801 X-Ray Stress A 28. X-Ray Stress Analysis Jonathan D. Almer, Robert A. Winholtz 28.1 Relevant Properties of X-Rays ................ 28.1.1 X-Ray Diffraction........................ 28.1.2 X-Ray Attenuation...................... 28.1.3 Fluorescence ............................. 802 802 802 803 28.2 Methodology ........................................ 28.2.1 Measurement Geometry .............. 28.2.2 Biaxial Analysis .......................... 28.2.3 Triaxial Analysis ......................... 28.2.4 Determination of Diffraction Peak Positions ........ 804 805 805 805 28.3 Micromechanics of Multiphase Materials. 28.3.1 Macrostresses and Microstresses... 28.3.2 Equilibrium Conditions ............... 28.3.3 Diffraction Elastic Constants......... 28.4 Instrumentation ................................... 28.4.1 Conventional X-Ray Diffractometers .......................... 28.4.2 Special-Purpose Stress Diffractometers .......................... 28.4.3 X-Ray Detectors ......................... 28.4.4 Synchrotron and Neutron Facilities 28.5 Experimental Uncertainties ................... 28.5.1 Random Errors ........................... 28.5.2 Systematic Errors ........................ 28.5.3 Sample-Related Issues................ 28.6 Case Studies ......................................... 28.6.1 Biaxial Stress ............................. 28.6.2 Triaxial Stress ............................ 28.6.3 Oscillatory Data Not Applicable to the Classic Model.................... 28.6.4 Synchrotron Example: Nondestructive, Depth-Resolved Stress........................................ 28.6.5 Emerging Techniques and Studies 807 807 808 808 809 809 810 810 810 810 811 811 812 813 813 814 815 815 816 806 28.7 Summary ............................................. 817 28.8 Further Reading ................................... 818 References .................................................. 818 X-rays are an important tool for measuring stresses, particularly residual stresses, in crystalline materials. x-ray stress measurements are used to help solve material failure problems, check quality control, verify computational results, and for fundamental materials research. X-ray diffraction can be used to precisely determine the distance between planes of atoms in crystalline materials through the measurement of peak positions. These positions can be used to determine elastic strains in each crystalline phase of the material. These strains can then be converted to stresses using appropriate elastic constants. Plastic deformation can be detected through changes in diffraction peak widths rather than peak shifts. Laboratory-based x-ray sources typically penetrate only a few tens of microns into common materials, yielding stresses averaged over the near-surface region. Deeper depths can be accessed using destructive layer-removal methods, with appropriate corrections. Alternatively, more highly penetrating neutrons or high- Part C 28 X-ray diffraction is a powerful non-destructive technique capable of measuring elastic strain in all crystalline phases of a material, which can be converted to stress using appropriate elastic constants. With laboratory sources typical penetration depths are on the micron-level, and deeper depths can be evaluated using (destructive) layer removal methods or higher-energy x-rays (esp. from synchrotron sources) or neutrons. Diffraction also provides complementary information on crystallographic texture and plastic deformation. Potential sources of errors in stress measurements are outlined. Finally, some case studies and emerging techniques and studies in this field are highlighted. 802 Part C Noncontact Methods energy x-rays [28.1, 2]. can be used to examine the interior of components nondestructively. The subject of stress measurement with x-ray diffraction has been treated in great depth by Noyan and Cohen [28.3] and more recently by Hauk [28.4]. In addition, related reports have been presented by the Society for Automotive Engineers [28.5], the Cen- tre Technique de Industries Méchaniques [28.6], the American Society for Metals [28.7], and most recently as a United Kingdom good practice guide [28.8]. Here we provide an updated account of the subject, including case studies from both laboratory- and synchrotron-based experiments, together with a summary of emerging studies in the field. 28.1 Relevant Properties of X-Rays Part C 28.1 In this section, several properties of x-rays relevant to diffraction stress analysis are summarized. 28.1.1 X-Ray Diffraction X-rays provide a means of examining the atomic-scale structure of materials because their wavelengths are similar to the size of atoms. The measurement of lattice spacings in crystalline materials with x-rays utilizes diffraction, the constructive interference of x-ray waves scattered by the electrons in the material. If the crystal is oriented properly with respect to the incident x-ray beam, constructive interference can be established in particular directions, giving rise to a diffracted x-ray beam. Constructive interference arises when the path length for successive planes in the crystal equals an integral number of wavelengths. This condition is summarized with Bragg’s law nλ = 2d sin θ . (28.1) Here, λ is the wavelength of the x-rays, d is the interplanar spacing in the crystal, θ is the diffraction angle, and n is the number of wavelengths in the path difference between successive planes. In practice, n is eliminated from (28.1) by setting it equal to 1 and considering the diffraction peaks arising with n > 1 to be occurring from higher-order lattice planes, i. e., diffraction from (100) planes with n = 4 are considered as arising from the (400) planes with n = 1. The interplanar spacing or d-spacing in the crystal is measured along a direction that bisects the incident and diffracted x-ray beams. A vector in this direction with length 2 sin θ/λ is termed the diffraction vector, as shown in Fig. 28.1. Equation (28.1) can be used in two ways to determine d-spacings in materials. First, a nominally monochromatic x-ray beam can be used and the diffraction angle θ precisely measured. Second, a white or polychromatic x-ray beam can be incident on the specimen and the wavelengths scattered at a particular angle measured. This is usually done with a solid-state detector that can precisely measure the energy of detected photons using the wavelength–energy relationship for electromagnetic radiation hc (28.2) , E where h is Planck’s constant, c is the speed of light, and E is the energy of the photons. Substituting numerical values for the constants, if λ is in Angstroms and E is in keV, the wavelength energy relation becomes λ= λ= 12.39842 . E (28.3) 28.1.2 X-Ray Attenuation As x-rays pass through matter they are attenuated. This is an important consideration in measuring stresses with x-rays. For typically used reflection geometries, the attenuation defines the depth being sampled in a stress measurement while, for transmission geometries, it will define the thickness of a specimen that can be examined. An x-ray beam of intensity I0 passing through a thickness t of a material will be attenuated to an intensity I given by I = I0 exp(−μl t) , (28.4) where μl is the linear attenuation coefficient for the material. The linear attenuation coefficient can be computed for a sample having N elements using μl = ρ N wi (μm )i , (28.5) i=1 where ρ is the sample density and wi and μm,i are, respectively, the weight fraction and mass attenuation coefficient of element i. The μm are inherent properties X-Ray Stress Analysis a) 28.1 Relevant Properties of X-Rays 803 b) X3 Ψ X3' σ Diffraction vector X1 θ Ψ X2 c) Ψ Diffraction vector Fig. 28.1 (a) Coordinate systems used in diffraction stress measurement, showing the specimen coordinates (X i ) and laboratory coordinates (X i ) and the manner in which the specimen is rotated to orient the diffraction vector along the laboratory X 3 axis; definition of tilt angle ψ for (b) Ω-gonio-metry and (c) Ψ -goniometry, with the in-plane stress direction indicated of a given element and commonly tabulated. Between absorption edges (which lead to fluorescence, see below) these values vary with wavelength λ and elemental atomic number Z as μm ∝ Z 4 λ3 , (28.6) which when combined with (28.2) indicates that absorption decreases considerably with x-ray energy. Table 28.1 lists μl values for selected materials and common laboratory x-ray energies. Also tabulated is the inverse value τ1/e , which represents the x-ray path length at which I = I0 /e ≈ 0.37I0 , and which is seen to be in the micron range. The actual penetration depth into the sample depends on the diffraction geometry used, and this is covered further in Sect. 28.2.1. 