Next Article in Journal
Shape-Memory Effect of 4D-Printed Gamma-Irradiated Low-Density Polyethylene
Previous Article in Journal
Study on Pulsed Gas Tungsten Arc Lap Welding Techniques for 304L Austenitic Stainless Steel
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Extended Caking Method for Strain Analysis of Polycrystalline Diffraction Debye–Scherrer Rings

1
Department of Engineering Science, University of Oxford, Oxford OX1 3PJ-1, UK
2
I15 Synchrotron Beamline, Diamond Light Source, Didcot OX11 0DE-2, UK
3
Department of Engineering, University of Cambridge, Cambridge CB2 1PZ-3, UK
*
Author to whom correspondence should be addressed.
Crystals 2024, 14(8), 716; https://doi.org/10.3390/cryst14080716 (registering DOI)
Submission received: 16 July 2024 / Revised: 4 August 2024 / Accepted: 7 August 2024 / Published: 9 August 2024

Abstract

:
Polycrystalline diffraction is a robust methodology employed to assess elastic strain within crystalline components. The Extended Caking (exCaking) method represents a progression of this methodology beyond the conventional azimuthal segmentation (Caking) method for the quantification of elastic strains using Debye–Scherrer 2D X-ray diffraction rings. The proposed method is based on the premise that each complete diffraction ring contains comprehensive information about the complete elastic strain variation in the plane normal to the incident beam, which allows for the introduction of a novel algorithm that analyses Debye–Scherrer rings with complete angular variation using ellipse geometry, ensuring accuracy even for small eccentricity values and offering greater accuracy overall. The console application of the exCaking method allows for the accurate analysis of polycrystalline X-ray diffraction data according to the up-to-date rules presented in the project repository. This study presents both numerical and empirical examinations and error analysis to substantiate the method’s reliability and accuracy. A specific validation case study is also presented to analyze the distribution of residual elastic strains in terms of force balance in a Ti-6Al-4V titanium alloy bar plastically deformed by four-point bending.

1. Introduction

The experiment conducted in 1912 by Friedrich and Knipping under the supervision of Laue [1] using a Crookes tube, a single crystal, and a photographic plate resulted in the discovery of X-ray diffraction patterns [2]. This breakthrough in radio crystallography enabled the measurement of X-ray wavelengths and the determination of crystal structures based on observed diffraction patterns. Further studies by Laue demonstrated that the diffraction spots are distributed along conic curves [3]. Later, W.L. Bragg analyzed Laue’s results and proposed the Bragg configuration, which involved the reflection of X-ray beams off crystal planes at a specific angle, known as twice the Bragg angle [4]. This innovative setup, differing from Laue’s, allowed for the measurement of diffracted beams on the same side as the X-ray source, proving Bragg’s theoretical concept and paving the way for advancements in X-ray diffraction studies. Debye and Scherrer also expanded X-ray diffraction studies by exploring diffraction patterns from polycrystalline samples, marking a departure from single-crystal analysis [5]. By irradiating cylindrical samples and measuring diffraction arcs, they applied the Bragg relation to estimate interplanar distances, broadening the applicability of X-ray diffraction beyond single crystals to more prevalent materials.
Synchrotron X-ray scattering [6,7,8] from a polycrystalline aggregate [9] forms a system of coaxial cones of nearly fixed semi-angles, with each cone corresponding to a particular peak with Miller indices ( h k l ) [10,11] that is found on the equivalent 1D powder diffraction pattern [12] for the same material. If a 2D pixelated detector [13] is placed normally to the incident beam, it captures the scattered beam intensity in the perpendicular plane in the form of a concentric pattern that consists of so-called Debye–Scherrer rings [14,15,16,17,18]. For a powder of entirely random orientation, these rings are truly circular [18,19]. If the detector plane is not normal to the incident beam, the conical section produced has the form of an ellipse, with the primary beam passing through one of the foci. If a mechanically loaded or residually stressed polycrystalline specimen [20,21,22,23] that contains internal elastic strains [24,25,26,27] is considered, the ring also becomes distorted into an ellipse, but in contrast to the previous case, this time the incident beam passes through the midpoint between foci and the detector plane is normal to the incident beam, i.e., centrally. Furthermore, the ellipticity parameter due to elastic strain does not deviate from unity by more than 0.02 [28,29] since elastic strain in crystalline metals and ceramics typically does not exceed 1%. Distinguishing between these two types of distortions could present a problem, but luckily the first type (geometric distortion) [30] leads to a radial distance variation of the type sin ( ϕ ) with an azimuthal angle ϕ , whilst the second type (strain distortion) [31] leads to an azimuthal variation given by sin 2 ϕ due to the absence of positive/negative directional dependence in the definition of strain. Therefore, the determination of precise experimental geometry (detector inclination) [32] and complete 2D variation in strain requires simultaneous refinement of the entire ring (or system of rings) [33] in a manner not entirely dissimilar to Fourier analysis [15]. The quantification of elastic strains for calculating residual stresses constitutes a significant proportion of polycrystalline diffraction applications [28,34]. This application is built on the premise that mechanical stress induces changes in the crystal lattice interplanar distances [35,36]. The fundamental relation between X-ray scattering geometry and crystal lattice spacing encapsulated in Bragg’s Law [37,38,39] facilitates the discernment of small alterations in lattice spacings [40,41]. Various diffraction techniques [42] that make use of this law include neutron [43,44,45,46,47,48,49,50,51,52], conventional laboratory X-ray [37,53,54,55,56,57], and synchrotron X-ray [58,59,60,61,62,63,64,65,66,67,68]. These observations prompted the use of the entirety of the Debye–Scherrer ring pattern as a single dataset serving as input to ellipse fitting [16,69,70]. The problem is somewhat complicated by the empirical evidence of the additional distortive influences linked to the topographical configuration of the specimen’s surface and variations in pixel dimensions [71].
In various industries, highly accurate measurements of residual stresses are crucial for ensuring the safety, reliability, and optimal performance of materials and structures [72,73,74,75,76]. While determining the distribution of residual stresses provides valuable insights, the precision offered by accurate measurements [77] is paramount in design, manufacturing, and analysis processes [78]. This precision is critical for quality assurance [79,80], as it helps identify and mitigate potential defects or failures. In safety-critical applications like aerospace [81,82], ultra-supercritical power plants [83,84,85,86], and automotive engineering [87], even minor variations in residual stress levels can have significant consequences, making highly accurate measurements essential [88]. Additionally, precise data aid in the optimization of manufacturing processes [89] and the validation of numerical models [90] and contribute to materials research and development. Accurate measurements serve as a cornerstone for cross-verifying results, standardizing testing procedures, and enhancing overall consistency in the industry [91]. Ultimately, the pursuit of highly accurate measurements is crucial in advancing the understanding and control of residual stresses [92,93], impacting the entire lifecycle of materials and structures [94]. The Caking method introduced by Korsunsky et al. [95] emerged as a practical and viable strategy involving the averaging of deviations of the Debye–Scherrer rings from their reference morphology [96] for a number of chosen azimuthal angular ranges to satisfy the requirement for quantifying residual stresses with high accuracy.
The Caking method has proved its value and reliability over the last three decades [97,98,99,100,101,102,103,104,105,106]. However, it relies on the conversion of 2D detector data via radial–azimuthal binning for chosen sectors separately [107,108] without attempting a full fitting of the ellipse due to its extremely small eccentricity [101,109]. Once the radial positions of the rings are known, this data may be interpreted in connection with the full 2D elastic strain variation as an additional analysis step. In contrast, the present research introduces a novel algorithm that performs a complete angular variation analysis of Debye–Scherrer rings by exploiting the geometric properties of an ellipse, maintaining accuracy even for small eccentricity values. The advantages of this approach include greater accuracy achieved via a faster, single-step analysis. In this study, numerical analyses demonstrate the error reduction achieved through the employment of the newly proposed algorithm. For validation, a case study was conducted using synchrotron X-ray diffraction data obtained from a Ti-6Al-4V alloy subjected to elastic-plastic four-point bending. This examination illustrates the accuracy and reliability of the exCaking method compared to conventional Caking utilizing a custom MATLAB code and GSAS-II ellipse-fitting analysis. This presentation also includes a comparison of force balances that validates the improvement in accuracy when compared to other methods.

