Fact-checked by Grok 2 weeks ago

Rutherford backscattering spectrometry

Rutherford backscattering spectrometry (RBS) is a non-destructive technique that determines the elemental composition, thickness, and depth profiles of thin films and surface layers in materials by measuring the energies of high-energy , typically , that are elastically scattered from the sample's nuclei through interactions. The method relies on the formula, which describes the probability of ion deflection based on the numbers of the incident and target atom, enabling without the need for reference standards. The technique traces its origins to early 20th-century experiments on scattering. In 1909, and observed the backscattering of alpha particles from thin gold foils, providing empirical data on nuclear interactions. interpreted these results in 1911, developing the nuclear model of the atom and the differential cross-section formula for Coulomb scattering, which forms the theoretical foundation of RBS. Modern RBS emerged in the mid-20th century with advancements in particle accelerators, and its first extraterrestrial application occurred in 1967 when the Surveyor V spacecraft used an alpha backscattering instrument to analyze the chemical composition of the lunar surface. In practice, RBS involves accelerating ions to energies of 1–3 MeV using a Van de Graaff or tandem accelerator, directing the beam onto the sample, and detecting backscattered ions with a solid-state detector positioned at a scattering of 90°–170° to resolve energy losses corresponding to depth and . The energy of scattered ions is governed by the kinematic factor k = \left[ \frac{M_1 \cos \theta + \sqrt{M_2^2 - M_1^2 \sin^2 \theta}}{M_1 + M_2} \right]^2 , where E_0 is the initial , M_1 and M_2 are the masses of the ion and target atom, and \theta is the scattering ; this allows identification of elements by their mass-dependent energy shifts. Depth information arises from energy straggling and as ions penetrate the sample, with surface depth resolution typically around 10–20 nm. RBS is particularly valuable for analyzing heavy elements in lighter matrices, such as transition metals in oxides or semiconductors, and is widely applied in for characterizing thin films in , superconductors, and magnetic multilayers. It offers high sensitivity for elements with atomic number Z > 10, quantitative depth profiling up to several micrometers, and compatibility with in-situ studies of interfaces, including solid-liquid systems. However, limitations include reduced sensitivity to light elements (Z < 10), potential overlap in spectra for adjacent elements, and the need for vacuum conditions, making it complementary to techniques like secondary ion mass spectrometry (SIMS) for lighter species. Channeling variants enhance its utility for assessing crystalline quality and defect analysis by aligning the ion beam with lattice directions.

History

Geiger–Marsden experiment

The Geiger–Marsden experiments, conducted from 1909 to 1913 under the supervision of at the University of Manchester, provided the foundational observations of alpha particle scattering that revealed the nuclear structure of the atom. Hans Geiger, a postdoctoral researcher, and , an undergraduate student, carried out the practical work, with directing the overall investigation and providing theoretical interpretation. The series began with initial tests in 1909, followed by refinements in 1910 and 1912–1913, culminating in quantitative confirmations of scattering laws. The experimental setup utilized alpha particles emitted from a radium bromide source, collimated into a narrow beam directed at thin gold foils roughly 0.00004 cm thick, equivalent to the stopping power of about 1.6 mm of air. The foils were mounted in a vacuum chamber to minimize air scattering, and scattered particles were detected using zinc sulfide scintillation screens viewed through a microscope, allowing manual counting of individual impacts. This geometry enabled measurement of scattering at various angles relative to the incident beam direction. Key observations included the passage of most alpha particles through the foil with little or no deflection, suggesting atoms consist largely of empty space. However, a notable fraction underwent large-angle deflections: approximately 1 in 8,000 particles scattered by more than 90 degrees from , with even rarer events showing near-backward scattering. In their 1909 report, described this as a "diffuse reflection," where a small proportion of incident alpha particles reversed direction and emerged from the same side as the beam, defying expectations of minimal interaction based on prior models. Rutherford's 1911 interpretation of these results overturned J.J. Thomson's plum-pudding model, in which positive charge was diffusely distributed throughout the atom. He proposed that the atom's positive charge and nearly all its mass are concentrated in a minuscule, dense nucleus, with electrons orbiting at a distance; large deflections arise from rare close approaches between the positively charged alpha particle and the nucleus via Coulomb repulsion. Further experiments in 1912–1913 by Geiger and Marsden verified the scattering probability's dependence on atomic number and particle velocity, aligning with Rutherford's nuclear hypothesis. Qualitatively, the low probability of backscattering reflects the nucleus's tiny size—about 1/10,000th the atom's diameter—making head-on collisions infrequent despite the foil containing billions of atoms. These findings established the principles of single-scattering events that underpin modern as a precursor technique.

Development as an analytical technique

Following the foundational observations of alpha particle scattering in the , Rutherford backscattering evolved into a practical analytical tool in the post-World War II era through advancements in particle accelerators that enabled controlled ion beams for materials studies. In the 1950s, cyclotrons and early electrostatic accelerators, such as those developed at laboratories like , facilitated experiments on ion-solid interactions, shifting focus from qualitative nuclear physics to quantitative surface characterization. A key milestone was the 1957 work by S. Rubin, T.O. Passell, and L.E. Bailey, who demonstrated the use of nuclear scattering methods, including backscattering, for chemical analysis of surfaces. By the 1960s, Rutherford backscattering spectrometry (RBS) gained traction in solid-state physics, particularly for semiconductor analysis, as researchers adapted low-energy accelerators to probe thin films and interfaces. Pioneers like James W. Mayer at utilized 2-3 MeV helium ions to measure film thicknesses and compositions non-destructively, leveraging the newly invented surface-barrier detectors for energy-resolved spectra. The transition from qualitative nuclear studies to quantitative surface analysis was accelerated by the widespread adoption of in the 1960s and 1970s, which provided stable, high-voltage beams up to several MV, enabling precise depth profiling in materials like silicon and alloys. Early applications included determining alloy compositions in evaporated metal films and thicknesses of oxide layers on semiconductors, with resolutions better than 100 nm for heavy elements. The establishment of RBS as a standard technique by the 1980s was bolstered by computing advancements in the 1970s, which allowed for accurate spectrum simulation and fitting. In 1976, J.F. Ziegler and collaborators introduced the IBA code, the first program for full Monte Carlo simulations of RBS spectra, incorporating stopping powers and cross-sections to model complex multilayer structures and extract quantitative depth profiles. This computational support, combined with the technique's non-destructive nature, solidified RBS's role in materials characterization, influencing fields from microelectronics to thin-film deposition.

