding charge carrier dynamics and optoelectronic properties. In ARPES, the momentum-resolved measurement of photoemitted electrons from the valence band provides a direct visualization of the electronic band dispersion across the , allowing precise determination of the Fermi surface topology in materials like topological insulators or transition metal dichalcogenides. For instance, in semiconductors such as or , ARPES has quantified band gaps by identifying the energy separation between the valence band maximum and the Fermi level, with values typically ranging from 0.7 eV to 3.4 eV depending on the material composition. UPS complements this by offering high-resolution spectra of the valence band density of states near the Fermi edge, facilitating the assessment of ionization potentials and work functions in semiconductor heterostructures. X-ray photoelectron spectroscopy (XPS) elucidates chemical bonding through core-level binding energy shifts, which reflect changes in electron density due to hybridization and local coordination environments, particularly in organic materials. These shifts arise from variations in the effective nuclear charge experienced by core electrons, influenced by the electronegativity and overlap of neighboring atoms; for example, in amorphous carbon materials, the C 1s core-level shift of approximately 1 eV distinguishes sp² from sp³ hybridization, indicating π-conjugation critical for charge transport. In organometallic complexes, XPS core-level analysis has revealed d-orbital hybridization in transition metal centers bonded to carbon ligands, with shifts up to 3 eV correlating to back-donation effects that alter reactivity. Such insights enable the characterization of bonding motifs in molecular electronics and catalysts, where hybridization dictates electronic properties like conductivity and stability. Electron energy loss spectroscopy (EELS) excels in identifying and differentiating carbon allotropes, such as and , by probing the fine structure of the carbon K-edge, which encodes information on local bonding and electronic states in nanomaterials. The π* and σ* resonances in the EELS spectrum at around 285 eV and 292 eV, respectively, allow discrimination of sp²-hybridized (with a prominent π* peak) from sp³-hybridized (dominated by σ* features), enabling nanoscale mapping of phase purity in hybrid carbon structures. In graphene-based composites, EELS has quantified defect-induced hybridization changes, such as nitrogen doping shifting the K-edge by 1-2 eV, which impacts electrical conductivity. This capability supports material identification in applications like energy storage devices, where allotrope-specific properties determine performance. Ambient pressure XPS (AP-XPS) facilitates operando studies of dynamic processes in energy materials, capturing real-time electronic changes in battery electrodes and photocatalysts under working conditions. In lithium-ion battery electrodes, such as Ni-rich cathodes, AP-XPS has observed valence state transitions (e.g., Ni²⁺ to Ni³⁺) and surface reconstruction during charge-discharge cycles, with oxygen 1s shifts indicating lattice oxygen participation post-2020 advancements in near-ambient setups. For photocatalysts like or NiO/Ni systems, operando AP-XPS under illumination reveals band bending and adsorbate-induced electronic restructuring, such as downward shifts in the valence band edge by 0.5-1 eV upon water adsorption, enhancing charge separation efficiency. These studies, enabled by differential pumping, provide direct evidence of active site evolution, guiding catalyst optimization for sustainable energy conversion. Beyond specialized applications, electron spectroscopy supports broader material characterization, including quality control in metal alloys and environmental analysis of pollutants. In alloys like stainless steel or aluminum, and detect elemental segregation and oxidation states at grain boundaries, ensuring compositional uniformity; for example, AES mapping identifies chromium enrichment in passive layers, preventing corrosion failures in industrial components. For environmental monitoring, and quantify heavy metal pollutants (e.g., lead or cadmium) on biological matrices like fish tissues, with detection limits below 0.1 at%, revealing bioaccumulation and speciation for risk assessment. These techniques thus underpin quality assurance and pollution mitigation strategies across industries.