2. Methodology

The Caking method evaluates elastic strains along a chosen direction through the determination of radial displacement of the peak center in relation to a reference position. This methodology involves polar rebinning of pixel intensity data into angular bins relative to the pattern center within an azimuthal angle range of ϕ at a transformation angle of ϕ T , as depicted in Figure 1a. Conversely, the exCaking method, which is formulated as an extension of the Caking method, utilizes the trigonometric averaging of the x - and y -components of distance from the ring center within the azimuthal span range of 2 π at a transformation angle of ϕ T , as depicted in Figure 1b. It is important to note that the azimuthal angle step size necessitates adjustment in accordance with the observed data scatter. This formulation allows extraction of the information about the elastic strain components from the whole ring in spite of a limited angular range.
The reference position, which is the center of the Debye–Scherrer 2D X-ray diffraction rings, is determined using calibration rings that are fitted into a circle. Subsequent to the determination of the coordinates of the center of the beam, the radial distance of peaks from the center, denoted as D ϕ at angles denoted as ϕ varying in a range of 360 degrees with a predetermined angle step size, is determined through Gaussian peak fitting, as depicted in Figure 2a. Finally, trigonometric averages of the x - and y -components of the radial distance from the predetermined ring center, denoted as D x and D y , are computed at a transformation angle of ϕ T utilizing Equations (1) and (2).
D x = 0 2 π D ϕ cos ϕ d ϕ / 0 2 π cos ϕ d ϕ = 0 2 π D x ϕ d ϕ / 0 2 π cos ϕ d ϕ
D y = 0 2 π D ϕ sin ϕ d ϕ / 0 2 π sin ϕ d ϕ = 0 2 π D y ϕ d ϕ / 0 2 π sin ϕ d ϕ
Trigonometric averaging is applied across the entire range of 360 degrees, encompassing the complete angular variation in the Debye–Scherrer ring. This comprehensive angular coverage ensures that data from all points around the ring contribute to the averaging calculations. As a result, any error in the determination of the coordinates of the ring’s center is rendered negligible in its influence on the final averaged values. This phenomenon can be understood from a mathematical perspective. Trigonometric averaging involves integrating the radial distances from the ring’s center over the full angular range. Since the integration process considers all points evenly distributed around the ring, errors in individual coordinate measurements are effectively distributed and balanced out over the entire range. Consequently, any inaccuracies in determining the center coordinates have a minimal impact on the overall averaging calculations, leading to robust and reliable results.
Based on empirical analysis utilizing the semi-minor and semi-major axes of the artificially strained ring illustrated in Figure 2b, trigonometric averaging values of distances corresponding to the transformed coordinate axes are updated to D x and D y , as outlined in Equations (3) and (4).
D x = 2 D x D y
D y = 2 D y D x
For monochromatic (fixed wavelength) diffraction, Bragg’s Law leads to the relation between elastic strain (relative lattice spacing change) and the scattering angle increment given by ε = δ d / d = cot θ δ θ . For the sample–detector distance L , the distance from the ring center is set by tan 2 θ = D / L . The variation in the scattering angle, 2 δ θ , is linked to the change in the distance from the ring center, δ D , as 2 δ θ / sin 2 2 θ = δ D / L . Hence, tan 2 θ δ D / D = 2 δ θ / sin 2 2 θ , and ε = ( cot θ sin 4 θ ) ( δ D / D ) . The geometric factor that appears above remains remarkably close to unity for small scattering angles. For the experimental parameters considered in the present study, ε = 0.993   δ D / D , i.e., <1% difference. Furthermore, the variation in the geometry factor as a function of the Bragg angle is exemplified in Figure 2c. In the limit of θ 0 , the multiplying factor converges to unity, so for small scattering angles, the following approximation is found to hold: ε = δ D / D . Finally, the quantification of the elastic strain’s x x - and y y -components at a transformation angle of ϕ T hinges upon the utilization of Equation (5), which employs the stress-free reference distance from the center, denoted as D 0 .
ε x x , y y D 0 D x , y / D 0
Numerical analyses were performed to determine elastic strains as a function of azimuthal angles over a range from 0 to 2 π , with an azimuthal angle step size of π /180. The reference unstrained ring was determined to have a radius of 100 pixels. For the ‘artificially strained’ ring, the semi-minor axis of the ellipse was set to 99 pixels, while the semi-major axis was set to 102 pixels. Numerical analysis involved calculating the direct elastic strain components within a transformation range from 0 to 2 π , with a step size of π /180 for both the exCaking and Caking methods. Concurrently, an azimuthal step size of π /180 was determined for the polar rebinning of the Caking method and for the trigonometric averaging of the exCaking method. The Caking method was applied according to the guidelines outlined in the study that introduced this method [95], and the Caking range was set to 15 π /180.
Debye–Scherrer data were collected from the middle segment of this titanium bar around the symmetry axis of four-point bending over a grid of 9 columns and 51 rows, with step sizes of 1 and 0.1 mm along the x - and y -axes, respectively, as illustrated in Figure 3b. The baseline state of the titanium bar’s ring structure was established through the determination of the mean D x and D y values calculated at an azimuthal angle of zero for all experimental measurements. A synchrotron X-ray beam with a photon energy of 72 keV was collimated to a beam spot size of 76 × 115 µm and a 2.856-degree diffraction angle. Diffraction patterns were recorded using the 2D detector Pilatus P3-2M (Dectris, Switzerland), with a matrix of 1679 × 1475 pixels.
The experimental assessment of elastic strain quantification methods followed the same transformation and azimuthal angle step sizes as determined in the numerical analysis using the selected ring illustrated in Figure 3a. For this purpose, a Debye–Scherrer ring with a Miller index of ( 102 ), indicated by the red arrows in Figure 3a, was obtained using synchrotron X-ray diffraction scanning of a four-point bent titanium alloy specimen and selected for experimental analysis of the Caking and exCaking methods. The selection of this ring was based on the sensitivity of lattice planes with low Miller indices to small changes in lattice spacing, making them suitable for detecting subtle strains within the crystal lattice. The incomplete and missing data on the ring were omitted after identifying them as outliers using standard deviation-based error analysis according to the three-sigma rule.
The bending of this specimen was performed to induce plastic deformation in regions close to the faces normal to the y -axis while keeping the central region free of plastic deformation. The resulting x x - and y y -components of elastic strain in the Cartesian coordinate system of both techniques within the selected ring were fitted against the transformation function formulated in Equation (6) and compared with calculations at transformation angles.
ε ( ϕ T ) = ε x x + ε y y 2 + ε x x ε y y 2 cos ( 2 ϕ T )
Residual stresses distributed over the measurement grid on the four-point bent titanium alloy bar were calculated using the planar stress assumption in the Cartesian coordinate system using Equation (7), with Young’s modulus ( E ) and Poisson’s ratio ( v ) set to 115 GPa and 0.34 [110].
σ x x , y y = E ε x x , y y + v ε y y , x x 1 v 2
The console application of the exCaking method was developed using MATLAB R2022a, a proprietary, multi-paradigm programming language and numeric computing environment developed by MathWorks. This application enables the accurate determination of trigonometric averaging values of distances corresponding to the transformed coordinate axes in the updated form. The flowchart of the application given in Figure 2d begins with the calibration step, the aim of which is to determine the center of the Debye–Scherrer ring. This stage is recommended to be completed using a calibration image of Debye–Scherrer rings obtained from the same setup of polycrystalline X-ray diffraction scans of the specimen. Following the calibration step, the user selects the rings to be analyzed from any image of the Debye–Scherrer ring data collected from the specimen. Finally, the solver of the console application is initiated, and the results are stored in plain text formats. The project repository contains the up-to-date rules of the exCaking console application.