28.1.3 Fluorescence When the incident x-ray energy is higher than the (element-specific) energy required to remove an electron (e.g., K-edge electron) from its shell, the electron will be ejected. This ejected electron is called a photoelectron and the emitted characteristic radiation is called fluorescent radiation. For strain measurements, fluorescence is unwanted as it increases x-ray background levels, thereby decreasing the signal-to-background ratio and reducing accuracy in peak position determination. The fluorescence signal can be eliminated by using x-ray energies below the fluorescence energies of the elements comprising the sample, though this is not always practical. For tube sources, containing both continuous (bremsstrahlung) radiation and characteristic (Kβ and Kα ) radiation, the fraction of higher-energy (fluorescence-forming) x-rays can be reduced relative to the signal-forming (typically Kα ) x-rays through the use of an incident-beam energy filter in the form of a metal or oxide foil or monochromator. Alternatively, fluorescence signals can be reduced by the use of a diffracted-beam energy filter (foil or monochromator) and/or through use of an energy-discriminating (e.g., solid-state) detector. Part C 28.1 σ 804 Part C Noncontact Methods Table 28.1 X-ray penetration depths, common reflection/radiation combinations for stress measurements, and elastic properties for selected materials Material (structure) μ (1/mm) and (τ(1/e), μm) Mo–Kα Cu–Kα λ = 0.71 Å λ = 1.54 Å Cr–Kα λ = 2.29Å 14.3 132 397 (69.9) (7.6) (2.5) Al (fcc) Elastic aniso- a tropy 1.2 310 2547 904 (3.3) (0.4) (1.1) Ferrite (bcc) 2.3 Part C 28.2 311 2633 934 (3.2) (0.4) (1.1) Austenite (fcc) 3.3 hkl b 2θ (deg)b 111 53.8 200 49.9 311 (Cr) 200 139.0 52.1 141.7 211 (Cr) 156.0 178.7 732 (Mo) 111 154.5 163.2 209.8 200 124.4 420 (Cu) 111 422 439 1291 (2.4) (2.3) (0.8) Nickel (fcc) 2.5 107 919 2718 (9.4) (1.1) (0.4) - b c d 155.4 200 52.3 163.6 145.7 152.1 164.0 90.9 83.0 139.5 E/(1 + ν) bulk (GPa) d 142.3 219.4 144.2 213 (Cu) a 147.0 200 420 (Cu) 002 Titanium (hcp) E/(1 + ν) Hill (GPa) c 87.6 83.6 Anisotropy is calculated for cubic materials as 2c44 /(c11 − c12 ), where cij are the elastic constants from [28.9] and [28.10]. Suggested reflection/radiation combinations and corresponding Bragg angles for laboratory stress measurements are shown. Calculated using cij from [28.9] and [28.10] and using the Hill–Neerfeld average (Sect. 28.3.3). From [28.11] 28.2 Methodology Stress measurement with diffraction involves measuring lattice d-spacings in different directions in a specimen utilizing Bragg’s law and then using them to compute strains through ε= sin θ0 d − d0 = −1 . d0 sin θ (28.7) Here d0 and θ0 are the unstressed d-spacing and the corresponding Bragg angle. We wish to obtain the stress and strain components in the specimen with respect to the specimen coordinate system shown in Fig. 28.1. By observing the diffraction with a diffraction vector oriented along the X 3 axis, we can determine the strain in this direction with (28.7). The X 3 axis is oriented with respect to the specimen coordinate system by the an- gles φ and ψ. Applying tensor transformation rules to the specimen coordinate system we may write the measured strain, εφψ , in terms of the strains in the specimen coordinate system: dφψ − d0 d0 = ε11 cos2 φ sin2 ψ + ε22 sin2 φ sin2 ψ εφψ = + ε33 cos2 ψ + ε12 sin 2φ sin2 ψ + ε13 cos φ sin 2ψ + ε23 sin φ sin 2ψ . (28.8) By measuring strains with diffraction in at least six independent directions, the strains in the specimen coordinate system can be determined by a least-squares procedure [28.12]. After the strains in the specimen co- X-Ray Stress Analysis ordinate system have been computed, the stresses in this coordinate system can be determined with Hooke’s law as S1 1 εij − δij 1 εii , (28.9) σij = 1 2 S2 2 S2 + 3S1 28.2.1 Measurement Geometry Placing the diffraction vector along a particular X 3 axis can be accomplished in an infinite number of ways by rotating the diffraction plane (containing the incident and diffracted beams) around the diffraction vector. In practice this is accomplished in two ways, termed Ω goniometry and Ψ goniometry, which are illustrated in Fig. 28.1b,c. In Ω goniometry the specimen is rotated by an angle ψ about an axis perpendicular to the diffraction plane, while for Ψ goniometry this rotation is performed about an axis within the diffraction plane. These rotations orient the diffraction vector to the specimen coordinate system by the angle ψ in Fig. 28.1a. For both goniometry methods the orientation angle φ is obtained by rotating the specimen about the X 3 axis. The penetration depth into the sample, z, depends on the x-ray path length into and out of the sample, and thus on measurement geometry. This depth is given for both types of goniometry as [28.13] σ23 , and σ33 , which have a component perpendicular to the surface. It is therefore common to assume a biaxial stress state and measure the stress in a particular in-plane direction (given by the angle φ in Fig. 28.1a). If we drop the σi3 components and define the stress in the φ direction as σφ = σ11 cos2 φ + σ12 sin 2φ + σ22 sin2 φ (28.11) and solve for dφψ we obtain dφψ = 1 S2 d0 σφ sin2 ψ + S1 d0 (σ11 + σ22 ) + d0 . 2 (28.12) This equation gives the measured d-spacing as a function of the tilt angle. For a stressed material, at ψ = 0◦ the strain is given by the Poisson effect of the stress. As the specimen is tilted, the measured strain varies according to strain transformation rules, varying linearly with sin2 ψ. If the measured d-spacing is plotted versus the variable sin2 ψ, there should be a linear relationship. It is good to check experimentally that this relationship is indeed linear by measuring at multiple ψ tilts. Nonlinear relations can reveal a number of problems and can invalidate the analysis. Case studies illustrating both linear and nonlinear behavior are presented below. If a line is fitted to the d versus sin2 ψ data, the resulting slope will be proportional to the stress σφ . The stress can be obtained by multiplying the experimentally determined slope by 12 S2 and d0 . The stress-free lattice spacing, d0 , is generally unknown. Since it is simply a multiplier in (28.12), the intercept of the d versus sin2 ψ plot or another approximation can be used with only a small error. 28.2.3 Triaxial Analysis ln II0 sin2 θ − sin2 ψ + cos2 θ sin2 ψ sin2 η , z= 2μl sin θ cos ψ (28.10) where η is the sample rotation around the diffraction vector and equal to 0◦ (90◦ ) for Ω (Ψ ) goniometry, respectively, and other quantities are as previously defined. The maximum penetration depth is thus at ψ = 0◦ and θ ⇒ 90◦ such that z max ⇒ ln(I0 /I )/2μl . 28.2.2 Biaxial Analysis Because conventional x-rays only penetrate a few tens of microns into most engineering materials, the stresses measured are usually biaxial, lacking components σ13 , 805 When using high-energy x-rays or neutrons, the penetration depth is larger, so that all the components of the stress tensor may be present. In addition, as we will see in Sect. 28.3, triaxial analysis may be needed for conventional laboratory-based x-ray sources if multiphase materials are investigated. To determine the complete stress tensor the lattice spacing is measured in a number of directions/angles, and strains in these directions are calculated with the stress-free lattice spacing d0 . The triaxial strain components εij are computed using (28.8), which can be rewritten as εφψ = 6 k=1 ak f k (φ, ψ) , (28.13) Part C 28.2 where δij is the Kronecker delta function and S1 = (ν/E)hkl and S2 /2 = ((1 + ν)/E)hkl are the diffraction elastic constants (DEC), which are discussed in Sect. 28.3.3. Equations (28.7)–(28.9) encapsulate the basic model of stress measurement by diffraction. Practical implementation of this model will be discussed in the following sections. 28.2 Methodology 806 Part C Noncontact Methods Part C 28.2 where ak are the six strain components determined by fitting the measured data, and the functions f k (φ, ψ) are functions of the two variables φ and ψ. At least six independent strain measurements are needed to solve (28.13) for the strain components. Greater accuracy is obtained by increasing the number of strain measurements and by using measurement directions that have a wide angular distribution. It is good practice to measure at a number of ψ values (both positive and negative) for several different constant φ values. Plots of the measured strain values for each of the φ values versus sin2 ψ help reveal the quality of the measured data. Examining (28.8), we see that if the shear strains ε13 and ε23 are zero, the ε versus sin2 ψ plots will be linear with the positive and negative ψ data overlapping on this line. If the shear strains ε13 and ε23 are present, the positive and negative branches of the plot will take different paths following two sides of an ellipse. If the data do not follow one of these shapes, it indicates a problem and fitting the data to (28.8) is unlikely to give accurate stress values. Plots of ε versus sin2 ψ which can be properly fitted are shown as a case study below. An important issue for accurate triaxial stress measurement is obtaining an accurate value of the stress-free lattice spacing. In contrast to the biaxial case where it is just a multiplier, the stress-free lattice spacing is subtracted from the measured d-spacings to obtain a strain. Consequently an accurate unstressed lattice spacing d0 is needed for triaxial analysis [28.14]. An error in the stress-free lattice spacing results in an error in all the resulting computed strain values, which yields an erroneous hydrostatic stress/strain component appearing in the final stress/strain tensor determination. If an accurate stress-free lattice spacing is unavailable, the deviatoric components of the stress and strain tensors will still be accurate even though the hydrostatic components are not [28.15]. 