Instrumentation and Analysis

Experimental Components

Electron spectroscopy experiments require a carefully assembled setup of hardware components to generate, collect, and analyze emitted electrons while maintaining conditions that preserve surface sensitivity and minimize contamination. The core elements include excitation sources for electron emission, energy analyzers for kinetic energy measurement, detectors for signal amplification, ultra-high vacuum (UHV) systems for environmental control, and sample handling mechanisms for precise manipulation and preparation. These components are typically integrated into modular chambers to enable versatile operation across techniques like , , and . Excitation sources provide the incident radiation or particles necessary to eject electrons from the sample. For (XPS), twin-anode X-ray sources using aluminum (Al Kα at 1486.6 eV) and magnesium (Mg Kα at 1253.6 eV) are standard, allowing quick switching between energies without breaking vacuum to verify spectral features. These anodes operate in a vacuum tube where electrons bombard the target to produce characteristic X-rays, with monochromators often employed to reduce satellite lines and improve resolution to about 0.4 eV full width at half maximum (FWHM). (UPS) utilizes helium discharge lamps, primarily He I (21.2 eV) and He II (40.8 eV), which emit resonant lines ideal for probing valence electrons near the . For (AES), field emission electron guns serve as the excitation source, delivering a focused beam (typically 1-30 keV) with currents of 1-100 nA and spot sizes down to 10 nm, enabling high spatial resolution surface analysis. Synchrotron sources offer tunable energies (e.g., 200-1250 eV) for advanced experiments but require specialized facilities. As of 2025, advancements in synchrotron beamlines and detector technologies have reduced experimental times and improved resolving powers in time-of-flight systems exceeding 10,000 (E/ΔE). Energy analyzers separate electrons based on their kinetic energy, a critical step for spectral resolution. The hemispherical sector analyzer (HSA), also known as a hemispherical deflection analyzer, is the most common type, featuring two concentric hemispherical plates with an electrostatic field that allows electrons of selected energy to reach the detector. The energy resolution in an HSA is given by ΔE / E_p, where ΔE is the energy width and E_p is the pass energy, typically set between 5-200 eV to balance resolution (e.g., 0.1 eV FWHM) and count rate; lower E_p yields higher resolution but reduces transmission efficiency. HSAs often include retarding lenses to slow electrons, improving throughput for low-energy signals, and can achieve angular acceptance up to ±3° for photoelectron diffraction studies. Cylindrical mirror analyzers (CMA) are alternatives used in AES for their azimuthal symmetry and 2π steradian collection efficiency, though they offer coarser resolution (~0.3-1% of kinetic energy). These analyzers operate in conjunction with the mean free path of electrons (~5-20 Å in solids), ensuring surface specificity. Detection systems amplify the low-intensity electron signals post-analysis. Channeltron electron multipliers, discrete single-channel devices, provide gains of 10^6 to 10^8 by cascading secondary electron emissions within a continuous dynode structure, making them suitable for sequential energy scanning in spectroscopy. For parallel detection and higher throughput, microchannel plate (MCP) detectors consist of arrays of microscopic channels (6-25 μm diameter) that multiply electrons via similar mechanisms, often coupled with position-sensitive anodes for 2D imaging; MCPs achieve gains up to 10^4 per plate (stacked for higher) and quantum efficiencies >50% for electrons above 100 . Time-of-flight (TOF) analyzers, using free-flight paths under pulsed excitation, pair with MCPs for fast, broadband detection in techniques like EELS, resolving energies via flight time differences with given by t = L √(m / (2 )) (where m is , L is path length, and E is ) and relative energy ΔE/E ≈ 2 (Δt / t) for timing resolution Δt. These detectors are positioned at the analyzer exit slit or focal plane to capture the dispersed electron beam. Vacuum systems are indispensable to prevent electron scattering by residual gas and surface contamination, as mean free paths are short (~10^{-6} cm at 1 atm but extend to centimeters or more in UHV). Ultra-high vacuum chambers maintain base pressures of 10^{-9} to 10^{-12} Torr using turbomolecular pumps backed by roughing pumps, ion getters, and non-evaporable getters for long-term stability. Differential pumping isolates regions like the excitation source (higher pressure, e.g., 10^{-6} Torr) from the main analysis chamber via small apertures and intermediate stages, preserving UHV in the sample area. Cryoshrouds, cooled by liquid nitrogen to ~77 K, surround the sample and optics to cryopump water vapor and hydrocarbons, reducing outgassing rates by factors of 10^3-10^4 and achieving pressures below 10^{-10} Torr. Bakeout procedures (150-250°C) further clean chamber walls, essential for monolayer-sensitive measurements. Energy resolution in electron spectroscopy depends on low background gas pressure to avoid inelastic collisions, typically requiring <10^{-8} Torr during operation. Sample handling ensures reproducible positioning and preparation under UHV to avoid adventitious contamination. Manipulators, often multi-axis (x, y, z, polar, azimuthal), allow precise sample alignment with resolutions of ±0.5° and translations up to 50 mm, enabling heating via electron bombardment or resistive elements (up to 2000 K) or cooling with liquid nitrogen cryostats (down to 100 K) for temperature-dependent studies. In-situ preparation tools include evaporators for metal deposition (e.g., Knudsen cells at 1000-1500 K for flux control to 10^{13}-10^{15} atoms/cm² s) and ion sputter guns for cleaning via Ar^+ bombardment (1-5 keV, 1-50 μA/cm²). Load-lock systems facilitate sample introduction without full chamber venting, while transfer rods move samples between preparation and analysis chambers. These components support diverse samples, from single crystals to thin films, with active volumes ~10^{-6} cm³ probed effectively.