3. Results

Polar representations depicted in Figure 4a demonstrate that the exCaking method yields direct elastic strain components for artificially strained rings with an error not exceeding 0.04%. By contrast, quantification via the Caking method results in errors greater than 7.3% for the y y -component and exceeding 3.5% for the x x -component within the same strained ring, as depicted in Figure 4b. These observations illustrate that the error percentage of the Caking method increases with diminishing magnitudes. Additionally, the highest error is noted in the x x - and y y -components of elastic strain in the Cartesian coordinate system aligned with the relative lattice rotation, while the error diminishes at intermediate angles.
Evident from the outcomes depicted in Figure 4c, the Caking method displayed a distribution characterized by noise, whereas the exCaking method exhibited a notably smoother polar distribution. The fitting mean error for the exCaking method was substantially lower in comparison to the Caking method’s solution, depicted in Figure 4d. In congruence with the numerical analysis, while the exCaking fitting error was anticipated to be under 0.04%, practical errors stemming from data collection procedures introduced deviations in the determination of distances from the center positions on the ring, thus amplifying the fitting error. Consequently, the mean error, quantified at 1.637% for the exCaking method, remains variable across different Debye–Scherrer ring data; nonetheless, the Caking method consistently yields a notably higher mean error of 13.071% in comparison to the numerical analysis-derived errors of the x x - and y y -components of elastic strain in the Cartesian coordinate system. Although both methods exhibit analogous distribution trends, their magnitudes diverge, culminating in the deduction that the exCaking method offers a substantively more accurate and reliable measure of elastic strain in contrast to the Caking method.
This case study, which involved a comprehensive grid, facilitated an assessment of residual elastic strain distribution in a titanium alloy bar after it underwent plastic four-point bending using both the exCaking and Caking methods. Elastic strains averaged across grid rows, as depicted in Figure 4e, reveal that the exCaking outcomes exhibit a smoother distribution closely resembling the theoretically anticipated distribution of bending residual elastic strains. By contrast, the Caking method, as depicted in Figure 4f, yields a discordant distribution characterized by noise and distortion attributable to errors elucidated via numerical and experimental analyses. Subsequent quantitative examination of the same dataset utilizing GSAS-II crystallography data analysis software [111] also involved ellipse fitting.
Eigenstrain reconstruction of residual elastic strains in a similar specimen of the same alloy [112] further corroborated the distribution of residual elastic strain due to elastic-plastic four-point bending. This investigation also revealed that the averages of the x x -components of bending residual stresses, which were distributed within the diffraction map grid, were determined to be −0.199, −13.404, and 2.231 MPa for the exCaking, Caking, and ellipse-fitting solutions, respectively. The relationship between force balance and the accuracy of strain quantification methods is integral to understanding the reliability of residual stress assessments in materials. Force balance, representing the equilibrium of internal stresses within a specimen, is directly influenced by the accuracy of residual elastic strain measurements. This study demonstrates that the exCaking method consistently provides more accurate results compared to the traditional Caking method. This higher accuracy translates to a more precise determination of residual stresses, thereby impacting the overall force balance within the material. Consequently, it can be concluded that the exCaking method provides the most reliable solution in terms of force balance. The high deviation from force balance achieved by the Caking method can be attributed to the elevated error range observed in the numerical experiments, while the low deviation from force balance of the exCaking method is consistent with the low error range presented in the polar representations shown in Figure 4a–d.
The exCaking method represents a significant advancement in polycrystalline diffraction analysis, particularly in the quantification of elastic strain within crystalline materials. This method extends beyond the conventional Caking approach by utilizing complete angular variation analysis of Debye–Scherrer rings, leveraging ellipse geometry properties for accurate strain determination. The exCaking algorithm ensures precision even for small eccentricity values and offers a faster, single-step analysis, enhancing overall accuracy. The numerical and experimental examinations presented in this study validated the reliability and accuracy of the exCaking method. Specifically, the numerical analyses demonstrated a reduction in error compared to the Caking method, showcasing the efficacy of the new algorithm. Additionally, a validation case study conducted using a titanium alloy bar subjected to elastic-plastic four-point bending confirmed the accuracy and reliability of the exCaking method. The comparison between the exCaking and Caking methods revealed that exCaking produced smoother polar distributions and lower fitting errors, indicating superior performance. The accuracy of the exCaking method underscores the notion that individual points across Debye–Scherrer rings contain valuable data about elastic strain components in polycrystalline materials. By extracting this information through trigonometric averaging, the exCaking method achieved increased accuracy in strain quantification. Experimental analyses further validated the method’s effectiveness, highlighting its potential for various applications requiring precise determination of residual elastic strains and stresses. Overall, the exCaking method offers a more coherent and accurate approach to strain analysis compared to conventional methods, making it a preferred choice in scenarios demanding high accuracy and reliability.

4. Conclusions

A comparative analysis of the exCaking method was undertaken through both numerical and experimental analyses. The numerical investigations revealed that the exCaking method led to negligible errors in quantifying the x x - and y y -components of elastic strains at transformation angles of ϕ T within the artificially strained ring. By stark contrast, the Caking method resulted in quantifications with a high error within the same ring. The effectiveness of the exCaking method substantiates the assertion that individual points across Debye–Scherrer rings harbour valuable data regarding the components of elastic strain within polycrystalline materials. Consequently, extracting this data through trigonometric averaging yields increased accuracy. Experimental analysis corroborated the numerical findings, as both methods were evaluated under the same conditions. The exCaking method exhibited smoother polar distributions, consistently outperforming the significantly higher mean error associated with the Caking method. The subsequent application of both methods to a titanium alloy bar subjected to four-point bending confirmed the exCaking method’s capacity to produce a more coherent and accurate distribution of residual elastic strains compared to the Caking and ellipse-fitting methods. The results of the case study also demonstrate that all three methods were successful in determining the expected distribution of bending residual stress but with varying deviations from force balance. Accordingly, the exCaking method and its console application (see Supplementary Materials) in the Supplementary Materials should be preferred in cases where an accurate determination of residual elastic strains and residual stresses is needed.

Supplementary Materials

The following supporting information can be downloaded at: https://github.com/fatihxuzun/exCaking (accessed on 6 August 2024), up-to-date versions of the MATLAB console application of the exCaking method and the documentation.

Author Contributions

Conceptualization, F.U.; methodology, F.U.; software, F.U.; validation, F.U.; formal analysis, F.U.; investigation, F.U., D.D., K.L., Z.I.W., J.C. and C.B.; resources, A.M.K.; data curation, F.U.; writing—original draft preparation, F.U.; writing—review and editing, F.U. and A.M.K.; visualization, F.U.; supervision, A.M.K. and F.U.; project administration, A.M.K.; funding acquisition, A.M.K. All authors have read and agreed to the published version of the manuscript.

Funding

This work, entitled Rich Nonlinear Tomography for Advanced Materials (EP/V007785/1), was funded by The Engineering and Physical Sciences Research Council (EPSRC).

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Acknowledgments

The authors acknowledge the support from EPSRC project EP/V007785/1 “Rich Nonlinear Tomography for advanced materials”, EuroHPC project grant EHPC-DEV-2022D10-054 for allowing the simulations to be performed on the Luxembourg national supercomputer MeluXina and are also grateful to the LuxProvide teams for their expert support. The authors thank Diamond Light Source (Didcot, UK) for beamtime allocation CY30712-2 to I15 beamline.