28.2.4 Determination of Diffraction Peak Positions Lattice spacings are determined from diffraction by careful determination of diffraction peak positions. As a stressed specimen is tilted and diffraction peaks are recorded, the position of the peak will shift as a different component of strain is resolved. Strains in crystalline materials are small (usually 10−4 or less) and, thus, peak shifts will be small, often less than half the width of the peak. Therefore, the diffraction peak must be carefully recorded and the peak position precisely determined. The position of the peak can be defined as the 2θ value of the centroid of the diffraction intensity or the position of maximum diffracted intensity. For symmetric diffraction peaks these quantities will be the same, but in general, they will be different. The same definition should be used throughout a stress measurement. Theoretically, the peak centroid is more appropriate, but is not widely used in practice. In order to determine the peak position, the intensity versus 2θ data is usually fitted to a mathematical function representing the peak profile [28.3,16]. A Gaussian peak with a linear background is given by 2θ − 2θp 2 + 2θm + b , I = I0 exp −4 ln 2 W (28.14) where I0 is the diffracted intensity over the background level, 2θp is the peak position, W is the peak full-width at half-maximum, m is the slope of the background, and b is the background intercept. A Gaussian often fails to fit the tails of the diffraction peak well. Adding a Lorentzian component to the peak function gives a pseudo-Voigt function, ⎧ ⎪ ⎨ 2θ − 2θp 2 I = I0 η exp −4 ln 2 ⎪ W ⎩ ⎫ ⎪ ⎬ 1 + (1 − η) 2 ⎪ + 2θm + b , 2θ−2θp ⎭ 1+ W which will usually better fit the tails. Here η is the fraction of the Gaussian component in the peak and (1 − η) is the fraction of the Lorentzian component. With conventional x-ray tubes the characteristic x-rays usually contain two closely spaced wavelengths, so the diffraction peaks are actually a doublet of two closely spaced peaks. For a Kα1 –Kα2 doublet, the higher-angle peak will have approximately half the height of the lower-angle peak due to the intensity ratio of the respective wavelengths in the incident beam. Constraining the peak height ratio to 0.5 can improve the fitting of the doublet. With sufficient broadening, the two peaks in the doublet can be treated as a single peak and fitted with one of the above equations. If the two peaks do not sufficiently overlap (i. e., they are distinct), they should be fitted with two peaks separated in 2θ by δ. Note that adding the second peak to the fit function does not require the addition of any new fit parameters. The separation between the two peaks δ is computed X-Ray Stress Analysis from the known x-ray wavelengths and Bragg’s law −1 λ2 sin(2θ1 /2) − 2θ1 . δ = 2 sin (28.15) λ1 Intensity (Counts) 450 I0 = 203.44 ± 2.67 2θ = 144.071 ± 0.00275 W = 0.463429 ± 0.00546 m = – 8.7897 ± 2.03 b = 146.46 ± 293 400 350 300 250 200 150 143.5 144 144.5 145 2θ (Degrees) Fig. 28.2 211 diffraction peak from the β phase of a 60–40 brass along with fit to a Gaussian doublet. The best fit parameters from (28.14) are given in the inset box. The value of δ was calculated prior to fitting to have a value of 0.60409 advantage is that the degree of doublet overlap can vary with ψ, causing shifts in the location of maximum intensity due to varying doublet overlap, which will erroneously be recorded as due to residual stress. Fitting a doublet peak function to the data automatically accounts for this potential problem. With linear or area detectors, the whole peak is automatically recorded and there is little need to fit just the top portion of the peak. 28.3 Micromechanics of Multiphase Materials In Sect. 28.2, methods for measuring stress states with diffraction were outlined. In order to properly interpret these stresses, we now distinguish between types of stresses, outline key equilibrium conditions, and provide information on diffraction elastic constants needed to compute stresses from measured strains. 28.3.1 Macrostresses and Microstresses Macrostresses are the stresses that appear in a homogeneous material. They vary slowly on the scale of the microstructure and are the stresses that are revealed by dissection techniques. Residual stresses, as they have been traditionally treated, are macrostresses and arise due to nonuniform deformations on a macro- 807 scopic scale. Microstresses, in contrast, arise from the microstructural inhomogeneities in the material and nonuniform deformations on this scale. While microstresses can often be an impediment to determining the true residual macrostresses in a material, they are increasingly becoming of interest in their own right, with the recognition that they can also influence material behavior [28.18]. Residual stresses have also been distinguished in the literature as types I, II, and III, as illustrated in Fig. 28.3. Type I stresses are macrostresses that are constant over many grains in the material. Type II stresses are microstresses and can vary between phases, as well as between grains within a given phase. Typical diffraction data provide the grain-averaged type II stress value within a given Part C 28.3 Figure 28.2 shows a 211 diffraction peak from the β phase of a 60–40 brass, measured with Cr Kα1 –Kα2 radiation and fitted with a Gaussian peak profile, including a second peak to account for the doublet. With an area or linear detector, the background away from the peak position is automatically recorded and should be included in the fitting. A simple line is usually sufficient, though a higher-order polynomial or other forms can also be used to represent the background accurately. If the diffraction peak is recorded by point counting, the background may not be sufficiently well recorded away from the peak to accurately fit. Recording the background can be time consuming and may not be worthwhile, particularly if the background is not changing as the specimen is tilted. In these instances the background should be assumed constant in 2θ and only this constant included in the fit function. If a background function with a slope is used and intensity data away from the peak that would allow an accurate determination of the slope is absent, erroneous peak positions can result from the fitting, due to correlations between the slope and peak position. Fitting the top 15% of the diffraction peak to a parabola has also been utilized for x-ray stress measurements [28.17]. For point counting this method has the advantage that time is not spent collecting the whole diffraction peak, speeding up the measurements. A dis- 28.3 Micromechanics of Multiphase Materials 808 Part C Noncontact Methods phase. Type III stresses are also microstresses but represent stress variations within single grains. Residual stresses of types I and II will result in diffraction peak shifts, as they change the mean lattice spacing, while type III stresses result in diffraction peak broadening rather than peak shifts. Measurement of type III stresses is outside the current scope and will not be further discussed. For a two-phase material, the measured stress from peak shifts can be written as t α I σij = σij + II σijα , t β I β σij = σij + II σij , (28.16) Part C 28.3 where the superscripts t, I, and II stand for the total stress in a phase (α or β), the type I macrostresses, and the type II microstresses, respectively. The total stresses and microstresses are represented as averages (denoted by carets) as they can vary significantly within the diffracting volume, whereas the macrostress does not. Note also that the macrostress component is common to both phases and thus does not have a phase designation. These equations hold for each component of the stress tensor. Using (28.16), the microstresses in the α and β phases must balance to zero when weighted by their volume fractions, thus β (28.17) (1 − f ) II σijα + f II σij = 0 . Here f is the volume fraction of the β phase. The total stress components in the α and β phases can be σ α σ β σ II I σ II β σ II measured with diffraction measurements as outlined in Sect. 28.2.3. If the volume fractions of the phases are known, the components of the macrostress tensor and microstress tensors can be then determined using (28.16) and (28.17). 