Data Interpretation Methods

Data interpretation in electron spectroscopy involves processing raw spectra to extract meaningful information about surface composition and electronic structure, primarily through steps that address instrumental and physical artifacts. This process ensures accurate identification of elemental peaks and their chemical states, such as shifts observed in () due to varying oxidation environments. Key procedures include background subtraction to isolate true photoemission signals, peak to resolve overlapping features, and accounting for instrumental responses. Background subtraction is essential for removing the inelastic scattering tail that contributes to the spectral baseline, particularly in and (). The Shirley method, introduced in 1972, iteratively subtracts a background proportional to the of the signal above and below the region, effectively modeling the secondary . This approach assumes a linear increase in background intensity with energy loss and is widely used for its simplicity in routine analysis. For more physically accurate modeling of , the Tougaard method employs a parameterized function, typically with parameters like C = 1643 eV², to deconvolute the background based on transport . This method better accounts for depth-dependent contributions in non-homogeneous samples, though it requires computational intensity. Peak fitting decomposes complex spectra into individual components, often using Voigt profiles that convolute Gaussian (instrumental and thermal broadening) and (lifetime broadening) functions. In core-level spectra, spin-orbit splitting must be incorporated, such as for p-orbitals where the 2p_{3/2} and 2p_{1/2} components follow branching ratios of 2:1 and energy separations around 0.6–20 depending on the . Constraints like equal full-width at half-maximum (FWHM) and fixed intensity ratios ensure physical realism, preventing in multiplet or chemically shifted peaks. For example, fitting the 4f requires linking the 4f_{7/2} and 4f_{5/2} peaks with a area ratio and ~3.7 separation. Quantification derives atomic concentrations from peak areas, but challenges arise from matrix effects that alter photoelectron escape depths and signal attenuation in layered or alloyed samples. Corrections for the analyzer's transmission function, typically T(E) ∝ E where E is , are applied to normalize intensities across elements, often using relative sensitivity factors (RSFs) calibrated against standards. Inhomogeneous matrices can introduce errors up to 20–50% without depth profiling, necessitating iterative models for accurate . Specialized software facilitates these analyses, with CasaXPS providing tools for Shirley/Tougaard backgrounds, Voigt-based , and RSF-integrated quantification, supporting database matching for chemical state identification. Similarly, Avantage software from Thermo Fisher enables automated peak fitting with spin-orbit constraints, transmission corrections, and exportable reports for checks. Error sources must be minimized for reliable interpretation, including statistical from Poisson-distributed electron counts, which scales as the of intensity and affects low-signal peaks. calibration using the Au 4f_{7/2} line at 84.0 eV ensures scale accuracy within ±0.1 eV, while issues from X-ray-induced sample damage or charging require multiple acquisitions and charge compensation verification.

References

  1. [1]
    Electron Spectroscopy - an overview | ScienceDirect Topics
    Measurement of the kinetic energy (KE) of the electrons that are emitted from a solid will lead to the measurement of the bound electron spectrum in that solid.
