Fact-checked by Grok 2 weeks ago

Electron backscatter diffraction

Electron backscatter diffraction (EBSD) is a scanning ()-based technique that analyzes the crystallographic , phase distribution, and microstructure of crystalline materials at the micro- to nanoscale by capturing and indexing patterns formed by backscattered electrons. In this method, an beam interacts with a tilted sample surface (typically at 70°), producing Kikuchi patterns that reveal the and at each scanned point, enabling automated mapping of grains, boundaries, textures, and strains across large areas. The origins of EBSD trace back to 1928, when Seishi Kikuchi and Shoji Nishikawa first observed and published electron backscatter patterns using crystals, laying the groundwork for understanding Kikuchi diffraction. Significant advancements occurred in the , with J.A. Venables introducing EBSD to by attaching video cameras for pattern capture, followed by David Dingley's development of the first commercial systems in the and automated mapping techniques in the early 1990s using Hough transforms for band detection. Modern EBSD systems, often integrated with () for compositional analysis, achieve high-speed data acquisition—over 3 million points in minutes—with spatial resolutions down to the nanometer scale, making it indispensable for materials characterization. EBSD finds broad applications in , , , and , including and analysis in metals, phase identification in minerals, deformation studies in semiconductors, and quality assessment of thin films and biomaterials. Key advantages include its non-destructive nature, ability to correlate microstructure with mechanical properties, and compatibility with in-situ experiments under varying conditions like or , thereby supporting optimization of manufacturing processes in industries such as , , and .

Principles and Fundamentals

Basic principles

Electron backscatter diffraction (EBSD) utilizes backscattered electrons generated in a (SEM) to produce Kikuchi diffraction patterns that reveal crystallographic information about the sample. When a focused beam of high-energy electrons (typically 10–30 keV) strikes a crystalline specimen tilted at approximately 70° to the beam, a portion of the electrons scatter back from the interaction volume, which extends roughly 10–100 nm into the sample depending on the material. These backscattered electrons, having wavelengths on the order of atomic spacings, diffract coherently from the crystal lattice planes, forming characteristic patterns projected onto a detector screen. The formation of Kikuchi patterns arises from the interaction of incident electrons with the crystal , involving both inelastic and processes. Inelastic predominates initially, as electrons lose through interactions with (phonons) or collective excitations (plasmons), randomizing their directions and creating a divergent source of lower- electrons within the crystal. Subsequent events, governed by interactions with atomic nuclei and electrons, lead to Bragg when the scattering angle satisfies the condition nλ = 2d_{hkl} sin θ_{hkl}, where n is the order, λ is the electron , d_{hkl} is the interplanar spacing for planes (hkl), and θ_{hkl} is the Bragg angle. Channeling effects further influence this process, as electrons can be temporarily trapped and guided along open planes or directions, enhancing the probability of from specific orientations and contributing to the directional emission of backscattered electrons. The de Broglie wavelength λ of the electrons, essential for , is calculated using the relativistic formula \lambda = \frac{h}{\sqrt{2 m e V \left(1 + 0.9788 \times 10^{-6} V\right)}}, where h is Planck's constant, m is the electron rest mass, e is the elementary charge, and V is the accelerating voltage in volts; this accounts for relativistic corrections at typical energies, yielding λ ≈ 0.01 nm. The resulting Kikuchi patterns feature prominent dark and bright bands (Kikuchi bands) that trace the traces of diffracting crystal planes, with band width equal to twice the Bragg angle and edges perpendicular to the (hkl) planes denoted by —a triplet of integers (hkl) specifying the and spacing of planes relative to the crystal axes. Intersections of these bands define zone axes, while the central zero-order Laue zone exhibits a network of low-index bands surrounding the transmitted beam direction, providing a direct map of the crystal's .

Historical development

The roots of electron backscatter diffraction (EBSD) trace back to the first observation of Kikuchi diffraction patterns, including backscatter configurations, in 1928 by Seishi Kikuchi and Shoji Nishikawa using crystals and high-energy electrons recorded on photographic plates. In 1937, Hans Boersch reported backscatter electron diffraction patterns from crystalline samples, capturing Kikuchi bands through an oblique incidence geometry in a setup akin to , which provided further insights into dynamical electron diffraction effects. During the 1940s, advanced the theoretical framework by extending his dynamical diffraction theory to electrons, explaining the formation of Kikuchi lines and bands observed in backscatter configurations. The technique gained practical momentum in the 1970s with the advent of the (SEM), enabling in-situ observation of Kikuchi patterns from bulk samples. A pivotal advancement occurred in 1973 when J.A. Venables and C.J. Harland demonstrated EBSD's potential for crystallographic analysis within an SEM, using screens to visualize patterns from tilted specimens. By the late 1970s and early 1980s, researchers like D.J. Dingley refined the method for sub-micron resolution, with Dingley's 1984 work highlighting its application for local orientation determination. The first automated EBSD systems emerged around 1980, transitioning from manual pattern interpretation to computer-assisted indexing, exemplified by early commercial efforts from companies like International Scientific Instruments (ISI). The marked EBSD's standardization and widespread adoption in . In 1991, S.I. and colleagues introduced fully automated lattice orientation determination, leading to the development of orientation imaging (OIM) in 1993 and the first comprehensive automated orientation maps in the early . HKL Technology launched its software in 1994, providing robust tools for pattern acquisition, indexing, and mapping that became industry standards for high-speed EBSD analysis. This era shifted EBSD from labor-intensive manual systems to automated, commercial platforms integrated with SEMs, facilitating routine microtexture studies. Further evolution in the and focused on enhanced resolution and speed. In 2004, D.J. Dingley, in collaboration with A.J. Wilkinson and G. Meaden, introduced high-resolution EBSD (HR-EBSD), leveraging of patterns to measure strains and rotations with sub-pixel . The brought a detector revolution with the adoption of complementary metal-oxide-semiconductor () sensors, starting around 2015 with direct electron detection systems that boosted pattern acquisition rates to thousands per minute, improving data quality and enabling large-area mapping.

Instrumentation and Setup

Microscope integration and geometry

Electron backscatter diffraction (EBSD) is integrated into scanning microscopes (SEMs) through specialized sample holders and detector assemblies that enable the precise of beam-sample-detector to maximize the collection of backscattered electrons for pattern formation. The standard configuration tilts the sample at approximately 70° from the horizontal, positioning the surface at a 20° incidence angle relative to the vertical beam, which optimizes the backscattering yield and enhances the in the resulting patterns. This tilt also facilitates the forward of diffracted electrons toward the detector, reducing within the sample and improving overall pattern contrast. The EBSD detector is mounted forward of the sample, typically at an angle of 20–30° from the sample normal, to subtend a large (often >0.5 sr) for efficient capture of the backscattered . Working distances of 10–25 mm are preferred, providing adequate clearance for the detector insertion while preserving resolution and minimizing lens aberrations that could degrade beam focus. Accelerating voltages between 15 and 30 kV are optimized for this setup, as they balance sufficient energy for with limited to maintain surface sensitivity; lower voltages (down to 5–10 kV) may be used for high-resolution applications but require adjustments to reduce noise from reduced signal intensity. In operation, the electron beam interacts with the tilted sample surface, forming a teardrop-shaped interaction volume where primary electrons penetrate to depths of tens of nanometers to a few micrometers, depending on material density and voltage, though the diffracting volume is confined to the uppermost 20–50 due to rapid energy loss and . For mapping, the beam is raster-scanned across the surface at step sizes from nanometers to micrometers, with the geometry ensuring consistent interaction conditions across the field of view. High vacuum (typically <10^{-5} mbar) is required in conventional SEMs to minimize beam scattering by residual gases and prevent sample contamination, but environmental SEM adaptations enable low-vacuum EBSD (around 0.1–1 mbar) for non-conductive or beam-sensitive samples, albeit with potential trade-offs in pattern sharpness from increased scattering. Key geometric factors influencing pattern resolution include foreshortening due to the high tilt, which compresses features along the tilt direction and alters apparent grain dimensions, and projection effects onto the detector plane, which can distort band widths and curvatures if the effective pattern center deviates from optimal positioning. These effects are mitigated through precise alignment and calibration, ensuring accurate representation of crystallographic information without introducing artifacts in orientation mapping.