Conflicts of Interest

Author Dominik Daisenberger was employed by the company Diamond Light Source. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Authier, A. Early Days of X-ray Crystallography; OUP Oxford: Oxford, UK, 2013. [Google Scholar] [CrossRef]
  2. Friedrich, W.; Knipping, P.; Laue, M. Interferenzerscheinungen Bei Röntgenstrahlen. Ann. Phys. 1913, 346, 971–988. [Google Scholar] [CrossRef]
  3. Laue, M. Eine Quantitative Prüfung Der Theorie Für Die Interferenzerscheinungen Bei Röntgenstrahlen. Ann. Phys. 1913, 346, 989–1002. [Google Scholar] [CrossRef]
  4. Bragg, W.L. The Specular Reflection of X-rays. Nature 1912, 90, 410. [Google Scholar] [CrossRef]
  5. Debye, P.; Scherrer, P. Interferenz an Regellos Orientierten Teilchen Im Röntgenlicht I. Phys. Z. 1916, 17, 277. [Google Scholar]
  6. Kohara, S.; Ohara, K.; Tajiri, H.; Song, C.; Sakata, O.; Usuki, T.; Benino, Y.; Mizuno, A.; Masuno, A.; Okada, J.T.; et al. Synchrotron X-ray Scattering Measurements of Disordered Materials. Z. Phys. Chem. 2016, 230, 339–368. [Google Scholar] [CrossRef]
  7. Li, Z.; Wu, Z.; Mo, G.; Xing, X.; Liu, P. A Small-Angle x-Ray Scattering Station at Beijing Synchrotron Radiation Facility. Instrum. Sci. Technol. 2014, 42, 128–141. [Google Scholar] [CrossRef]
  8. Reinartz, I.; Sarter, M.; Otten, J.; Höfig, H.; Pohl, M.; Schug, A.; Stadler, A.M.; Fitter, J. Structural Analysis of a Genetically Encoded Fret Biosensor by Saxs and Md Simulations. Sensors 2021, 21, 4144. [Google Scholar] [CrossRef] [PubMed]
  9. Bernier, J.V.; Park, J.S.; Pilchak, A.L.; Glavicic, M.G.; Miller, M.P. Measuring Stress Distributions in Ti-6Al-4V Using Synchrotron X-ray Diffraction. Metall. Mater. Trans. A Phys. Metall. Mater. Sci. 2008, 39, 3120–3133. [Google Scholar] [CrossRef]
  10. Miller, W.H. A Treatise on Crystallography; Online access with subscription: JISC Historical Texts; For J. & J.J. Deighton: Cambridge, UK, 1839. [Google Scholar]
  11. Muslih, M.R.; Nishida, M.; Sugeng, B.; Sadeli, Y. Improvements of the X-ray Diffractometer (XRD) to Become Small Angle X-ray Scattering (SAXS) and Residual Stress Diffractometer. AIP Conf. Proc. 2021, 2381, 020047. [Google Scholar] [CrossRef]
  12. Fauth, F.; Peral, I.; Popescu, C.; Knapp, M. The New Material Science Powder Diffraction Beamline at ALBA Synchrotron. Powder Diffr. 2013, 28, 360–370. [Google Scholar] [CrossRef]
  13. Straas, T.; Becker, J.; Iversen, B.B.; Als-Nielsen, J. The Debye–Scherrer Camera at Synchrotron Sources: A Revisit. J. Synchrotron Radiat. 2012, 20, 98–104. [Google Scholar] [CrossRef]
  14. MacDonald, M.J.; Vorberger, J.; Gamboa, E.J.; Drake, R.P.; Glenzer, S.H.; Fletcher, L.B. Calculation of Debye-Scherrer Diffraction Patterns from Highly Stressed Polycrystalline Materials. J. Appl. Phys. 2016, 119, 215902. [Google Scholar] [CrossRef]
  15. Miyazaki, T.; Fujimoto, Y.; Sasaki, T. Improvement in X-ray Stress Measurement Using Debye-Scherrer Rings by in-Plane Averaging. J. Appl. Crystallogr. 2016, 49, 241–249. [Google Scholar] [CrossRef]
  16. Karamched, P.S.; Xiong, Y.; Nguyen, C.-T.; Collins, D.M.; Magazzeni, C.M.; Wilkinson, A.J. Weighted Ellipse Fitting Routine for Spotty or Incomplete Debye-Scherrer Rings on a 2D Detector. arXiv 2021, arXiv:2110.05467. [Google Scholar]
  17. Sirhindi, R.; Khan, N. Clustering-Based Detection of Debye-Scherrer Rings. J. Comput. Inf. Sci. Eng. 2023, 23, 041013. [Google Scholar] [CrossRef]
  18. Shahzad, S.; Khan, N.; Nawaz, Z.; Ferrero, C. Automatic Debye-Scherrer Elliptical Ring Extraction via a Computer Vision Approach. J. Synchrotron. Radiat. 2018, 25, 439–450. [Google Scholar] [CrossRef] [PubMed]
  19. Li, C.; Xiao, P.; Cernik, R. The Nondestructive Measurement of Strain Distributions in Air Plasma Sprayed Thermal Barrier Coatings as a Function of Depth from Entire Debye-Scherrer Rings Strain Mapping in APS TBCs from Debye-Scherrer Rings. J. Appl. Crystallogr. 2020, 53, 69–75. [Google Scholar] [CrossRef] [PubMed]
  20. Keckes, J.; Bartosik, M.; Daniel, R.; Mitterer, C.; Maier, G.; Ecker, W.; Vila-Comamala, J.; David, C.; Schoeder, S.; Burghammer, M. X-ray Nanodiffraction Reveals Strain and Microstructure Evolution in Nanocrystalline Thin Films. Scr. Mater. 2012, 67, 748–751. [Google Scholar] [CrossRef]
  21. Uzun, F.; Basoalto, H.; Liogas, K.; Fares Slim, M.; Lik Lee, T.; Besnard, C.; Ivan Wang, Z.; Chen, J.; Dolbnya, I.P.; Korsunsky, A.M. Tomographic Eigenstrain Reconstruction for Full-Field Residual Stress Analysis in Large Scale Additive Manufacturing Parts. Addit. Manuf. 2024, 81, 104027. [Google Scholar] [CrossRef]
  22. Akrivos, V.; Smith, M.C.; Muransky, O.; Ohms, C.; Youtsos., A. A residual stress measurement and numerical analysis round robin on a three-pass slot nickel-base repair weld. Procedia Manuf. 2020, 51, 779–786. [Google Scholar] [CrossRef]
  23. Smith, M.C.; Smith, A.C.; Wimpory, R.; Ohms, C. A Review of the NeT Task Group 1 Residual Stress Measurement and Analysis Round Robin on a Single Weld Bead-on-Plate Specimen. Int. J. Press. Vessel. Pip. 2014, 120–121, 93–140. [Google Scholar] [CrossRef]
  24. Statnik, E.S.; Salimon, A.I.; Uzun, F.; Korsunsky, A.M. Polar Transformation of 2D X-ray Diffraction Patterns for 2D Strain Evaluation. Proc. World Congr. Eng. 2019, 2019, 2–6. [Google Scholar]
  25. Uzun, F.; Salimon, A.I.; Statnik, E.S.; Besnard, C.; Chen, J.; Moxham, T.; Salvati, E.; Wang, Z.; Korsunsky, A.M. Polar Transformation of 2D X-ray Diffraction Patterns and the Experimental Validation of the HDIC Technique. Measurement 2019, 151, 107193. [Google Scholar] [CrossRef]
  26. Mukhopadhyay, D. Identifying the Causes of Residual Stress in Polycrystalline Diamond Compact (PDC) Cutters by X-ray Diffraction Technique. Results Mater. 2021, 11, 100216. [Google Scholar] [CrossRef]
  27. Ortiz, A.L.; Tian, J.W.; Villegas, J.C.; Shaw, L.L.; Liaw, P.K. Interrogation of the Microstructure and Residual Stress of a Nickel-Base Alloy Subjected to Surface Severe Plastic Deformation. Acta Mater. 2008, 56, 413–426. [Google Scholar] [CrossRef]
  28. Vorster, W.J.J.; Zhang, S.Y.; Golshan, M.; Laundy, D.; Dini, D.; Korsunsky, A.M. Comparison of X-ray Diffraction Measurement of Residual Elastic Strains: Monochromatic Beam and Image Plate versus White Beam Energy-Dispersive Analysis. J. Strain Anal. Eng. Des. 2007, 42, 23–37. [Google Scholar] [CrossRef]