  2. [2]
    Electron Spectroscopy - an overview | ScienceDirect Topics
    Electron spectroscopy is a technique for the investigation of the energy distribution of electrons ejected from a material when irradiated by a source of ...
  3. [3]
    Electron Spectroscopy - Fritz Haber Institute
    Electron spectroscopy anaylses the electron energies emitted from a sample to provide surface-sensitive chemical and physical information. This section provides ...
  4. [4]
    None
    Summary of each segment:
  5. [5]
    (PDF) A Review on Surface Analysis Techniques - ResearchGate
    Apr 15, 2020 · A focused review on the fundamentals of Surface Analysis and related characterisation techniques. Topics covered include: X-ray Photoelectron Spectroscopy (XPS ...
  6. [6]
    Photoelectron spectroscopy under ambient pressure and ...
    We describe the development and applications of novel instrumentation for photoemission spectroscopy of solid or liquid surfaces in the presence of gases ...
  7. [7]
    Surface Analysis II. Electron Spectroscopy Methods - SpringerLink
    Electron spectroscopy probes the electronic structure of the surface through analysis of the energy spectra of the secondary electrons emitted from the ...
  8. [8]
    Principle - West Campus Materials Characterization Core
    In photoelectron spectroscopy (PES) recorded electrons are emitted when the surface is irradiated with photons. The technique is based on the photoelectric ...
  9. [9]
    Auger Electron Spectroscopy - SERC (Carleton)
    May 10, 2021 · Auger Electron Spectroscopy (AES) is a surface-sensitive analytical method the provides compositional information in the near-surface environment of materials.
  10. [10]
    [PDF] Basic Concepts of X-Ray Photoelectron Spectroscopy
    X-ray photoelectron spectroscopy studies atoms, molecules, solids, and surfaces by exposing them to radiation and observing emitted photoelectrons. It is also ...
  11. [11]
    A hemispherical retarding field energy analyzer to characterize ...
    Abstract. We have designed and built a hemispherical retarding field energy analyzer in order to facilitate characterization of large area electron emitters ...
  12. [12]
    A novel design of a retarding field electron energy analyzer with a ...
    Jun 8, 2023 · This paper describes the principle of the newly developed energy analyzer and shows the obtained energy spectra from a Schottky emitter ZrO/W(100)INTRODUCTION · Electron energy spectra at... · CONCLUSIONS
  13. [13]
    X-ray photoelectron spectroscopy: Towards reliable binding energy ...
    As electrons absorb over the entire photon energy, the conservation of energy requires that their kinetic energy is equal to h ν – W, where W is the work ...
  14. [14]
    [PDF] NIST Electron Inelastic-Mean-Free-Path Database
    This database provides values of electron inelastic mean free paths (IMFPs) in solid elements and compounds at selected electron energies between 50 eV and ...Missing: specificity | Show results with:specificity
  15. [15]
    Auger Electron Spectroscopy (AES) Tutorial | EAG Laboratories
    The mean free paths of electrons depend on their energies and on the sample material. The minimum mean free path (~0.5 nm) occurs at about 80 eV. Under ...
  16. [16]
    Angular distribution and spin polarization of Auger electrons ...
    The influence of different factors on the angular distribution of the Auger electron intensity and its spin polarization is studied and discussed.
  17. [17]
    Low-energy spin-polarized electrons: their role in surface physics
    Nowadays, we can rely on a full portfolio of diffraction, spectroscopy and microscopy approaches using low-energy electrons on various length and time scales.
  18. [18]
    [PDF] CHAPTER 10 AUGER ELECTRON SPECTROSCOPY
    Auger electron spectroscopy (AES) is a nondestructive technique for determining elemental composition of surfaces, thin films, and interfaces. It uses a three- ...
  19. [19]
    [PDF] Bambynek et al: X-ray fluorescence yields, Auger, and Coster ...
    An Auger event resulting from an inner-shell vacancy produces a doubly ionized atom; successive radiationless transitions can lead to highly ionized atomic ...