Detectors and detection methods

Electron backscatter diffraction (EBSD) patterns are captured using specialized detectors integrated into scanning electron microscopes, with traditional systems relying on phosphor screens coupled to charge-coupled device (CCD) or complementary metal-oxide-semiconductor (CMOS) cameras. These indirect detection methods, pioneered in the 1980s, convert backscattered electrons into visible light via a scintillator phosphor screen, which is then optically imaged by the camera to form a digital pattern. Early implementations used low-light CCD cameras for pattern acquisition, enabling automated orientation mapping but limited by light conversion inefficiencies and readout speeds. By the 1990s and 2000s, advancements included binary detectors, such as forward scatter detectors (FSDs), which combine phosphor-based EBSD with solid-state diodes for simultaneous orientation and compositional contrast. High-speed variants emerged in the 2010s, incorporating faster CMOS sensors to achieve frame rates exceeding 1000 patterns per second, facilitating large-area mapping in materials like metals and composites. The introduction of direct electron detectors (DEDs) in 2013 marked a significant evolution, utilizing pixel array detectors like the Timepix hybrid pixel sensor to directly convert incoming electrons into electrical signals without intermediate light production. This approach, first demonstrated for energy-filtered EBSD patterns, enhances sensitivity and reduces noise by eliminating phosphor-related losses, enabling higher signal-to-noise ratios at lower beam currents. DEDs, such as those based on , support readout speeds suitable for dynamic experiments and improve overall detection efficiency, with quantum efficiencies exceeding 90% compared to approximately 10-20% for traditional phosphor-CCD systems due to the avoidance of optical coupling inefficiencies. Modern DED implementations achieve angular resolutions as low as 0.1° through refined pattern quality and cross-correlation analysis, surpassing the ~0.5-1° typical of earlier indirect detectors. Recent advancements, including static and stage-mounted DEDs introduced in 2023, further minimize geometric distortions by co-aligning the detector and sample in a tilt-free configuration, reducing mean angular errors to below 0.005° and enhancing sensitivity to low-energy backscattered electrons (threshold ~3 keV). These systems support frame rates up to 1000 Hz while maintaining high dynamic range, allowing for rapid indexing with exposures as short as 0.02 seconds without frame averaging. In March 2025, Bruker introduced the eWARP detector, a pioneering system that advances technology for enhanced materials characterization in scanning electron microscopes. Overall, DEDs offer superior detective quantum efficiency over phosphor-based detectors, particularly for low-signal conditions, though indirect systems remain advantageous for energy-weighted detection in high-throughput applications.

Sample preparation techniques

Sample preparation is crucial for electron backscatter diffraction (EBSD) analysis, as the technique is highly sensitive to surface quality, requiring a flat, strain-free surface to generate clear Kikuchi patterns from the top 20-40 nm of the material. Improper preparation can introduce artifacts such as distorted patterns due to residual stresses or charging effects, compromising orientation mapping accuracy. Mechanical polishing forms the foundation of EBSD sample preparation, typically involving sequential grinding with silicon carbide papers (from coarse 180-400 grit to fine 1200-2500 grit) followed by diamond lapping on cloths with suspensions decreasing from 9 μm to 1 μm particle size. The process culminates in a final polish to a surface finish better than 0.05 μm using 0.04-0.06 μm diamond or alumina suspensions on soft cloths, often under automated conditions to ensure uniformity and minimize scratches or relief. For enhanced results, this is followed by electropolishing in electrolytic solutions such as 90% acetic acid with 10% perchloric acid at 20-40 V for 10-30 seconds, or ion milling with an argon beam at 2-6 kV and 4-30° incidence angles for 5-30 minutes, both of which remove the deformed Beilby layer (typically 10-100 nm thick) to yield strain-free surfaces ideal for high-indexing-rate EBSD. Non-conductive samples, such as ceramics or polymers, require a thin conductive coating to prevent electron charging under the electron beam, which would otherwise degrade pattern quality. Carbon sputtering or evaporation at 2-5 nm thickness is preferred, as it minimally attenuates the backscattered signal while providing sufficient conductivity; alternatives like gold or tungsten can be used but may require higher beam energies if thicker (up to 10 nm). For metals and composites, vibratory polishing has seen advances in the 2020s, particularly for alloys like or , where horizontal vibration at 90 Hz with 0.05 μm colloidal silica for 1-8 hours produces deformation-free surfaces by gently abrading without introducing new stresses. (FIB) milling is employed for preparing cross-sections, using Ga+ ions at 5-30 kV to create site-specific trenches 10-50 μm wide, enabling EBSD on internal microstructures while avoiding bulk mechanical damage. To avoid artifacts like residual stresses that distort lattice parameters and degrade pattern contrast, preparation protocols emphasize low-force polishing, water or alcohol lubrication during grinding, and post-mechanical chemical etching (e.g., Keller's reagent for 4 seconds on Al alloys). Recent 2023-2024 reviews highlight tailored protocols for additively manufactured samples, such as sequential vibratory polishing followed by electrochemical polishing in perchloric acid mixtures for AlSi10Mg, achieving >95% indexing rates by mitigating porosity-induced stresses and surface oxidation. Depth considerations in EBSD preparation focus on removing the surface layer to access representative bulk material, as patterns originate from ~20-100 nm depth; thus, final ion milling or targets 50-200 nm removal to eliminate preparation-induced damage without altering subsurface composition.

Data Acquisition

Pattern formation and collection

In electron backscatter diffraction (EBSD), the incident electron beam from a () interacts with the crystalline sample, which is typically tilted at approximately 70° to optimize the escape of diffracted electrons. This interaction generates backscattered electrons, comprising roughly 20-50% of the incident beam depending on the material's , with the diffracted portion forming Kikuchi patterns through Bragg from planes within a shallow interaction volume of 10-50 nm depth. The Kikuchi bands in these patterns arise from pairs of diffracted beams, appearing as high-intensity zones separated by dark minima, and their formation occurs over interaction depths influenced by the accelerating voltage (typically 15-30 kV). Pattern generation at each measurement point requires a beam of approximately 100-500 ms to accumulate sufficient signal intensity for clear band definition, balancing acquisition speed with quality. The collection workflow begins with positioning the stationary electron beam on the sample surface, where backscattered electrons are projected onto a screen or located 10-30 mm from the specimen to form a visible diffraction . This is then captured by a high-sensitivity camera, such as a (CCD) or electron-multiplying CCD (EMCCD), often with on-chip binning (e.g., 2x2 or 4x4) to reduce noise and readout time by combining adjacent pixels, thereby improving without sacrificing essential resolution. The process emphasizes high beam currents in the 1-15 nA range to ensure adequate diffracted electron flux, though excessive current can induce sample heating or damage. Detector roles, such as phosphor excitation for indirect imaging or direct electron detection in CMOS-based systems, facilitate efficient capture but are optimized separately for overall system performance. Several factors influence pattern quality, including beam current, which must be sufficient (typically several ) to produce sharp, high-contrast bands but not so high as to cause beam-induced artifacts like charging or drift. Sample plays a role, as elevated temperatures (e.g., above 100-200°C) can broaden bands and reduce sharpness due to thermal vibrations disrupting coherent scattering, necessitating cryogenic or controlled heating stages for sensitive materials. Accelerating voltage also affects and pattern clarity, with higher voltages (20-30 kV) yielding broader interaction volumes but potentially weaker relative band intensities. Simulation tools are essential for predicting and validating , employing dynamical diffraction models such as the Bloch wave method to compute wave propagation through the crystal lattice, accounting for multiple events. These simulations, often implemented in software like EMsoft or DREAM.3D, reproduce experimental Kikuchi patterns with high fidelity by solving the for waves, enabling virtual experiments to optimize parameters before physical acquisition. The raw output from pattern collection consists of digital images, typically in 400x400 pixel resolution, encoding intensity gradients across the diffraction field as grayscale bitmap files for subsequent processing. These patterns capture the full angular range (often 60-120° half-angle) with 8- or 16-bit depth to preserve subtle variations in band contrast and background.

Mapping procedures

Mapping procedures in electron backscatter diffraction (EBSD) involve systematically scanning a defined area of the sample to collect orientation data at multiple points, enabling the construction of spatial maps of crystallographic orientations. The process typically employs a step-scan approach, where the electron beam is rastered across the sample surface in a grid pattern, with step sizes commonly ranging from 0.05 to 10 μm depending on the desired resolution and microstructure features. For each grid point, an EBSD pattern is acquired with exposure times of 50-500 ms, resulting in total acquisition durations of several hours for large-area maps, such as those covering hundreds of micrometers. This raster scanning builds on single-point pattern collection by automating the beam positioning to cover the area efficiently. Indexing success rates in these mappings exceed 95% under optimal conditions, such as well-prepared samples and appropriate parameters, though non-indexed pixels—often occurring at boundaries or in deformed regions—are handled through from neighboring points to fill gaps in the map. Automated software workflows, including Oxford Instruments' AZtec and EDAX TSL's OIM, control the acquisition by integrating scanning, pattern capture, and initial indexing in , streamlining the process from setup to data output. Parameter optimization is crucial for effective mapping, with step size selected relative to average grain size—ideally 1/10th to 1/5th of the grain diameter—to balance spatial resolution and acquisition efficiency; for instance, a 2.5 μm grain size might use a 250 nm to 500 nm step for detailed boundary resolution without excessive time. High-speed variants, enabled by complementary metal-oxide-semiconductor (CMOS) detectors, achieve rates over 1000 points per second, significantly reducing map times for large areas while maintaining quality. The resulting maps are typically output in formats such as Euler angle representations for precise data or inverse (IPF) color coding, where colors correspond to crystallographic directions relative to the sample , facilitating visual interpretation of and orientations.