  29. Higginbotham, A.; McGonegle, D. Prediction of Debye-Scherrer Diffraction Patterns in Arbitrarily Strained Samples. J. Appl. Phys. 2014, 115, 174906. [Google Scholar] [CrossRef]
  30. Smith, N.L.; Coukouma, A.; Dubnik, S.; Asher, S.A. Debye Ring Diffraction Elucidation of 2D Photonic Crystal Self-Assembly and Ordering at the Air-Water Interface†. Phys. Chem. Chem. Phys. 2017, 19, 31813. [Google Scholar] [CrossRef]
  31. Gelfi, M.; Bontempi, E.; Roberti, R.; Depero, L.E. X-ray Diffraction Debye Ring Analysis for Stress Measurement (DRAST): A New Method to Evaluate Residual Stresses. Acta Mater. 2004, 52, 583–589. [Google Scholar] [CrossRef]
  32. Mildner, D.F.R.; Cubitt, R. The Effect of Gravity on the Debye-Scherrer Ring in Small-Angle Neutron Scattering. J. Appl. Crystallogr. 2012, 45, 124–126. [Google Scholar] [CrossRef]
  33. Thompson, P.; Wood, I.G. X-ray Rietveld Refinement Using Debye-Scherrer Geometry. J. Appl. Cryst. 1983, 16, 458–472. [Google Scholar] [CrossRef]
  34. Guinebretière, R. X-ray Diffraction by Polycrystalline Materials; ISTE Ltd: London, UK, 2007; ISBN 9781905209217. [Google Scholar]
  35. Salvalaglio, M.; Voigt, A.; Elder, K.R. ARTICLE Closing the Gap between Atomic-Scale Lattice Deformations and Continuum Elasticity. Npj Comput. Mater. 2019, 5, 48. [Google Scholar] [CrossRef]
  36. Zhang, H.; Sui, T.; Salvati, E.; Daisenberger, D.; Lunt, A.; Fong, K.; Song, X.; Korsunsky, A. Digital Image Correlation of 2D X-ray Powder Diffraction Data for Lattice Strain Evaluation. Materials 2018, 11, 427. [Google Scholar] [CrossRef]
  37. Korsunsky, A.M. A Critical Discussion of the Sin2 ψ Stress Measurement Technique. Mater. Sci. Forum 2008, 571–572, 219–224. [Google Scholar] [CrossRef]
  38. Withers, P.J. Synchrotron X-ray Diffraction. In Practical Residual Stress Measurement Methods; Schajer, G.S., Ed.; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2013; pp. 163–194. ISBN 9781118402832. [Google Scholar]
  39. Hammond, C. X-ray Diffraction of Polycrystalline Materials. In The Basics of Crystallography and Diffraction; Oxford Academic: Oxford, UK, 2015. [Google Scholar]
  40. Dolabella, S.; Borzì, A.; Dommann, A.; Neels, A.; Dolabella, S.; Borzì, A.; Dommann, A.; Neels, A. Lattice Strain and Defects Analysis in Nanostructured Semiconductor Materials and Devices by High-Resolution X-ray Diffraction: Theoretical and Practical Aspects. Small Methods 2022, 6, 2100932. [Google Scholar] [CrossRef] [PubMed]
  41. Farajian, M.; Nitschke-Pagel, T.; Wimpory, R.C.; Hofmann, M.; Klaus, M. Residual Stress Field Determination in Welds by Means of X-ray, Synchrotron and Neutron Diffraction (Ermittlung Des Schweißeigenspannungsfeldes Mittels Röntgen-, Synchrotron-Und Neutronenbeugungsverfahren). Materialwiss. Werkstofftech. 2011, 42, 996–1002. [Google Scholar] [CrossRef]
  42. Maurya, P.; Kota, N.; Gibmeier, J.; Wanner, A.; Roy, S. Review on Study of Internal Load Transfer in Metal Matrix Composites Using Diffraction Techniques. Mater. Sci. Eng. A 2022, 840, 142973. [Google Scholar] [CrossRef]
  43. Huang, Q.; Shi, R.; Muránsky, O.; Beladi, H.; Kabra, S.; Schimpf, C.; Volkova, O.; Biermann, H.; Mola, J. Neutron Diffraction Analysis of Stress and Strain Partitioning in a Two-Phase Microstructure with Parallel-Aligned Phases. Sci. Rep. 2020, 10, 13536. [Google Scholar] [CrossRef]
  44. Uzun, F.; Papadaki, C.; Wang, Z.; Korsunsky, A.M. Neutron Strain Scanning for Experimental Validation of the Artificial Intelligence Based Eigenstrain Contour Method. Mech. Mater. 2020, 143, 103316. [Google Scholar] [CrossRef]
  45. Wimpory, R.C.; Ohms, C.; Hofmann, M.; Schneider, R.; Youtsos, A.G. Statistical Analysis of Residual Stress Determinations Using Neutron Diffraction. International J. Press. Vessel. Pip. 2009, 86, 48–62. [Google Scholar] [CrossRef]
  46. Woo, W.; An, G.B.; Kingston, E.J.; Dewald, A.T.; Smith, D.J.; Hill, M.R. Through-Thickness Distributions of Residual Stresses in Two Extreme Heat-Input Thick Welds: A Neutron Diffraction, Contour Method and Deep Hole Drilling Study. Acta Mater. 2013, 61, 3564–3574. [Google Scholar] [CrossRef]
  47. Brown, D.W.; Bernardin, J.D.; Carpenter, J.S.; Clausen, B.; Spernjak, D.; Thompson, J.M. Neutron Diffraction Measurements of Residual Stress in Additively Manufactured Stainless Steel. Mater. Sci. Eng. A 2016, 678, 291–298. [Google Scholar] [CrossRef]
  48. Pratihar, S.; Turski, M.; Edwards, L.; Bouchard, P.J. Neutron Diffraction Residual Stress Measurements in a 316L Stainless Steel Bead-on-Plate Weld Specimen. Int. J. Press. Vessel. Pip. 2009, 86, 13–19. [Google Scholar] [CrossRef]
  49. Jiang, W.; Woo, W.; An, G.-B.; Park, J.-U. Neutron Diffraction and Finite Element Modeling to Study the Weld Residual Stress Relaxation Induced by Cutting. Mater. Des. 2013, 51, 415–420. [Google Scholar] [CrossRef]
  50. Wang, Z.; Denlinger, E.; Michaleris, P.; Stoica, A.D.; Ma, D.; Beese, A.M. Residual Stress Mapping in Inconel 625 Fabricated through Additive Manufacturing: Method for Neutron Diffraction Measurements to Validate Thermomechanical Model Predictions. Mater. Des. 2017, 113, 169–177. [Google Scholar] [CrossRef]
  51. Akrivos, V.; Wimpory, R.C.; Hofmann, M.; Stewart, B.; Muransky, O.; Smith, M.C.; Bouchard, J. Neutron Diffraction Measurements of Weld Residual Stresses in Three-Pass Slot Weld (Alloy 600/82) and Assessment of the Measurement Uncertainty. J. Appl. Crystallogr. 2020, 53, 1181–1194. [Google Scholar] [CrossRef]
  52. Kelleher, J.; Prime, M.B.; Buttle, D.; Mummery, P.M.; Webster, P.J.; Shackleton, J.; Withers, P.J. The Measurement of Residual Stress in Railway Rails by Diffraction and Other Methods. J. Neutron Res. 2003, 11, 187–193. [Google Scholar] [CrossRef]
  53. Azanza Ricardo, C.L.; D’Incau, M.; Scardi, P. Revision and Extension of the Standard Laboratory Technique for X-ray Diffraction Measurement of Residual Stress Gradients. J. Appl. Crystallogr. 2007, 40, 675–683. [Google Scholar] [CrossRef]
  54. Pineault, J.A.; Belassel, M.; Brauss, M.E. X-ray Diffraction Residual Stress Measurement in Failure Analysis. In Failure Analysis and Prevention; ASM International: Detroit, MN, USA, 2002; pp. 484–497. [Google Scholar] [CrossRef]