28.3.2 Equilibrium Conditions As measured with conventional (low-energy) x-rays, residual macrostresses with a component perpendicular to the surface, σ13 , σ23 , and σ33 (with the x3 axis normal to the surface) are typically negligible. This can be seen via the following equilibrium relations for residual macrostresses: ∂σ11 ∂σ12 ∂σ13 + + =0, ∂x1 ∂x2 ∂x3 ∂σ21 ∂σ22 ∂σ23 + + =0, ∂x1 ∂x2 ∂x3 ∂σ31 ∂σ32 ∂σ33 + + =0. (28.18) ∂x1 ∂x2 ∂x3 These relationships show that the stress components σi3 can only change from their value of zero at the free surface if there are gradients in the other components of stress within the surface. These components of stress cannot normally have high enough gradients within the surface to cause the components σi3 to reach measurable levels within the penetration depth of conventional x-rays (the biaxial assumption, see Table 28.1). Another equilibrium relation that proves useful is that the stress integrated over a cross section of a solid body must be zero σij dA = 0 . (28.19) A This relation can be used as a check on experimental results and can also be used to determine the stress-free lattice spacing. σ III 28.3.3 Diffraction Elastic Constants Fig. 28.3 Definition of macrostress (σ I ) and microstresses in a two-phase material. The measured peak shift in a given phase is proportional to the total mean elastic stress in that phase (σ I + σ II ). Variations in stress within a grain (σ III ) cause peak broadening rather than peak shifts Elastic constants are needed to compute stresses from measured diffraction strains in polycrystalline materials. Due to the selective nature of the diffraction process and presence of crystalline elastic anisotropy, the so-called diffraction elastic constants (DEC) differ, in general, from the bulk elastic constants. Here, two methods for attaining DEC are described: 1. computing them using single-crystal elastic constants and a given micromechanical (grain-averaging) model, and X-Ray Stress Analysis 2. measuring the constants in situ using known elastic loads. • • The higher the crystalline anisotropy, the larger the deviation between DEC and bulk values for a given diffraction peak. Thus, use of bulk values will give a corresponding error in absolute stress value for a given diffraction measured strain (28.12). These deviations are especially important for lowmultiplicity peaks (e.g., 111/200). For highermultiplicity peaks (e.g., 732 in ferrite) the DEC approach the bulk values. Such peaks have the added benefit of being less sensitive to texture, as shown by Hauk in textured steel [28.25]. A second method is to measure the DEC in situ using known loads. This represents the inverse problem to stress analysis described throughout the rest of the chapter, where stresses are unknown and DEC are assumed known. Loads should be applied using either uniaxial tension or four-point bending, since the applied stress distributions in such cases can be determined accurately from elasticity theory, and the diffraction strains are recorded for reflections of interest. The American Society for Testing and Materials has developed a standard test method for the measurement of the diffraction elastic constant S2 /2 for a biaxial stress measurement, which can give some practical guidance [28.26]. Such methods are particularly important when elastic constants of the material are unknown or in question. For example, alloying can modify DEC considerably from elemental values, as found by Dawson et al. [28.27] on aluminum alloys. Tables of measured DEC for various materials (including common alloys) can be found in [28.3, 7]. Taking this method one step further, Daymond [28.28] found that, by using strain from multiple diffraction peaks and a weighted Rietveld method, it was possible to calculate bulk elastic properties E and ν in polycrystals. He found good matches to bulk properties for both textured and untextured cubic (steel) and hexagonal (titanium) crystal symmetries. We note that the use of bulk elastic constants (E, v) in place of hkl-dependent DEC is common. While this can cause an error in the absolute stress values, relative changes in stress (e.g., between samples of the same composition, or positions in a given sample) may still be valid. Finally, we note that the grain-interaction models described above are appropriate for bulk samples, but van Leeuwen et al. [28.29] found that in the case of thin films (e.g., vapor-deposited nickel) the Vook– Witt [28.30] model is more appropriate to describe the measured strains. The latter model takes the special geometry of thin films into account for specifying the strain and stress state, and can be used to calculate nonlinear d–sin2 ψ behavior even in the absence of crystallographic texture. For additional information on strain measurements in thin films two reviews on the subject are noted [28.31, 32]. 28.4 Instrumentation A variety of equipment can be used for diffraction stress measurements. Conventional x-ray diffractometers are often used, as well as special-purpose stress diffractometers which may be laboratory-based or portable. At synchrotron x-ray and neutron facilities the working area is typically larger than with laboratory-based diffraction systems, permitting larger samples and/or ancillary equipment. Associated instrumentation is generally specialized and outside the scope of this chapter. 28.4.1 Conventional X-Ray Diffractometers Conventional x-ray diffractometers are widely available and used for a variety of applications. These can of- 809 Part C 28.4 The most widely used micromechanical models are those from Voigt [28.19], Reuss [28.20], Neerfeld– Hill [28.21,22], and Eshelby–Kroner [28.23]. The Voigt and Reuss models assume that all grains in a polycrystalline aggregate have, respectively, the same strain and stress, and represent, respectively, the upper and lower bounds of the elastic constants. It is generally found that the Neerfeld–Hill (which is the arithmetic average of the Voigt and Reuss values), and Eshelby–Kroner models best match experimentally determined DEC. Equations for determining DEC for various crystal symmetries can be found in [28.3, 4], and stress-analysis software packages may provide computational tools (e.g., [28.24]). In Table 28.1, Hill model values of the inverse DEC, 2/S2 = (E/(1 + ν))hkl , are compared with bulk values (i. e., those expected for an isotropic body) for various materials and hkl peaks. We note the following trends from the table: 28.4 Instrumentation 810 Part C Noncontact Methods ten be utilized for stress measurement, though software to control the data acquisition and analysis will often be deficient for this task, requiring a greater degree of operator expertise. Conventional x-ray diffractometers may have several limitations, making stress measurement difficult or impossible. They often have limited space for specimens, have limited or no ability to achieve ψ-tilts, have incompatible beam defining slits, and provide poor access to the raw data for analysis. If Ψ -goniometry is to be used with a diffractometer having a χ-circle, the x-ray tube needs to be oriented for point focus to limit the length of the beam along the diffractometer axis. Part C 28.5 28.4.2 Special-Purpose Stress Diffractometers Various stress diffractometers have been specifically designed to perform these measurements. They have the ability to handle larger and heavier specimens and have software to automate the data collection and analysis. Most modern stress diffractometers use linear or area detectors to increase the measurement speed, and utilize the Ψ -goniometry because it does not defocus the x-ray optics. More information and a photograph of a stress diffractometer are given in a case study in Sect. 28.6.1. Laboratory-based units will be more flexible diffraction instruments, usually having texture measurement capabilities and the ability to do general powder diffraction. Portable stress diffractometers can be taken into the field to make stress measurements on very large objects without having to move them. 28.4.3 X-Ray Detectors X-ray detectors can be broadly classified into point, line, or area detectors. Point detectors include proportional, scintillation, and solid-state detectors. For each the x-ray energy is proportional to the measured pulse height, so that energy discrimination can be accomplished through electronic filtering. The energy resolution is highest for solid-state detectors (ΔE/E ∼ 3–5%) and lowest for scintillation counters (ΔE/E ∼ 30–40%). Thus, solid-state detectors generally have lower background levels and are the standard choice for energy-dispersive measurements. If necessary, secondary monochromators and/or foils can be used in conjunction with scintillation or proportional detectors to reduce background levels. As their name implies, linear and area detectors measure, respectively, one- and two-dimensional angular sections of x-ray diffraction cones. As this information is collected simultaneously, these detectors allow for significantly faster strain measurements than point counters, which require scanning over the 2θ region of interest. Area detectors have the added benefit of providing simultaneous grain size and texture information through intensity variations around a diffraction cone. These detectors generally have limited energy resolution, and incident-beam filters can be employed to limit background levels. 