Fundamental Principles

Scattering kinematics

In Rutherford backscattering spectrometry (RBS), the interaction between incident ions and target atoms is modeled using the binary collision approximation, which treats each scattering event as an isolated elastic collision between the projectile ion and a single target nucleus, neglecting many-body effects and electronic interactions for high-energy ions (typically MeV range). This approximation holds because the ion energies are much greater than atomic binding energies, allowing collisions to be analyzed as classical head-on or glancing encounters governed by conservation of energy and momentum. The energy of the backscattered ion is determined by the kinematic factor k, which relates the post-collision energy E_3 to the incident energy E_1 through the masses of the projectile m_1 and target m_2, and the laboratory-frame scattering angle \theta_2: k = \frac{E_3}{E_1} = \left( \frac{\sqrt{m_2^2 - m_1^2 \sin^2 \theta_2} + m_1 \cos \theta_2}{m_1 + m_2} \right)^2 This factor arises from solving the conservation equations for an elastic collision and quantifies the fractional energy retained by the projectile after scattering. For backscattering near 180°, k approaches \left[ (m_2 - m_1)/(m_2 + m_1) \right]^2 when m_2 > m_1, resulting in minimal energy loss for light ions (e.g., He⁺) incident on heavy targets like , where k \approx 0.92, enabling clear . Conversely, when the projectile mass exceeds the target mass (m_1 > m_2), the maximum scattering angle is limited to \arcsin(m_2 / m_1), preventing full backscattering and leading to greater energy transfer to the target. Scattering trajectories are often analyzed by transforming between the laboratory (where detectors are positioned) and the center-of-mass (CM) , where the total is zero, simplifying the collision to isotropic for interactions. In the lab , the observed scattering angle \theta_2 relates to the CM angle \theta_{CM} via \tan \theta_2 = \sin \theta_{CM} / [\cos \theta_{CM} + (m_1 / m_2)], allowing conversion of measured energies and angles to predict ion paths. This transformation is crucial for interpreting near-surface collisions, where small deviations in can affect depth sensitivity. Additionally, the recoil energy imparted to light atoms, given by E_r = E_1 (4 m_1 m_2 / (m_1 + m_2)^2) \sin^2 (\theta_2 / 2), enables detection of low-mass like oxygen or carbon through forward-recoil spectrometry, distinguishing them from heavier matrix contributions in the backscattered spectrum.

Rutherford scattering cross-section

The Rutherford scattering cross-section provides the fundamental probabilistic description of the interaction between incident ions and target atoms in backscattering spectrometry, quantifying the likelihood of events at a given . The classical differential cross-section, derived from the interaction between charged particles, is given by \frac{d\sigma}{d\Omega} = \left( \frac{Z_1 Z_2 e^2}{8\pi \epsilon_0 E} \right)^2 \frac{1}{\sin^4(\theta/2)}, where Z_1 and Z_2 are the atomic numbers of the incident and target atom, respectively, [e](/page/E!) is the , \epsilon_0 is the , [E](/page/Kinetic_energy) is the of the incident , and \theta is the scattering angle in the center-of-mass frame. This formula, originally developed for scattering, assumes a purely repulsive point-like potential with no nuclear contact or quantum effects, valid for high incident energies where the ion trajectories are and the distance of closest approach exceeds the . The assumptions underlying this cross-section hold for typical Rutherford backscattering conditions, such as incident energies above 100 keV for ions, where electronic screening by the target atom's electron cloud is negligible, allowing the interaction to be treated as unscreened. At these energies, the ion penetrates the electron cloud without significant deviation, and the scattering is dominated by the Coulomb field. The backscattered yield Y(\theta, E), or number of detected ions at \theta and E, is then proportional to the product of the target atom density N, the differential cross-section, and the \Omega subtended by the detector: Y(\theta, E) \propto N \cdot \frac{d\sigma}{d\Omega} \cdot \Omega. This relation enables quantitative prediction of scattering intensities from known sample compositions. Deviations from the classical Rutherford cross-section occur at lower energies, typically below 100 keV for light ions, due to electronic screening effects that soften the Coulomb potential. The Thomas-Fermi model describes this screening through a potential of the form \phi(r) = \frac{Z_1 Z_2 e^2}{4\pi \epsilon_0 r} \exp(-r/a), where a is the screening length, leading to a reduced cross-section compared to the unscreened case; corrections are often applied using the Molière potential for more accurate low-energy modeling. Quantum mechanically, the Rutherford formula remains valid for heavy ions like those used in backscattering spectrometry, as relativistic and spin effects are minimal; the Mott cross-section, which includes quantum corrections for identical particles, is primarily relevant for electron scattering rather than ion-nucleus interactions.

Instrumentation

Ion sources and accelerators

Rutherford backscattering spectrometry (RBS) primarily employs light ions such as ions (He⁺ or He²⁺, often referred to as alpha particles) with energies typically in the range of 1–4 MeV, which provide optimal and scattering for surface and near-surface analysis of most materials. Protons (H⁺) are also commonly used, particularly for deeper in lighter matrices due to their lower mass and reduced scattering cross-section compared to . For enhanced sensitivity to specific elements, such as light impurities in heavy matrices, heavier ions like (N⁺) or oxygen (O⁺) are selected, as their higher mass improves mass resolution and increases the backscattering yield from low-Z elements. Ion sources in RBS systems generate stable, high-brightness beams of these through plasma-based techniques. The duoplasmatron , which uses a low-pressure arc discharge between a hot filament and an with a for plasma confinement, is widely adopted for producing positive ions from gases like or , yielding currents up to several milliamperes. Radio-frequency (RF) and () offer alternatives for higher charge states and longer-lived beams, with ECR particularly suited for heavier ions due to its ability to sustain dense, hot via microwave heating in a . In accelerators, negative ions (e.g., He⁻) are produced via charge exchange with vapors like , enabling efficient acceleration. Accelerators provide the high voltages needed to energize the ion beams for effective . Single-ended Van de Graaff accelerators, using a moving belt or chain ( variant) to transport charge within an insulating gas enclosure like SF₆, routinely achieve energies up to 5 MeV with stable operation. Tandem Van de Graaff systems extend this to higher energies (up to several MeV for RBS) by injecting negative ions, stripping them to positive charge at a central high-voltage terminal, and accelerating them twice, which is ideal for beams at 2–3 MeV. For specialized applications requiring pulsed or higher-energy beams, cyclotrons or radio-frequency quadrupoles (RFQs) are employed in modern facilities, offering versatility for heavier ions. Compact tandetron accelerators, using solid-state voltage generation without moving parts, are increasingly used for routine RBS setups. Beam characteristics are critical for achieving precise depth and compositional in RBS. Energy better than 0.1% (often ~10⁻⁴ relative) is maintained through stable acceleration and magnetic analysis, minimizing energy spread in the incident beam. Typical beam currents range from nanoamperes to microamperes, sufficient for accumulating spectra in minutes to hours without sample damage. Spot sizes are collimated to 1 × 1 mm² for standard analysis but can be focused to micrometer scales using lenses and slits, enabling micro-RBS for spatially resolved measurements. RBS requires high-vacuum environments to prevent beam scattering or neutralization, with pressures of 10⁻³ to 10⁻⁴ Pa maintained in beam lines and target chambers using turbomolecular or diffusion pumps backed by rotary-vane fore-pumps. Beam transport from to sample involves a 90° analyzing to select mass and energy, followed by doublets for focusing and steering, and adjustable slits for collimation and emittance control, ensuring a clean, divergent-free beam at the target.