  55. Korsunsky, A.M.; Brandt, L.R. The Effect of Deposition Parameters on the Mechanical and Transport Properties in Nanostructured Cu/W Multilayer Coatings. In Functional Thin Films Technology; CRC Press: Boca Raton, FL, USA, 2021; pp. 287–318. ISBN 9780367541774. [Google Scholar]
  56. Macdonald, C.A. Structured X-ray Optics for Laboratory-Based Materials Analysis. Annu. Rev. Mater. Res. 2017, 47, 115–134. [Google Scholar] [CrossRef]
  57. Fultz, B.; Howe, J.M. Diffraction and the X-ray Powder Diffractometer. In Transmission Electron Microscopy and Diffractometry of Materials; Springer: Berlin/Heidelberg, Germany, 2001; pp. 1–61. [Google Scholar] [CrossRef]
  58. Xu, L.; Zhang, S.Y.; Sun, W.; McCartney, D.G.; Hyde, T.H.; James, J.; Drakopoulos, M. Residual Stress Distribution in a Ti-6Al-4V T-Joint Weld Measured Using Synchrotron X-ray Diffraction. J. Strain Anal. Eng. Des. 2015, 50, 445–454. [Google Scholar] [CrossRef]
  59. Korsunsky, A.M.; Regino, G.M.; Latham, D.P.; Li, H.Y.; Walsh, M.J. Residual Stresses in Rolled and Machined Nickel Alloy Plates: Synchrotron X-ray Diffraction Measurement and Three-Dimensional Eigenstrain Analysis. J. Strain Anal. Eng. Des. 2007, 42, 1–12. [Google Scholar] [CrossRef]
  60. Korsunsky, A.M.; James, K.E. Residual Stresses in an Induction Hardened Gear Tooth Mapped by Synchrotron X-ray Diffraction. J. Neutron Res. 2003, 11, 241–245. [Google Scholar] [CrossRef]
  61. Uzun, F.; Basoalto, H.; Liogas, K.; Chen, J.; Dolbnya, I.P.; Ivan, Z.; Korsunsky, A.M. Voxel-Based Full-Field Eigenstrain Reconstruction of Residual Stresses in Additive Manufacturing Parts Using Height Digital Image Correlation. Addit. Manuf. 2023, 77, 103822. [Google Scholar] [CrossRef]
  62. Korsunsky, A.M. Variational Eigenstrain Analysis of Synchrotron Diffraction Measurements of Residual Elastic Strain in a Bent Titanium Alloy Bar. J. Mech. Mater. Struct. 2006, 1, 259–277. [Google Scholar] [CrossRef]
  63. Capria, E.; Ciuffini, A.; Drnec, J.; De Nolf, W.; Fares-Slim, M.; Frey, J.; Hinrichsen, B.; Honkimäki, V.; Levantino, M.; Mathon, O.; et al. Adapting the European Synchrotron to Industry. Synchrotron Radiat. News. 2024, 37, 10–15. [Google Scholar] [CrossRef]
  64. Fauth, F.; Boer, R.; Gil-Ortiz, F.; Popescu, C.; Vallcorba, O.; Peral, I.; Fullà, D.; Benach, J.; Juanhuix, J. The Crystallography Stations at the Alba Synchrotron. Eur. Phys. J. Plus. 2015, 130, 160. [Google Scholar] [CrossRef]
  65. Korsunsky, A.M.; Collins, S.P.; Owen, R.A.; Daymond, M.R.; Achtioui, S.S.S.; James, K.E.; Alexander Owen, R.; Daymond, M.R.; Achtioui, S.S.S.; James, K.E.; et al. Fast Residual Stress Mapping Using Energy-Dispersive Synchrotron X-ray Diffraction on Station 16.3 at the SRS. J. Synchrotron. Radiat. 2002, 9, 77–81. [Google Scholar] [CrossRef] [PubMed]
  66. Malmelöv, A.; Hassila, C.J.; Fisk, M.; Wiklund, U.; Lundbäck, A. Numerical Modeling and Synchrotron Diffraction Measurements of Residual Stresses in Laser Powder Bed Fusion Manufactured Alloy 625. Mater. Des. 2022, 216, 110548. [Google Scholar] [CrossRef]
  67. Valentine, M.D.A.; Dhokia, V.; Flynn, J.; McNair, S.A.M.; Lunt, A.J.G. Characterisation of Residual Stresses and Oxides in Titanium, Nickel, and Aluminium Alloy Additive Manufacturing Powders via Synchrotron X-ray Diffraction. Mater. Today Commun. 2023, 35, 105900. [Google Scholar] [CrossRef]
  68. Daniels, J.E.; Drakopoulos, M. High-Energy X-ray Diffraction Using the Pixium 4700 Flat-Panel Detector. J. Synchrotron Radiat. 2009, 16, 463–468. [Google Scholar] [CrossRef]
  69. Hart, M.L.; Drakopoulos, M. Weighted Least Squares Fit of an Ellipse to Describe Complete or Spotty Diffraction Rings on a Planar 2D Detector. arXiv 2013, arXiv:1311.5430. [Google Scholar] [CrossRef]
  70. Kanatani, K.; Rangarajan, P. Hyper Least Squares Fitting of Circles and Ellipses. Comput. Stat. Data Anal. 2011, 55, 2197–2208. [Google Scholar] [CrossRef]
  71. Withers, P.J.; Preuss, M.; Webster, P.J.; Hughes, D.J.; Korsunsky, A.M. Residual Strain Measurement by Synchrotron Diffraction. Mater. Sci. Forum 2002, 404-407, 1–12. [Google Scholar] [CrossRef]
  72. Korsunsky, A.M. A Teaching Essay on Residual Stresses and Eigenstrains; Butterworth-Heinemann: Oxford, UK, 2017; ISBN 978-0-12-810990-8. [Google Scholar]
  73. Nelson, D.V.; Ricklefs, R.E.; Evans, W.P. Residual Stresses in Quenched and Tempered Plain Carbon Steels. SAE Tech. Pap. 1971, 12, 5–9. [Google Scholar] [CrossRef]
  74. Uzun, F.; Korsunsky, A.M. The Use of Eigenstrain Theory and Fuzzy Techniques for Intelligent Modeling of Residual Stress and Creep Relaxation in Welded Superalloys. Mater. Today Proc. 2020, 33, 1880–1883. [Google Scholar] [CrossRef]
  75. Uzun, F.; Korsunsky, A.M. The OxCM Contour Method Solver for Residual Stress Evaluation. Eng. Comput. 2024. [Google Scholar] [CrossRef]
  76. Uzun, F.; Korsunsky, A.M. Voxel-Based Full-Field Eigenstrain Reconstruction of Residual Stresses. Adv. Eng. Mater. 2023, 25, 2201502. [Google Scholar] [CrossRef]
  77. Uzun, F.; Korsunsky, A.M. On the Analysis of Post Weld Heat Treatment Residual Stress Relaxation in Inconel Alloy 740H by Combining the Principles of Artificial Intelligence with the Eigenstrain Theory. Mater. Sci. Eng. A 2019, 752, 180–191. [Google Scholar] [CrossRef]
  78. Soyama, H.; Korsunsky, A.M. A Critical Comparative Review of Cavitation Peening and Other Surface Peening Methods. J. Mater. Process. Tech. 2022, 305, 117586. [Google Scholar] [CrossRef]
  79. Everton, S.K.; Hirsch, M.; Stavroulakis, P.I.; Leach, R.K.; Clare, A.T. Review of In-Situ Process Monitoring and in-Situ Metrology for Metal Additive Manufacturing. Mater. Des. 2016, 95, 431–445. [Google Scholar] [CrossRef]
  80. Wang, Z.; Chen, J.; Magdysyuk, O.V.; Uzun, F.; Korsunsky, A.M. Ultra-Fast Quantification of Polycrystalline Texture via Single Shot Synchrotron X-ray or Neutron Diffraction. Mater. Charact. 2022, 186, 111827. [Google Scholar] [CrossRef]
  81. Salvati, E.; Lunt, A.J.G.; Ying, S.; Sui, T.; Zhang, H.J.; Heason, C.; Baxter, G.; Korsunsky, A.M. Eigenstrain Reconstruction of Residual Strains in an Additively Manufactured and Shot Peened Nickel Superalloy Compressor Blade. Comput. Methods Appl. Mech. Eng. 2017, 320, 335–351. [Google Scholar] [CrossRef]