  20. [20]
    None
    ### Summary on Auger Parameter and Chemical Shifts
  21. [21]
    Introduction to x-ray photoelectron spectroscopy - AIP Publishing
    Sep 24, 2020 · Plasmons, Data acquisition, Auger effect, Surface analysis techniques, Ultraviolet photoelectron spectroscopy ... Bremsstrahlung radiation ...
  22. [22]
    (PDF) Concerning Hertz' photoelectric effect - ResearchGate
    Experimental evidence of the photoelectric effect goes back to H. Hertz. It occurred during the famous confirmation experiments of the Maxwellian theory.
  23. [23]
    Heinrich Hertz and the Foundations of Electromagnetism
    Jun 1, 2011 · In 1887, Hertz also discovered the photoelectric effect; he used a prism to show that ultraviolet light enhanced the generation of sparks.
  24. [24]
    [PDF] lenard-lecture.pdf - Nobel Prize
    Hertz, whose assistant I then was, had found that thin metal leaf transmits cathode rays (15). He used quite thin, very soft and porous gold, silver and ...
  25. [25]
    [PDF] 3 Experiments that Shook the World of Physics.
    The first quantitative experiment on the photoelectric effect was conducted in 1902 by. Philipp Lenard [14]. The maximum energy of the emitted photoelectrons ...
  26. [26]
    [PDF] Einstein's Proposal of the Photon Concept-a Translation
    The American Journal of Physics is publishing the following translation in recognition of the sixtieth anniversary of the appearance of the original work.
  27. [27]
    Quantum Milestones, 1905: Einstein and the Photoelectric Effect
    Jan 22, 2025 · Einstein explained the photoelectric effect by considering light as a stream of energy quanta—which were later called photons ...
  28. [28]
    [PDF] A Direct Photoelectric Determination of Planck's `h'. - MIT
    These authors found the value of h in the Einstein photoelectric equation varying in the eight metals studied from 3.55 X 10-27 to 5.85 X 10-27 Planck's h was ...
  29. [29]
    Millikan's struggle with theory - Europhysics News
    In short, Millikan's classic. 1916 paper was purely intended to be the verification of Einstein's equation for the photoelectric effect and the determination.
  30. [30]
    [PDF] Thermionic Phenomena and the Laws which Govern Them
    It became apparent at a very early stage that the emission of electrons from conductors at a given temperature was very susceptible to the influence of.Missing: 1920s | Show results with:1920s
  31. [31]
    The birth of electronics: Thermionic emission and vacuum
    Owen Willans Richardson. Richardson published many papers35 between 1900 and 1920 on thermionic emission and gave theoretical and ex- perimental evidence that ...
  32. [32]
    Kai M. Siegbahn – Facts - NobelPrize.org
    In the 1950s Kai Siegbahn developed methods for achieving highly accurate measurements of energy levels in atoms by irradiating them with photons and ...Missing: XPS | Show results with:XPS<|separator|>
  33. [33]
    Electronic structure of BN-aromatics: Choice of reliable ...
    Oct 26, 2017 · For the purpose, Ultra-Violet Photoelectron Spectroscopy (UV-PES), developed in 1962 by David W. Turner,8,9 provides the experimental values ...
  34. [34]
    [PDF] Synchrotron radiation study of free and adsorbed organic molecules
    Ultraviolet Photoelectron Spectroscopy (UPS) is a technique developed by. David W. Turner at 1960's for valence photoelectron ... band photoelectron spectra.
  35. [35]
  36. [36]
    Photoemission spectroscopy—from early days to recent applications
    Apr 29, 2005 · Other important advantages of the synchrotron light are, e.g., high intensity and brightness, variable polarization, small photon spots, or ...
  37. [37]
    The evolution of a dedicated synchrotron light source - Physics Today
    The success of those machines paved the way in the 1990s and 2000s for third- and fourth-generation light sources that rely heavily on insertion devices ...
  38. [38]
    Operando and in situ APXPS investigation of the atomic layer ...
    Oct 14, 2025 · In this study, synchrotron-based ambient pressure X-ray photoelectron spectroscopy (APXPS) is employed to investigate the surface evolution of a ...
  39. [39]
    Efficiency improvement of spin-resolved ARPES experiments using ...