Detection and data acquisition systems

In Rutherford backscattering spectrometry (RBS), the primary detectors for measuring the of backscattered are surface barrier (SSB) diodes, which operate as detectors with a thin depleted layer for detection. These detectors provide high efficiency for in the MeV range typically used in RBS, achieving an of approximately 10–20 keV (FWHM) for alpha particles, depending on the and incident . The arises from the statistical fluctuation in production and collection within the , enabling separation of signals from different elements based on kinematic shifts. Detector geometry is optimized for kinematic separation of backscattered ions from different target elements, with placement typically at scattering angles of 120–170° relative to the incident beam direction. Common configurations include a single detector at around 165° for standard RBS, or annular detectors surrounding the beam axis to increase coverage up to near 180° while maintaining angular selectivity. Position-sensitive designs, such as resistive-charge division photodiodes, allow mapping of ion impacts across the detector face, providing additional for improved depth profiling in complex samples. Energy measurement relies on pulse-height analysis, where the backscattered ion generates electron-hole pairs in the SSB detector, producing a voltage pulse proportional to the ion's energy. This pulse is amplified by a charge-sensitive and shaped by a spectroscopic before digitization via an (ADC), with the resulting spectrum accumulated in a (MCA). Calibration of the system ensures accurate energy-to-channel mapping, often using known peaks. Data acquisition systems incorporate nuclear instrumentation module (NIM)-standard electronics to handle and storage, including pile-up rejection circuits to suppress overlapping pulses at high count rates and minimize spectral distortion. For heavy recoils or low-energy ions where energy loss in foils degrades performance, time-of-flight (TOF) options employ start-stop detectors with microchannel plates and carbon foils to measure velocity, enabling mass identification complementary to energy . Modern enhancements include large-area pixelated arrays and position-sensitive microchannel plate systems, which offer improved and higher throughput for multidimensional RBS mapping.

Data Analysis

Spectrum interpretation for composition

In Rutherford backscattering spectrometry (RBS), the energy spectrum of backscattered ions provides key insights into the of a sample. The high-energy edge of the spectrum represents backscattering events from atoms at the sample surface, where the incident ions undergo head-on collisions without prior loss. This edge occurs at the maximum possible backscattered , determined by the of the collision between the incident and the target atom. The yield, or number of backscattered ions detected at this edge, is proportional to Z_2^2, the square of the of the target atom, as dictated by the cross-section. This strong dependence on Z_2 enhances the sensitivity of RBS for heavier elements relative to lighter ones in the sample. Elemental identification in RBS relies on the unique positions of these kinematic edges in the , which arise from the mass-dependent during collisions. The backscattered E for a surface is given by E = K E_0, where E_0 is the incident and K is the kinematic , K = \left[ \frac{M_1 \cos \theta + \sqrt{M_2^2 - M_1^2 \sin^2 \theta}}{M_1 + M_2} \right]^2 for angle \theta, with M_1 and M_2 being the masses of the incident and target , respectively. For example, in a 2 MeV ^4He beam incident on a surface at 180° , the kinematic K \approx 0.92, placing the edge at approximately 1.84 MeV, distinctly higher than edges from lighter elements like (around 1.1 MeV for the same conditions). This separation allows unambiguous assignment of elements based on edge positions, particularly effective for heavy elements where K approaches 1. Quantification of elemental composition from RBS spectra involves calculating the areal density Nt (atoms per unit area) of a given element using the measured yield at its kinematic edge. For a uniform thin film or surface layer, this is expressed as Nt = \frac{Y}{\frac{d\sigma}{d\Omega} \Omega}, where Y is the number of detected backscattered ions (yield), \frac{d\sigma}{d\Omega} is the differential Rutherford cross-section at the scattering angle, and \Omega is the solid angle of the detector. This formula assumes single scattering and neglects energy loss for thin, uniform samples, enabling absolute concentration measurements without standards. Challenges arise when signals from elements of similar atomic masses overlap in the spectrum, as their kinematic edges may coincide or partially superimpose, complicating and quantification. To resolve such overlaps, spectra are often acquired at multiple incident ion energies or scattering angles, which shift the relative positions of the edges due to changes in the kinematic factor and cross-section. For instance, varying the energy from 2 MeV to 3 MeV can separate overlapping signals from elements like and in alloys. Accurate interpretation requires of the to the total dose delivered to the sample. This is typically achieved using a to integrate the incident charge, ensuring Y reflects the true scattering probability rather than beam fluctuations. Alternatively, for , the from a known element (e.g., in a Si wafer) can serve as an , leveraging its well-characterized . These methods provide precision on the order of 5% for elemental areal densities in simple compositions.

