Fact-checked by Grok 2 weeks ago

Wide-angle X-ray scattering

Wide-angle X-ray scattering (WAXS), also referred to as wide-angle diffraction (WAXD), is an analytical technique that investigates the atomic and molecular structure of materials by measuring the of at wide angles, typically ranging from 4° to 60° (2θ), which corresponds to interplanar spacings on the sub-nanometer scale (approximately 0.5–10 Å). The method relies on the interaction of monochromatic with electrons in the sample, producing patterns governed by (n\lambda = 2d \sin\theta), where the Bragg angle θ reveals the distance d between atomic planes, enabling quantitative analysis of crystal lattice parameters. In practice, WAXS patterns are recorded using detectors positioned close to the sample (e.g., 20–300 mm), capturing high-angle Bragg peaks that provide insights into crystallinity, phase identification, crystallite size (via the ), and molecular orientation (via Herman's orientation function). This distinguishes WAXS from (SAXS), which uses longer sample-to-detector distances and smaller angles (up to ~1–10°) to probe larger structural features in the nanometer range (10–100 nm), such as domain sizes or particle distributions. WAXS finds extensive applications in for characterizing crystalline polymers, nanocomposites, and thin films—such as determining polymorphism in polyesters or clay dispersion in composites—as well as in for analyzing protein conformations like those in or . In biopolymer research, it resolves hierarchical structures in materials like , silk , and at atomic resolution, supporting studies of scaffolds and nanofibers. Advances in , including high-brilliance microfocus sources (e.g., Cu Kα radiation at λ = 0.154 nm), have enhanced accessibility and enabled observations of phase transitions without relying on facilities.

Fundamentals

Definition and principles

Wide-angle X-ray scattering (WAXS) is a structural characterization technique that involves the of X-rays by matter at scattering angles 2θ typically greater than 5°, enabling the probing of and molecular arrangements on length scales of 0.1–2 . This method is particularly suited for investigating both crystalline and amorphous materials, where the scattered intensity provides information about short-range order and local environments. The fundamental principle of WAXS relies on the interaction of X-rays with the electrons in a sample, leading to coherent that generates patterns. These patterns arise from the differences in scattered , which constructively or destructively interfere depending on the distribution within the material, thereby revealing the underlying structural motifs such as spacings in crystals or pair correlations in disordered systems. Unlike (SAXS), which focuses on larger-scale features, WAXS emphasizes higher-angle to access finer details of atomic-scale organization. In WAXS, measurements are conducted in reciprocal space, where wide scattering angles correspond to high values of the momentum transfer, typically q > 0.1 ⁻¹, allowing resolution of sub-nanometer features. The scattering vector is defined as the difference between the incident and scattered wavevectors, with its magnitude given by q = \frac{4\pi}{\lambda} \sin\theta where \lambda is the wavelength of the X-rays and \theta is half the scattering angle (2θ). This formulation maps the angular scattering data directly to real-space structural parameters via the relation d \approx 2\pi / q, facilitating the interpretation of atomic arrangements.

Relation to other scattering techniques

Wide-angle X-ray scattering (WAXS) is closely related to small-angle X-ray scattering (SAXS), as both techniques utilize X-ray scattering to probe material structures, but they differ fundamentally in the length scales examined and the scattering angles involved. WAXS focuses on atomic and molecular order, capturing scattering at wide angles (typically 5° to 60° 2θ), which corresponds to small interatomic d-spacings on the order of 0.1–1 nm and scattering vectors q up to ~50 nm⁻¹. In contrast, SAXS investigates larger nanoscale domains, such as particles or pores with d-spacings greater than 10 nm, using very small angles (up to 1° 2θ) and low q values (<0.1 nm⁻¹). This distinction allows WAXS to reveal crystalline phases, lattice parameters, and molecular arrangements, while SAXS provides information on overall morphology and aggregation. WAXS shares significant overlap with powder X-ray diffraction (XRD), often considered a form of wide-angle XRD applied to polycrystalline or amorphous samples rather than perfect single crystals. While traditional powder XRD emphasizes sharp Bragg diffraction peaks to index crystal structures in finely ground powders, WAXS adopts a broader scattering perspective, accommodating diffuse patterns from partially ordered or disordered materials like polymers and glasses, without requiring strict peak resolution. This makes WAXS particularly suited for in situ studies of non-ideal samples, where it interprets both Bragg reflections and amorphous halos to assess crystallinity and phase composition. WAXS complements other diffraction techniques such as electron diffraction and neutron scattering, each offering unique advantages based on probe-sample interactions. Electron diffraction, which uses high-energy electrons for atomic-resolution imaging, is limited to thin samples (typically <100 nm thick) due to strong electron-matter interactions and shallow penetration depth, making it ideal for nanoscale surfaces or nanocrystals but less practical for bulk analysis. In comparison, WAXS benefits from X-rays' deeper penetration (up to millimeters in soft matter), enabling non-destructive probing of bulk samples without sectioning. Neutron scattering provides orthogonal sensitivity to light elements like hydrogen and isotopes, with even greater penetration than X-rays for bulky samples, but requires specialized facilities; WAXS thus offers a more accessible alternative for routine bulk characterization where electron density contrast dominates. In modern synchrotron facilities, combined WAXS/SAXS setups are increasingly standard, allowing simultaneous acquisition across multiple length scales in a single experiment to capture hierarchical structures, such as in polymers or biomaterials. These integrated systems, like those at the or beamlines, use multiple detectors to bridge q-ranges from 0.01 nm⁻¹ (SAXS) to over 10 nm⁻¹ (WAXS), facilitating time-resolved studies of dynamic processes without sample repositioning.