28.4.4 Synchrotron and Neutron Facilities Both synchrotron and neutron facilities provide unique capabilities for strain and stress analysis. Synchrotrons offer x-ray fluxes several orders of magnitude higher than x-ray tubes, and as such can be used for timeresolved experiments. In addition, these sources have much lower divergence and source sizes, so focal spots from mm down to tens of nm can be used for spatially resolved measurements. Furthermore, thirdgeneration facilities [the Advanced Photon Source (APS), USA; the European Synchrotron Radiation Facility, France; and SPRing-8, Japan] are excellent sources of high-energy x-rays, with concomitant high penetration depths (over 1 mm in most materials). This, in turn, enables many in situ and bulk studies that are either not practical or possible with laboratory-based sources. Even higher penetration depths are provided by neutrons (up to the cm level), with similar in situ capabilities. Access to synchrotron and neutron facilities is typically granted (for free) based on peer-reviewed proposals, but proprietary (fee-based) work may be arranged at certain facilities. 28.5 Experimental Uncertainties With proper care, diffraction stress measurements can readily be made with sufficient accuracy and precision to solve many problems. There are, however, ran- dom and systematic errors in the measurement process that should be carefully considered when evaluating diffraction stress measurement results. The best way X-Ray Stress Analysis 28.5.1 Random Errors Random errors in x-ray measurements arise from the statistical nature of x-ray counting and the random component of proper alignment of the specimen and instrument. The statistical errors will typically be the dominant error in measurements with laboratory xrays or neutrons. These errors can be driven arbitrarily low by collecting data for longer periods of time, but the improvement varies as the square root of the time. When a diffraction peak is recorded and fitted to find its position, there will be an uncertainty associated with this position due to the statistical nature of recording x-ray intensities. Fitting programs usually give an estimate of the uncertainty in the fitted parameters, including the peak position. This uncertainty can be propagated through the analysis to estimate the uncertainty in the stresses. The uncertainty in a d-spacing is given by differentiating Bragg’s law [28.3] STD(dψ ) = dψ cot θ STD(2θp ) 2 π 180 . (28.20) Here, STD(dψ ) represents an estimated standard deviation in the value of dψ and STD(2θp ) is an estimated uncertainty in the peak position arising from the fitting of the diffraction peak. The π/180 factor is a conversion of degrees to radians, assuming that the uncertainties in peak position are in degrees. Similarly, the uncertainty in a strain is given by STD(εφψ ) = cot θ STD (2θ) 2 π 180 811 . (28.21) These uncertainties can then be used in the fitting of the d-spacing or strains εφψ to the biaxial or triaxial models ((28.12) and (28.13)). A proper fitting package will propagate the uncertainties in the input data to the fitted parameters. A goodness-of-fit parameter can be computed to help determine if the measured data suitably fits the model used [28.36]. The goodness of fit is the probability that the observed deviations from the model fit arise from the estimated errors. A low goodness of fit can indicate that the errors are underestimated or that the data do not fit the model and the resulting stresses should not be considered valid. It has been recommended that goodness-of-fit parameters should be better than 10−3 –10−5 before considering the model fit to be valid [28.36]. Poor goodness-of-fit parameters may indicate the presence of systematic errors and/or samplerelated issues, both of which are described below. Measurements made under computer control can take some quick preliminary measurements of the x-ray intensities and control the measurement time to give a specified statistical error in the final measurements. 28.5.2 Systematic Errors Systematic errors in diffraction stress measurement consist primarily of imperfect optics and misalignments in the diffractometer, both of which cause shifts in the diffraction peak position from its true position. For conventional x-ray stress measurement using reflection geometries, a number of researchers have given equations to estimate the uncertainties in stress due to these sources, and these are summarized in Table 28.2. The size of these systematic peak shifts generally increases Table 28.2 References for systematic instrumental errors in diffraction stress measurements Source of error Refs. Specimen displacement (ω-goniometry) Specimen displacement (Ψ -goniometry) Horizontal beam divergence Vertical beam divergence Missetting of diffractometer angles ψ-axis displacement (ω-goniometry) ψ-axis displacement (Ψ -goniometry) Curvature of specimen surface [28.3, 6, 34] [28.3, 6] [28.3] [28.3, 34] [28.35] [28.3, 6, 34] [28.3, 6] [28.3, 6] Part C 28.5 to evaluate the errors present is to repeatedly measure a stress-free standard, such as an annealed powder from the material of interest. The average stress measured should give a measure of the systematic errors present, while the variation in results should give a measure of the random errors present. The American Society for Testing and Materials ASTM outlines a standard test method for verifying the alignment of an x-ray stress diffractometer [28.33] and specifies that an annealed powder measured according to accepted procedures should give a value of stress less than 14 MPa (2 ksi). Measuring an annealed powder is a time-consuming process and may not be practical. It is desirable to obtain good estimates of the random and systematic errors in a measurement based on just that single measurement. It is also desirable to be able to estimate the errors so that the measurements can be made to a specified accuracy. Finally, it can be a good practice to repeat measurements to check reproducibility. 28.5 Experimental Uncertainties 812 Part C Noncontact Methods Part C 28.5 with the tilt angle ψ and with decreasing diffraction peak position 2θ. Care should be taken if extreme values of ψ are to be used (greater than 60◦ ) or with diffraction peaks below 2θ of 130◦ . For transmission methods with high-energy x-rays or neutrons, successful measurements can be successfully made with low-2θ peaks, in part because they often have limited (or no) requisite sample movement. Generally the largest instrumental error is that due to specimen displacement, i. e., not having the sample surface at the center of the diffractometer axis. This results in a diffraction peak shift that depends on ψ and thus introduces an error in the measured stresses. This error can be minimized by using parallel beam geometry in which vertically aligned Soller slits are used to define the diffraction angles. This geometry only works for point detectors that are scanned across the diffraction peak, which is much slower than using linear or area detectors. This component of error can change as specimens are moved for measurements at different locations or as specimens are changed. This error will also change for a specimen from the value recorded on a stress-free powder. Placement of the specimen surface on the diffractometer axis is typically achieved with special alignment gages for dedicated instruments. Additionally, one can measure the specimen displacement for cubic materials with diffraction [28.3]. The lattice parameter a is determined from a series of diffraction peaks and plotted versus cos2 θ/ sin θ. The slope of this plot gives the specimen displacement Δx according to the relation Δx cos2 θ ahkl − a0 = , a0 RG sin θ (28.22) where RG is the goniometer radius and a0 is the true lattice parameter, which can be approximated with the intercept of the plot. For triaxial analysis an accurate value of the stressfree lattice spacing must be used. This should be measured on the same diffractometer used for the stress measurements. Inaccuracies in the stress-free lattice spacing lead to an error in the hydrostatic component of the stress tensor [28.15]. If a single stress-free lattice spacing is used for a series of measurements, this error will be systematic and changes in the hydrostatic component between measurements will be meaningful. In some situations, notably welds, the stress-free lattice spacing can vary from point to point in the specimen due to composition gradients. If the full tensor is then desired, stress-free reference values must be measured for each point sampled, which can be accomplished by sample sectioning. 