Simulation and quantitative modeling

Quantitative modeling in Rutherford backscattering spectrometry (RBS) relies on computational simulations to accurately interpret experimental spectra, enabling the extraction of precise information on sample composition, thickness, and depth profiles. These simulations account for trajectories, loss, and events within the target material, addressing the limitations of analytical approximations in complex structures. methods are particularly effective for this purpose, as they statistically track large numbers of ion paths to generate simulated spectra that can be directly compared to experimental data. In simulations, ion trajectories are followed through the sample, incorporating energy loss due to electronic and nuclear interactions, as well as angular deflections from . Energy loss straggling, which broadens the energy distribution of backscattered ions, is modeled using the Bohr approximation, treating it as a classical diffusive process arising from multiple small-angle collisions. This approach captures the statistical fluctuations in energy deposition, essential for simulating the tails and widths of spectral features in thick or multilayered samples. Validation of these simulations often involves comparing outputs to known standards, such as thin films of uniform composition, to ensure accuracy in trajectory predictions. Specialized software tools facilitate these simulations and fitting processes. SRIM/TRIM provides comprehensive databases and calculations for stopping powers and straggling parameters, serving as a foundational input for RBS models by estimating loss in various materials. For full and fitting, programs like SIMNRA and RUMP employ or analytical algorithms to generate theoretical spectra, allowing users to adjust parameters iteratively until the matches the experimental yield across the entire range. SIMNRA, in particular, integrates advanced physics modules for handling diverse ion-target combinations, making it a standard for quantitative RBS analysis. Layered model fitting involves representing the sample as a stack of homogeneous layers and optimizing parameters such as thickness, elemental composition, and density to minimize the difference between simulated and experimental spectra, often using least-squares or evolutionary algorithms. This iterative process starts with an initial guess based on kinematic edges and refines it by adjusting layer properties, ensuring the simulated backscattering yield aligns with observed peak heights and shapes. Such fitting is crucial for resolving overlapping signals in multicomponent systems, providing absolute quantification without destructive sample preparation. To enhance realism, simulations incorporate non-Rutherford effects that deviate from pure , including multiple which causes angular spreading and energy broadening, and detector which convolves the ideal with instrumental broadening. Multiple is modeled by tracking successive small-angle deflections, using approximations like the theory integrated into codes, while detector effects are added as Gaussian broadening post-simulation. These inclusions are vital for high-fidelity fits in deeper or denser samples where such effects can shift peak positions by several keV. Uncertainty analysis in RBS modeling quantifies errors from sources like energy straggling, which introduces statistical broadening, and beam energy spread, typically 0.1-1% of the incident energy, affecting depth resolution. Error propagation is performed by varying these parameters in simulations and assessing their impact on fitted values, often using resampling to generate confidence intervals for thickness or composition. Validation against , such as NIST thin-film standards, confirms the reliability of these models, with typical uncertainties below 5% for well-resolved layers.

Applications

Depth profiling and elemental distribution

In Rutherford backscattering spectrometry (RBS), the primary mechanism enabling depth profiling is the progressive energy loss of incident ions as they penetrate the sample material, quantified by the stopping power \frac{dE}{dx}. This energy loss arises predominantly from inelastic collisions with electrons in the target atoms and is described by the Bethe-Bloch formula, which provides the mean energy loss per unit path length for charged particles traversing matter: -\frac{dE}{dx} = \frac{4\pi z^2 e^4 N Z}{m_e v^2} \left[ \ln \left( \frac{2 m_e v^2}{I (1 - \beta^2)} \right) - \beta^2 \right], where z and v are the charge and velocity of the incident ion, N Z is the electron density of the target, m_e is the electron mass, \beta = v/c, and I is the mean excitation energy of the target atoms. As ions backscattered from deeper layers have undergone greater energy loss both inbound and outbound, their detected energies are shifted downward relative to surface-scattered ions, allowing reconstruction of elemental distributions as a function of depth. This non-destructive approach typically probes depths up to several micrometers, depending on ion energy and material composition. The depth d corresponding to a given energy shift \Delta E in the backscattered spectrum is approximated by d = \frac{\Delta E}{\epsilon}, where \epsilon is the effective energy loss factor, often calculated as \epsilon \approx \frac{dE/dx_{\text{in}} + k \cdot dE/dx_{\text{out}}}{\cos \phi} with k as the kinematic scattering factor and \phi the ion trajectory angle. Depth resolution, limited by energy straggling (the statistical fluctuation in energy loss) and detector resolution, achieves 10–50 nm for MeV helium ions near the surface, degrading slightly with depth due to increased straggling. For rough interfaces, straggling broadens spectral features, but this can be accounted for in analysis to estimate interface roughness without assuming perfect layering. In multilayered structures, RBS spectra exhibit characteristic edge shifts for elements at buried interfaces; for instance, in a SiO₂/Si stack, the oxygen edge appears at lower energy than the edge, with the shift proportional to the oxide thickness, enabling precise layer sequencing up to ~500 nm total depth. For thicker films exceeding the ion's range, the yield progressively drops off as fewer ions backscattered from depth reach the detector, providing an indirect measure of film thickness. Heavy ions enhance isotopic depth profiling by increasing the Rutherford cross-section (\sigma \propto Z_1^2 Z_2^2) for better sensitivity to or isotopically similar elements. Representative applications include mapping dopant distributions in semiconductors, such as arsenic implants in silicon, where RBS reveals concentration profiles with ~20 resolution to assess implantation uniformity and diffusion. Similarly, oxidation layers in alloys are profiled to evaluate growth kinetics and composition gradients, with edge analysis distinguishing oxide from substrate signals up to 100 depths. Surface elemental composition can be inferred from the high-energy edges of the spectrum, providing initial boundary conditions for full depth reconstruction. Recent applications as of 2025 include the use of RBS for evaluating elemental depth profiles and radiation hardness in cells for space photovoltaics, where ion migration under stressors is assessed to improve device durability. Additionally, RBS has been applied to analyze the formation and composition of superconducting V₃Si thin films on substrates, supporting integration into technologies.

Structural characterization via channeling

In ion channeling during Rutherford backscattering spectrometry (RBS), a of energetic ions aligned parallel to a major crystallographic or in a is steered by the electrostatic potential of the atomic rows or planes, minimizing close encounters with target atoms and substantially reducing the backscattering yield relative to randomly oriented incidences. This steering effect, arising from correlated multiple scattering within the continuum approximation of the , confines ions to open channels between lattice sites, enabling depth-resolved assessment of crystalline perfection. The technique exploits this reduction in yield—often by a factor of 10 to 100 compared to isotropic scattering—to probe subsurface structure without significant disruption to the sample. Blocking complements channeling by creating enhanced scattering or shadowing patterns when ions or backscattered particles align with surface rows directed toward the detector, forming shadows that dip the yield from aligned surface atoms. In perfect , this blocking reduces the detected signal from the topmost layer, providing sensitivity to surface and . The effect is particularly pronounced for outgoing trajectories, where atoms obstruct ions from deeper layers, allowing isolation of contributions in multilayer structures. Yield measurements in channeling RBS quantify structural through the minimum yield χ_min, defined as the of aligned backscattering to random at the sample surface, which is typically below 5% for high-quality crystals due to the low fraction of ions undergoing close collisions. Dechanneling, the progressive increase in with depth from scattering events that eject ions from channels, is characterized by the dechanneling length, which shortens in the presence of defects and reveals their profile. For instance, in irradiated materials, elevated χ_min values above 10% indicate significant disorder, while the slope of the dechanneled traces defect . Applications of channeling RBS focus on semiconductors, where it evaluates ion implant damage by mapping increased dechanneling from displaced atoms. The technique also assesses strain in epitaxial films, as tensile or compressive distortions broaden dechanneling or shift yield profiles, aiding optimization of heterostructures like SiGe on . In damaged layers from implantation, channeling distinguishes amorphous regions (χ_min ≈ 100%) from partially recovered crystals, guiding annealing processes. Angular scans in channeling RBS involve tilting the across a channeling to measure the width of the dip, governed by the ψ_{1/2} ≈ \sqrt{\frac{2 Z_1 Z_2 e^2}{d E}}, where Z_1 and Z_2 are the numbers of the incident and target atom, respectively, d is the spacing between atoms along the , and E is the . This angle, typically 0.5° to 2° for MeV He ions in , marks the transition from channeled to random scattering and sensitively detects misalignments from or mosaicity, with narrowing indicating improved perfection. Scans thus provide a direct metric for atomic row straightness, essential for validating .