    Sep 23, 2024 · We focus on a representative topological insulator , which has well-understood spin-polarized electronic states. We employ Gaussian process ...Missing: post- | Show results with:post-
  40. [40]
    Machine Learning Data Augmentation Strategy for Electron Energy ...
    Apr 29, 2024 · This article introduces a data augmentation (DA) strategy for electron energy loss spectroscopy (EELS) data, employing generative adversarial networks (GANs).Materials And Methods · Experimental Data Set... · Spectra Generation...<|separator|>
  41. [41]
    Recent progress in electron energy loss spectroscopy with ...
    Aug 29, 2025 · Recent breakthroughs in aberration correction technology have propelled STEM spatial resolution to sub-angstrom levels [12, 13], enabling EELS ...
  42. [42]
    [PDF] Kai Siegbahn - Nobel Lecture
    I now got the idea that I should instead use an X-ray tube to expel photoelectrons from ordinary materials, in order to measure their binding energies to the.Missing: seminal | Show results with:seminal
  43. [43]
    X-Ray Photoelectron Spectroscopy (XPS - SERC (Carleton)
    Aug 13, 2021 · X-ray photoelectron spectroscopy (XPS) is a surface sensitive, non-destructive technique used routinely to analyze the outermost ~10 nm (~30 atomic layers) of ...
  44. [44]
    Quantitative AES IX and quantitative XPS II: Auger and x‐ray ...
    Aug 2, 2001 · ... sensitivity factors from spectral digital databases reanalysed using a REELS database. M. P. Seah. Corresponding Author. martin.seah@npl.co.uk.
  45. [45]
    Ultraviolet photoelectron spectroscopy: Practical aspects and best ...
    Ultraviolet photoelectron spectroscopy (UPS) is an important technique for measuring the energies of the valence states of metallic, semiconducting and ...
  46. [46]
    Angle-resolved Photoemission Spectroscopy - Shen Laboratory
    A generic ARPES measurement involves a single crystalline sample irradiated by monochromatic light of energy hν, resulting in photoemission of electrons in all ...Missing: 2020 | Show results with:2020
  47. [47]
    UV Photoelectron Spectroscopy of Aqueous Solutions
    Nov 28, 2022 · One of the most direct ways of determining electronic structure is to use photoelectron spectroscopy to measure electron binding energies.
  48. [48]
    Penning ionization electron spectroscopy of CO and Fe(CO)5. Study ...
    The He* Penning ionization electron spectra (PIES) and He I UV photoelectron spectra (UPS) of CO and Fe(CO)5 were measured. The relative intensity of the bands ...
  49. [49]
    Extreme ultraviolet time-resolved photoelectron spectroscopy of ...
    Time-resolved photoelectron spectroscopy using an extreme-ultraviolet (XUV) probe pulse was used to investigate the UV photoinduced dynamics of adenine ...
  50. [50]
    Auger Electron Spectroscopy (AES) - EAG Laboratories
    Auger Electron Spectroscopy (AES) is a surface-sensitive analytical technique that utilizes a high-energy electron beam as an excitation source.Aes Key Applications · Auger Strengths · Limitations
  51. [51]
    (PDF) Auger Electron Spectroscopy-A Review
    ### Summary of Key Sections on AES Principles
  52. [52]
    Field Emission Auger Electron Spectroscopy with Scanning ... - NREL
    Jan 6, 2025 · Auger electron spectroscopy (AES) uses high-energy electrons to generate Auger electrons for elemental surface analysis, depth profiling, and ...How It Works · Applications · Depth Profiling<|control11|><|separator|>
  53. [53]
    What is EELS?
    EELS uses the energy distribution of electrons passing through a thin sample to analyze its content and create images, typically in a TEM or STEM.
  54. [54]
    Electron Energy Loss Spectroscopy - an overview - ScienceDirect.com
    Electron energy-loss spectroscopy (EELS) is an analysis of the distribution of electron energy emergent from a thin specimen due to inelastic interactions. An ...
  55. [55]
    Fine structure of ionization edges in EELS - Electron microscopy
    Fine structures in an EEL spectrum fundamentally originate from the transition from a core level to unoccupied states under the dipole selection rule.