Resolution and limitations

The of electron backscatter (EBSD) is typically on the order of 20–50 laterally, primarily limited by the electron interaction volume within the sample, which determines the size of the region contributing to the pattern. This resolution can vary with accelerating voltage and material properties, achieving values as low as 40 in optimized conditions for metals. Angular resolution in standard EBSD is approximately 0.5–1°, reflecting the with which orientations can be determined from indexing. This limit arises from factors such as quality and the accuracy of band detection algorithms, though it suffices for most microstructural analyses. Depth , or the information depth from which backscattered s contribute meaningfully to the , ranges from 10–40 at typical operating voltages around 20 , with shallower depths in denser materials due to reduced . Lower voltages, such as 5 , can further restrict this to about by minimizing the interaction volume. Common limitations of EBSD include challenges in low-indexed regions, such as grain boundaries or twins, where exhibit reduced contrast and band visibility, leading to indexing errors or non-indexed pixels. Charging artifacts are particularly problematic for non-conductive samples, causing distortion and beam deflection that degrade data quality. In deformed samples, overlap from strained lattices can further complicate indexing, resulting in artifacts like false boundaries. Mitigations for these issues include operating at low accelerating voltages below 5 kV to achieve nanometer-scale by shrinking the interaction volume, though this requires careful to maintain pattern quality. correction in the enhances pattern sharpness, improving overall resolution and reducing distortions. Quantitative metrics for assessing EBSD data reliability include the hit rate, defined as the percentage of successfully indexed patterns, which can drop below 80% in challenging regions like deformed or low-indexed areas. Confidence index thresholds in analysis software, typically set above 0.1–0.5 on a 0–1 scale, help filter low-quality data by evaluating pattern matching reliability, with higher values indicating robust indexing.

Data Analysis Methods

Pattern indexing algorithms

Pattern indexing algorithms in electron backscatter diffraction (EBSD) are computational techniques designed to extract and information from individual diffraction patterns by identifying and interpreting key features such as Kikuchi bands. These algorithms process the raw pattern images to determine the orientation matrix, which describes the rotation relating the crystal lattice to the sample , enabling subsequent microstructural analysis. The primary approaches include edge-detection methods like the and template-matching strategies such as dictionary indexing, each balancing speed, accuracy, and robustness against pattern quality variations. The Hough transform-based indexing, introduced in the , remains a cornerstone for automated EBSD analysis due to its efficiency in detecting linear features in patterns. It applies to highlight Kikuchi band edges, followed by the —a feature extraction technique that maps potential lines from image space to peaks in a parameter space (ρ, θ), where ρ is the from the origin and θ is the angle of the normal to the line. Detected peaks correspond to band positions, and intersections of multiple bands (typically 7–9) are used to vote for candidate orientations via interplanar angle matching. Refinement then optimizes the initial solution by minimizing projection errors, where the orientation matrix R (a 3×3 ) is adjusted to reduce the least-squares discrepancy between observed band traces and their theoretical projections onto the detector plane, often formulated as: \min_R \sum_i \left( \mathbf{p}_i^{\text{obs}} - \mathbf{p}_i^{\text{proj}}(R) \right)^2 Here, \mathbf{p}_i^{\text{obs}} are observed band positions, and \mathbf{p}_i^{\text{proj}}(R) are projections based on the gnomonic model of the diffraction geometry. This method achieves indexing rates exceeding 90% hit rates on high-quality patterns but can struggle with low-contrast or noisy data. Dictionary-based approaches, refined in the 2010s, enhance accuracy by comparing the entire experimental pattern to a precomputed dictionary of simulated patterns rather than relying solely on band edges. Simulations are generated using dynamical electron scattering models for a dense sampling of orientations (e.g., millions of entries) and phases, accounting for detector geometry and pattern center. Matching employs normalized cross-correlation to identify the best-fit simulation, yielding sub-degree orientation precision and improved performance on deformed or multiphase materials. For instance, dictionaries can incorporate band contrast and zone axis intensities for robust indexing, with computational efficiency boosted by subspace projections or GPU acceleration in recent implementations. Commercial and implements these algorithms for practical EBSD workflows. TSL OIM from EDAX supports both Hough-based triplet indexing and dictionary matching with dynamic simulations for high-fidelity pattern generation, enabling refinement and confidence indexing. Open-source alternatives like PyEBSD provide Python-based tools for processing and custom dictionary indexing, facilitating reproducible research and integration with scientific computing ecosystems. To handle errors such as , pseudosymmetry, or multiphase scenarios, algorithms incorporate schemes like triplet voting, where candidate orientations are scored based on consistent matches of band triplet angles across phases (tolerance ~0.5–1°). This reduces false positives by requiring consensus from multiple band combinations, improving indexing reliability to >95% in challenging datasets while distinguishing phases through reflector list comparisons.

and phase mapping

Once indexed electron backscatter diffraction (EBSD) patterns are obtained, orientation mapping reconstructs the crystallographic orientations across the scanned area by assigning each pixel an orientation descriptor, typically in the form of or quaternions. , defined by three rotation parameters (φ₁, Φ, φ₂) in the Bunge convention, quantify the rotation from a reference frame to the sample , enabling precise representation of orientations. Quaternions, as four-dimensional vectors, offer a compact and singularity-free alternative for orientation calculations, particularly useful in avoiding issues during processing. These orientations are commonly visualized using inverse pole figures (IPFs), which project the distribution of orientations onto a of the sample's normal, tangential, or loading directions, with colors corresponding to specific crystallographic planes aligned with the measurement axis. For texture analysis, pole figures plot the intensity of specific crystallographic planes (e.g., {111} or {100}) as a function of their orientation relative to the sample axes, revealing preferred orientations such as or sheet textures in polycrystalline materials. Phase mapping in EBSD discriminates between multiple crystalline phases by comparing experimental patterns to simulated diffraction patterns for known crystal structures, allowing identification based on band positions, widths, and intensities unique to each phase. This dictionary-based approach simulates Kikuchi patterns using dynamical electron scattering models for candidate phases, then matches them to acquired patterns via cross-correlation or voting schemes to assign phase labels pixel-by-pixel. Band contrast, a measure of the sharpness and visibility of Kikuchi bands in the pattern, provides complementary phase discrimination by highlighting regions of lower pattern quality often associated with secondary phases or defects, as phases with different atomic numbers or structures exhibit varying backscattering efficiencies. For instance, in multiphase alloys like steels, this enables mapping of ferrite, austenite, and carbide phases with spatial resolutions down to the sub-micrometer scale. Grain boundaries are detected in orientation maps by calculating the misorientation between adjacent pixels, with thresholds typically set at 5° to delineate high-angle grain boundaries (HAGBs, >5° misorientation) from low-angle boundaries or subgrain structures. This misorientation is computed using the smallest rotation between the quaternions of neighboring pixels, often averaged over multiple neighbors to reduce noise sensitivity. Quantitative outputs from these maps include distributions, derived from boundary tracing to measure equivalent circle diameters or area-based metrics, providing statistics such as mean and log-normal distributions that characterize recrystallization or deformation states. is further quantified via the distribution function (ODF), reconstructed from data using series expansion methods (e.g., or Bunge series), yielding coefficients that describe the volume fraction of orientations in Euler space and enable prediction of anisotropic properties. Post-processing enhances map reliability through noise filtering, such as neighbor or wild-spike removal, which interpolates or discards outlier pixels based on misorientation consistency, improving hit rates from ~80% to over 95% in challenging scans. tracing algorithms then refine outlines by segmenting regions of continuous below the misorientation , using flood-fill or methods to compute boundary lengths, curvatures, and types (e.g., twin vs. random), facilitating automated extraction of microstructural . These steps, often implemented in like TSL OIM or Oxford Channel 5, ensure robust mapping for materials .

Pattern center determination

In electron backscatter diffraction (EBSD), the pattern center is defined as the point where the incident beam intersects the plane of the screen, representing the foot of the perpendicular from the beam interaction volume on the tilted sample surface to the screen. This location serves as the origin for the of the Kikuchi diffraction pattern, enabling accurate transformation of pattern features into crystallographic orientations relative to the sample coordinates. Precise determination of the pattern center is essential, as it directly influences the reliability of subsequent indexing and mapping processes. Methods for pattern center determination include both manual and automated approaches. Manual techniques typically involve projecting lines from identified Kikuchi bands in the pattern to their intersection point, often using known geometric features or shadow-casting aids like grids or wires on the sample. Automated methods, which are more commonly used in modern systems, rely on detecting and fitting imaging lines or bands via algorithms such as the Hough or , followed by refinement through of pattern features or least-squares optimization of band positions. For instance, commercial software like Oxford Instruments' Channel 5 implements automated band detection and fitting routines to compute the pattern center coordinates efficiently during . Accuracy in pattern center determination is paramount, with errors below 0.1% of the detector width required for high-fidelity measurements, as larger deviations can introduce angular inaccuracies exceeding 1° and compromise analysis. Advanced algorithms, such as techniques that simultaneously refine the pattern center and crystal across multiple patterns, can achieve errors as low as 0.05% or better, even in noisy or low-resolution data. These methods often model the pattern center in normalized coordinates (x*, y*, z*) to account for projection geometry. Challenges in pattern center determination arise primarily from sample tilt distortions, which can shift the effective projection if the tilt angle is not precisely known, and from or binning in captured patterns that obscure features. Validation typically involves checking for in indexed orientations across an EBSD map, where systematic variations indicate errors.

Advanced Measurement Techniques

Strain and stress analysis

Early efforts to measure using electron backscatter diffraction (EBSD) in the pre-2000s relied on analyzing relative misorientations within grains to provide qualitative assessments of plastic deformation. These methods examined intragranular orientation variations arising from residual dislocations introduced by straining, often mapping misorientation gradients to infer local distortion levels. However, such approaches were limited in precision, typically achieving strain resolutions around $10^{-3}, due to orientation measurement accuracies of 0.5° to 1° and the empirical nature of correlating misorientations to quantitative without direct . Subsequent advancements introduced reference pattern subtraction techniques, where experimental EBSD patterns are compared to a strain-free pattern via to detect shifts in axes and Kikuchi band positions indicative of lattice . This method, pioneered in the mid-2000s, quantifies small in pattern features, enabling the extraction of the tensor that decomposes into and components. By selecting a nearby unstrained region as the , these shifts reveal distortions with high sensitivity, though the requires careful alignment to avoid artifacts from large misorientations. To infer stress from these measured strains, is applied using the material's elastic stiffness tensor C, where the stress tensor \sigma = C : \varepsilon. This conversion accounts for anisotropic elastic properties, allowing mapping in polycrystalline materials, such as near grain boundaries or defects. For instance, in metallic alloys, tensile strains of $10^{-4} might correspond to stresses on the order of tens of , depending on the . However, assumptions of and known C values are critical, and reference selection remains a key challenge to ensure accurate stress inference without confounding plastic effects.

High-resolution EBSD

High-resolution electron backscatter diffraction (HR-EBSD) is an advanced variant of conventional EBSD that employs analysis to achieve sub-pixel precision in measuring strains and rotations, enabling the mapping of deformations and local misorientations at the nanoscale. Introduced in 2006 by Wilkinson, Meaden, and Dingley, this technique involves normalizing experimental EBSD patterns to remove intensity gradients and noise, followed by remapping the patterns to a common orientation using initial Hough-based indexing results. is then applied between corresponding regions of interest in the remapped experimental patterns and a reference pattern, quantifying sub-pixel shifts that correspond to displacements in the diffracting . The core procedure relies on a dictionary of simulated Kikuchi patterns generated from dynamical electron scattering models, which serve as idealized, distortion-free references for comparison. These simulations account for the sample's , orientation, and detector geometry, allowing the decomposition of measured shifts into components of the displacement gradient tensor, including elastic (with precision around 10^{-4}) and (with precision around 10^{-4} rad or approximately 0.006°). The shifts are derived from the full-field , providing maps of relative and rotation fields across the scanned area, typically at spatial resolutions of 20–100 depending on beam conditions. Selecting an appropriate reference pattern is critical, as it defines the zero-strain baseline; ideally, stress-free simulated patterns are used to enable absolute measurements, though experimental references from low-strain regions in the sample are common for relative assessments. To mitigate artifacts from noisy or distorted references, neighborhood averaging techniques combine multiple adjacent patterns to generate a more robust reference, improving and reliability. However, inaccuracies in reference selection can propagate errors, particularly in heterogeneous microstructures. Subsequent developments have enhanced HR-EBSD's utility, including the integration of kernel average misorientation (KAM) metrics, which calculate the average misorientation between a and its neighbors to estimate local and derive densities of geometrically necessary dislocations (GNDs) on the order of 10^{14}–10^{16} m^{-2}. Precision has improved in the through refined remapping algorithms and corrections, achieving resolutions around 10^{-4} (approximately 0.006°) or better while maintaining below 10^{-4}. As of 2025, further enhancements include the use of monolithic active sensors (MAPS) for direct detection, improving and in , alongside GPU-based software for . These advances, exemplified by pattern-matching refinements and faster computational pipelines, have broadened applications in quantifying deformation mechanisms. Despite these gains, HR-EBSD faces limitations, including high computational demands from intensive and processes, often requiring offline analysis on high-performance . Additionally, the technique's sensitivity diminishes in severely deformed microstructures, where large rotations or plastic strains exceed the linear approximation of the , leading to reduced quality and mapping accuracy.

Transmission EBSD

Transmission electron backscatter diffraction (t-EBSD), also known as transmission Kikuchi diffraction (TKD), is a variant of electron backscatter diffraction performed in geometry, typically within a (). This technique involves preparing electron-transparent samples, such as foils thinned to less than 100 , and collecting patterns formed by electrons transmitted through the sample to a detector positioned below it. The method was introduced around 2011, enabling the analysis of nanoscale features in SEM without the need for a dedicated transmission electron microscope (TEM). Subsequent developments in the 2020s have integrated t-EBSD with scanning transmission electron microscopy in SEM (STEM-in-SEM), allowing combined imaging and diffraction for enhanced characterization of thin specimens. t-EBSD offers significant advantages over conventional reflection-mode EBSD, particularly in achieving higher of approximately 5-10 , which is suitable for mapping sub-10 domains. This improvement arises from the reduced interaction volume in thin samples, minimizing lateral and channeling effects that degrade pattern quality in bulk materials. Applications of t-EBSD are prominent in the study of , including nanoparticles and thin films, where it facilitates orientation mapping at scales inaccessible to standard EBSD. It is also effective for phase identification and crystallographic analysis in two-dimensional () materials, such as dichalcogenides, providing insights into grain structure and defects. Despite these benefits, t-EBSD presents challenges, including the need for precise sample thinning, often achieved via (FIB) milling, which can introduce artifacts if not optimized. Additionally, the lower intensity of transmitted electron signals compared to backscattered ones necessitates longer exposure times, potentially limiting throughput for large-area mapping.

Emerging Developments

Machine learning integration

has significantly enhanced electron backscatter diffraction (EBSD) analysis, particularly in pattern indexing and , by enabling direct prediction of crystallographic orientations from raw patterns without relying on traditional dictionary-based methods. Convolutional neural networks (CNNs) have emerged as a key tool for this purpose, trained to regress directly from EBSD patterns, achieving mean disorientation errors as low as 1.77° on simulated datasets and extending to experimental data via . These models, developed in advances from 2023 to 2025, address limitations of classical Hough-based indexing by processing high-noise patterns that would otherwise yield zero solutions, thus improving indexing success rates in challenging materials like those under large . Supervised learning approaches predominantly utilize vast simulated EBSD datasets to train these networks, bypassing the need for extensive experimental labeling while capturing the physics of Kikuchi . For instance, CNNs trained on over 300,000 simulated patterns demonstrate robust generalization to real-world scans, enabling dictionary-free indexing that eliminates the computational overhead of precomputing dictionaries. In multiphase alloys, models for phase identification, such as those segmenting EBSD maps in dual-phase steels, achieve accuracies up to 95.92% by classifying phases based on pattern features, outperforming traditional thresholding methods in complex microstructures. Generative models, including conditional variational autoencoders (CVAEs) and GANs, further support this by parametrically simulating patterns for training , enhancing model performance on underrepresented phases. A notable 2025 advancement is the Latice method, a (VAE) that compresses EBSD patterns into a 16-dimensional for rapid orientation mapping, delivering 99.9% data compression and a 7.5-fold over dictionary indexing while maintaining sub-1° disorientation accuracy on recrystallized stainless steel.00470-9) models have also been integrated for pattern refinement, denoising low-quality acquisitions to boost indexing reliability; the DIFFRACT approach, , uses a UNet-based to clarify Kikuchi bands, preserving surface features and enabling indexing of patterns that traditional methods fail. These techniques collectively offer up to 10-fold processing compared to Hough transforms and excel at handling low-quality patterns from deformed or multiphase samples, facilitating high-throughput EBSD in materials research.00470-9)

Compressive and accelerated imaging

Compressive sensing techniques in electron backscatter diffraction (EBSD) enable efficient by probe positions on the sample grid, reducing the number of collected patterns while reconstructing full datasets through advanced algorithms. For instance, uniform density sampling can acquire patterns at rates as low as 1% to 25% of a complete grid, such as 10% for band contrast maps or 5% for inverse pole figure (IPF) maps, significantly lowering data volume and exposure time. This subsampling generates incomplete 4D EBSD datasets, analogous to those in 4D-STEM, which are reconstructed using iterative algorithms that exploit sparsity in crystallographic orientations across the map. A key method employs (BPFA), a dictionary-learning approach that sparsely codes patterns in a shared basis, iteratively missing regions while preserving structural details like grain boundaries. Binning and direct electron detectors (DEDs) further accelerate acquisition, with CMOS-based DEDs achieving speeds exceeding 5000 indexed patterns per second by directly detecting without scintillator distortion, and region-of-interest (ROI) scanning focusing the beam on targeted areas to minimize unnecessary exposures. These approaches yield 10- to 100-fold reductions in mapping time for large areas, maintaining through priors that ensure high structural similarity (SSIM) values—plateauing at 10% for band contrast and 5% for IPF Z-maps—thus enabling of beam-sensitive materials without quality loss. However, challenges persist, including artifacts such as blurring at boundaries in complex microstructures, particularly under low rates or noise levels below 10 dB , as observed in 2024-2025 studies.

Multimodal and 3D extensions

Three-dimensional electron backscatter diffraction (3D EBSD) extends traditional two-dimensional EBSD by incorporating serial sectioning techniques, primarily using focused -scanning electron microscopy (FIB-SEM) systems, to reconstruct volumetric microstructural data. This approach, often termed slice-and-view , involves alternating cycles of ion beam milling to remove thin material layers (typically on the order of 1 μm per slice) and subsequent EBSD mapping of the exposed surface. Developed in the early with initial automated implementations around , the method has been refined in the 2020s to enable higher-throughput acquisitions through improved software and hardware integration. Multimodal integrations enhance 3D EBSD by combining it with complementary techniques for richer datasets. EBSD paired with energy-dispersive X-ray spectroscopy () allows simultaneous mapping of crystallographic orientation and chemical composition, revealing correlations between microstructure and elemental distribution in complex alloys. Similarly, integrating EBSD with digital image correlation () facilitates in-situ deformation studies, where surface patterns enable strain tracking alongside orientation evolution during mechanical loading. These combined mappings are typically performed in a single session, leveraging shared scanning parameters for pixel-level registration. Key techniques in EBSD include , which stacks and aligns sequential EBSD slices into a volume using fiducial markers or image correlation algorithms, followed by analysis of derived properties such as texture gradients or intragranular fields. This enables quantification of volumetric features like morphologies and boundary networks that are inaccessible in sections. For nanoscale applications, EBSD (t-EBSD), also known as transmission Kikuchi diffraction (TKD), supports nano-analysis by imaging thin electron-transparent samples, achieving sub-100 nm resolution for reconstructing deformation in nanostructured devices. Recent advancements focus on automated workflows to handle large datasets exceeding 10^6 voxels, as demonstrated in 2024 studies on comprehensive characterization of materials, incorporating correlative EBSD-EDS for efficient post-processing and artifact correction. These pipelines reduce manual intervention, enabling analysis of microstructures with millions of data points in hours rather than days. Despite these progresses, 3D EBSD remains time-intensive due to the sequential nature of sectioning and mapping, often requiring hours to days for volumes beyond 100 slices, and susceptible to alignment errors from milling artifacts or drift, which can distort reconstructed geometries if not mitigated by advanced registration software.

Applications

Microstructural characterization

Electron backscatter diffraction (EBSD) enables detailed mapping of structures in polycrystalline materials by acquiring data across scanned areas, typically resolving s as small as 0.2 μm in scanning electron microscopes (FEGSEMs). is quantified using metrics such as equivalent (ECD) and aspect ratios, derived from reconstructed areas, with accuracy improving when s encompass at least 8 pixels to achieve within 5% error. Shape factors, including Feret s and elongation ratios, reveal morphological variations, such as elongated s in deformed samples versus equiaxed forms post-annealing. Grain boundaries are classified by misorientation , distinguishing low-angle boundaries (LABs, <15°) from high-angle boundaries (HABs, >15°), with LABs often representing subgrain walls formed during deformation. Coincident site (CSL) boundaries, such as Σ3 twins, are identified using the Brandon criterion (misorientation tolerance θ = 15° Σ^{-1/2}), enabling quantification of special boundaries that influence material properties like resistance; for instance, in α-brass, Σ3 boundaries constitute up to 67% of the total. EBSD maps visualize these boundaries as lines overlaid on images, facilitating analysis of boundary character distribution (BCD) across microstructures. Texture analysis via EBSD involves computing the orientation distribution function (ODF) from Euler angle datasets to detect recrystallization, where random orientations indicate recrystallized grains contrasting with deformation textures. In rolled low-carbon steels, ODFs reveal weakened <100> fiber textures post-recrystallization, with machine learning-enhanced sampling ensuring accurate representation from reduced datasets (e.g., 350 grains capturing full ODF fidelity). For interstitial-free (IF) steels, EBSD identifies nucleation at γ-fiber (<111>//ND) orientations during annealing at 650–680°C, with recrystallized fractions tracked via misorientation gradients and ODF evolution from α- and γ-fibers. Deformation microstructures exhibit subgrain formation through dislocation rearrangement, visualized as LAB networks in EBSD maps with misorientations of 1–4.5°. Dynamic recovery promotes subgrain growth and polygonization, reducing stored energy, as seen in high stacking fault energy (SFE) stainless steels (~30 mJ/m²) where cross-slip limits LAB density compared to lower SFE alloys (~17 mJ/m²) that sustain higher dislocation densities. In rolled 316L stainless steel, subgrain sizes stabilize at ~0.9 μm after multiple passes, with bimodal lognormal distributions reflecting heterogeneous fragmentation. Case studies highlight EBSD's utility in welds and additive manufacturing. In ferritic steel welds, EBSD maps reveal grain refinement via acicular ferrite in -hybrid processes, yielding volume-weighted average sizes finer than welds, with bi-modal distributions (coarse primary ferrite alongside fine secondary phases) influencing toughness. For (LPBF) 316L , (lack-of-fusion defects with aspect ratios 2.4–3.1) orients parallel to build layers, correlating with anisotropic but minimal yield strength impact, as columnar s (10–500 μm) and cellular substructures persist post-heat treatment. Quantitative metrics from EBSD include neighbor distribution functions, quantifying correlated misorientations (e.g., twin pairs at Σ3), and aspect ratios averaging 1.5–2.0 in deformed grains, providing statistical insights into without exhaustive enumeration. These , often via software like MTEX, support scalable of neighbor pairing and trends in large datasets.

and analysis

Electron backscatter diffraction (EBSD) facilitates discrimination in multiphase materials by comparing experimental Kikuchi patterns to simulated patterns stored in crystallographic databases, allowing identification of phases based on matching quality and parameters. This pattern-matching approach, often implemented via dictionary indexing, refines assignment by evaluating band detection and alignment across multiple crystal structures. In dual-phase steels, for instance, EBSD distinguishes (face-centered cubic) from ferrite (body-centered cubic) by leveraging differences in pattern symmetry and intensity, enabling mapping of distributions at micrometer . Texture quantification in EBSD involves constructing orientation distribution functions (ODFs) from mapped , which reveal preferred crystallographic orientations and distinguish between fiber textures—such as <110>//rolling direction in rolled sheets—and random distributions indicative of isotropic materials. The quantifies texture strength through indices like the J coefficient, defined as the integral of the squared ODF over the Euler space, where values near 1 denote randomness and higher values signal strong . In applications like of ferritic stainless steels, EBSD-derived textures correlate with anisotropic mechanical properties, such as reduced r-values in γ-fiber dominated microstructures, which influence formability and earing behavior during forming. Advanced EBSD analysis extends to variant selection in martensitic transformations, where prior orientations are reconstructed from variants using the Kurdjumov-Sachs relationship, revealing deformation-induced preferences for specific habit planes and directions. Phase fractions are determined by measuring the area coverage of indexed pixels in EBSD maps, providing quantitative volume percentages accurate to within 1-2% for well-resolved . In Ni-based superalloys, EBSD combined with complementary techniques discriminates γ from γ' precipitates despite similar patterns, mapping their and interfaces in heat-treated samples. Recent studies on , such as CoCrFeMnNi, use EBSD to analyze stability and evolution during deformation, highlighting FCC-to-HCP transformations and their impact on . EBSD outputs for phase and texture include boundary statistics, such as total length and misorientation distributions at phase interfaces, which inform segregation and coherency effects. Texture indices like the J coefficient further enable comparison of processing routes, with machine learning enhancements improving sampling efficiency for large datasets in textured polycrystals.

Industrial and research uses

Electron backscatter diffraction (EBSD) plays a critical role in industrial quality control, particularly in aerospace manufacturing where it is employed to assess the microstructural integrity of turbine blades made from nickel-based superalloys like CMSX-4. By mapping grain orientations and boundaries, EBSD helps detect defects such as misoriented grains or recrystallization zones that could compromise high-temperature performance, enabling process optimization during cladding and repair operations. In the automotive sector, EBSD is utilized to evaluate weld integrity in high-strength steels and alloys, analyzing microstructure evolution across weld zones to identify issues like coarsening or variations that affect joint strength and . For instance, EBSD inspections reveal crystallographic changes in resistance spot welds, supporting improvements in weld parameter selection to enhance overall vehicle safety and durability. In geological , EBSD elucidates deformation mechanisms in rocks by quantifying recrystallization textures and misorientation distributions, providing insights into tectonic processes and histories in minerals such as or . This technique has been instrumental in mapping lattice rotations in deformed samples, revealing how dynamic recrystallization influences rock during faulting or folding events. Biological applications of EBSD focus on mineral phases in biominerals, such as the crystallographic organization in shells or bone , where it characterizes preferred orientations that contribute to mechanical resilience and pathways. Studies using EBSD have demonstrated how or layering in shells aligns with growth directions, informing evolutionary adaptations in natural armor structures. Integrated EBSD with () is widely applied in to investigate crack initiation and propagation, correlating crack paths with grain boundaries and phase distributions in alloys. For example, combined EBSD/EDS mapping has traced intergranular cracking in superalloys, highlighting how local misorientations accelerate under cyclic loading. In-situ EBSD during heating experiments monitors phase transformations in real time, such as the alpha-to-beta transition in , by tracking evolving grain structures and boundary mobility under . This approach has validated models of prior-austenite in steels, linking transformation to microstructural heritage. As of 2025, the global EBSD systems market is valued at approximately $777 million, driven by demand in advanced and sectors, with projections to reach $1.2 billion by 2035 due to expanded adoption in . In battery research, EBSD reveals how grain boundaries in lithium-ion cathode materials like Li(Ni,Mn,Co)O2 influence Li-ion diffusion pathways, showing that high-angle boundaries can impede transport and degrade cycle life. Such analyses guide design to minimize and enhance ionic . EBSD-derived distributions directly correlate with mechanical properties through the Hall-Petch relationship, where finer grains measured via EBSD increase yield strength in polycrystalline metals by impeding motion. This linkage has been confirmed in welds and deformed alloys, underscoring EBSD's value in predicting performance from microstructure.

References

  1. [1]
    Oxford Instruments - What is Electron Backscatter Diffraction (EBSD)?
    EBSD is a scanning electron microscope (SEM) based technique, which enables a sample's microstructure to be analysed, visualised and quantified.
  2. [2]
    What is EBSD? | Electron BackScatter Diffraction - Bruker
    Electron Backscatter Diffraction is a SEM-based technique that measures the interaction between the electron beam of the SEM with a sample's atomic structure.
  3. [3]
    How Does EBSD Work? | Electron Backscatter Diffraction
    Learn how EBSD works: electron backscatter diffraction is a complementary technique to scanning electron microscopy for the analysis of crystalline ...How Does Ebsd Work? · Ebsd Data Acquisition · What Is Electron Backscatter...
  4. [4]
    History and advancements of EBSD and BKD
    The Diffraction of Cathode Rays by Calcite. ... The first EBSP was published by S. Kikuchi and S. Nishikawa, his supervisor, in 1928, as a compliment to Kikuchi's ...
  5. [5]
    Strains, planes, and EBSD in materials science - ScienceDirect.com
    A brief history of EBSD. In 1954 Alam, Pashley and Nicholson recorded onto EM film patterns which we now call EBSD patterns (see Fig. 2a) and observed that ...
  6. [6]
    Electron Backscatter Diffraction (EBSD) - SERC (Carleton)
    What is EBSD? Accelerated electrons in the primary beam of a scanning electron microscope (SEM) can be diffracted by atomic layers in crystalline materials.
  7. [7]
    None
    ### Summary of EBSD Content from https://www.osti.gov/servlets/purl/964094
  8. [8]
    Review Electron backscatter Kikuchi diffraction in the scanning ...
    The technique of electron backscatter Kikuchi diffraction patterns (BKDPs) in the scanning electron microscope is reviewed. The paper focuses mainly on the.
  9. [9]
    [PDF] Electron Microscopy in Mineralogy - University of California, Berkeley
    ... de Geologie-CO No 140, Universite de. Nancy I, 54000 Nancy Cedex ... 0.9788 x 10-6 E)1/2 (A). (6) where h=Planck's constant, mo is rest mass and e ...
  10. [10]
    [PDF] The application of electron backscatter diffraction and orientation ...
    This paper aims to review (1) the principles behind EBSD and OC imaging; (2) analytical procedures appropriate to rock specimens; and (3) some applications to ...
  11. [11]
    Electron Backscatter Diffraction in Materials Science
    Jun 10, 2025 · Boersch, H., 1937, Z.Tech.Phys. 18:574. Booker, G.R., Shaw, A.M.B. ... Taylor, G.I., 1938, Plastic strain in metals, J. Inst. Met. 62 ...
  12. [12]
    Useful Electron Backscatter Diffraction (EBSD) References
    Below are some of the key papers that outline the major historical developments of the EBSD technique. ... D.G. Coates (1967). Kikuchi-like Reflection Patterns ...
  13. [13]
    Electron backscatter diffraction - EBSD
    The first on-line working (automated) EBSD systems were developed in 1980. In 1993 the Orientation Imaging Microscopy (OIM) or orientation mapping was ...
  14. [14]
  15. [15]
    AZtecCrystal: Meeting the challenges of an ever-evolving technique
    Many years ago, I was fortunate enough to be working for HKL Technology at the time when we developed and launched the Channel5 EBSD acquisition and data ...Missing: 1994 | Show results with:1994
  16. [16]
    The story behind Crosscourt & BLG Vantage - HR-EBSD.com
    In 2004 Prof David Dingley , Dr Angus Wilkinson and Dr Graham Meaden collaborated in order to try to measure Elastic Strains in semiconductors using EBSD.
  17. [17]
  18. [18]
    EBSD Geometry | EBSD Hints & Tips - Oxford Instruments
    Modern EBSD systems perform the geometry calibration automatically, but we still use a tilt angle of ~70° from horizontal for 2 reasons:
  19. [19]
    Electron backscatter diffraction and orientation imaging microscopy
    An overview of electron backscatter diffraction is presented in which experimental procedures are reviewed together with a basic theoretical description of ...
  20. [20]
    Experimental evidence concerning the significant information depth ...
    A statement that the information depth of EBSD ranges from 10 to 40 nm [2] is frequently cited in the literature. However, it is based on the assumption that ...
  21. [21]
    [PDF] Electron Backscatter Diffraction in Low Vacuum Conditions - OSTI
    Jul 22, 2008 · In this chapter we discuss the requirements, advantages and limitations of low vacuum EBSD (LV-EBSD), and ... environmental SEM (part 1). Scanning ...
  22. [22]
    Generalizable image segmentation for microstructure ... - Nature
    Oct 29, 2025 · Differences in tilt angles between SEM and EBSD lead to foreshortening of grains and varying interaction volumes, while differences in ...<|separator|>
  23. [23]
  24. [24]
  25. [25]
  26. [26]
    [PDF] Electron backscatter diffraction (EBSD) sample preparation revisited
    The present article intends to present to the reader a small survey of the most common EBSD sample preparation techniques and some of the newest advances in the ...
  27. [27]
    Advancements in EBSD Techniques: A Comprehensive Review on ...
    This comprehensive review focuses on the most recent advances in electron backscatter diffraction ... phosphor screen of the EBSD detector. Using a CCD ...
  28. [28]
    Polishing for EBSD Sample Preparation - Oxford Instruments
    The final polishing can be done using a mechanical or vibrational polishing step using colloidal silica, however there are also other methods which can be used ...
  29. [29]
    Preventing Charging/Coating for EBSD Sample Preparation
    Ideally the coat should be carbon either sputtered or evaporated onto the sample, but it is possible to use other coating materials such as gold or tungsten.
  30. [30]
    [PDF] Electron backscatter diffraction (EBSD) of nonconductive samples ...
    In conventional SEMs, EBSD of nonconductive materials can only be done after deposition of a conductive coating of,. e.g., a few nm of carbon or gold.
  31. [31]
    [PDF] EBSD Analyses On Metallic Fuel Samples - INL Digital Library
    Feb 27, 2024 · From this review it was evaluated that one limitation for achieving consistent EBSD analyses was related to sample preparation techniques. Thus, ...<|control11|><|separator|>
  32. [32]
    Developing a best practice for the final step of sample preparation of ...
    Sep 10, 2025 · Despite the wide use of EBSD analysis in AM, there is no universally accepted sample preparation protocol and the variety and sequence of the ...
  33. [33]
    3D EBSD Analysis - Oxford Instruments
    A solution is to use an SEM combined with a focused ion beam (FIB-SEM), where the ion column is used to mill away the surface between each EBSD map. This ...<|control11|><|separator|>
  34. [34]
    [PDF] The Use of Ion Milling for Surface Preparation for EBSD Analysis
    Jul 16, 2021 · polishing [5–8], electropolishing [9] or ion milling [10]. Plasma ... EBSD Sample Preparation: Techniques, Tips, and Tricks. Microsc ...
  35. [35]
    [PDF] Present State of Electron Backscatter Diffraction and Prospective ...
    Oct 24, 2008 · The geometry of a Kikuchi pattern can be interpreted as a gnomonic projection of the crystal lattice on the flat phosphor screen. The point of ...
  36. [36]
    [PDF] Introduction to EBSD
    May 2, 2013 · EBSD occurs due to Bragg diffraction. Bragg's law: nl = 2d sinq where n is the order of reflection, l is the electron wavelength, d is the ...
  37. [37]
    Introduction to Orientation Imaging Microscopy
    The diffracted beam produces a pattern composed of intersecting bands, termed electron backscatter patterns, or EBSPs. The patterns can be imaged by placing a ...
  38. [38]
    [PDF] Application-Note-EBSD.pdf - Struers.com
    As EBSD is a surface-sensitive technique, with the diffraction signal coming from the top few nanometres (5-50 nm) of the crystal lattice, it is es-.Missing: 1994 | Show results with:1994
  39. [39]
    Electron Backscatter Diffraction (EBSD) - Bowdoin College
    In geology, electron backscatter diffraction (EBSD) is a powerful tool for the observation and analysis of microstructures and for phase identification.
  40. [40]
    Optimising SEM Condition | EBSD Hints & Tips - Oxford Instruments
    In the tabs below you can read about how to set up and to optimise the SEM conditions in order to get the best EBSD results for a range of applications.
  41. [41]
    Many-beam dynamical simulation of electron backscatter diffraction ...
    Experiment. The experimental EBSPs were taken using a Nordlys II detector with CHANNEL 5 software (HKL technology) on a Carl Zeiss Supra 55VP FEGSEM. For EBSD ...
  42. [42]
    Bohmian trajectory-bloch wave approach to dynamical simulation of ...
    Nov 6, 2018 · Both methods result in intuitive and accurate electron trajectories for the simulation of the electron backscatter diffraction (EBSD) pattern.
  43. [43]
    [PDF] Automated Determination of Crystal Orientations from EBSPs
    of pixel values and correction for non-square pixels (400x400 pixels). As the next step of the preprocessing procedure, the resolution of the image is ...
  44. [44]
  45. [45]
    [PDF] Methods for Conducting Electron Backscattered Diffraction (EBSD ...
    The acceleration voltage was varied between 2 and 20 kV, beam current between 500 pA to a few nA and detector exposure time varied in the range of 20-500 ms.
  46. [46]
    None
    ### EBSD Mapping Procedures Summary
  47. [47]
    [PDF] Review Grain and subgrain characterisation by electron backscatter ...
    The application of automated Electron Backscatter Diffraction (EBSD) in the scanning ... well indexed (e.g. >95%) and deliberately de-indexing pixels in ...
  48. [48]
    (PDF) Orientation microscopy with fast EBSD - ResearchGate
    Aug 5, 2025 · With actually >1000 acquired patterns per second it is significantly faster than conventional electron backscattering diffraction. The data can ...
  49. [49]
    electron backscatter diffraction, EBSD | Glossary | JEOL Ltd.
    The specimen depth to generate the EBSD pattern ranges down to 30 to 50 nm from the specimen surface, though it depends on the accelerating voltage and the ...
  50. [50]
    Electron Backscatter Diffraction - ScopeM
    This measurement geometry results in an anisotropic interaction volume and subsequent resolution of about 20 nm by 50 nm [2]. Recently, a setup has been ...
  51. [51]
    Improved high-resolution EBSD analyses by correcting radial ...
    Standard EBSD usually provides results with an angular resolution of about 0.5°. Such an uncertainty is sufficient to evaluate local orientation maps but cannot ...
  52. [52]
    [PDF] High Resolution Mapping of Strains and Rotations using Electron ...
    Angular Resolution. The angular resolution of EBSD orientation measurements is often quoted as. 1° to 0.5°, however it is possible to achieve higher ...
  53. [53]
    Optimising the angular resolution of EBSD - ResearchGate
    Aug 5, 2025 · The spatial resolution of EBSD is around 40 nm (Chen et al., 2011), which can be greatly improved to reach below 10 nm utilizing transmission ...
  54. [54]
    Depth Sensitivity & Spatial Resolution of EBSD - Electron microscopy
    ... depth sensitivity) of the diffraction volume is better than 50 nm. ... [1] Hjelen J, Nes E: Spatial resolution measurements of electron backscatter diffraction ...
  55. [55]
    [PDF] Principles of depth-resolved Kikuchi pattern simulation for electron ...
    Oct 29, 2009 · Summary. This paper presents a tutorial discussion of the principles underlying the depth-dependent Kikuchi pattern formation of.
  56. [56]
    Orientation Effects on Indexing of Electron Backscatter Diffraction ...
    Aug 7, 2025 · Automated Electron Backscatter Diffraction (EBSD) has become a well-accepted technique for characterizing the crystallographic orientation ...Missing: Duncumb 1977
  57. [57]
    Compressive electron backscatter diffraction imaging - PMC - NIH
    Electron backscatter diffraction (EBSD) is a scanning electron microscopy (SEM) technique, providing important crystallographic information about the sample.Missing: Boersch 1938
  58. [58]
    Introduction and comparison of new EBSD post-processing ...
    However, some EBSD maps can be noisy, with a large fraction of non-indexed or mis-indexed points in the scan grid. Noisy maps generally arise when patterns ...Missing: extrapolation | Show results with:extrapolation
  59. [59]
    [PDF] EBSD Characterization of Tubular Cladding Plasticity after Bend ...
    The confidence index ranges from 0 to 1. The confidence index can be a bit misleading and a low CI value does not necessarily mean the EBSD data is not valuable ...Missing: limitations hit
  60. [60]
    [PDF] Application of Electron Backscatter Diffraction to Phase Identification
    Jul 22, 2008 · One reason is that band contrast depends on the phase orientation, since this dictates which Kikuchi bands are present in an EBSD pattern.
  61. [61]
    Large Area Stress Distribution in Crystalline Materials Calculated ...
    Aug 5, 2014 · We report a method to obtain the stress of crystalline materials directly from lattice deformation by Hooke's law.
  62. [62]
  63. [63]
  64. [64]
    An iterative method for reference pattern selection in high resolution ...
    Jun 21, 2022 · An iterative algorithm to identify the optimal reference pattern that maximises the precision of HR-EBSD.Missing: averaging | Show results with:averaging
  65. [65]
    Transmission Kikuchi Diffraction - Oxford Instruments
    The spatial resolution of conventional Electron Backscatter Diffraction (EBSD) is inherently limited by the pattern source volume to effective resolutions ...
  66. [66]
    Transmission Kikuchi Diffraction TKD - EBSD.info
    In practical work, the smallest probe size and hence spatial resolution is limited to about 10 nm with TKD, due to the decrease in image brightness as the probe ...
  67. [67]
    [PDF] Transmission EBSD from 10 nm domains in a scanning electron ...
    Oct 9, 2011 · A new form of electron backscatter diffraction (EBSD) has been demonstrated in the SEM. Transmission-EBSD presents two primary advantages over ...
  68. [68]
    Analytical Scanning Transmission Electron Microscopy in a ...
    Nov 27, 2012 · STEM-in-SEM uses transmission electron imaging in an SEM, enabling analysis of nanomaterials and soft matter, and measuring crystal phase and ...
  69. [69]
    STEM-in-SEM: A re-emerging material measurement approach - 2022
    Jun 1, 2022 · Scanning transmission electron microscopy performed in a scanning electron microscope (STEM-in-SEM) is a powerful method in and of itself ...
  70. [70]
    Transmission Kikuchi diffraction in a scanning electron microscope
    Electron backscatter diffraction (EBSD) is an orientation mapping technique that has typically been used to characterise materials with grain sizes in the order ...
  71. [71]
    Transmission Kikuchi diffraction: 10 years on - Oxford Instruments
    Feb 24, 2021 · Routinely it is possible to demonstrate a resolution of 5-10 nm, and this has been confirmed in numerous subsequent publications. This makes the ...
  72. [72]
    Transmission EBSD in the Scanning Electron Microscope
    May 1, 2013 · Lateral spatial resolution of less than 10 nm has been demonstrated for particles and better than 5 nm for orientation mapping of thin films.
  73. [73]
    Transmission EBSD from 10 nm domains in a scanning electron ...
    Aug 6, 2025 · We collected t-EBSD patterns from Fe-Co nanoparticles of diameter 10 nm and from 40 nm-thick Ni films with in-plane grain size 15 nm. The ...
  74. [74]
    Low-energy transmission electron diffraction and imaging of large ...
    Sep 1, 2017 · This work presents a simple and convenient method for characterizing the macroscopic structures of 2D materials, and the instrument we ...
  75. [75]
    Surface damage during ion beam thinning of samples for ...
    Aug 7, 2025 · The thinnest damaged layer formed on a silicon surface was measured as 1.5 nm thick, while an optimized FIB sample preparation process results ...
  76. [76]
    Thinning - FIB - MyScope
    The sample is thinned to <100nm using a larger current to thin to 600nm, then a smaller current to achieve electron transparency. The sample is tilted to ...
  77. [77]
    Indexing high-noise electron backscatter diffraction patterns using ...
    Jan 30, 2024 · This method uses a CNN with transfer learning for automatic, end-to-end crystal orientation indexing of EBSD patterns, even with high noise and ...
  78. [78]
    Deep Learning for automated phase segmentation in EBSD maps. A ...
    Deep Learning for automated phase segmentation in EBSD maps. A case study in ... The method allowed accurate classification of ferrite (94% accuracy) ...
  79. [79]
    Parametric simulation of electron backscatter diffraction patterns ...
    Oct 23, 2023 · In this study, we propose a framework (EBSD-CVAE/GAN) with great flexibility and scalability to realize parametric simulation of EBSD patterns.
  80. [80]
    [PDF] arXiv:2504.16951v1 [eess.IV] 22 Apr 2025
    Apr 22, 2025 · This paper presents a novel approach for denoising Elec- tron Backscatter Diffraction (EBSD) patterns using diffu- sion models.
  81. [81]
  82. [82]
    EBSD Detectors - Oxford Instruments
    The chart above shows how the analysis speeds of EBSD detectors have improved since the first commercial EBSD systems were developed at the start of the 1990s.Missing: ISI 1980s
  83. [83]
  84. [84]
    based large volume three-dimensional microstructure mapping ...
    Sep 29, 2022 · We report the development of a fully automatic large-volume 3D electron backscatter diffraction (EBSD) system (ELAVO 3D), consisting of a ...
  85. [85]
    EBSD Integration with EDS - Oxford Instruments
    Discover how and why electron backscatter diffraction (EBSD) can be integrated with energy dispersive X-ray spectrometry (EDS) in the SEM.
  86. [86]
    Accelerating Multi-Modal Materials Characterization with Plasma FIB ...
    Nov 16, 2020 · We use both EDS and EBSD to do this, with EDS revealing the ... composition and EBSD revealing the grains based on their orientation.
  87. [87]
    Surface patterning allowing combined EBSD and DIC investigations ...
    Feb 1, 2022 · A demonstration based on in-situ tensile deformation of aluminum reveals significant heterogeneities in several EBSD- and DIC-derived parameters ...
  88. [88]
    3D electron backscatter diffraction characterization of fine α titanium ...
    3D-EBSD reconstruction. The reconstruction is relatively complex, involving several steps including importing the raw data, thresholding, alignment, cropping ...
  89. [89]
    Tomographic orientation microscopy ( 3D EBSD ) on steels
    We present our recent progress in 3D high resolution characterization of crystalline microstructures. Our approach uses automated serial sectioning.Missing: iterative sparsity
  90. [90]
    Strain Measurement of 3D Structured Nanodevices by EBSD - PMC
    Aug 31, 2017 · We present a new methodology to accurately measure strain magnitudes from 3D nanodevices using Electron Backscatter Diffraction (EBSD).
  91. [91]
    Comprehensive 3D Characterization Workflow for Solid Oxide Cells
    Comprehensive 3D Characterization Workflow for Solid Oxide Cells. October 2024 ... 3D tomography, 3D EDX, 3D EBSD as well as correlative tomography. View.
  92. [92]
    Q-RBSA: high-resolution 3D EBSD map generation using ... - Nature
    Jan 30, 2024 · Our framework uses a quaternion residual block self-attention network (QRBSA) to generate high-resolution 3D EBSD maps from sparsely sectioned EBSD maps.Missing: iterative incomplete
  93. [93]
    On the alignment of 3D EBSD data collected by serial sectioning ...
    Aug 7, 2025 · In order to address the inherent limitations of two-dimensional imaging and enhance statistical robustness, serial sectioning-assisted EBSD can ...Missing: intensive | Show results with:intensive
  94. [94]
    Application of Electron Backscatter Diffraction to Grain Boundaries
    Aug 9, 2025 · This review documents the application of electron backscatter diffraction (EBSD) in a scanning electron microscope to studies of grain ...
  95. [95]
    Machine learning enhanced analysis of EBSD data for texture ...
    Jun 26, 2024 · This study introduces a machine-learning-based approach, termed texture adaptive clustering and sampling (TACS), for generating representative Euler angle ...Performance Analysis · Methods · Kernel Density Estimation...
  96. [96]
    [PDF] High-Resolution EBSD Investigation of Deformed and Partially ...
    Abstract: The recrystallization of titanium-alloyed intersti- tial-free steel (IF steel) has been investigated by high-resolu- tion electron backscattered ...Missing: seminal paper
  97. [97]
    [PDF] Experimental Studies of Deformation Structures in Stainless Steels ...
    Electron backscatter diffraction (EBSD) has been used to study the deformation structures, both qualitatively and quantitatively. EBSD is a scanning electron ...
  98. [98]
    EBSD characterisation of grain size distribution and grain sub ...
    Jan 4, 2022 · This paper implements EBSD-based linear intercept measurement of spatial grain size variation for ferritic steel weld metals, making analysis flexible and ...
  99. [99]
    Revealing relationships between porosity, microstructure and ...
    The understanding of relationships between processing, microstructure and mechanical properties in laser powder bed fusion is currently incomplete.
  100. [100]
    Qualities of electron backscatter diffraction patterns and image ...
    The acceleration voltage and the probe current of the electron beam was 20 kV and 9.8 nA respectively, while the specimen was tilted by 70°. The map covered ...
  101. [101]
    [PDF] An EBSD study on microstructure and texture ... - DiVA portal
    Dec 23, 2024 · The texture index, often referred to as the “J-index”, quan- tifies the degree of texture in a polycrystalline sample. A high texture index ...
  102. [102]
    Microstructure, Texture, and Anisotropic Properties of High-Strength ...
    Aug 16, 2023 · The EBSD characterization of ferritic stainless steel showed that γ-fiber {111} texture formed after rolling and annealing, which was more ...
  103. [103]
    Determining phase volume fraction in steels by electron ...
    EBSD pattern quality (PQ) maps are shown to improve imaging of martensite and ferrite over optical and secondary electron imaging. Ferrite volume fraction in ...
  104. [104]
    EBSD Study of Delta-Processed Ni-Based Superalloy - MDPI
    Nov 3, 2020 · In this study, an Inconel 718 (IN718) Ni-based superalloy was subjected to a delta-processing treatment (DP718) and subsequently deformed at high temperature.
  105. [105]
    In-situ EBSD investigation of excellent strength-ductility synergy ...
    Oct 19, 2025 · The development of high-entropy alloys (HEAs) is plagued by a trade-off between strength and ductility. In this study, we investigated the ...
  106. [106]
    EBSD analysis of CMSX-4 clad - ResearchGate
    The aim of the subproject is the restoration of defect, single-crystal high-pressure turbine blades. This objective was achieved by developing a novel two-stage ...
  107. [107]
    Automotive & Aerospace | EBSD Applications - Oxford Instruments
    EBSD is used for rapid texture and grain size analyses, characterizing microstructures of lightweight alloys, phase analyses of high strength steels, and ...Missing: control | Show results with:control
  108. [108]
    Electron Backscatter Diffraction Inspection | NDE - AWS CWI
    Aug 10, 2024 · Electron Backscatter Diffraction (EBSD) is a powerful technique used in weld inspection to characterize the microstructure of welds.
  109. [109]
    Geology, Petrology & Mining | EBSD Applications - Oxford Instruments
    Learn how Electron Backscatter Diffraction (EBSD) is being used in geology and the planetary sciences, and for ore characterisation in the mining industry.
  110. [110]
    EBSD in the Earth Sciences: Applications, Common Practice, and ...
    1. EBSD data can be collected from most minerals. · 2. Step sizes as small as 50 nm give 50 nm resolutions · 3. Orientation mapping using automated EBSD is pos-.
  111. [111]
    The application of electron backscatter diffraction and orientation ...
    Mar 2, 2017 · EBSD and OC imaging are SEM-based crystallographic tools. EBSD enables measurement of the crystallographic orientation of individual rock-forming minerals.Missing: key milestones<|control11|><|separator|>
  112. [112]
  113. [113]
    Quantifying microscale drivers for fatigue failure via coupled ... - Nature
    Jun 24, 2020 · They are, however, known to form fatigue cracks near coherent twin boundaries (low energy configuration grain boundaries separating twin and ...
  114. [114]
    In-situ EBSD - Oxford Instruments
    In-situ EBSD is not always simple: there are likely space limitations in the SEM chamber, restrictions on the sample size and shape and numerous other ...Missing: artifacts | Show results with:artifacts<|control11|><|separator|>
  115. [115]
    In-situ heating-stage EBSD validation of algorithms for prior ...
    Mar 15, 2024 · The use of algorithms that reconstruct PAGs from electron backscatter diffraction (EBSD) measurements of martensitic or bainitic microstructures have become ...
  116. [116]
    Electron Backscatter Diffraction (Ebsd) System Market: Trends ...
    The Electron Backscatter Diffraction (EBSD) System Market is expected to grow from 776.9 USD Million in 2025 to 1,200 USD Million by 2035. The Electron ...
  117. [117]
    Electron Backscatter Diffraction for Investigating Lithium-Ion ...
    Electron backscatter diffraction (EBSD) is a promising technique for acquiring grain orientations and boundaries. EBSD spatially resolves the orientation of ...
  118. [118]
    Application of electron backscatter diffraction techniques to quantify ...
    In this study, electron backscatter diffraction (EBSD) is employed to track sub-grain lattice structure across a statistically relevant number of Li(Ni 0.33 Mn ...
  119. [119]
    Correlation between local mechanical properties and corresponding ...
    The microstructural examinations support the notion that the strength of the SZ is correlated with the grain size, in line with the Hall-Petch equation that ...Original Article · 3. Results And Discussion · 3.1. Ebsd Analysis