  82. Chen, J.; Salvati, E.; Uzun, F.; Papadaki, C.; Wang, Z.; Everaerts, J.; Korsunsky, A.M. An Experimental and Numerical Analysis of Residual Stresses in a TIG Weldment of a Single Crystal Nickel-Base Superalloy. J. Manuf. Process. 2020, 53, 190–200. [Google Scholar] [CrossRef]
  83. Uzun, F.; Korsunsky, A.M. On the Identification of Eigenstrain Sources of Welding Residual Stress in Bead-on-Plate Inconel 740H Specimens. Int. J. Mech. Sci. 2018, 145, 231–245. [Google Scholar] [CrossRef]
  84. Uzun, F.; Korsunsky, A.M. On the Application of Principles of Artificial Intelligence for Eigenstrain Reconstruction of Volumetric Residual Stresses in Non Uniform Inconel Alloy 740H Weldments. Finite Elem. Anal. Des. 2019, 155, 43–51. [Google Scholar] [CrossRef]
  85. Uzun, F.; Everaerts, J.; Brandt, L.R.; Kartal, M.; Salvati, E.; Korsunsky, A.M. The Inclusion of Short-Transverse Displacements in the Eigenstrain Reconstruction of Residual Stress and Distortion in In740h Weldments. J. Manuf. Process. 2018, 36, 601–612. [Google Scholar] [CrossRef]
  86. Uzun, F.; Lee, T.L.; Wang, Z.I.; Korsunsky, A.M. Full-Field Eigenstrain Reconstruction for the Investigation of Residual Stresses in Finite Length Weldments. J. Mater. Process. Tech. 2024, 325, 118295. [Google Scholar] [CrossRef]
  87. Uzun, F.; Bilge, A.N. Non-Destructive Investigation of Bulk Residual Stress in Automobile Brake Pads with Its Service Life. J. Found. Appl. Phys. 2016, 3, 94–102. [Google Scholar]
  88. Jászfi, V.; Prevedel, P.; Raninger, P.; Todt, J.; Mevec, D.; Godai, Y.; Maawad, E.; Ebner, R. Residual Stress Distribution of a Locally and Inductively Quenched and Tempered 50CrMo4 Steel Analysed by Synchrotron Transmission Techniques. Mater. Des. 2022, 221, 110936. [Google Scholar] [CrossRef]
  89. Guo, D.; Yan, K.; Callaghan, M.D.; Daisenberger, D.; Chatterton, M.; Chen, J.; Wisbey, A.; Mirihanage, W. Solidification Microstructure and Residual Stress Correlations in Direct Energy Deposited Type 316L Stainless Steel. Mater. Des. 2021, 207, 109782. [Google Scholar] [CrossRef]
  90. Statnik, E.S.; Uzun, F.; Lipovskikh, S.A.; Kan, Y.V.; Eleonsky, S.I.; Pisarev, V.S.; Somov, P.A.; Salimon, A.I.; Malakhova, Y.V.; Seferyan, A.G.; et al. Comparative Multi-Modal, Multi-Scale Residual Stress Evaluation in SLM 3D-Printed Al-Si-Mg Alloy (RS-300) Parts. Metals 2021, 11, 2064. [Google Scholar] [CrossRef]
  91. Nierlich, W.; Gegner, J. X-ray Diffraction Residual Stress Analysis: One of the Few Advanced Physical Measuring Techniques That Have Established Themselves for Routine Application in Industry. Adv. Solid State Phys. 2008, 47, 301–314. [Google Scholar] [CrossRef]
  92. Chason, E.; Guduru, P.R. Tutorial: Understanding Residual Stress in Polycrystalline Thin Films through Real-Time Measurements and Physical Models. J. Appl. Phys. 2016, 119, 191101. [Google Scholar] [CrossRef]
  93. Uzun, F.; Bilge, A.N. Ultrasonic Investigation of the Effect of Carbon Content in Carbon Steels on Bulk Residual Stress. J. Nondestr. Eval. 2015, 34, 11. [Google Scholar] [CrossRef]
  94. Su, Y.; Oikawa, K.; Shinohara, T.; Kai, T.; Horino, T.; Idohara, O.; Misaka, Y.; Tomota, Y. Residual Stress Relaxation by Bending Fatigue in Induction-Hardened Gear Studied by Neutron Bragg Edge Transmission Imaging and X-ray Diffraction. Int. J. Fatigue 2023, 174, 107729. [Google Scholar] [CrossRef]
  95. Korsunsky, A.M.; Wells, K.E.; Withers, P.J. Mapping Two-Dimensional State of Strain Using Synchroton X-ray Diffraction. Scr. Mater. 1998, 39, 1705–1712. [Google Scholar] [CrossRef]
  96. Hasan, M.; Schmahl, W.W.; Hackl, K.; Heinen, R.; Frenzel, J.; Gollerthan, S.; Eggeler, G.; Wagner, M.; Khalil-Allafi, J.; Baruj, A. Hard X-ray Studies of Stress-Induced Phase Transformations of Superelastic NiTi Shape Memory Alloys under Uniaxial Load. Mater. Sci. Eng. A 2008, 481–482, 414–419. [Google Scholar] [CrossRef]
  97. Statnik, E.S.; Uzun, F.; Salimon, A.I.; Korsunsky, A.M. New Approach for Fast Residual Strain Estimation through Rational 2D Diffraction Pattern Processing. In Proceedings of the Communications in Computer and Information Science (CCIS); Springer: Berlin/Heidelberg, Germany, 2020; Volume 1086, pp. 282–288. [Google Scholar]
  98. Korsunsky, A.M.; Baimpas, N.; Song, X.; Belnoue, J.; Hofmann, F.; Abbey, B.; Xie, M.; Andrieux, J.; Buslaps, T.; Neo, T.K. Strain Tomography of Polycrystalline Zirconia Dental Prostheses by Synchrotron X-ray Diffraction. Acta Mater. 2011, 59, 2501–2513. [Google Scholar] [CrossRef]
  99. Brokmeier, H.G.; Maawad, E.; Bolmaro, R.E.; Zhong, Z.Y.; Schell, N. Combined Materials Characterization by Area Detector Investigations Using Hard X-rays. IOP Conf. Ser. Mater. Sci. Eng. 2015, 82, 012104. [Google Scholar] [CrossRef]
  100. Savage, D.J.; Lutterotti, L.; Biwer, C.M.; Mckerns, M.; Bolme, C.; Knezevic, M.; Vogel, S.C.; Borbély, A. MILK: A Python Scripting Interface to MAUD for Automation of Rietveld Analysis. J. Appl. Crystallogr. 2023, 56, 1277–1286. [Google Scholar] [CrossRef]
  101. Kieffer, J.; Karkoulis, D. PyFAI, a Versatile Library for Azimuthal Regrouping. J. Phys. Conf. Ser. 2013, 425, 202012. [Google Scholar] [CrossRef]
  102. Li, C.; Zhang, X.; Chen, Y.; Carr, J.; Jacques, S.; Behnsen, J.; di Michiel, M.; Xiao, P.; Cernik, R. Understanding the Residual Stress Distribution through the Thickness of Atmosphere Plasma Sprayed (APS) Thermal Barrier Coatings (TBCs) by High Energy Synchrotron XRD; Digital Image Correlation (DIC) and Image Based Modelling. Acta Mater. 2017, 132, 1–12. [Google Scholar] [CrossRef]
  103. Kieffer, J.; Valls, V.; Blanc, N.; Hennig, C. New Tools for Calibrating Diffraction Setups. J. Synchrotron Radiat. 2020, 27, 558–566. [Google Scholar] [CrossRef]
  104. Riedl, A.; Daniel, R.; Todt, J.; Stefenelli, M.; Holec, D.; Sartory, B.; Krywka, C.; Müller, M.; Mitterer, C.; Keckes, J. A Combinatorial X-ray Sub-Micron Diffraction Study of Microstructure, Residual Stress and Phase Stability in TiAlN Coatings. Surf. Coat. Technol. 2014, 257, 108–113. [Google Scholar] [CrossRef]