  56. [56]
    Electron Energy Loss Near Edge Structure - ScienceDirect.com
    The fine structure on an EELS edge arises because, during the excitation process, the incident electron can impart more energy than the critical energy required ...
  57. [57]
    Electron energy loss spectrum of silver - GPAW - Read the Docs
    ω p = n e 2 ϵ 0 m . In GPAW, the EELS can be calculated with the dielectric response module, where it is obtained from the macroscopic dielectric function:.Missing: ω_p = √( e² /
  58. [58]
    Assess Sample Thickness | EELS.info
    To determine relative sample thickness, you can use a variety of algorithms. However, the log-ratio (relative) method is most commonly used.Missing: ln( I_total /
  59. [59]
    EELS log-ratio technique for specimen-thickness measurement in ...
    We discuss measurement of the local thickness t of a transmission microscope specimen from the log-ratio formula t = lambda ln (It/I0)Missing: I_total / | Show results with:I_total /
  60. [60]
    TEM-EELS: A personal perspective - ScienceDirect.com
    The development of electron energy-loss spectroscopy in a transmission electron microscope (TEM-EELS) is illustrated through personal anecdote, highlighting ...
  61. [61]
    Advances in momentum-resolved EELS of phonons, excitons ... - arXiv
    Oct 10, 2025 · An additional advantage of q-EELS is the flexibility of STEM optics, which can be tuned to optimize for momentum, spatial and energy resolutions ...
  62. [62]
    Review on surface-characterization applications of X-ray ...
    This is a pedagogical review on XPS based surface characterization. Applications in materials, nanotechnology and corrosion evaluation are reviewed.
  63. [63]
    A Review of X-ray Photoelectron Spectroscopy Technique to ... - NIH
    The XPS can be used to analyze the elemental composition and oxidation state of cathode layers from the surface to a depth of approximately 5–10 nm. Any ...
  64. [64]
    (PDF) A quantitative analysis of AES and XPS specifically applied in ...
    Oct 13, 2020 · A quantitative analysis of AES and XPS specifically applied in adsorption systems at submonolayer regime. October 2020; Surface and Interface ...
  65. [65]
    AES and EELS study of alumina model catalyst supports
    The processes studied so far include the kinetics of adsorption and desorption involving various adsorbates and the reaction kinetics of several simple ...Missing: seminal paper
  66. [66]
    Band alignments at Ga 2 O 3 heterojunction interfaces with Si and Ge
    Jun 12, 2018 · X-ray photoelectron spectroscopy (XPS) was used to investigate the band alignments at the interfaces using the Kraut Method. The valence band ...
  67. [67]
    Photoelectron spectroscopy studies of SiO2/Si interfaces
    We describe state-of-the-art photoelectron spectroscopy studies of SiO2/Si interfaces that play fundamental roles in metal-oxide-semiconductor (MOS) ...Missing: AES | Show results with:AES
  68. [68]
    AES and XPS analysis of the interaction of Ti with Si and SiO2 ...
    The “peak-to-peak-minus-3-eV-height” method is used for the quantification with AES, while XPS is mainly used to characterize the oxide and nitride states.
  69. [69]
    The role of electron spectroscopy in corrosion science
    The two principal electronspectroscopies – photoelectron and Auger electron – now represent the mature methods of surface analysis and both have been used ...Missing: monitoring | Show results with:monitoring
  70. [70]
    Challenges in surface analysis - Frontiers
    Jun 28, 2023 · The most commonly used technique for surface analysis is XPS as it provides the simplest spectrum and is the easiest to quantify.Introduction · Instrumentation · Software · Publishing in the surface...
  71. [71]
    Excited-state band structure mapping | Phys. Rev. B
    Feb 17, 2022 · In this work we extend band structure mapping into the unoccupied states and across the entire Brillouin zone by using a state-of-the-art high repetition rate, ...
  72. [72]
    Three-dimensional band structure and surface electron ... - Nature
    May 29, 2019 · Three bands are distinguished in the ARPES mapping for Γ-K direction. The upper band with a maximum midway along the Γ-K line of the valence ...