Surface and interface analysis

Rutherford backscattering spectrometry (RBS) enables precise analysis of atomic-scale surface modifications through the enhanced backscattering signal from the topmost atomic layer, known as the surface peak, which arises from the higher probability of at low penetration depths. This peak is particularly sensitive to coverages as low as ~1 (ML), allowing detection of adsorbates or contaminants without significant shadowing from subsurface atoms. High-resolution RBS (HRBS), achieved via energy-sensitive detectors and incidence or exit angles, resolves individual monolayers by minimizing depth-of-field broadening. Surface relaxation and reconstruction alter the positions of top-layer atoms, manifesting as shifts in RBS blocking patterns or peak intensities under channeling conditions. For instance, the clean Au(110) surface exhibits a (1×2) reconstruction characterized by a missing-row structure, where every second chain of top-layer atoms is absent, accompanied by a lateral displacement of 0.12 ± 0.02 in the second layer and an outward vertical relaxation of the top layer by >0.25 . These structural changes increase the RBS surface peak yield beyond simple missing-row predictions, requiring modeling of interlayer displacements to fit experimental spectra and confirm boundaries separated by monatomic steps exposing (111) facets. Medium-energy ion scattering (MEIS), a high-resolution variant of RBS using 20-200 keV s, further refines sub-nm analysis of such relaxations by enhancing angular yield variations. RBS excels in probing interfaces within thin films, quantifying interdiffusion and elemental segregation by tracking peak broadening or shifts in elemental yields across the boundary. In polycrystalline metal films, thermal annealing induces grain-boundary diffusion, with RBS revealing penetration depths and phase formation, such as Pd₃Si at Pd/β-SiC interfaces after 400°C treatment, where the interface widens without complete degradation. For alloys, RBS detects surface segregation of lighter elements, as seen in Y-Ba-Cu-O films on MgO substrates, where Ba and Cu diffuse into the substrate, forming a ~10 nm intermixed layer. In electronics, RBS monitors oxide growth on semiconductors, with interface peaks shifting to deeper energies as SiO₂ layers thicken during oxidation, providing non-destructive thickness measurements up to 100 nm with 1-2 nm resolution at the interface. In catalysis, RBS determines adsorbate binding sites and coverages, such as Cs on Si(111)-7×7, where saturation at 0.51 ± 0.02 ML occurs preferentially at adatom and rest-atom sites, influencing work function reduction and reaction promotion. Impact collision ion scattering spectroscopy (ICISS), an extension of RBS using low-energy (keV) ions in impact mode, maps atomic trajectories via azimuthal and polar angle scans of backscattered yields, revealing real-space positions of surface atoms with sub-Ångström precision; for example, it delineates shadow cones around Pt(111) sites to quantify relaxations. These techniques support electronics applications by characterizing oxide-semiconductor interfaces for device reliability.

Limitations and Comparisons

Resolution and sensitivity constraints

The depth resolution in Rutherford backscattering spectrometry (RBS) is primarily limited by energy straggling, which arises from statistical fluctuations in the energy loss of incident ions as they penetrate the sample, typically resulting in a of about 1–5% of the probed depth or 5–20 near the surface. This effect worsens for deeper layers and is exacerbated by the finite energy of detectors, often around 10–20 keV (FWHM) for surface barrier detectors, leading to broader spectral peaks and reduced ability to resolve thin layers or interfaces. For elements, depth resolution is particularly poor due to lower backscattered energies and greater relative straggling, often requiring complementary techniques for accurate profiling. Mass resolution in RBS is constrained by kinematic factors in , where the of backscattered ions depends on the of the and atoms, causing overlaps in signals for elements of similar mass, such as carbon (Z=6) and (Z=14) when contributions from different depths coincide. While RBS provides good mass separation for light elements (e.g., , , ) due to significant differences in from projectiles, resolution deteriorates for heavy elements (e.g., , ) as the kinematic shift becomes minimal, limiting discrimination between nearby masses. Sensitivity to low-Z elements is further reduced because they produce low-energy backscattered signals that overlap with noise or signals from heavier matrix elements, often necessitating forward detection for better analysis. Detection limits in RBS typically range from 0.1–1 at.% for heavy elements in light matrices, owing to the large Rutherford cross-sections that yield strong signals, but rise to 1–10 at.% or higher for light elements due to smaller cross-sections and increased background. Background contributions from multiple events, where ions undergo several small-angle deflections, further degrade by broadening low-energy tails in the , particularly at depths beyond 0.5–1 μm. Overall can reach parts per million for heavy impurities, but light element detection remains challenging without enhancements like resonant . The technique operates effectively in the 1–5 MeV energy range for ions, where the approximation holds, but deviates into non-Rutherford regimes below ~100 keV due to electron screening effects that alter the cross-section and above ~10 MeV from nuclear interactions that introduce resonances and non-elastic processes. Sample requirements impose additional constraints: analyses assume flat, smooth surfaces for accurate depth scaling, as roughness can broaden resolution by 10–20%; insulating materials may accumulate charge under bombardment, distorting beam trajectories unless mitigated by conductive coatings or flooding; and high fluences (>10^16 ions/cm²) can induce , such as defect creation or amorphization, especially in sensitive materials like polymers or semiconductors.

Complementary ion beam techniques

Rutherford backscattering spectrometry (RBS) is often complemented by other analysis techniques to address its limitations, particularly in detecting light elements and providing high sensitivity for traces or isotopes. These methods, such as analysis (NRA) and detection (ERD), extend the analytical capabilities for comprehensive material characterization, especially in thin films and multilayers. Nuclear reaction analysis (NRA) utilizes nuclear reactions induced by ion beams, typically protons or deuterons, to detect light elements (atomic number Z = 1–15) and isotopes in heavy matrices, where RBS exhibits low sensitivity due to its reliance on Rutherford scattering from heavier atoms. For instance, reactions like ¹⁴N(d,α)¹²C enable quantification of nitrogen in titanium nitride multilayers with sensitivities down to 0.1% or 1 × 10¹⁵ atoms/cm², complementing RBS's strength in heavy-element depth profiling. NRA provides isotope-specific information without the high-Z background interference common in RBS, making it ideal for applications like analyzing deuterium in tungsten or oxygen in silicon carbide oxidation studies. Elastic recoil detection (ERD), often implemented as time-of-flight ERD (TOF-ERDA), detects light recoiling atoms (e.g., and ) forward-scattered from the sample using heavy incident ions, offering superior sensitivity for (0.1%) compared to RBS, which is insensitive to it. This technique achieves depth resolutions of ~20 nm near the surface and profiles up to 2.5 µm, enabling quantitative distribution in materials like films or interfaces. ERD complements RBS by resolving light-element profiles that overlap or are undetectable in backscattering spectra, as demonstrated in combined analyses of solar cells. Particle-induced X-ray emission (PIXE) employs proton beams to excite characteristic X-rays from elements ( ≥ 13), providing non-destructive trace elemental identification with sensitivities of 0.1–100 ppm, far exceeding RBS for mid-to-high traces but lacking inherent depth resolution beyond tens of µm. Unlike RBS, which offers model-independent depth profiling, PIXE yields integrated signals requiring complementary techniques for layering, as in or biological samples. It is particularly useful when RBS signals overlap, such as in analysis in thin foils. In comparison to electron-based surface techniques, RBS probes deeper (nm to µm scale) for quantitative, non-destructive bulk and depth analysis, while () and () excel in Ångstrom-level surface sensitivity and chemical speciation but are limited to ~10 nm depths and may require for profiling. detects elements via photoelectrons with ~10 Å resolution, ideal for contaminants, whereas provides ppm isotopic sensitivity but introduces matrix effects; RBS avoids these for heavier elements in thicker films. Hybrid approaches integrate RBS with other modalities for enhanced structural insights; for example, RBS in channeling mode quantifies crystallographic defects non-destructively, complementing (TEM), which provides atomic-scale imaging but is sample-destructive and limited to small areas. Studies on ion-bombarded layers use RBS/channeling to measure damage buildup (e.g., dechanneling yields) alongside TEM for visualizing dislocations. Recent efforts explore RBS integration with scanning tunneling microscopy (STM) for in-situ surface monitoring during analysis, though such setups remain experimental.

References

  1. [1]
    1.15: Rutherford Backscattering of Thin Films - Chemistry LibreTexts
    Aug 28, 2022 · Rutherford backscattering analysis is a straightforward technique to determine the thickness and composition of thin films (< 4000 Å).Introduction · Basic Concept of Rutherford... · Experimental Set-up
  2. [2]
    [PDF] Ion Beam Analysis in Materials Science - Nuclear Data Program
    ▫ Rutherford backscattering spectrometry (RBS). ▫ Introduction-history. ▫ Basic concepts of RBS. ▫ Kinematic factor (K). ▫ Scattering cross-section.
  3. [3]
    Rutherford Backscattering Spectrometry - Dresden - HZDR
    Rutherford Backscattering Spectrometry (RBS) is a physical method for the for the determination of the elemental composition of thin films, especially for ...
  4. [4]
    [PDF] The Scattering of α and β Particles by Matter and the Structure of the ...
    Aug 21, 2002 · The Scattering of α and β Particles by Matter and the Structure of the Atom. E. Rutherford, F.R.S.*. Philosophical Magazine. Series 6, vol. 21.Missing: text | Show results with:text
  5. [5]
    On a diffuse reflection of the α-particles - Journals
    In the following experiments, however, conclusive evidence was found of, the existence of a diffuse reflection of the α-particles. A small fraction of the α- ...Missing: setup | Show results with:setup
  6. [6]
    Geiger and Marsden's Alpha deflection article of 1913 - chemteam.info
    The 1913 article showed that alpha particles can be deflected through large angles, and that the results supported Rutherford's theory of a central charge in ...Missing: timeline | Show results with:timeline
  7. [7]
    Chemical Analysis of Surfaces by Nuclear Methods - ACS Publications
    Chemical Analysis of Surfaces by Nuclear Methods ... Rubin · T. O. Passell · L. E. Bailey. ACS Legacy Archive. Open PDF. Analytical Chemistry. Cite this: Anal.
  8. [8]
    None
    ### Key Milestones in RBS Development Post-WWII (1950s-1980s)
  9. [9]
    Quantitative Rutherford Backscattering from Thin Films | MRS Bulletin
    Nov 29, 2013 · In the early 1960s, several of the low-energy atomic and nuclear physics laboratories started to use their accelerators to do research in solid- ...
  10. [10]
    [PDF] Brief introduction to Rutherford Backscattering Spectrometry - KTH
    Rutherford backscattering spectrometry (RBS), as a technique to address thin film properties in materials research, has evolved from classical nuclear.
  11. [11]
    Computer methods for analysis and simulation of RBS and ERDA ...
    Since the seventies a large number of computer methods have been developed for data analysis in Rutherford backscattering spectrometry (RBS).
  12. [12]
    [PDF] BACKSCATTERING SPECTROMETRY - Sites do IFGW
    ln Appenelix F, we proviele tables of kinematic factors, scattering cross sections, anel various forms of energy loss anel energy loss factors. Analyses or ...
  13. [13]
    Backscattering Spectrometry - Google Books
    Title, Backscattering Spectrometry ; Authors, Wei-Kan Chu, James W. Mayer, Marc-A. Nicolet ; Edition, 4, illustrated ; Publisher, Academic Press, 1978 ; Original ...
  14. [14]
    [PDF] LXXIX. The scattering of α and β particles by matter and the structure ...
    To cite this Article Rutherford, E.(1911) 'LXXIX. The scattering of α and β particles by matter and the structure of the atom', Philosophical Magazine ...
  15. [15]
    Rutherford Scattering Cross-Section
    The probability per unit time of the particle being scattered into the range of solid angle ${\mit\Omega}$ to ${\mit\Omega}+d{\mit\Omega}$ then the ...
  16. [16]
    [PDF] RBS - Rutherford Backscattering Spectrometry - IF-UFRJ
    The methods are applicable to the detection of all elements to a depth of several µm, with sensitivities in the range of 10-8 to 10-6 g/cm2.” Page 7. © Matej ...Missing: paper | Show results with:paper
  17. [17]
    [PDF] Scattering in Quantum Mechanics - UF Physics
    Sep 25, 2013 · We thus obtain the well-known Rutherford differential cross section: ... general formula for the differential cross section: dσ. dΩ. = M fi q ...
  18. [18]
    [PDF] Instrumentation for PIXE and RBS
    The BSCAT code was developed for analysis and simulation of complex spectra recorded in RBS, NRA and PIXE experiments [146]. It runs on MS. Windows 95 and ...
  19. [19]
  20. [20]
    Rutherford Backscattering Spectrometry (RBS) Tutorial
    Rutherford Backscattering (RBS) is based on collisions between atomic nuclei and derives its name from Lord Ernest Rutherford, who in 1911 was the first to ...
  21. [21]
    [PDF] Introduction to Charged Particle Detectors
    The advantage of surface barrier technology is that it allows production of transmission detectors ... Representation of Electric Field E in Silicon Surface ...
  22. [22]
    [PDF] Performance of resistive-charge position sensitive detectors for RBS ...
    The performance of two types of 1 1 cm2 photodiode position sensitive detectors (PSDs) based on resistive charge division was evaluated for their use in ...
  23. [23]
    High Resolution Rutherford Backscattering (HR RBS) Detector
    A position sensitive micro channel plate detection system to count the ions as a function of position/energy. Mounted on tracks, the system can easily be ...Missing: modern spectrometry
  24. [24]
    [PDF] ion scattering techniques - Encyclopedia of Materials Characterization
    They are Rutherford. Backscattering Spectrometry, RBS, which typically utilizes high-energy He or H ions (usually 1-3.4 MeV energies), Medium-Energy Ion ...
  25. [25]
    In Situ Rutherford Backscattering Spectrometry for Electrochemical ...
    May 28, 2019 · This work demonstrates the potential for in situ RBS to become a powerful tool for the investigation of a wide range of electrochemical processes.
  26. [26]
    (PDF) Backscattering Spectrometry, Wei-Kan Chu, James W. Mayer ...
    PDF | This book is a prime reference book on ion beam ... Backscattering Spectrometry, Wei-Kan Chu, James W. Mayer, and Marc-A. Nicolet, Academic Press, 1978.
  27. [27]
    Evolution of Rutherford's ion beam science to applied research ...
    The kinematic factor depends on the Mi which allows backscattered ions from different elements to appear at different backscattered energies. Elemental ...<|separator|>
  28. [28]
    Scattering Cross Section - an overview | ScienceDirect Topics
    The T-RBS ... The areal density, the number of atoms per unit area, NS determined from the scattering cross section σ(ϑ) by. [5] N S = Y σ ( ϑ ) Q d Ω.
  29. [29]
    Research Methods | Louisiana Accelerator Center
    Rutherford Backscattering Spectrometry (RBS) is based on measuring the energy of elastically backscattered ions. This method builds upon Rutherford's classic ...
  30. [30]
    a case study of RBS analysis of multinary materials - Nature
    Apr 8, 2024 · RBS is ideally suited for the study of multilayered materials with sufficiently different atomic masses (hence sufficiently different kinematic ...
  31. [31]
    Simple and accurate spectra normalization in ion beam analysis ...
    Total beam charge integration from an isolated chamber enclosing the sample. The chamber acts a Faraday cup. •. Total charge integration from a positively ...
  32. [32]
    [PDF] Calibrating the Faraday Cup for Ion Beams Using Rutherford ...
    RBS and Faraday cup measurements were conducted varying parameters such as type of ion beam, beam energy, puck insertion and focus of the beam. Results of ...Missing: normalization dose
  33. [33]
    [PDF] Monte Carlo simulation of Ion Beam Analysis spectra using Corteo.
    Feb 27, 2009 · The current version of Corteo is intended to simulate Rutherford Backscattering Spectrometry. (RBS), Elastic Recoil Detection (ERD) and ...
  34. [34]
    [PDF] Monte Carlo Simulation of Electron Backscattering in Solids ... - CORE
    Mar 9, 1996 · A Monte Carlo study of backscattering of kilovolt electrons in solids, a process of primary importance in electron microscopy and surface ...
  35. [35]
    [PDF] Combining RBS/Channeling, X-ray diffraction and atomic-scale ...
    May 6, 2021 · represents the RBS spectrum obtained from a simulation of 1.4 MeV He ions on an Fe ... Computer simulation of atomic-displacement cascades.
  36. [36]
    James Ziegler - SRIM & TRIM
    SRIM includes quick calculations which produce tables of stopping powers, range and straggling distributions for any ion at any energy in any elemental target.Download SRIM-2013 · SRIM / TRIM · SRIM Tutorials · Details of SRIM-201Missing: RBS | Show results with:RBS
  37. [37]
    [PDF] SIMNRA User's Guide
    May 21, 2010 · SIMNRA simulates charged particle energy spectra with Rutherford scattering cross-sections (RBS), non-Rutherford cross- sections (EBS), nuclear ...
  38. [38]
    SIMNRA, a Simulation Program for the Analysis of NRA, RBS and ...
    SIMNRA is a Microsoft Windows 95/Windows NT program with fully graphical user interface for the simulation of non-Rutherford backscattering, nuclear reaction ...
  39. [39]
    Differential evolution optimization of Rutherford backscattering spectra
    Oct 26, 2022 · Common fit parameters in Rutherford backscattering analysis are the target model comprising the thicknesses of all layers as well as the ...Ii. The Differential... · Vi. Differential Evolution... · Vii. Uncertainty And Error...
  40. [40]
    Rutherford backscattering from layered structures beyond the single ...
    Rutherford backscattering spectroscopy (RBS) is a powerful and widely used technique for the quantitative surface layer analysis of solids [1], [2], [3]. The ...
  41. [41]
    [PDF] Improved Physics in SIMNRA 7
    SIMNRA is an analytical code for the simulation of ion beam analysis energy spectra obtained by Rutherford backscattering, non-Rutherford scattering, ...
  42. [42]
    Simulation of multiple scattering background in heavy ion ...
    It is also possible for multiple scattering from the lighter substrate atoms to lead to scattered projectiles with an energy comparable to the energy of ...
  43. [43]
    Depth profile determination with confidence intervals ... - IOP Science
    The contribution of geometrical straggling due to finite beam size and width of the detector aperture is negligible in the RBS geometry, unlike in ERDA ...
  44. [44]
    [PDF] IMPROVEMENT OF THE RELIABILITY AND ACCURACY OF ...
    Then, the uncertainty in the energy determination was considered, again. 5 keV for the beam energy spread, plus the uncertainty stemming from the time.<|control11|><|separator|>
  45. [45]
    Uncertainty minimization in electronic stopping cross-section ... - arXiv
    Oct 17, 2025 · These include uncertainties related to the angular alignment of the beam and detector, variations in the material density, energy calibration, ...
  46. [46]
    Thin film depth profiling by ion beam analysis - RSC Publishing
    Aug 31, 2016 · We have at last demonstrated what was long promised, that RBS can be used as a primary reference technique for the best traceable accuracy ...
  47. [47]
    Ion Beam Analysis | College of Science and Engineering
    Rutherford backscattering spectrometry (RBS) to probe elements from Be on up through the heaviest elements in the periodic table; Forward recoil spectrometry ...
  48. [48]
    Detailed arsenic concentration profiles at Si/SiO2 interfaces
    Aug 20, 2008 · The concentration was calibrated by normalizing the whole profile to the measured RBS dose of 1.1 × 10 16 cm − 2 ⁠. This protocol has been ...
  49. [49]
    Heavy ion Rutherford scattering for target profiling, and ...
    Also, heavy ion RBS can be used to profile multilayer targets. The high of heavy ion beams gives good depth resolution even on targets consisting entirely of ...
  50. [50]
    [PDF] NOTES ON CHANNELING
    A decisive step in the development was Jens Lindhard's theoretical analysis in ... excitation, or Rutherford backscattering, for example. The most direct ...
  51. [51]
    Materials Analysis by Ion Channeling - ScienceDirect.com
    Materials Analysis by Ion Channeling. Submicron Crystallography. Book • 1982. Authors: LEONARD C. FELDMAN, JAMES W. MAYER and S. THOMAS PICRAUX ...
  52. [52]
    [PDF] Channeling analysis in studying ion irradiation damage in materials ...
    [1-3] Among various types of characterization techniques, Rutherford backscattering spectrometry in its channeling mode (RBS/C, or ion channeling) has been ...
  53. [53]
    RBS with monolayer resolution - ScienceDirect.com
    It is shown that both high-energy resolution and the grazing angle technique are essential to the monolayer resolution. Results of some applications of HRBS are ...
  54. [54]
    [PDF] Domain structure of the clean reconstructed Au(110) surface
    Apr 15, 1984 · Rutherford-backscattering. (He+) measurements of the Au(110)-(110) surface peak were significantly larger than expected for a simple missing ...
  55. [55]
    (PDF) Medium-energy ion-scattering analysis of the c(2×2) structure ...
    The clean Au(110) surface exhibits a (1×2) reconstruction with an atomic arrangement corresponding to a missing-row (MR) structure. By depositing Cs on this ...
  56. [56]
    SIMS and RBS study of thermally annealed Pd/β-SiC interfaces
    ### Summary of RBS Studies on Interfaces in Thin Films: Interdiffusion Examples
  57. [57]
    Study of interdiffusion between thin Y–Ba–Cu–O films and MgO ...
    Study of interdiffusion between thin Y–Ba–Cu–O films and MgO substrates by applying Rutherford backscattering spectrometry combined with scanning tunneling ...
  58. [58]
    RBS and ERD analysis of semiconductor device materials
    This surface peak, which is not detected when measuring in a UHV scattering chamber, originates from adsorbed hydrocarbons that are cracked at the surface under ...
  59. [59]
  60. [60]
    Ion scattering spectroscopy in the impact collision mode (ICISS)
    The MD-FAN map had stronger agreement with the experimental ICISS measurements than did the BCA map, despite containing nearly 100-fold fewer ion trajectories.
  61. [61]
    [PDF] Rutherford Backscattering Spectrometry
    History. Sir Ernest Rutherford (1871 - 1937). • 1911: Rutherford's ... RBS as materials analysis method. • 1957: S. Rubin, T.O. Passell, E. Bailey ...
  62. [62]
    Charging effects in the ion beam analysis of insulating polymers
    ... Rutherford Backscattering Spectrometry (RBS). The potential that develops ... damage by penetrating ionizing radiation [7]. The charge accumulation ...
  63. [63]
    Radiation induced electrical degradation in ceramics insulators | AIP ...
    Results from Rutherford Backscattering Spectroscopy (RBS) confirm that radiation enhanced diffusion of the gold contact is responsible for the degradation ...
  64. [64]
    Techniques | Dutch Institute for Fundamental Energy Research
    Our ion beam can use six different techniques: RBS, NRA, ERD, PIXE, PIGE and Irradiation. The following paragraphs give an overview of each technique.
  65. [65]
    [PDF] Ion beam techniques for the analysis of light elements in thin films ...
    Ion beam techniques are used to analyze light elements in thin films, including depth profiling, to quantify their composition and distribution.
  66. [66]
    Comparison with Other Techniques: Surface Analysis - HORIBA
    For XPS and AES, this value is approximately 3 monolayers (≈10 angstroms), SIMS and SNMS, 10 monolayers, and on the order of 100 monolayers for RBS.
  67. [67]
    Selecting the best surface analysis method for your materials/samples
    May 30, 2024 · For this reason, these methods are generally complementary rather than redundant with XPS providing a greater surface sensitivity (relevant for ...
  68. [68]
    SIMS vs. Other Surface Analysis Techniques - Atomfair
    XPS and AES provide superior chemical state information and are better suited for surface studies, while RBS offers non-destructive, quantitative bulk analysis ...
  69. [69]
    (PDF) RBS/channeling and TEM study of damage buildup in ion ...
    Jan 9, 2015 · A systematic study on structural defect buildup in 320 keV Ar-ion bombarded GaN epitaxial layers has been reported, by varying ion fluences ...
  70. [70]
    RBS/Channeling and TEM Analysis of Thin Sandwiched EPI-Layers ...
    Feb 28, 2011 · The layers were analyzed by Rutherford backscattering(RBS)-channeling and transmission electron microscopy (TEM). For structures containing 4 or ...Missing: characterization | Show results with:characterization
  71. [71]
    RBS studies of the absolute coverage of Cs on the Si(1 1 1)-7×7 ...
    They concluded that Cs atoms initially adsorb on the adatom sites (A-type in Fig. 1) and then on the rest atoms (B-type sites in Fig. 1) to form an ordered ...