Theoretical foundations

Bragg diffraction

Bragg diffraction arises from the constructive interference of X-rays scattered by the periodic arrangement of atoms in a crystal lattice, where the scattered waves reinforce each other only when the path length difference between waves reflected from adjacent lattice planes satisfies specific geometric conditions. This phenomenon, first described by in 1913, models the crystal planes as partially reflecting mirrors, leading to enhanced intensity at particular scattering angles. The condition for constructive interference is encapsulated in : n\lambda = 2d \sin\theta where n is a positive integer representing the diffraction order, \lambda is the wavelength of the incident X-rays, d is the spacing between the lattice planes, and \theta is the angle between the incident beam and the planes (known as the ). The derivation considers two parallel rays incident on successive planes separated by d; the extra path length traveled by the second ray is $2d \sin\theta, which must equal an integer multiple of \lambda for the waves to be in phase and interfere constructively. This law provides a direct relationship between measurable diffraction angles and the underlying atomic-scale structure of the crystal. In polycrystalline samples, where crystallites are randomly oriented, Bragg diffraction produces characteristic Debye-Scherrer rings on a detector, as each orientation contributes to a conical diffraction envelope that intersects the detection plane in circular patterns. This method, developed by Peter Debye and Paul Scherrer in 1916, enables the analysis of powder or finely ground samples without requiring single-crystal alignment, with ring radii corresponding to the $2\theta angles satisfying Bragg's law for various hkl planes. The rings' positions allow determination of lattice parameters, while their intensities reflect the multiplicity of equivalent planes and atomic scattering factors. In wide-angle X-ray scattering (WAXS), Bragg diffraction at large scattering angles (typically $5^\circ < 2\theta < 50^\circ) is essential for resolving small interplanar spacings d on the order of 1–5 Å, which correspond to atomic bond lengths and short-range structural features in materials. According to , higher \theta values probe smaller d, making WAXS particularly suited for studying crystalline phases in metals, semiconductors, and polymers where atomic-scale order dominates. However, for amorphous materials lacking long-range periodicity, the diffraction condition cannot be strictly met, resulting in broad, diffuse halos rather than sharp peaks; these halos arise from short-range order and pair correlations, with peak positions indicating average intermolecular distances but without the discrete d-spacing resolution of crystalline .

Intensity and structure factors

In wide-angle X-ray scattering (WAXS), the structure factor F_{hkl} quantifies the amplitude and phase of the scattered X-ray wave from a set of crystal planes characterized by Miller indices (hkl). It arises from the coherent interference of waves scattered by electrons around individual atoms in the unit cell, providing a direct link between the observed scattering pattern and the atomic arrangement. The structure factor is given by F_{hkl} = \sum_j f_j \exp \left[ 2\pi i (h x_j + k y_j + l z_j) \right], where the sum is over all atoms j in the unit cell, f_j is the atomic scattering factor (or form factor) for atom j, and (x_j, y_j, z_j) are the fractional coordinates of that atom relative to the unit cell axes. The scattered intensity I_{hkl} for a reflection from planes (hkl) is proportional to the square of the structure factor magnitude, I_{hkl} \propto |F_{hkl}|^2, which encodes information about atomic positions and types through constructive and destructive interference. In single-crystal diffraction, this relation holds directly under the conditions of Bragg's law, but for powder samples common in WAXS, additional geometric factors must be accounted for due to the random orientation of crystallites. Specifically, the observed intensity includes a multiplicity factor p_{hkl}, which counts the number of equivalent planes contributing to the reflection (e.g., p_{100} = 6 for a cubic lattice), and the Lorentz-polarization factor L_p = \frac{1 + \cos^2 2\theta}{2 \sin \theta \cos \theta} for unpolarized incident radiation, yielding I_{hkl} \propto p_{hkl} |F_{hkl}|^2 L_p. These corrections arise from the statistical averaging over all crystallite orientations and the angular dependence of X-ray polarization during scattering. The atomic scattering factor f_j represents the scattering efficiency of atom j and depends on the scattering vector magnitude q = \frac{4\pi}{\lambda} \sin \theta, where \lambda is the X-ray wavelength and \theta is half the scattering angle. At low q (small angles), f_j approaches the atomic number Z_j as all electrons scatter in phase, but it decreases at higher q (wide angles) because the finite size and distribution of the electron cloud around the nucleus cause phase differences among the scattered waves from individual electrons. This q-dependence is typically parameterized using Gaussian or tabulated functions, such as those based on , leading to a falloff that is more pronounced for lighter atoms with more diffuse electron densities. In WAXS, where wide angles probe atomic-scale structures, this attenuation must be corrected to accurately interpret peak intensities. Thermal motion of atoms further modulates the structure factor and observed intensities through the Debye-Waller factor, which accounts for the time-averaged displacement of atoms from their equilibrium positions due to vibrational energy. This effect broadens diffraction peaks and reduces their intensity, particularly at higher angles, via the multiplicative factor \exp \left( -2B \frac{\sin^2 \theta}{\lambda^2} \right), where B = 8\pi^2 \langle u^2 \rangle and \langle u^2 \rangle is the mean-square atomic displacement. The parameter B increases with temperature and decreases with atomic mass, reflecting stronger vibrations in lighter atoms or at elevated temperatures; for example, typical B values range from 0.5 to 3 Ų for metals at room temperature. In the full intensity expression, the Debye-Waller factor modifies |F_{hkl}|^2 as |F_{hkl}|^2 \exp(-2B \sin^2 \theta / \lambda^2), enabling extraction of thermal parameters from WAXS data when combined with known structures.

Experimental methods

Instrumentation and setup

Wide-angle X-ray scattering (WAXS) experiments require specialized instrumentation to generate, collimate, and detect X-rays scattered at angles typically ranging from 5° to 120° (2θ). Laboratory-based setups commonly employ sealed X-ray tubes or rotating anode generators, while synchrotron sources provide higher flux and tunability for advanced studies. In laboratory environments, X-ray sources often utilize copper (Cu) Kα radiation with a wavelength of approximately 1.54 Å (energy ~8.04 keV), generated by microfocus rotating anode systems or sealed tubes to achieve sufficient intensity for crystalline structure analysis. Rotating anodes, such as those in the , enhance flux up to 5.7 × 10^9 photons s⁻¹ through continuous rotation, minimizing anode damage and enabling longer exposures compared to fixed-target tubes. Alternative lab sources include liquid metal jets, like gallium-based MetalJet systems operating at 9.24 keV, which provide stable, high-brightness beams (~3.7 × 10^6 photons s⁻¹) suitable for combined small- and wide-angle measurements. Recent advancements include the , offering high-speed and high-resolution analysis for polymers as of 2024. Synchrotron facilities, in contrast, offer tunable energies (e.g., 10–12.9 keV) from undulators or bending magnets, delivering fluxes orders of magnitude higher (up to 10^12 photons s⁻¹) with low divergence, ideal for time-resolved WAXS on dynamic samples. Optics and goniometers ensure beam purity and precise angular control in WAXS setups. Monochromators, such as multilayer confocal mirrors or Si(111) crystals, filter the primary beam to select specific wavelengths and reduce background, while adjustable slits or pinholes collimate the beam to sizes as small as 140 × 140 µm for high resolution. Multi-functional goniometer platforms, like the ScatterX 78 system, allow rapid switching between configurations via exchangeable modules, supporting 2θ scanning for point-collimated measurements or fixed orientations for area detection. In synchrotron beamlines, such as ESRF's BM01, additional optics like elliptically bent mirrors and Soller slits further minimize divergence, enabling grazing-incidence geometries with incidence angles below 1°. Detection in WAXS relies on area detectors to capture 2D scattering patterns, with hybrid pixel detectors like (1M or 100k variants) or providing high dynamic range (~10^6:1) and low noise (<6 counts per second per pixel) for wide q-ranges (0.06–51.4 nm⁻¹). Newer detectors, such as the , extend capabilities for high-energy applications as of 2025. These detectors, often positioned 100–500 mm from the sample, support fast acquisitions (<0.1 s) essential for in situ experiments. To mitigate air scattering, which intensifies at wide angles and obscures high-q signals, setups incorporate evacuated or helium-filled beam paths. Laboratory systems use vacuum chambers (<0.1 mbar) with thin windows (e.g., 6 µm PET or 75 µm Kapton) to enclose the flight tube, reducing parasitic scattering by over 90%. Synchrotron lines similarly employ helium paths or guarding slits with beamstops to suppress background. Typical WAXS configurations include transmission geometry for bulk samples like powders or liquids, where the beam passes through the specimen to an area detector, facilitating isotropic pattern collection. Reflection geometry, often as grazing-incidence WAXS (GIWAXS), suits thin films or surfaces, with the beam at shallow angles (~0.15°–1°) to enhance penetration while probing orientation, as in CVD chambers or polymer substrates. These geometries are selected based on sample morphology, with transmission preferred for volumetric averaging and reflection for interface sensitivity.

Sample requirements and preparation

Wide-angle X-ray scattering (WAXS) experiments require samples that can produce sufficient scattering intensity at scattering angles typically ranging from 5° to 120°, making it particularly suitable for crystalline materials where sharp Bragg peaks are observed, though it is also applicable to semi-crystalline and amorphous systems through diffuse scattering patterns. Common sample types include powders, which provide isotropic scattering for bulk structural analysis; thin films, often used in grazing-incidence configurations to probe surface and interface structures; and bulk solids, which may be sectioned for uniform exposure. For biological or soft matter applications, such as atherosclerotic plaques, thin sections (3–100 μm thick) embedded in resins or fixed tissues are employed to maintain structural integrity. Preparation techniques emphasize achieving random orientation and minimal artifacts to ensure reliable data. For powders, gentle grinding in an agate mortar with a solvent like acetone promotes particle uniformity and randomness, preventing preferred orientation that could distort peak intensities. Samples are then mounted in quartz capillaries (for liquids or fine powders, typically 1–2 mm diameter) or specialized holders with thin polymer foils to contain the material without introducing strain. Thin films are deposited via methods like spin-coating on flat substrates (e.g., silicon wafers or glass) and may require edge cleaving to eliminate diffraction artifacts from substrate boundaries. Bulk solids or fibers are clamped in holders or adhered to low-absorbing sheets (e.g., beryllium or Kapton tape) to facilitate alignment in the beam path. In all cases, handling minimizes mechanical stress, and for temperature-sensitive samples, controlled environments (e.g., 5–70°C with 0.1°C stability) are used during mounting. At wide angles, specific challenges arise due to the geometry and interaction of s with matter. Thick samples (>50 μm) necessitate absorption corrections, as varying depths can lead to uneven profiles, particularly in high-density materials; these are calculated based on sample and thickness to normalize data. Organic and biological samples are prone to from high-flux beams, manifesting as structural degradation (e.g., in perovskite films after prolonged exposure); mitigation involves selecting energies below absorption edges (e.g., 10–12.9 keV for lead-containing materials) and using fast or purging with inert gases. Sample quantities depend on the source brightness and detection sensitivity, typically ranging from micrograms to milligrams. For sources, milligram-scale powders provide adequate from ~10^20 electrons, while setups enable microgram amounts or microliter volumes (40–50 μL) in capillaries for weakly solutions. Thin films require only nanoscale thicknesses (100–500 nm) over areas of a few square millimeters, minimizing material use while achieving high signal-to-noise ratios.

Data analysis

Pattern interpretation

Pattern interpretation in wide-angle X-ray scattering (WAXS) involves the qualitative examination of patterns to identify crystalline phases, assess structural features, and recognize potential distortions or artifacts. The resulting patterns, typically plotted as intensity versus scattering angle (2θ) or scattering vector (q), reveal characteristic features such as sharp Bragg peaks for crystalline components and broad diffuse for amorphous regions. This initial reading guides subsequent quantitative refinement by providing insights into material and microstructure. Peak identification begins with measuring the positions of diffraction maxima, which correspond to interplanar spacings (d-spacings) via : n\lambda = 2d \sin\theta, where n is the reflection order, \lambda is the , and \theta is the Bragg angle. These d-spacings are indexed to Miller indices (hkl) by comparing observed positions to known structures, enabling phase assignment. Software tools like GSAS-II facilitate this by automating detection, indexing, and matching against such as the Centre for Data (ICDD) Powder File (PDF), which contains over 1 million entries for inorganic and organic phases relevant to WAXS . Phase analysis relies on distinguishing sharp, discrete crystalline peaks from broad amorphous humps in the pattern; crystalline materials produce well-defined Bragg reflections due to long-range order, while amorphous components yield diffuse indicative of short-range correlations. For quantification, the Rietveld method provides an overview by fitting the entire pattern to structural models, estimating phase fractions without internal standards through refinement of scale factors and profile parameters, as originally developed for nuclear structure analysis and extended to multiphase mixtures. Peak intensities in these analyses are influenced by structure factors, which account for atomic contributions. Texture and strain effects manifest as deviations from ideal powder-averaged patterns. Preferred orientation, or , arises from aligned crystallites and leads to enhanced or suppressed peak intensities, particularly in processed materials like fibers or thin films. Strain, including lattice distortions from defects or , causes peak broadening (via microstrain contributions) or shifts in position, with broadening proportional to the angle and distinguishable from size effects through angular dependence. Common artifacts in WAXS patterns of complex mixtures include overlapping peaks, where contributions from multiple phases superimpose, complicating individual identification and requiring careful or higher-resolution data. Such overlaps are prevalent in multiphase samples like alloys or composites, potentially mimicking single-phase features if not resolved.

Quantitative analysis techniques

Quantitative analysis techniques in wide-angle X-ray scattering (WAXS) enable the extraction of precise structural parameters, such as constants, fractions, and atomic coordinates, from patterns by employing advanced computational modeling and fitting methods. These techniques go beyond qualitative interpretation by incorporating least-squares optimization and transforms to quantify both crystalline and amorphous components in materials. Rietveld refinement is a cornerstone method for analyzing crystalline structures in WAXS data, involving the least-squares fitting of the entire diffraction pattern to a calculated profile based on a structural model. The objective is to minimize the chi-squared statistic, defined as \chi^2 = \sum_i w_i (I_{\text{obs},i} - I_{\text{calc},i})^2, where I_{\text{obs},i} and I_{\text{calc},i} are the observed and calculated intensities at data point i, and w_i are weighting factors typically proportional to $1/\sigma_i^2 with \sigma_i being the uncertainty in I_{\text{obs},i}. This approach refines parameters including scale factors, lattice parameters, atomic positions, thermal displacements, and microstructural effects like preferred orientation, providing quantitative phase abundances with uncertainties often below 1 wt%. Originally developed for neutron diffraction, it has been widely adapted for X-ray powder diffraction, including WAXS, due to its ability to handle overlapping peaks and instrumental broadening. A key challenge in Rietveld refinement arises from parameter correlations, where variables such as lattice parameters and thermal factors may covary, leading to inflated uncertainties or unstable convergence if not constrained. To mitigate this, refinements often proceed sequentially—starting with scale and background, then profile parameters— and may incorporate restraints or combined datasets from multiple techniques. Software implementations like FullProf and TOPAS facilitate these processes; FullProf, a free suite for Rietveld and pattern matching, supports sequential and global refinements for multi-phase systems, while TOPAS employs fundamental parameters modeling for accurate peak shapes across wide angular ranges in WAXS. Both tools output refined parameters with error estimates, but users must validate fits using goodness-of-fit metrics like \chi^2 < 2 and profile R-factors below 10%. For amorphous or poorly crystalline materials, pair distribution function (PDF) analysis extracts local atomic structure from WAXS total scattering data via a Fourier transform of the reduced structure factor. The PDF, G(r), is given by G(r) = 4\pi r [\rho(r) - \rho_0], where \rho(r) is the local atomic density at distance r from a reference atom, and \rho_0 is the average number density. This yields bond lengths and coordination numbers up to 10–20 Å, revealing short-range order invisible in Bragg diffraction alone. PDF is particularly suited to WAXS due to its sensitivity to wide momentum transfers (Q up to 30 Å^{-1}), enabling resolution of atomic pairs in disordered systems like glasses or nanoparticles. Crystallite size and microstrain quantification in WAXS often employs the Scherrer equation, which relates peak broadening to domain size via D = \frac{K\lambda}{\beta \cos\theta}, where D is the average crystallite size, K is a shape factor (typically 0.9 for spherical particles), \lambda is the X-ray wavelength, \beta is the full width at half maximum (FWHM) in radians after instrumental correction, and \theta is the Bragg angle. This method assumes broadening is dominated by finite size effects, providing sizes from 3 nm to 100 nm; microstrain (\epsilon) is incorporated by analyzing \beta \cos\theta versus $4\sin\theta plots (Williamson-Hall method), where the y-intercept yields strain values on the order of 0.1–1%. Applications in WAXS confirm sizes in nanomaterials, with errors reduced by multi-peak averaging.

Applications

In materials science

Wide-angle X-ray scattering (WAXS) plays a pivotal role in materials science by enabling the detailed characterization of crystalline structures in inorganic and metallic materials, focusing on phase composition, defects, and microstructural evolution. This technique leverages the diffraction of X-rays at wide angles (typically 5° to 120° 2θ) to produce patterns that reveal atomic arrangements, allowing researchers to identify phases, quantify impurities, and assess defects without destructive sampling. In alloys and ceramics, WAXS is particularly valuable for detecting subtle phase transformations and minor constituents that influence mechanical, thermal, and electrical properties. Phase identification in alloys and ceramics is a primary application of WAXS, where diffraction peaks are compared against reference databases to pinpoint crystal structures and detect impurities or phase changes. For instance, in ferritic-martensitic steels like Grade 91, in situ WAXS has been used to monitor microstructural evolution during tensile deformation, revealing changes in dislocation density and lattice strains within the martensitic laths that contribute to strengthening mechanisms, while identifying defects such as dislocations. In ceramics, such as silica-zirconia composites, WAXS distinguishes between amorphous and crystalline phases, aiding in the detection of zirconia polymorphs (e.g., tetragonal vs. monoclinic) that affect toughness and stability during sintering or thermal cycling. These analyses often integrate quantitative Rietveld refinement to determine phase fractions with high precision, essential for optimizing material performance in high-temperature environments. Texture analysis in metals utilizes WAXS to generate pole figures, which map the preferred crystallographic orientations arising from processing like rolling or forging, directly linking texture to anisotropic properties such as ductility and fatigue resistance. By collecting diffraction data at varying sample orientations, inverse pole figures can quantify texture strength. This approach is crucial for deformation studies, where strong textures (e.g., <110> fiber in body-centered cubic metals) are evaluated to predict formability and failure modes. For nanocrystalline materials, WAXS excels in analyzing peak broadening to estimate grain sizes in nanoparticles via the , D = \frac{K \lambda}{\beta \cos \theta}, where D is the average crystallite size, K is the (typically 0.9), \lambda is the , \beta is the , and \theta is the Bragg angle. This method has been applied to metal oxide nanoparticles, such as or ceria, yielding sizes from 2–50 and highlighting how nanoscale grains enhance catalytic activity or hardness while introducing defect-related broadening beyond pure size effects. Accurate interpretation requires deconvoluting instrumental and contributions to ensure reliable size distributions. In-situ WAXS studies facilitate real-time observation of phase changes in catalysts, capturing dynamic transformations that govern activity and deactivation. For oxide-derived copper catalysts under CO₂ reduction, synchrotron-based in-situ WAXS tracked the reduction from Cu₂O to metallic Cu phases, correlating structural changes with active site formation and morphological evolution leading to deactivation. Such experiments, often combined with environmental cells, provide kinetic insights into phase stability, enabling the design of robust catalysts for energy applications.

In polymer and biological systems

Wide-angle X-ray scattering (WAXS) is extensively applied to characterize the semi-crystalline nature of , where it quantifies the degree of crystallinity by analyzing patterns from crystalline and amorphous regions. The degree of crystallinity, denoted as X_c, is calculated as X_c = \frac{I_c}{I_c + I_a} \times 100\%, where I_c and I_a represent the integrated intensities of the crystalline and amorphous peaks, respectively, often derived from peak areas after background and for instrumental factors. This method, refined by Ruland to incorporate disorder corrections, enables precise assessment of how processing conditions like or shear influence morphology, as demonstrated in studies of where shear-induced increased X_c from approximately 40% to over 60%. In biological systems, WAXS excels in fiber diffraction studies of biomolecular structures, revealing atomic-scale features in oriented samples such as proteins and DNA. For instance, the meridional reflection at 5.1 Å in alpha-helical proteins like keratin confirms the helical pitch, providing evidence for secondary structure motifs that underpin protein function and assembly. Similarly, in DNA fibers, WAXS patterns display layer lines corresponding to the 3.4 Å base-pair spacing in B-form DNA, allowing differentiation between conformational states like A- and B-forms during hydration changes, as observed in nonoriented fiber melting transitions. These insights have been pivotal since early fiber diffraction work, offering a non-destructive probe for hierarchical ordering in natural biopolymers. Time-resolved WAXS facilitates the observation of dynamic processes in and biological systems, capturing transitions and conformational changes on timescales. In liquid crystalline polymers, it tracks smectic-to-nematic transitions by monitoring shifts in wide-angle reflections, revealing kinetic pathways influenced by cooling rates, as seen in polyesters where structural rearrangements occur within seconds post-quench. For , time-resolved WAXS detects unfolding intermediates in , with signal changes in the 4-10 range indicating tertiary structure collapse on sub-microsecond scales, complementing spectroscopic methods to map energy landscapes. Combining WAXS with (SAXS) provides a comprehensive view of hierarchical structures in block copolymers, where SAXS probes microphase separation (e.g., lamellar domains at 10-100 nm) and WAXS resolves local crystallinity within those domains. This dual approach has elucidated in polystyrene-block-polyethylene oxide copolymers, showing how block incompatibility drives ordered nanostructures with crystalline cores, essential for applications in . Sample preparation for such organic systems typically involves thin films or solutions to ensure orientation and minimal , aligning with general requirements for .

History

Early developments

The discovery of X-ray diffraction by crystals marked the foundational step in wide-angle X-ray scattering (WAXS), originating from Max von Laue's experiment in 1912. In this work, conducted with assistants Walter Friedrich and Paul Knipping, from a tube source were passed through a and recorded on a , producing a pattern that confirmed the wave nature of and their interaction with periodic atomic arrangements. This transmission geometry Laue method laid the groundwork for probing structures at wide angles, though initial patterns were complex due to the polychromatic beam. Building on Laue's findings, and William Lawrence Bragg developed a reflection-based interpretation in 1913, introducing to quantitatively relate angles to interplanar spacings in . Their approach simplified analysis by treating as from atomic planes, enabling the determination of simple structures like those of and using monochromatic X-rays. This advancement shifted focus toward wide-angle scattering for structural elucidation, with the Braggs earning the 1915 for their contributions. A pivotal innovation for WAXS came in 1916 with the Debye-Scherrer powder method, developed by and to extend studies to polycrystalline materials. By grinding samples into fine powders and rotating them in a cylindrical camera, they produced concentric ring patterns on , capturing wide-angle diffractions from randomly oriented crystallites and making the technique accessible for non-single-crystal samples like metals and minerals. Independently, Albert Hull reported similar results in 1917, further popularizing . In the 1920s, William Lawrence Bragg applied these methods to analyze the structures of minerals and metals, determining atomic arrangements in silicates and alloys that advanced understanding of material properties. His work at the utilized ionization chambers for intensity measurements, bridging early experiments to practical materials characterization. Prior to synchrotron sources, early WAXS was constrained by the low intensity of tubes, which required long exposure times—often hours or days—to produce detectable patterns on photographic plates. These plates, while sensitive, offered limited quantitative precision due to non-uniform development and difficulty in digitizing data, restricting studies to relatively simple systems and hindering real-time or in-situ observations.

Key advancements and modern techniques

The advent of synchrotron radiation sources in the marked a transformative era for wide-angle X-ray scattering (WAXS), providing unprecedented brilliance and coherence that surpassed laboratory sources by orders of magnitude. This "synchrotron revolution" enabled the development of microbeam WAXS techniques, allowing at sub-micrometer scales with high flux densities. For instance, the ID11 at the European Radiation Facility (ESRF), operational since the early 1990s, delivers beams as small as 15 μm with fluxes up to 2 × 10¹³ photons/s, facilitating in-situ studies of heterogeneous materials under extreme conditions. In the , the introduction of two-dimensional (2D) detectors, particularly (CCD) arrays at third-generation synchrotrons like the ESRF, revolutionized for time-resolved WAXS by capturing full rings in a single exposure. These pixel array detectors supported fast imaging with frame rates approaching 100 ms, enabling the study of rapid dynamic processes such as propagation in materials. This advancement shifted WAXS from static to kinetic analyses, with readouts fast enough to track transient structural changes in . The 2000s saw the popularization of the atomic pair distribution function (PDF) method in WAXS, which extracts local atomic arrangements from total scattering data, proving invaluable for disordered and nanocrystalline materials where Bragg diffraction alone is insufficient. Driven by improved high-energy beamlines and rapid-acquisition PDF (RAPDF) protocols using image plate detectors like the MAR345, data collection times dropped from hours to seconds, spurring widespread adoption. Seminal work, including the 2003 RAPDF implementation at , coupled with user-friendly software like PDFgui, led to an exponential increase in PDF applications, particularly for nanoscale disorder in glasses and alloys. Modern WAXS has increasingly integrated with complementary techniques for multimodal, real-time structural insights. In-situ combinations with (TEM) allow correlated nanoscale imaging and scattering, as demonstrated in studies of battery electrode evolution where synchrotron WAXS probes bulk phase changes alongside TEM's local morphology. Similarly, coupling WAXS with setups, such as extensional rheometers, enables tracking flow-induced in polymers during processing, with temporal resolutions below 1 second to capture viscoelastic transitions. These hybrid approaches, facilitated by synchrotron versatility, have become standard for investigating coupled structural-mechanical behaviors in complex systems.

References

  1. [1]
    Wide Angle X-Ray Diffraction - an overview | ScienceDirect Topics
    Wide-angle X-ray diffraction (WAXD) is defined as a technique used to study crystal structure and form by observing diffraction maxima at larger angles, ...Missing: review | Show results with:review<|control11|><|separator|>
  2. [2]
    WAXS and SAXS - Chemistry LibreTexts
    Aug 21, 2022 · WAXS (Wide-angle X ray scattering): In X-ray crystallography, is the analysis of Bragg peaks scattered to wide angles, which are caused by sub-nanometer sized ...
  3. [3]
    Wide-Angle X-Ray Scattering - an overview | ScienceDirect Topics
    WAXS, or wide angle X-ray scattering, refers to a technique that collects scattering patterns at higher angles to analyze secondary structures and their ...
  4. [4]
    Wide Angle X-Ray Scattering to Study the Atomic Structure of ... - MDPI
    Apr 4, 2020 · In this paper, we report the WAXS investigation performed with a table-top high brilliance X-ray micro-source on three different types of ...
  5. [5]
    Small and Wide Angle X-Ray Scattering Studies of Biological ... - NIH
    Jan 8, 2013 · SWAXS is a technique where X-rays are elastically scattered by an inhomogeneous sample in the nm-range at small angles (typically 0.1 - 5°) and ...
  6. [6]
    SAXS | LSU CAMD - Louisiana State University
    Depending on the angles studies, X-ray scattering probes different length scales. WAXS (5-40°) probes structures ~0.1-2 nm, SAXS (0.1-10°) probes 0.5-100 nm ...
  7. [7]
    and wide-angle X-ray scattering (SAXS/WAXS) experiments on a ...
    Aug 22, 2018 · ... wide-angle X-ray scattering experiments (SAXS/WAXS). There is ... In practice, the actual smallest achievable scattering vector qmin ...
  8. [8]
    X-Ray Diffraction - an overview | ScienceDirect Topics
    X-ray diffraction techniques are often categorized as wide-angle X-ray scattering (WAXS) and small-angle X-ray scattering (SAXS). WAXS is typically used to ...
  9. [9]
    Imaging of Biological Materials and Cells by X-ray Scattering and ...
    Aug 8, 2017 · In this review, we describe the advances in X-ray scattering and diffraction techniques for imaging biological systems at the nanoscale.
  10. [10]
    Hybridization of Wide-Angle X-ray and Neutron Diffraction ... - NIH
    Jan 16, 2023 · The neutron diffraction method is one of the most useful techniques complementary to the X-ray method. We can use the bulky sample and ...
  11. [11]
    BM26: SAXS/WAXS - European Synchrotron Radiation Facility (ESRF)
    If required, a SAXS/WAXS setup without gap between the two measured patterns is possible for a SAXS camera length of up to 1.5 m. SAXS-WAXS-samplechanger-700px.
  12. [12]
    I22: SAXS/WAXS beamline at Diamond Light Source - IUCr Journals
    Feb 23, 2021 · I22 was designed as a multipurpose, undulator driven, combined SAXS–WAXS facility to serve the needs of a broad. UK community of SAXS/WAXS users ...
  13. [13]
    [PDF] Chapter 3 Diffraction (Basic Idea) We will develop Bragg's Law by ...
    In derivation of Bragg's law we must consider the phase difference between two waves to tell if the amplitude interference is constructive (added) or ...
  14. [14]
    [PDF] The diffraction of X-rays by crystals - Nobel Lecture, September 6 ...
    The. Page 3. 372. 1915 W.L.BRAGG points of a space lattice may be arranged in series of planes, parallel and equidistant from each other. As a pulse passes over ...
  15. [15]
    Calculation of Debye-Scherrer diffraction patterns from highly ...
    Jun 6, 2016 · FELs produce nearly monochromatic x-rays, requiring polycrystalline samples to produce Debye-Scherrer diffraction rings from a compressed ...
  16. [16]
    [PDF] Calculation of Debye-Scherrer diffraction patterns from highly ...
    May 19, 2016 · FELs produce nearly monochromatic x-rays, requir- ing polycrystalline samples to produce Debye-Scherrer diffraction rings from a compressed ...<|separator|>
  17. [17]
    [PDF] Quantification of amorphous phases - theory
    – Broad diffraction halos resulting in an increased peak overlap problem. – Discrimination of peak tail / amorphous band / background intensities.
  18. [18]
    [PDF] Elements of X-Ray Diffraction BD Cullity SR Stock Third Edition
    ... Factor. 635. B.D. Cullity/S.R. Stock. Appendix: Data for Calculation of the ... rays and provided a new method for investigating the fine structure of matter.
  19. [19]
    [PDF] Chapter 4 Development of a Model for Diffracted Intensity
    The diffracted intensity can be calculated from a series of terms related to 1) what hkl reflection is being observed, 2) what wavelength is being used, 3) the ...
  20. [20]
    NIST Atomic Form Factors: Form factors and standard definitions
    The (x-ray) atomic form factor f is the resonant scattering amplitude of x-rays by charge (primarily electron) density.
  21. [21]
    Serial small- and wide-angle X-ray scattering with laboratory sources
    Jul 26, 2022 · This setup delivers an almost parallel beam that minimizes smearing of the scattering intensity at small angles. Samples were introduced to the ...
  22. [22]
  23. [23]
    Laboratory vs Synchrotron SAXS - Xenocs
    Sep 23, 2025 · ... Wide Angle X-ray Scattering (WAXS). Instruments at a synchrotron can offer collimated beams down to 100 x 100 nm² with a flux as high as 10 ...
  24. [24]
    A wide-angle X-ray scattering laboratory setup for tracking phase ...
    Nov 14, 2022 · The developed GIWAXS setup uses a highly-sensitive 2D X-ray detector with fast read-out time, which allows observation of crystal structure ...
  25. [25]
    [PDF] X-ray Scattering Techniques for Characterization of Nanosystems in ...
    Apr 14, 2005 · This range of domain size requires both the wide-angle x-ray scattering (WAXS) and small-angle. (SAXS) x-ray scattering techniques. Roughly WAXS ...<|control11|><|separator|>
  26. [26]
    [PDF] Sample Preparation of Atherosclerotic Plaque for SAXS/WAXS ...
    Apr 4, 2023 · Small- and wide-angle X-ray scattering (SAXS/. WAXS) holds the ... all sample preparation methods were able to produce samples of the same ...
  27. [27]
  28. [28]
    and wide-angle X-ray scattering (SAXS/WAXS) experiments on a ...
    Aug 22, 2018 · wide-angle X-ray scattering (WAXS) are enabled by the exper- imental ... Repeated sample preparation and measurements of a given sample ...
  29. [29]
  30. [30]
    X-ray reflection in accordance with Bragg's Law
    Solving Bragg's Equation gives the d-spacing between the crystal lattice planes of atoms that produce the constructive interference. A given unknown crystal ...Missing: derivation | Show results with:derivation
  31. [31]
    Safinya Group Research Small Angle X-Ray Scattering and Diffraction
    Wide angle X-ray diffraction further allows to determine structures at a finer resolution than any microscopy technique. Bragg's law, which is a fundamental ...Missing: scientific review
  32. [32]
    GSAS-II
    GSAS-II is a tool for x-ray and neutron diffraction data analysis, used for crystal structure determination and materials characterization.Missing: WAXS ICDD
  33. [33]
    PDF-5+ - - International Centre for Diffraction Data
    ICDD's search indexing software, SIeve+ for PDF-5+ is now included for FREE. · SIeve+ offers a variety of algorithms and options that allow users to optimize ...
  34. [34]
    [PDF] X-Ray Diffraction (XRD) - IIT Kanpur
    spacing, as calculated by Bragg's law. • Bragg's law relates the diffraction angle, 2θ, to d hkl. – In most diffractometers, the X-ray wavelength λ is fixed.
  35. [35]
    Deriving Amorphous Component Abundance and Composition of ...
    Sep 14, 2018 · Because the peaks are broad and low, XRD patterns of amorphous materials are significantly less distinct than for crystalline phases. However, ...Introduction · Background · Methods · DiscussionMissing: distinction | Show results with:distinction
  36. [36]
    [PDF] quantitative phase-analysis by the rietveld method using x-ray
    The method is based on a least-squares fit between step-scan data of a measured diffraction pattern and a simulated X-ray-diffraction pattern. The simulated XRD ...
  37. [37]
    Preferred orientation and its effects on intensity-correlation ... - NIH
    Macrotexture occurs when the preferred orientation of crystal grains is correlated to the macroscopic dimensions of the sample. This type of texture is ...
  38. [38]
    X-ray diffraction study on residual stress and preferred orientation in ...
    Dec 15, 2004 · Like residual stress, preferred orientation may significantly influence thin-film properties, which is not only known from nitride hard coatings ...<|separator|>
  39. [39]
    X-Ray Diffraction Line Broadening: Modeling and Applications ... - NIH
    Strain broadening is angle dependent. Therefore, the angle dependence of the line broadening gives a possibility to distinguish between contributions of size ...
  40. [40]
    Factors effecting peak width - XRD - MyScope
    There are three contributions to peak broadening in the XRD machine: Instrumental broadening; Grain-size broadening; Micro-strain broadening (proportional to ...
  41. [41]
    Multiscale Description of Shale Pore Systems by Scanning SAXS ...
    Nov 2, 2016 · While different minerals may show overlapping peaks, other single peaks may be indicative for a specific mineral. Minerals can thus be ...
  42. [42]
    Rietveld Refinement - an overview | ScienceDirect Topics
    Rietveld refinement is conducted by fitting a calculated diffraction pattern to the observed data by adjusting each of the variables that describe the ...
  43. [43]
    Challenges in Rietveld Refinement and Structure Visualization in ...
    The method now known as Rietveld refinement was the first step towards full profile whole powder pattern fitting method for x-ray and neutron diffraction data.
  44. [44]
    FullProf Suite
    The FullProf Suite is formed by a set of crystallographic programs mainly developed for Rietveld analysis (structure profile refinement) of neutron (constant ...
  45. [45]
    DIFFRAC.TOPAS - Bruker
    TOPAS is a profile fitting based software for quantitative phase analysis, microstructure analysis and crystal structure analysis.
  46. [46]
    Structural Analysis of Molecular Materials Using the Pair Distribution ...
    Nov 17, 2021 · This is a review of atomic pair distribution function (PDF) analysis as applied to the study of molecular materials.
  47. [47]
    The Scherrer equation versus the 'Debye-Scherrer equation' - Nature
    Aug 28, 2011 · Scherrer derived his equation for the ideal condition of a perfectly parallel, infinitely narrow and monochromatic X-ray beam incident on a monodisperse powder ...
  48. [48]
    Scherrer Equation - an overview | ScienceDirect Topics
    The Debye Scherrer equation, D = Kλ / βCosθ, is used to calculate the crystalline size of the nanoparticles, where D is the nanoparticles crystalline size.
  49. [49]
    Investigation of deformation and microstructural evolution in Grade ...
    For WAXS, diffraction peaks from multiple phases within the illuminated material volume are recorded simultaneously on the detector.
  50. [50]
    XRD, WAXS, FTIR, and XANES studies of silica-zirconia systems
    Aug 15, 2019 · An analysis of the structure through XRD, WAXS, FTIR, and XANES was carried out to examine the characteristics of the ceramic products. Section ...
  51. [51]
    High energy X-ray diffraction and small-angle scattering ...
    Here we present simultaneous High Energy X-ray Diffraction (HEXRD) and Small-Angle X-ray Scattering (SAXS) during fatiguing of steel in hydrogen.
  52. [52]
    Tutorial on Powder X-ray Diffraction for Characterizing Nanoscale ...
    Jul 23, 2019 · When the crystallite size decreases from bulk to nanoscale dimensions, the XRD peaks broaden. The Scherrer equation, , quantitatively describes ...Missing: WAXS | Show results with:WAXS
  53. [53]
    Multiscale X-ray scattering elucidates activation and deactivation of ...
    Jan 3, 2025 · We present a multiscale in situ investigation of activation and deactivation pathways of oxide-derived copper electrocatalysts under CO 2 reduction conditions.
  54. [54]
    Melting of DNA Nonoriented Fibers: A Wide-Angle X-ray Diffraction ...
    Mar 14, 2014 · The melting transition of A- and B-DNA has been investigated by wide-angle X-ray diffraction. A significant crystalline phase is present in both the systems.
  55. [55]
    Max von Laue – Facts - NobelPrize.org
    In 1912, Max von Laue came upon the idea that X-rays passing through crystals might create similar patterns. That is, that a crystal's structure would ...
  56. [56]
    [PDF] The discovery of the diffraction of X-rays by crystals
    The idea that crystals ought to diffract X-rays arose in Munich, LADE maintained, because of two unique factors in the intellectual milieu: i) confidence in the ...
  57. [57]
    Perspectives: X-ray's identity becomes crystal clear - NobelPrize.org
    von Laue designed an experiment in which he placed a copper sulphate crystal between an X-ray tube and a photographic plate. His assistants, Walter Friedrich ...
  58. [58]
    The reflection of X-rays by crystals - Journals
    William Henry Bragg, Proceedings A, 1913. The analysis of crystals by the X-ray spectrometer. William Lawrence Bragg, Proceedings A, 1914. III. On the ...
  59. [59]
    The birth of X-ray crystallography - Nature
    Nov 7, 2012 · A century ago this week, physicist Lawrence Bragg announced an equation that revolutionized fields from mineralogy to biology.
  60. [60]
    100 years of the X-ray powder diffraction method | OUPblog
    Sep 20, 2016 · In a series of papers published in 1913, Debye calculated the influence of lattice vibrations on the diffracted intensity (the Debye factor), ...Missing: original | Show results with:original
  61. [61]
    [PDF] Powder Methods of X-Ray Analysis - Department of Physics
    The use of powder in X-ray analysis was first demonstrated about 1916 when Debye and Scherrer in Germany and Hull in the United States independently obtained.Missing: original | Show results with:original
  62. [62]
    Research Profile - William Bragg | Lindau Mediatheque
    The line of research on organic crystals started in the 1920s confirmed the structural models of organic chemistry. X ray crystallography soon informed ...
  63. [63]
    X-Ray Crystallography Is Developed by the Braggs | Research Starters
    Lawrence Bragg and William Henry Bragg founded the science of X-ray crystallography, verified the very short wavelength nature of X radiation, developed ...
  64. [64]
    Demystifying the synchrotron trip: a first time user's guide
    The X-ray radiation produced at synchrotrons is very intense and thousands of times brighter than that produced by a home laboratory source. Therefore, the ...Missing: Early | Show results with:Early
  65. [65]
    XRD Basics - UPenn Physics
    Jul 16, 2025 · X-ray diffraction (XRD) is a non-destructive technique for analyzing the structure of materials, primarily at the atomic or molecular level.What are the components of... · How are x-ray area detector...
  66. [66]
    Synchrotron Radiation X-ray Scattering Approaching Real Industrial ...
    Apr 20, 2025 · An instrument for the collection of simultaneous small and wide angle x-ray scattering and stress-strain data during deformation of polymers ...
  67. [67]
    ID11 - Materials science beamline
    Scientists synthesise new materials at terapascal pressures for the first time · ESRF user awarded ERC advanced grant.
  68. [68]
    X-ray imaging at synchrotron research facilities - ScienceDirect.com
    In the late 1990s most large-area 2D X-ray detectors at ESRF and at other 3rd generation synchrotron facilities were based on charge-coupled (CCD) sensors ...
  69. [69]
    The rise of the X-ray atomic pair distribution function method - Journals
    Apr 29, 2019 · The atomic pair distribution function (PDF) technique is a powerful approach to gain quantitative insight into the structure of materials.Abstract · Introduction · The atomic pair distribution... · Appendix A. The PDF method
  70. [70]
    In situ TEM and synchrotron SAXS/WAXS study on the impact of ...
    Jul 16, 2025 · A potential solution is catalytic graphitization, where catalysts from abundant elements such as iron are used to convert amorphous carbon to ...
  71. [71]
    Extensional rheometer for in situ x-ray scattering study on flow ...
    Apr 13, 2011 · We designed and constructed an extensional rheometer for in situ small and wide angle x-ray study on flow-induced crystallization of polymer.Missing: TEM studies advancements