  73. [73]
    Momentum-resolved electronic structure and band offsets in ... - NIH
    The offset in electron energies between the Fermi level (EF) of the metal and the conduction band minimum (CNM) of the semiconductor was found to be responsible ...
  74. [74]
    Understanding Chemical versus Electrostatic Shifts in X-ray ...
    Jan 25, 2016 · (97, 98)z is the position of the atom whose core level is excited. The screening shifts core-level energies to less negative values (smaller ...
  75. [75]
    from photoelectron spectroscopy data to energy-level diagrams
    Nov 14, 2022 · The chemical shift of the C 1s core level between carbon atoms in the conjugated backbone and in the functional groups of an organic molecule ...
  76. [76]
    Probing the Bonding in Nitrogen-Doped Graphene Using Electron ...
    Simultaneous recording of the EELS spectra has provided complementary bonding information, and there is a change in the carbon K-edge from the region ...Figure 1 · Figure 3 · Materials And Methods
  77. [77]
    Magic angle and STEM-EELS mapping of the sp2/sp3 hybridization ...
    Scanning transmission electron microscopy-electron energy loss spectroscopy is a powerful technique that enables the identification of carbon allotropes with ...
  78. [78]
    Operando and in situ APXPS investigation of the atomic layer ...
    Oct 30, 2025 · Operando and in situ APXPS investigation of the atomic layer deposition of a metal oxide coating on a Ni-rich cathode. October 2025; Applied ...
  79. [79]
    Photocatalytic setup for in situ and operando ambient-pressure X ...
    APXPS has proven to be a very powerful tool for studying in situ and operando studies of ALD cycles and for revealing reaction steps otherwise unobservable ...
  80. [80]
    Solar light driven atomic and electronic transformations in a ...
    Apr 22, 2024 · This work employs ambient pressure X-ray photoelectron spectroscopy (APXPS) to delve into the atomic and electronic transformations of a core–shell Ni@NiO/NiCO ...<|separator|>
  81. [81]
    XPS/AES Studies of Aluminum Fracture Surfaces. - DTIC
    Abstract: X-ray photoemission XPS, or ESCA and Auger electron scattering AES studies were made of aluminum alloy surfaces after polishing, etching, and ion ...
  82. [82]
    Applications of surface analysis in the environmental sciences
    In this paper, we illustrate how surface analytical techniques, including Auger electron spectroscopy (AES), X-ray photoelectron spectroscopy (XPS), attenuated ...
  83. [83]
    (PDF) Direct XPS (X-ray photoelectron spectroscopy) analysis of ...
    Oct 7, 2025 · Fish tissues were analyzed directly with X-ray photoelectron spectroscopy (XPS) in order to find pollutant elements that may be ...
  84. [84]
    [PDF] Electron Spectroscopy: Theory, Techniques and Applications
    This surface sensitivity of XPS (or any form of electron spectroscopy) can ... surface of aluminium (which exhibits well-defined surface- and bulk-plasmon.
  85. [85]
    A step-by-step guide to perform x-ray photoelectron spectroscopy
    Jul 1, 2022 · The most common x-ray sources are Kα lines from Mg- and Al-coated electrodes (referred to as anodes, since they are at positive potential in the ...Missing: lamps | Show results with:lamps
  86. [86]
  87. [87]
    16.4: Ion Detectors - Chemistry LibreTexts
    Aug 21, 2020 · Another ion detector related to the channeltron is the microchannel plate (MCP) detector. A microchannel plate is a thin disc composed of many ...
  88. [88]
    Ultrahigh Vacuum - an overview | ScienceDirect Topics
    XPS measurements are conducted under ultra-high vacuum (< 10- 8 Torr) in order to avoid collision between photoelectrons and gas molecules in the spectrometer, ...<|control11|><|separator|>
  89. [89]
    Instrumentation and analytical methods of an x-ray photoelectron ...
    Aug 7, 2006 · The manipulator in the analysis chamber allows samples to be cooled using a L N 2 cryostat to 100 K and heated by electron beam bombardment to ...
  90. [90]
    Electron Spectroscopy - Scienta Omicron
    It provides ports for various excitation sources, evaporators a sputter cleaning source and sample heating. In addition, the system is also designed to house a ...