  105. Du, H.; Gong, Y.; Xu, Y.; Zeng, Q.; Xiong, L.; Li, Y.; Nie, Y.; Wang, J.; Jin, X. Obtaining Ultrastrong and Ductile Steel with Hierarchical Lamellar Duplex Phase Microstructure by Two-Stage Martensitic Transformation Mechanism. Metall. Mater. Trans. A Phys. Metall. Mater. Sci. 2022, 53, 1613–1629. [Google Scholar] [CrossRef]
  106. Li, C.; Jacques, S.D.M.; Chen, Y.; Daisenberger, D.; Xiao, P.; Markocsan, N.; Nylen, P.; Cernik, R.J. A Synchrotron X-ray Diffraction Deconvolution Method for the Measurement of Residual Stress in Thermal Barrier Coatings as a Function of Depth. J. Appl. Crystallogr. 2016, 49, 1904–1911. [Google Scholar] [CrossRef]
  107. Öztürk, H.; Yan, H.; Hill, J.P.; Noyan, I.C. Correlating Sampling and Intensity Statistics in Nanoparticle Diffraction Experiments. J. Appl. Crystallogr. 2015, 48, 1212–1227. [Google Scholar] [CrossRef]
  108. Sulyanov, S.N.; Popov, A.N.; Kheiker, D.M. Using a Two-Dimensional Detector for X-ray Powder Diffractometry. J. Appl. Crystallogr. 1994, 27, 934–942. [Google Scholar] [CrossRef]
  109. Xie, M.Y.; Baimpas, N.; Reinhard, C.; Korsunsky, A.M. Texture Analysis in Cubic Phase Polycrystals by Single Exposure Synchrotron X-ray Diffraction. J. Appl. Phys. 2013, 114, 163502. [Google Scholar] [CrossRef]
  110. Tumer, D.; Gungorurler, M.; Havitcioglu, H.; Arman, Y. Investigation of Effective Coating of the Tie6Ale4V Alloy and 316L Stainless Steel with Graphene or Carbon Nanotubes with Finite Element Methods. J. Mater. Res. Technol. 2020, 9, 15880–15893. [Google Scholar] [CrossRef]
  111. Toby, B.H.; Von Dreele, R.B. GSAS-II: The Genesis of a Modern Open-Source All Purpose Crystallography Software Package. J. Appl. Crystallogr. 2013, 46, 544–549. [Google Scholar] [CrossRef]
  112. Everaerts, J.; Salvati, E.; Uzun, F.; Romano Brandt, L.; Zhang, H.; Korsunsky, A.M. Separating Macro- (Type I) and Micro- (Type II+III) Residual Stresses by Ring-Core FIB-DIC Milling and Eigenstrain Modelling of a Plastically Bent Titanium Alloy Bar. Acta Mater. 2018, 156, 43–51. [Google Scholar] [CrossRef]
Figure 1. Schematic representations illustrating the procedural framework of the (a) Caking and (b) exCaking methods. Blue and red lines represent the Caking range, while blue and red dots depict trigonometric averaging points covering the entire ring with a predetermined angle step size.
Figure 1. Schematic representations illustrating the procedural framework of the (a) Caking and (b) exCaking methods. Blue and red lines represent the Caking range, while blue and red dots depict trigonometric averaging points covering the entire ring with a predetermined angle step size.
Crystals 14 00716 g001
Figure 2. (a) Illustration of Gaussian fitting along a profile line of pixel intensities generated by cubic interpolation. (b) The artificially strained ring and its reference state, along with the updated distances through the transformed coordinate axes, denoted as D x and D y . (c) Exemplification of the variation in the geometry factor as a function of the Bragg angle. (d) Flowchart of the exCaking console application.
Figure 2. (a) Illustration of Gaussian fitting along a profile line of pixel intensities generated by cubic interpolation. (b) The artificially strained ring and its reference state, along with the updated distances through the transformed coordinate axes, denoted as D x and D y . (c) Exemplification of the variation in the geometry factor as a function of the Bragg angle. (d) Flowchart of the exCaking console application.
Crystals 14 00716 g002
Figure 3. (a) Debye–Scherrer ring pattern with the selected ring used for analysis using the exCaking method. (b) Dimensions of the grid used for diffraction mapping of the four-point bent titanium alloy specimen.
Figure 3. (a) Debye–Scherrer ring pattern with the selected ring used for analysis using the exCaking method. (b) Dimensions of the grid used for diffraction mapping of the four-point bent titanium alloy specimen.
Crystals 14 00716 g003
Figure 4. Polar representations of elastic strains corresponding to numerical (a,b) and experimental (c,d) analyses of the exCaking (a,c) and Caking (b,d) methods. Numerical analyses were conducted using an artificially strained ring with a semi-minor axis of 99 pixels and a semi-major axis of 102 pixels, while experimental analyses were performed using the Debye–Scherrer ring with a Miller index of (102). Elastic strains assessed using (e) exCaking and (f) Caking methods within the diffraction map grid for azimuthal angles of 0 and 90 degrees that correspond to x x - and - y y -components, respectively, along with their comparison to GSAS-II ellipse-fitting calculations.
Figure 4. Polar representations of elastic strains corresponding to numerical (a,b) and experimental (c,d) analyses of the exCaking (a,c) and Caking (b,d) methods. Numerical analyses were conducted using an artificially strained ring with a semi-minor axis of 99 pixels and a semi-major axis of 102 pixels, while experimental analyses were performed using the Debye–Scherrer ring with a Miller index of (102). Elastic strains assessed using (e) exCaking and (f) Caking methods within the diffraction map grid for azimuthal angles of 0 and 90 degrees that correspond to x x - and - y y -components, respectively, along with their comparison to GSAS-II ellipse-fitting calculations.
Crystals 14 00716 g004
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Uzun, F.; Daisenberger, D.; Liogas, K.; Wang, Z.I.; Chen, J.; Besnard, C.; Korsunsky, A.M. Extended Caking Method for Strain Analysis of Polycrystalline Diffraction Debye–Scherrer Rings. Crystals 2024, 14, 716. https://doi.org/10.3390/cryst14080716

AMA Style

Uzun F, Daisenberger D, Liogas K, Wang ZI, Chen J, Besnard C, Korsunsky AM. Extended Caking Method for Strain Analysis of Polycrystalline Diffraction Debye–Scherrer Rings. Crystals. 2024; 14(8):716. https://doi.org/10.3390/cryst14080716

Chicago/Turabian Style

Uzun, Fatih, Dominik Daisenberger, Konstantinos Liogas, Zifan Ivan Wang, Jingwei Chen, Cyril Besnard, and Alexander M. Korsunsky. 2024. "Extended Caking Method for Strain Analysis of Polycrystalline Diffraction Debye–Scherrer Rings" Crystals 14, no. 8: 716. https://doi.org/10.3390/cryst14080716

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop