Fact-checked by Grok 2 weeks ago

Rietveld refinement

Rietveld refinement is a least-squares for refining crystal structures from data, where a calculated diffraction profile, based on a structural model, is fitted to the entire observed to determine parameters such as positions, occupancies, constants, and thermal displacements. Developed by Hugo M. Rietveld in 1969 while working at the Reactor Centrum Nederland, the technique was initially applied to for nuclear and magnetic structure analysis but quickly extended to and data. By modeling overlapping peaks and background contributions through analytical profile functions—such as Gaussian, , or pseudo-Voigt—the overcomes limitations of traditional single-peak extraction, enabling precise structural characterization even from polycrystalline samples. The Rietveld method revolutionized powder crystallography by providing a full-pattern fitting approach that incorporates instrumental resolution, sample microstructure (e.g., size and microstrain broadening), and preferred orientation effects, resulting in more reliable quantitative phase analysis and structural refinements. Key to its success is the iterative refinement , which minimizes the weighted difference between observed and calculated intensities using figures of merit like the weighted profile R-factor (R_wp) and goodness-of-fit (χ²), with convergence typically assessed when parameter shifts are less than 0.1 times their estimated standard deviations. Originally implemented in code for mainframe computers, modern software packages like GSAS, FullProf, and TOPAS facilitate its use across diverse applications, from to pharmaceutical development. Over the past five decades, Rietveld refinement has become the standard for analyzing complex multiphase materials, with its original publication cited more than 22,000 times as of 2025, underscoring its foundational role in advancing structural science. Guidelines from the International Union of emphasize high-quality —such as step sizes such that there are approximately 5–10 steps across the width of each peak (or about 10–20% of the ) and to minimize and effects (ideally 1–5 µm)—to ensure accurate results. Despite challenges like correlated parameters or amorphous content, restraints on bond lengths and angles, derived from known chemistry, enhance reliability in underdetermined cases.

History and Overview

Development and Origins

The Rietveld method, building on earlier profile refinement work by colleagues Bert O. Loopstra and Bob van Laar, was developed by Hugo M. Rietveld starting around 1965 at the Reactor Centre Netherlands in Petten, , where he worked in the neutron diffraction group following his . Initially created to analyze powder data, the approach addressed key challenges in extracting accurate structural parameters from patterns featuring overlapping reflections, which traditional methods struggled to resolve. This development marked an evolution from conventional single-peak fitting techniques, which relied on isolating and integrating individual Bragg , to a full-profile refinement strategy that modeled the entire pattern holistically. The motivation stemmed from the inherent limitations of , such as peak overlap and low signal-to-noise ratios in experiments, prompting Rietveld to leverage computational capabilities available at the time, including the Electrologica X8 computer. The foundational description of peak line profiles appeared in a 1967 note in Acta Crystallographica, but the first full implementation of the profile refinement method was detailed in Rietveld's 1969 publication in the Journal of Applied Crystallography. In this seminal work, the method was presented as a least-squares procedure using step-scanned profile intensities rather than integrated intensities. Early applications focused on simple crystalline structures, including the refinement of nuclear parameters in alkaline-earth metal uranates such as Sr₂UO₅. These initial tests highlighted the method's potential to improve precision in structure determination beyond what was achievable with prior techniques.

Basic Principles

Rietveld refinement is a full-pattern fitting technique that employs a least-squares approach to minimize the difference between an observed profile and a calculated profile generated from a structural model. This method directly utilizes the step-scanned intensities of the pattern, rather than relying on extracted integrated intensities from individual peaks, enabling the refinement of complex structures where overlaps are common. Developed initially for neutron data on nuclear and magnetic structures, it has since been extended to , providing a robust framework for quantitative and structural determination. The primary goal of Rietveld refinement is to simultaneously optimize structural parameters—such as lattice constants, atomic positions, and site occupancies—along with instrumental factors like peak shape and width parameters, and sample-related effects including background and preferred orientation. By modeling the entire pattern, the extracts more reliable information from the data than methods limited to isolated reflections, particularly in cases of low-resolution or highly overlapping peaks typical of samples. This holistic fitting process enhances the accuracy of refined parameters, making it indispensable for in fields like and . The workflow begins with an initial structural model, often derived from single-crystal data or predictions, which is used to compute the theoretical profile based on the crystal's , atomic coordinates, and scattering factors. This calculated profile is then compared to the experimental data, and parameters are iteratively adjusted via least-squares minimization until the residual difference converges to a minimum, typically assessed by metrics like the profile R-factor. Constraints, such as those ensuring physical plausibility (e.g., lengths), may be applied to guide the refinement and prevent divergence. In contrast to single-reflection methods, such as profile-independent analysis that decomposes peaks to obtain individual intensities before refinement, Rietveld refinement leverages the full pattern to inherently account for overlaps and correlations between reflections, improving precision without requiring peak isolation. This distinction addresses the inherent limitations of data, where many reflections contribute to the observed , allowing for more comprehensive structural insights.

Powder Diffraction Fundamentals

Peak Positions and Shapes

In powder diffraction patterns, the positions of Bragg peaks are fundamentally determined by , expressed as n\lambda = 2d_{hkl} \sin\theta, where n is a positive representing the diffraction order, \lambda is the of the incident , d_{hkl} is the interplanar spacing for the set of planes with (hkl), and \theta is the Bragg diffraction angle. The interplanar spacing d_{hkl} is directly related to the unit cell parameters of the through geometric relations specific to the , such as \frac{1}{d_{hkl}^2} = \frac{h^2}{a^2} + \frac{k^2}{b^2} + \frac{l^2}{c^2} for an orthorhombic . In Rietveld refinement, these theoretical peak positions are computed from an initial structural model and serve as the foundation for fitting the entire profile, enabling the refinement of parameters to match observed data. Diffraction peak shapes in powder patterns deviate from ideal symmetry due to both instrumental and sample-related effects. Instrumental asymmetry often arises from the use of flat-plate samples in Bragg-Brentano geometry, where the sample surface is not perfectly aligned with the diffractometer circle, leading to defocusing and angular shifts that elongate peaks toward lower $2\theta angles, particularly for low-angle reflections. Physical contributions to peak shape include broadening from finite crystallite sizes, which produces Lorentzian-like tails due to the limited number of coherently scattering domains, and microstrain from lattice distortions, which adds Gaussian broadening through variations in interplanar spacings. These factors collectively result in asymmetric profiles that must be accurately parameterized for reliable structural analysis in Rietveld methods. Peak shapes are modeled using analytical profile functions to capture both symmetric broadening and asymmetry. A widely adopted function is the pseudo-Voigt profile, defined as I(2\theta) = \eta \, L(2\theta; \Gamma_L) + (1 - \eta) \, G(2\theta; \Gamma_G), where L and G represent the Lorentzian and Gaussian components, respectively, \Gamma_L and \Gamma_G are their respective full widths at half maximum, and \eta (ranging from 0 to 1) is a refinable mixing parameter that adjusts the relative contributions, often varying with $2\theta to account for evolving asymmetry. This function approximates the true Voigt convolution while being computationally efficient for whole-pattern fitting. In the seminal work introducing the Rietveld method, peak profiles were modeled using a Gaussian function.

Integrated Intensities and Widths

In , the integrated intensity of a Bragg represents the total area under the peak and serves as a key observable for . This intensity I_{hkl} is proportional to the square of the magnitude of the , |F_{hkl}|^2, which encodes information about atomic positions, site occupancies, and thermal displacements within the crystal lattice. Additionally, I_{hkl} incorporates geometric and experimental factors, including the multiplicity m_{hkl} (accounting for equivalent reflections) and the Lorentz-polarization factor L_p (arising from the scanning geometry and beam polarization). The full expression is thus I_{hkl} \propto m_{hkl} L_p |F_{hkl}|^2, enabling the extraction of refined structural parameters during Rietveld refinement by comparing modeled intensities to observed data. Peak widths in diffraction patterns, quantified by the (FWHM, denoted \beta), provide insights into microstructural features such as size and . For size D, the relates \beta to the average domain size via D = \frac{K \lambda}{\beta \cos \theta}, where K is a (typically 0.9–1.0), \lambda is the , and \theta is the Bragg ; narrower peaks indicate larger . To disentangle size and strain effects, the Williamson-Hall method employs a linear plot of \beta \cos \theta versus \sin \theta: \beta \cos \theta = \frac{K \lambda}{D} + 4 \epsilon \sin \theta, where \epsilon is the microstrain; the y-intercept yields size information, while the slope reflects strain broadening. These analyses are integrated into Rietveld refinement to quantify microstructural properties alongside structural refinement. Rietveld's original method modeled peak widths using a Gaussian profile to account for overall broadening from instrumental and sample effects. This unified approach avoids isolating individual peaks for width measurement, instead refining width parameters globally across the pattern to simultaneously optimize intensities and shapes. Through such modeling, site occupancies (affecting scattering power per site) and thermal parameters (via the Debye-Waller factor in F_{hkl}) are determined by minimizing discrepancies between calculated and observed integrated intensities, enhancing the accuracy of atomic-scale models.

Core Refinement Techniques

Least Squares Method

The method serves as the foundational optimization algorithm in Rietveld refinement, enabling the adjustment of structural and instrumental parameters to achieve the best fit between observed and calculated powder diffraction profiles. Developed by Rietveld in his seminal 1969 paper, this technique adapts classical principles to the entire diffraction pattern, allowing for the extraction of detailed structural information from powder data where single-crystal methods are infeasible. At its core, the method minimizes the chi-squared (\chi^2) statistic, which quantifies the discrepancy between experimental and modeled data: \chi^2 = \sum_i w_i (y_{obs,i} - y_{calc,i})^2, where y_{obs,i} and y_{calc,i} represent the observed and calculated step intensities at the i-th point across the profile, and the weights w_i = 1/\sigma_i^2 are inversely proportional to the variances \sigma_i^2 of the measurements, ensuring higher-precision data points exert greater influence on the fit. This formulation assumes Gaussian errors in the intensity measurements, a common approximation in powder diffraction analysis. The refinement is inherently nonlinear due to the complex dependence of calculated intensities on parameters, necessitating an iterative solution. Parameter updates \Delta \mathbf{p} are obtained by solving the normal equations: \Delta \mathbf{p} = (A^T W A)^{-1} A^T W \Delta \mathbf{y}, where \mathbf{A} is the Jacobian matrix containing partial derivatives \partial y_{calc,i}/\partial p_j of the calculated profile with respect to each refinable parameter p_j, W is the diagonal weight matrix, and \Delta \mathbf{y} = \mathbf{y}_{obs} - \mathbf{y}_{calc}. To enhance numerical stability, especially with ill-conditioned matrices from correlated parameters or noisy data, implementations often employ gradient-based algorithms like the Levenberg-Marquardt method, which blends steepest descent and Gauss-Newton approaches. Convergence is monitored through criteria such as parameter shifts falling below 0.1 times their estimated standard deviations (esds) or the \chi^2 value stabilizing across iterations, preventing overfitting while ensuring a reliable minimum is reached. The final variance-covariance matrix, computed as the inverse of the weighted curvature matrix (A^T W A) scaled by the reduced \chi^2 (i.e., \chi^2 / (N - P), with N observations and P parameters), yields the esds along its diagonal and quantifies correlations off-diagonal, highlighting potential parameter interdependencies that may require constraints or sequential refinement strategies. Unlike traditional crystallographic least squares refinements, which operate on extracted integrated intensities of discrete Bragg reflections and thus discard profile shape information, the Rietveld method leverages the full intensity profile—including overlaps and backgrounds—for a more holistic and statistically robust optimization.

Profile Function Modeling

In the Rietveld refinement method, the theoretical diffraction profile is constructed by summing the contributions from all Bragg reflections across the entire powder diffraction pattern. The calculated intensity at each point $2\theta_i is given by y_{calc}(2\theta_i) = \sum_{hkl} I_{hkl} \, P(2\theta_i - 2\theta_{hkl}) + BG(2\theta_i), where I_{hkl} represents the integrated intensity of the (hkl) reflection, P is the normalized profile function describing the shape of an individual peak, $2\theta_{hkl} is the calculated Bragg position, and BG(2\theta_i) accounts for the background intensity. This approach models the entire observed pattern y_{obs}(2\theta_i) without isolating individual peaks, enabling a least-squares fit to refine structural and other parameters simultaneously. Rietveld's original formulation, developed for neutron powder diffraction data, employed analytical profile functions to describe shapes, assuming Gaussian profiles determined primarily by instrumental resolution. Peak widths were parameterized using the Caglioti relation, \mathrm{FWHM}^2 = U \tan^2\theta + V \tan\theta + W, where U, V, and W are refinable coefficients capturing angular dependence. Later adaptations incorporated more flexible analytical profile functions, such as the pseudo-Voigt or split-Pearson functions, to model deviations from simple Gaussian or shapes, including asymmetry and sample effects, without requiring explicit separation of overlapping contributions. Modern extensions to profile modeling have shifted toward the fundamental parameters approach (FPA), which physically parameterizes the instrumental resolution based on optical elements like source emission profiles, monochromator characteristics, and detector responses, rather than empirically fitting parameters. In FPA, the overall peak profile is obtained by convolving the intrinsic sample (e.g., a delta at $2\theta_{hkl}) with the instrumental and any sample broadening terms, such as size or microstrain effects. This method, pioneered for data, reduces the number of refinable parameters and improves transferability across instruments, enhancing accuracy in quantitative phase analysis and microstructural characterization. The Rietveld approach inherently addresses overlapping peaks by refining all profile parameters globally across the pattern, avoiding the need for decomposition into individual reflections. This simultaneous fitting leverages the structural model to correlate intensities and positions, allowing resolution of severe overlaps that would be intractable in traditional peak-by-peak methods.

Refinement Parameters and Strategies

Structural and Instrumental Parameters

In Rietveld refinement, structural parameters define the crystal lattice and atomic arrangement within the unit cell, enabling the calculation of Bragg peak positions and intensities from the known atomic scattering factors. Lattice parameters, including the cell lengths a, b, c and angles \alpha, \beta, \gamma, are refined to align calculated peak positions with observed diffraction data, often starting with values from single-crystal studies or indexing results. Accurate refinement of these parameters requires high-resolution data and correction for instrumental shifts, as small errors can propagate to significant deviations in derived volumes or densities. Atomic coordinates, typically expressed as fractional positions x, y, z for each in the asymmetric unit, are refined to optimize the contributions to peak intensities, with heavier atoms refined before lighter ones to minimize correlations. These parameters account for site occupancies in cases of substitutional , ensuring the model reflects the average structure observed in data. parameters, which describe motion or static , are refined as isotropic values B_{iso} for simpler models or anisotropic tensors U_{ij} for more detailed analyses, with providing superior sensitivity due to from nuclei rather than electrons. Refinement proceeds cautiously, often grouping similar atoms to avoid divergence from correlated shifts in coordinates and displacements. Instrumental parameters correct for systematic errors in the setup, ensuring the calculated matches the experimental across the angular range. The zero-point error, representing the angular offset at 2θ = 0°, is refined alongside lattice parameters to adjust peak positions, typically calibrated using a like NIST SRM 640b for precision better than 0.01°. calibration accounts for any deviation from the nominal value in the incident beam, essential for d-spacing determination. The scale factor normalizes the overall intensity of the calculated to the observed , while corrections address sample transparency or μρ effects, particularly in transmission geometry for accurate intensity scaling. For multi-phase samples, phase-related parameters include weight fractions, calculated as W_\alpha = \frac{S_\alpha Z_\alpha [M_\alpha](/page/Molecular_mass) V_\alpha}{\sum S_i Z_i M_i V_i} where S_\alpha is the refined scale factor, Z_\alpha the number of formula units, M_\alpha the , and V_\alpha the unit cell volume for phase \alpha, enabling without external standards. Unit cell variations across phases are refined independently to capture compositional differences, with the sum of fractions constrained to unity. To maintain physical realism and convergence in refinements with limited data, constraints fix relationships between parameters, such as linking occupancies in solid solutions to maintain charge balance or . Restraints, applied as soft penalties, enforce expected bond lengths or angles based on prior chemical knowledge, weighted by an empirical factor to balance geometric and data without over-constraining the model. These techniques are particularly vital for complex or low-symmetry structures, reducing parameter correlations and improving precision.

Background and Preferred Orientation

In Rietveld refinement, the background scattering component arises from various sources such as air scattering, sample holder contributions, and incoherent processes like or , necessitating explicit modeling to isolate the Bragg signals accurately. Empirical polynomial functions, particularly shifted Chebyshev polynomials of the first kind, are widely employed for this purpose due to their flexibility in fitting smooth, slowly varying backgrounds without introducing unphysical oscillations. The background intensity is typically expressed as BG(2\theta) = \sum_{k=1}^{n} c_k T_k(2\theta - \gamma), where T_k are the shifted Chebyshev polynomials, c_k are refinable coefficients, $2\theta is the scattering angle, \gamma is a shift to ensure over the data range, and n is the number of terms selected based on the pattern's complexity to avoid . For cases involving significant incoherent scattering, physical models incorporating Compton profiles or fluorescence corrections can supplement or replace polynomials, particularly in high-energy experiments where such contributions are pronounced. Detection of background inadequacies during refinement involves of residuals between observed and calculated profiles, where systematic positive or negative deviations in inter-peak regions signal the need for adjustment; strategies typically begin with a fixed low-order (e.g., 3-6 terms) refined after initial scale and zero-angle parameters, progressing to higher orders or physical terms only if residuals persist. Refinable coefficients c_k allow adaptation to experimental variations, but fixed constraints on higher-order terms prevent in noisy data, ensuring while maintaining physical realism. These corrections directly influence the scaling of integrated intensities, as unmodeled inflates or suppresses areas, thereby affecting quantification and structural reliability. Preferred orientation, or texture, introduces intensity distortions in powder diffraction patterns when crystallites align preferentially due to particle morphology or preparation methods, leading to over- or under-representation of specific hkl reflections. The March-Dollase model addresses uniaxial textures by applying a correction factor to the structure factor multiplicity, assuming ellipsoidal distribution of orientations around a preferred axis, and is particularly effective for moderate anisotropies where the direction is known a priori from morphology. For more complex, fiber-like or multi-axial textures, spherical harmonics expansions provide a general framework, modeling the orientation distribution function as \alpha_{hkl} = \sum_{m=0}^{l_{\max}} g_m Y_m(\gamma), where Y_m are spherical harmonics, g_m are refinable coefficients up to order l_{\max} (often 4-8 for convergence), and \gamma is the angle between the scattering vector and the sample normal. Detection relies on discrepancies in relative peak intensities compared to a random powder standard, with initial indicators from enhanced basal peaks in platy materials or axial reflections in elongated ones. Refinement strategies for preferred orientation commence after profile and lattice parameters are stable, introducing fixed initial values (e.g., March parameter G = 1 for ) and refining stepwise to avoid correlations with scale factors; higher-order harmonics require constraints to reduce parameters and ensure stability. In anisotropic materials like clays, where platy habits induce strong (00l) alignment, or fibers exhibiting axial textures, uncorrected preferred orientation can bias site occupancies and thermal parameters by up to 20-50% in affected reflections, underscoring the need for these models to achieve quantitative accuracy in phase analysis and microstructure inference.

Evaluation and Applications

Figures of Merit

In Rietveld refinement, figures of merit provide quantitative assessments of the agreement between the observed pattern and the model calculated from the refined structural parameters, as well as the overall statistical reliability of the fit. These metrics are derived from the least-squares minimization process and are essential for evaluating refinement quality, though they must be interpreted alongside of difference plots. The profile R-factor (R_p) measures the unweighted discrepancy between observed and calculated intensities across the entire profile: R_p = 100 \sqrt{ \frac{ \sum (y_{obs} - y_{calc})^2 }{ \sum y_{obs}^2 } } where y_{obs} and y_{calc} represent the observed and calculated step intensities, respectively. A weighted counterpart, the weighted profile R-factor (R_{wp}), incorporates measurement uncertainties via weights w_i = 1 / \sigma_i^2: R_{wp} = 100 \sqrt{ \frac{ \sum w_i (y_{obs,i} - y_{calc,i})^2 }{ \sum w_i y_{obs,i}^2 } } These profile factors quantify how well the entire pattern, including background and peak shapes, matches the model; R_{wp} is preferred as it penalizes larger errors more heavily. Typical values for R_{wp} below 10% indicate a good fit for standard data, though this can vary with instrumental and sample quality—lower thresholds (e.g., 2–5%) are common for high-resolution data. The Bragg R-factor (R_B) focuses specifically on the agreement between observed and calculated integrated Bragg intensities, which relate directly to factors: R_B = 100 \sqrt{ \frac{ \sum |I_{obs} - I_{calc}|^2 }{ \sum I_{obs}^2 } } where I_{obs} and I_{calc} are the extracted or modeled integrated intensities for each reflection. This metric is particularly useful for assessing the structural model's accuracy, independent of profile shape details, and values under 5–8% are generally indicative of reliable structure factor refinement. The goodness-of-fit (GoF) evaluates the fit relative to the expected statistical variance and is defined as GoF = R_wp / R_exp, where R_exp = 100 \sqrt{(N - P) / \sum w_i y_{obs,i}^2} is the expected weighted profile R-factor, with N as the number of data points and P as the number of refined parameters. An ideal GoF near 1 suggests the model adequately explains the data without overfitting or underfitting; values significantly above 1 may indicate unmodeled systematic errors, while those below 1 could reflect overestimated uncertainties. These figures of merit are sensitive to background modeling, as incorrect background subtraction can inflate R_p and R_{wp} by including non-structural contributions in the profile comparison. Additionally, they do not capture preferred orientation or sample-related artifacts, necessitating complementary diagnostics.

Practical Applications and Limitations

Rietveld refinement is widely applied in quantitative phase analysis (QPA) of multiphase mixtures, enabling precise determination of phase fractions without internal standards by scaling structure factors to match observed intensities. This excels in complex systems where traditional single-peak methods fail due to overlapping reflections, achieving accuracies within ±1-3 wt% for crystalline components in geological and samples. In , it models diffraction peak broadening to quantify crystallite size and microstrain, providing insights into defect distributions and anisotropic broadening via integral breadth or Williamson-Hall approaches integrated into the refinement. For instance, refinements of ceria nanoparticles yield crystallite sizes on the order of 15 , correlating with conditions. In-situ studies leverage Rietveld refinement to track dynamic transformations in , particularly in materials under electrochemical cycling. High-energy combined with refinement reveals lithium insertion mechanisms in cathodes like Li-rich layered oxides, quantifying evolving fractions during charging. Integration with pair distribution function (PDF) analysis extends Rietveld to total scattering data, coupling long-range crystallographic models with local atomic correlations for and disordered systems. This hybrid approach refines both Bragg and diffuse scattering, elucidating short-range order in illite clays or zeolite-encapsulated clusters. Practical case studies highlight its utility in pharmaceuticals for polymorph identification and quantification, where form stability affects . In famotidine tablets, Rietveld QPA detects trace polymorphic impurities at levels of a few percent, ensuring without destructive sampling. In , it quantifies fractions in Fe-B-Nb alloys to optimize magnetic properties. Figures of merit like the weighted profile R-factor (R_wp < 10%) validate these analyses, confirming model reliability. Despite its strengths, Rietveld refinement requires an accurate starting structural model; poor initial parameters lead to failures or artifacts in refined occupancies. Amorphous content complicates fits by elevating the background, often necessitating separate modeling or underestimating crystalline phases in nanocomposites. Peak overlap in complex, low-symmetry systems amplifies errors from profile asymmetry, while during prolonged exposure induces phase changes, degrading data quality. Modern advancements address these challenges through total scattering refinements that unify Rietveld with PDF for comprehensive structural probing, as in high-pressure studies of layered chalcogenides. Post-2020 developments incorporate to optimize starting models and automate phase quantification, reducing refinement time for large datasets from hours to minutes while improving accuracy in noisy data from various applications.

References

  1. [1]
  2. [2]
  3. [3]
  4. [4]
    Powder diffraction | Nature Reviews Methods Primers
    Nov 18, 2021 · Powder diffraction is a non-destructive technique, which is experimentally simple in principle. Because the physics behind diffraction is well understood.
  5. [5]
    Profile fitting of powder diffraction patterns - GeoScienceWorld
    Mar 3, 2017 · S. A. Howard, K. D. Preston; Profile fitting of powder diffraction patterns. Reviews in Mineralogy and Geochemistry 1989;; 20 (1): 217–275.
  6. [6]
  7. [7]
    [PDF] Rietveld Profile Functions
    Hugo Rietveld's low-resolution neutron diffraction peaks were Gaussian (determined mainly by the neutron spectral distribution, the monochromator response ...
  8. [8]
    The Rietveld Refinement Method: Half of a Century Anniversary
    Aug 18, 2021 · In 1969, Hugo Rietveld published the seminal article (2) on what has become known as the Rietveld refinement method. As a result, in the ...Author Information · References
  9. [9]
    [PDF] Rietveld re®nement guidelines
    The topics covered include (i) data collection, (ii) back- ground contribution, (iii) peak-shape function, (iv) refinement of profile parameters, (v) Fourier ...<|separator|>
  10. [10]
    (PDF) Quantitative Phase Analysis Using the Rietveld Method
    Aug 7, 2025 · Subsequently, the quantitative phase analysis by the Rietveld method (Hill and Howard, 1987; Bish and Howard, 1988; O'Connor and Raven, 1988) ...
  11. [11]
    Evaluation of the Rietveld method for determining content and ...
    Dec 1, 2024 · Combined results of Rietveld and XRF analyses can yield accurate results for the chemical composition of iXAMs within a relative error of 10% ...
  12. [12]
    Analyzing Microstructure by Rietveld Refinement | NIST
    Jun 1, 2005 · The review discusses determination of these properties in Rietveld ... crystallite size, diffraction, microstructure, Rietveld refinement, strain.
  13. [13]
    Strain and Grain Size Determination of CeO2 and TiO2 Nanoparticles
    Sep 6, 2021 · With this method, the estimated crystallite size was 23.6 (1) and 18.5 (8) nm, with R2 values of 0.75 and 0.08 for CeO2 and TiO2, respectively.
  14. [14]
    Application of high energy X-ray diffraction and Rietveld refinement ...
    Apr 29, 2023 · This review summarizes the application of high energy XRD and Rietveld refinement in LTMO cathode materials, including the brief introduction of ...
  15. [15]
    Rietveld and pair distribution function analysis using synchrotron X ...
    Jun 20, 2025 · This study employs integrated synchrotron XRD and pair distribution function (PDF) analysis to elucidate the crystal structure of the 1M illite polytype.
  16. [16]
    A Rietveld and an Atomic Pair Distribution Function (PDF) Study of ...
    The structure of Se particles in the ∼13 Å diameter α-cages of zeolite NdY has been determined by Rietveld refinement and pair distribution function (PDF) ...
  17. [17]
    Rietveld refinement in the routine quantitative analysis of famotidine ...
    Accurate, precise and reliable X-ray powder diffraction method was developed for the quantitative determination of famotidine polymorphic forms in their ...Missing: identification | Show results with:identification
  18. [18]
    Applications of Rietveld refinement in Fe–B–Nb alloy structure studies
    The paper presents in a short view the theoretical background of the Rietveld refinement method and some application of it in the structure studies of the ...
  19. [19]
    R factors in Rietveld analysis: How good is good enough?
    Mar 1, 2012 · The least-squares minimization method used for Rietveld allows the statistical uncertainty in the data to be extrapolated to statistical ...Missing: seminal | Show results with:seminal
  20. [20]
    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.
  21. [21]
    [PDF] The use of the Rietveld method and pairs distribution function ... - arXiv
    The PDF functions can give a more accurate insight structure of the effect of high-pressure. We are using the refined structural data for SnSe2 to calculate the.
  22. [22]
    Phase quantification using deep neural network processing of XRD ...
    Aug 12, 2024 · To circumvent the problems of Rietveld refinement and the processing of large datasets, deep learning (DL) methods have emerged as a promising ...
  23. [23]
    Powder diffraction crystal structure determination using generative ...
    Aug 11, 2025 · Although Rietveld refinement generally reduces the RMSE for most structures, a small fraction (<5%) exhibited increased RMSE after refinement ( ...