28.5.3 Sample-Related Issues The error analysis described above applies to ideally behaved specimens, and thus represents a lower limit on the error in stress values. Here we describe potential sample-related issues that can modify strain data relative to this ideal behavior. Texture Effects Crystallographic texture is common in materials and is due to directional processing methods, such as rolling in bulk materials or directional deposition in thin films and coatings. The presence of texture causes systematic variations in the peak intensity as a function of tilt. This can be distinguished from random (stochastic) variations in intensity, which indicate improper grain averaging (addressed below). If strong texture is known or suspected to be present, it can be beneficial to measure a pole figure on the peak intended for strain analysis, which both provides quantitative texture information and aids determination of the most appropriate orientations to measure. Pole figure measurements are outside the scope of this chapter; the reader is referred to [28.37] for details. The presence of texture, coupled with elastic anisotropy, can lead to oscillations in d versus sin2 ψ plots. In such cases the diffraction elastic constants vary on a macroscopic scale with orientation, so grains with different DEC are sampled at each (φ and ψ) tilt setting. Methods to account for texture have been developed by various researchers [28.4, 38, 39]. One particularly straightforward method is to use highmultiplicity planes, such as the (732) reflection in steel, which Hauk [28.25] showed provides more linear d versus sin2 ψ behavior than lower-multiplicity peaks (e.g., (211)). Drawbacks to the use of these peaks are that they are less intense, and thus subject to higher statistical error, than higher multiplicity peaks, and require higher-energy x-ray sources (e.g., Mo tubes) than those commonly used and available. Grain Averaging As illustrated in Fig. 28.3, the measured strain can vary substantially from grain to grain in multiphase materials, and even within a single phase due to elastic and/or plastic anisotropy. Thus, in (typical) cases where the macrostress is the desired quantity, it is desirable to sample many grains such that these intergrain variations X-Ray Stress Analysis Stress Gradients The fact that the penetration depth changes with tilting (28.10) can be important if strain/stress gradients exist within the penetration depth, as different (depthaveraged) strain values will be sampled at different tilts. This effect can cause curvature in d versus sin2 ψ plots, which can be analyzed by assuming a particu- Stress error (MPa) 200 180 160 Error Envelope 140 120 100 80 60 40 20 0 400 900 1400 1900 Number of grains sampled in x-ray spot Fig. 28.4 Difference between measured (using x-ray diffraction) and known stresses as a function of the number of sampled grains (after [28.40]). Each point represents a single diffraction measurement, taken across a bent austenitic stainless-steel specimen containing a gradient both in residual stress and grain size (variation from 62–248 grains/mm2 ). Note the inverse relationship between sampled grains and stress error lar depth dependence of the stress [28.3]. This issue can be of particular concern for coatings where microstructural gradients on the μm level are common, and measurement schemes have been developed for such cases [28.41]. If the stress variations over larger ranges than the penetration depth are of interest, these can be measured by destructive layer-removal methods (Sect. 28.6.2) or by nondestructive profiling with higher-energy x-ray or neutron sources (Sect. 28.6.4). 28.6 Case Studies In this section a series of case studies are presented. Laboratory x-ray diffraction is used for biaxial and triaxial stress analysis, for both well-behaved and oscillatory d versus sin2 ψ data. In addition, synchrotron high-energy x-ray diffraction is used for nondestructive stress profiling. 28.6.1 Biaxial Stress Biaxial residual stress measurements with the sin2 ψ method are the most commonly performed diffraction 813 stress measurements. In this example, biaxial x-ray stress measurements were used to help optimize the processing of coil springs for improved fatigue resistance. Large coil springs were heat treated to form tempered martensite, shot peened, and then preset. Figure 28.5 shows the stress diffractometer used to make biaxial residual stress measurements on the springs. Cr-Kα x-rays were used and the tempered martensite 211 diffraction peak was recorded with a linear position-sensitive detector. Specimen tilting was performed with the Ω-goniometry. The positions of the Part C 28.6 are averaged out. The need for averaging is especially important in the case of large grain sizes and/or small x-ray beam sizes, and is exacerbated by the fact that only a small fraction of the irradiated grains diffract. Common methods to enhance averaging are to oscillate (in φ and/or ψ) and/or translate the sample during data collection. The importance of grain averaging was illustrated by Prime et al. [28.40], who examined a bent austenitic steel sample using both neutron and x-ray diffraction. The bending provided a gradient in both residual stress and grain size (determined from optical microscopy), and laboratory x-ray stress measurements were taken along the gradient direction. These x-ray stresses were compared with the calculated macrostresses (which agreed well with bulk neutron results), and the error in x-ray stresses (difference from calculations) is plotted versus number of sampled grains in Fig. 28.4. While there is considerable scatter in the data, there is clearly an inverse trend between stress error and the number of grains sampled. Finally it is noted that the other extreme to polycrystalline averaging is to measure strain in single grains. Such measurements provide fundamental insights into polycrystalline deformation behavior, as well as unique input to deformation models. New methods are being developed to perform such measurements in individual grains within polycrystals (Sect. 28.6.5). 28.6 Case Studies 814 Part C Noncontact Methods Part C 28.6 These measurements were used to study processing methods for spring manufacturing. Figure 28.7 shows the residual stresses with depth below the shot-peened surface for two different heat treatment methods. The stresses below the surface were measured after layer removal and then later corrected for the stress relaxation that occurs with layer removal. One heat treatment method leads to decarburization of the surface, which diminishes the compressive residual stresses the material will support. It should be noted that, with decarburization, the stress-free lattice parameter will vary with depth because of its dependence on the carbon content. Because the stresses are biaxial in nature, accurate stress-free lattice parameters are not needed, simplifying the analysis. 28.6.2 Triaxial Stress Fig. 28.5 Stress diffractometer making measurements on a segment of a coil spring (photo by Yunan Prowato of NHK International Corporation) diffraction peaks were determined from the average of the two half-maximum points on the peak. The d versus sin2 ψ plot is shown in Fig. 28.6 and is seen to be linear. The goodness-of-fit is 0.57, indicating that the estimated errors from the diffraction peak position determinations account for the observed error to a high level of certainty. The slope and intercept from the d versus sin2 ψ plot were used, along with the DEC E/(1 + ν) = 176 GPa, to compute the residual stress as −329 MPa. The data plotted in Fig. 28.8 were obtained from the surface of a ground steel bar [28.36]. The data were collected with a conventional x-ray diffractometer utilizing Cr-Kα radiation. The grinding direction is aligned along the φ = 0◦ direction. Along the grinding direction we see ψ-splitting, indicating the presence of shear stresses, σ13 . Perpendicular to the grinding direction, φ = 90◦ , ψ-splitting is absent, indicating that the shear stresses σ23 are negligible. Along φ = 45◦ , there is moderate ψ-splitting, indicating a component of shear stress perpendicular to the surface along this direction. From the discussion in Sect. 28.3, we know that these stresses arise as microstresses balanced by stresses of opposite sign in another phase. Here, the shear microstresses d-spacing (Å) 1.1706 1.1704 sin2 ψ 2θ d-spacing (Å) 0.041819 0.189426 0.325714 0.479062 156.296 156.439 156.598 156.742 1.170401 1.170096 1.169758 1.169455 Slope Intercept Stress –0.0021916 1.1704953 –328.71 MPa 1.1702 1.17 1.1698 1.1696 1.1694 GOF = 0.573 1.1692 0 0.1 0.2 0.3 0.4 0.5 0.6 sin2 ψ Fig. 28.6 Plot of d versus sin2 ψ for a shot-peened steel coil spring, with data and linear fit parameters included (data by Yunan Prowoto of NHK International Corporation) X-Ray Stress Analysis are balanced by stresses in the carbide phases of the steel. They arise from the shear deformations imposed on the material from the grinding process. Low volume fraction, low crystal symmetry, and considerable peak broadening from the large deformations make measurement of the stresses in the carbide phase extremely difficult and they were not measured here. Fitting this data along with the known d0 and DEC to (28.8)–(28.9) gives the stress tensor [28.12]: ⎛ ⎞ ⎛ ⎞ 527 −8 −40 7 5 1 ⎜ ⎟ ⎜ ⎟ σij = ⎝ −8 592 5 ⎠ ± ⎝5 7 1⎠ MPa . −40 5 102 1 1 3 28.6 Case Studies 815 Residual stress (MPa) 0 –100 –200 –300 – 400 Decarburized –500 –600 –700 (28.23) Here we confirm the presence of the shear stress σ13 and the negligible σ23 . 28.6.3 Oscillatory Data Not Applicable to the Classic Model Frequently, the strains measured in specimens will not behave with orientation as predicted by (28.8). This equation assumes that the strain measured in different directions transforms as a tensor quantity. Because the volume of material diffracting in each orientation is different, this does not have to hold true. If the stresses and strains partition themselves inhomogeneously on average, the measured strains will not behave according to (28.8) and numerically forcing the fit to compute stresses is questionable. Materials with texture due to deformation or growth, such as cold-rolled sheet or thin films [28.42], will frequently show this behavior. Figure 28.9 shows a set of biaxial data on a specimen of ground, cold-rolled steel [28.36]. The data clearly does not show the linear behavior as a function of sin2 ψ, which is partly attributed to the presence of sample texture (not shown). A goodness-of-fit value of 6.2 × 10−49 was computed, indicating an extremely low probability that the deviations from the model fit can be accounted for by the estimated errors on the data points [28.36]. It should be noted that, even if the data do not adequately fit the models, the computed stresses may prove useful as a quality control tool or nondestructive test, even though the interpretation of the results as stresses may not be warranted. Advanced analysis methods, described in Sect. 28.6.5, may in certain cases be used to more accurately determine mean stresses from such data. –900 Not decarburized 0 0.05 0.1 0.15 0.2 Depth (mm) Fig. 28.7 Residual stresses measured with x-ray diffraction on the outside of two coil springs as a function of depth. One spring has experienced decarburization during heat treatment while the other did not. Both raw data (solid lines) and layer-removal-corrected data (dashed lines) are shown 28.6.4 Synchrotron Example: Nondestructive, Depth-Resolved Stress Here we illustrate the use of high-energy x-rays from a synchrotron source to nondestructively measure strain and stress versus depth [28.43]. A schematic of the experiments is shown in Fig. 28.10a. Cylindrical steel specimens of 9 mm diameter were heat treated to form a tempered martensite matrix and nanosize M2 C strengthening precipitates. After heat treatment, specimens were laser peened and then subjected to rolling contact fatigue (RCF). Measurements were performed at the 1-ID beamline at the APS, Argonne National Laboratory. An x-ray energy of 76 keV and a conical slit [28.44] were used to create a diffraction volume of ≈ 20 × 20 × 150 μm3 . The high penetration power at this x-ray energy (τ1/e ≈ 3 mm) allowed for transmission measurements, wherein an area detector was placed after the conical slit to collect diffraction over a plane encompassing (nearly) the axial (ε11 ) and normal (ε33 ) strain directions in a single exposure. Thus, sample tilting was not required to evaluate two principal strain components, as opposed to the reflection-geometry methods cited above. Specimens were translated along the vertical (x3 ) direction, in 20 μm increments, relative to the fixed probe volume, to obtain strain and stress information versus depth, and translated in the Part C 28.6 –800 816 Part C Noncontact Methods d-spacing (Å) 1.1705 1.17 1.1695 1.169 Part C 28.6 θ = 0°, θ = 0°, θ = 90°, θ = 90°, θ = 45°, θ = 45°, 1.1685 1.168 0 0.1 0.2 0.3 0.4 ψ (+) ψ (–) ψ (+) ψ (–) ψ (+) ψ (–) 0.5 sin2 ψ Fig. 28.8 Plot of d versus sin2 ψ for a ground steel spec- imen showing the presence of ψ-splitting, indicating the presence of shear stresses with a component normal to the surface horizontal (x1 ) direction to evaluate different RCF wear tracks. The axial stress σ11 was determined from the strains from the martensite (211) reflection, using DEC for martensite and assuming an equibiaxial strain state (ε11 = ε22 ) (Fig. 28.10b). Significant compressive residd-spacing (Å) 1.1714 1.17135 ual stresses were observed near the surface after heat treatment, and these stresses were further increased after peening, with a maximum value near the surface. Furthermore, the RCF was found to change the residual stress profile, with a subsurface maximum (≈ 100 μm deep) observed under a wear track. It should be noted that both white-beam (energydispersive) x-ray [28.45] and neutron [28.46] techniques can also be used for such deeply penetrating, nondestructive measurements, with some limitations. As in the case study shown, three-dimensional volumes are defined by the incident- and diffracted-beam slits, with neutrons generally having larger probe volumes ( 1 mm3 ) due to flux limitations. White-beam measurements can suffer from high background levels, which can limit the measurable signal due to the limited dynamic range of typical solid-state detectors. 28.6.5 Emerging Techniques and Studies Both neutrons and synchrotron x-ray sources carry the obvious disadvantages of not being located on-site and having limited beamtime availability. However, the results of these measurements can be useful in a general sense, both to check against conventional x-ray stress measurements [28.40,47] and to provide critical tests of, and input to, deformation models [28.48]. Furthermore, they possess unique aspects such as being well suited for in situ measurements, especially under mechanical and/or thermal loading, and permitting nondestructive strain mapping such as in the case study described above. Here, we note emerging studies and techniques in the field of diffraction strain analysis, largely made possible by these sources. 1.1713 1.17125 1.1712 1.17115 1.1711 1.17105 1.171 GOF = 6.2 ×10– 49 1.17095 1.1709 0 0.1 0.2 0.3 0.4 0.5 sin2 ψ Fig. 28.9 Plot of d versus sin2 ψ for a ground steel spec- imen, illustrating data that does not fit the biaxial stress model, with a correspondingly low goodness-of-fit (GOF) value Single-Grain Studies In these studies individual grains within polycrystalline aggregates are evaluated (e.g., [28.49, 50]). The basic concept is to use beam sizes on the order of the grain size (typically 1–100 μm), spatially locate diffracting grains using single-crystal orientation methods, and determine strain using peak shifts. As these techniques rely on small beam sizes, they have so far been done at synchrotron sources. These techniques can also be used for structural characterization, including grain boundary mapping [28.51] and evaluating dislocations quantitatively [28.52] as well as dynamically [28.53]. Finally, these techniques can be combined with spectroscopy techniques such as extended x-ray absorption fine structure (EXAFS) and/or fluorescence for both materials and environmental science applications [28.54]. X-Ray Stress Analysis Combined Strain and Imaging Studies For these studies, typically a relatively large (mm-sized) x-ray beam and an area detector are used to image heterogeneities/defects, and then a smaller beam is used for localized diffraction/strain measurements. For x-rays, the typical imaging contrast mechanism is absorption (radiography), but the coherence of synchrotron sources also allows for more sensitive phase-contrast imaging [28.55]. Examples of work in this area includes deformation studies on composite systems [28.56] and evaluation of creep damage [28.57]. Advanced Quantitative Analysis There have been several recent efforts to extend some fundamental concepts of quantitative texture analysis to describe orientation-dependent elastic strains/stresses in polycrystalline materials [28.39, 65, 66]. Strain measurements from multiple sample orientations are represented on so-called strain pole figures, which are analogous to pole figures used in texture analysis, and which may be inverted to obtain the underlying strain/stress distribution [28.39, 65, 66]. Using these Top view, specimen x1 Peened surface x2 9 mm Conical slit RCF wear tracks 20 µm APS x-rays E = 76 keV x3 2θ ~ 8° x2 Beam stop 100 µm 9 mm Side view, experiment Area detector b) Axial residual stress σ11 (MPa) 0 –200 – 400 –600 –800 –1000 –1200 Unpeened, outside wear track Peened, outside wear track Peened, under wear track –1400 –1600 –1800 0 0.5 1 1.5 Depth into sample x2 (mm) Fig. 28.10 (a) Setup for three-dimensional spatially resolved strain measurements using high-energy synchrotron x-rays, a conical slit, and area detector, with specimen photo shown in inset. (b) Measured residual stresses in the axial (x1 ) direction for three cases techniques, it is possible to examine the micromechanical states of specific crystallite populations within the aggregate in greater detail than with standard methods. Such techniques also provide the modeling community with a powerful motivation and validation tool. 28.7 Summary • X-rays are capable of measuring elastic strain in all crystalline phases present, using shifts in diffraction peak positions. These strains can be converted to elastic stresses using elastic constants. Diffraction also provides information on peak intensities and widths that can be used to analyze texture and plastic deformation, respectively. 817 • • With laboratory x-rays the probed depth is typically microns. Deeper depths can be sampled either by (destructive) layer-removal methods or (nondestructively) using higher-energy x-rays or neutrons. Errors in strain values can arise from many sources, including counting statistics, instrument misalignment, and sample-related sources. Methods to assess experimental error include measuring an Part C 28.7 Strain in Nontraditional Materials While diffraction has traditionally only been used to measure strain in crystalline materials, Windle et al. [28.58] showed that strain can be extracted from amorphous materials (polymers), by measuring changes in peak positions and/or radial distribution functions with applied load. This work has recently been extended with synchrotron studies [28.59, 60] on bulk metallic glasses. Another emerging area is biomaterials studies, including bone [28.61, 62], teeth [28.63], and synthetic coatings on implants [28.64]. These materials (e.g., bone) often contain additional ordering on longer length scales than typically evaluated by wide-angle scattering (up to the μm level [28.62]), and deformation on these levels can be evaluated by small-angle scattering [28.61] (28.1). a) 28.7 Summary 818 Part C Noncontact Methods annealed sample and/or repeating a given measurement. In addition, the use of incorrect diffraction elastic constants to convert strain to stress will give a proportional error in absolute stress values. • With the maturation and anticipated growth of both synchrotron and neutron facilities, their use for advanced strain and stress analysis is expected to continue well into the future. 28.8 Further Reading General information on diffraction stress measurements can be found in [28.3–8]. Part C 28 References 28.1 28.2 28.3 28.4 28.5 28.6 28.7 28.8 28.9 P.J. Withers: Use of synchrotron x-ray radiation for stress measurement. In: Analysis of Residual Stress by Diffraction Using Neutron and Synchrotron Radiation, ed. by M.E. Fitzpatrick, A. Lodini (Taylor and Francis, London 2004) pp. 170–189 H.F. Poulsen, S. Garbe, T. Lorentzen, D. Juul Jensen, F.W. Poulsen, N.H. Andersen, T. Frello, R. Feidenhans’l, H. Graafsma: Applications of high-energy synchrotron radiation for structural studies of polycrystalline materials, J. Synchrotron Radiat. 4, 147–54 (1997) I.C. Noyan, J.B. Cohen: Residual Stress: Measurement by Diffraction and Interpretation (Springer, New York 1987) V. Hauk: Structural and Residual Stress Analysis by Nondestructive Methods (Elsevier Science B.V., Amsterdam 1997) M.E. Hilley, J.A. Larson, C.F. Jatczak, R.E. Ricklefs (Eds.): Residual Stress Measurement by X-ray Diffraction. In: SAE Information Report J784a, ed. by M.E. Hilley (Society of Automotive Engineers, Warrendale 1971) F. Lecroisey, B. Miege, A. Saint-Etienne: La Mesure de contraintes residuelles: methode de determination par rayons X, Memoires Techniques du CETIM, Vol. 33 (Centre Technique de Industries Mechaniques, 1978) P. Prevey: X-ray diffraction residual stress techniques. In: Metals Handbook 9th Edition, Vol. 10 (American Society for Metals: Metals Park, Ohio 1986) pp. 380–392 M.E. Fitzpatrick, A.T. Fry, P. Holdway, F.A. Kandil, J. Shackleton, L. Suominen: Good Practice Guide No. 52: Determination of Residual Stresses by X-ray Diffraction (National Physical Laboratory, Teddington, UK 2005) pp. 1–68 R.W. Hertzberg: Deformation and Fracture Mechanics of Engineering Materials (Wiley, New York 1996) 28.10 28.11 28.12 28.13 28.14 28.15 28.16 28.17 28.18 28.19 28.20 28.21 K. Inal, R. Pesci, J.L. Lebrun, O. Diard, R. Masson: Grain and phase stress criteria for behaviour and cleavage in duplex and bainitic steels, Fatigue Fract. Eng. Mater. Struct. 29(9,10), 685–696 (2006) G.W.C. Kaye, T.H. Laby: Tables of Physical and Chemical Constants (Longman, London 1995) R.A. Winholtz, J.B. Cohen: Generalized leastsquares determination of triaxial stress states by x-ray diffraction and associated errors, Aust. J. Phys. 41, 189–199 (1988) C. Genzel: X-ray stress gradient analysis in thin layers – problems and attempts at their solution, Phys. Status Solidi A 159, 283–296 (1997) I.C. Noyan: Determination of the unstressed lattice parameteraofor (triaxial) residual stress determination by x-rays, Adv. X-ray Anal. 28, 281–288 (1985) R.A. Winholtz, J.B. Cohen: Separation of macroand micro-stresses in plastically deformed 1080 steel, Adv. X-ray Anal. 32, 341–353 (1989) W.H. Press, B.P. Flannery, S.A. Teukolsky, W.T. Vetterling: Numerical Recipes: The Art of Scientific Computing (Cambridge University Press, Cambridge 1986) M.R. James, J.B. Cohen: Study of the precision of x-ray stress analysis, Adv. X-ray Anal. 20, 291–307 (1977) R.A. Winholtz: Macro and micro-stresses. In: Encyclopedia of Materials, ed. by K.H.J. Buschow et al. (Pergamon, Oxford 2001) W. Voigt: Lehrbuch der Kristallphysik (Teubner, Leipzig, Berlin 1910) A. Reuss, Z. Agnew: Calculation of flow limits of mixed crystals on the basis of plasticity of single crystal, Math. Mech. 9, 49–52 (1929) H. Neerfeld: Zur Spannungsberechnung aus rontgenographische Dehnungsmessungen, In: Mitt. K.-Wilh.-Inst. Eisenforschg. 24, 61–70 (1942) X-Ray Stress Analysis 28.22 28.23 28.24 28.25 28.26 28.28 28.29 28.30 28.31 28.32 28.33 28.34 28.35 28.36 28.37 28.38 28.39 28.40 28.41 28.42 28.43 28.44 28.45 28.46 28.47 28.48 28.49 28.50 28.51 J.V. Bernier, M.P. Miller: A direct method for the determination of the mean orientationdependent elastic strains and stresses in polycrystalline materials from strain pole figures, J. Appl. Cryst. 39, 358–368 (2006) M.B. Prime, P. Rangaswamy, M.R. Daymond, T.G. Abeln: Several methods applied to measuring residual stress in a known specimen. In: SEM Spring Conference on Experimental and Applied Mechanics (Society for Experimental Mechanics, Houston 1998) pp. 497–499 C. Genzel, W. Reimers: Depth-resolved X-ray residual stress analysis in PVD (Ti, Cr) N hard coatings, Z. Metallkunde 94(2), 655–661 (2003) J. Almer, U. Lienert, R.L. Peng, C. Schlauer, M. Oden: Strain and texture analysis of coatings using high-energy X-rays, J. Appl. Phys. 94(2), 697–702 (2003) Y. Qian, J. Almer, U. Lienert, G.B. Olson: Nondestructive Residual Stress Distribution Measurement in Nano-structured Ultra-high Strength Gear Steels, Fifth Int. Conf. Sychrotron Radiat. Mater. Sci. (Chicago 2006) pp. 27–28 U. Lienert, S. Grigull, Å. Kvick, R.V. Martins, H.W. Poulsen: Three Dimensional Strain Measurements in Bulk Materials with High Spatial Resolution, ICRS-6 (Oxford 2000) pp. 1050–1057 A. Steuwer, J.R. Santisteban, M. Turski, P.J. Withers, T. Buslaps: High-resolution strain mapping in bulk samples using full-profile analysis of energy-dispersive synchrotron X-ray diffraction data, J. Appl. Cryst. 37, 883–889 (2004) M.A.M. Bourke, D.C. Dunand, E. Ustundag: SMARTS – a spectrometer for strain measurement in engineering materials, Appl. Phys. A-Mater. Sci. Process. 74, S1707–S1709 (2002) D. Stefanescu, A. Bouzina, M. Dutta, D.Q. Wang, M.E. Fitzpatrick, L. Edwards: Comparison of residual stress measurements using neutron and X-ray diffraction around cold expanded holes, J. Neutron Res. 9(2), 399–404 (2001) P.R. Dawson, D.E. Boyce, R. Rogge: Issues in modeling heterogeneous deformations in polycrystalline metals using multiscale approaches, CMES-Comput. Model. Eng. Sci. 10(2), 123–141 (2005) G.E. Ice, B.C. Larson: 3D X-ray crystal microscope, Adv. Eng. Mater. 2(2), 643–646 (2000) H.F. Poulsen, S.F. Nielsen, E.M. Laurdisen, S. Schmidt, R.M. Suter, U. Lienert, L. Margulies, T. Lorentzen, D. Juul Jensen: Three-dimensional maps of grain boundaries and the stress state of individual grains in polycrystals and powders, J. Appl. Cryst. 34, 751–756 (2001) H.F. Poulsen, X. Fu, E. Knudsen, E.M. Lauridsen, L. Margulies, S. Schmidt: 3DXRD – Mapping grains and their dynamics in 3 dimensions, Mater. Sci. Forum 467, 1363–1372 (2004) 819 Part C 28 28.27 R. Hill: The elastic behavior of polycrystalline aggregate, Proc. Phys. Soc. London 65, 349–354 (1952) F. Bollenrath, V. Hauk, E.H. Muller: On the calculation of polycrystalline elasticity constants from single crystal data, Z. Metallkunde 58, 76–82 (1967) A.C. Vermeulen: An elastic constants database and XEC calculator for use in XRD residual stress analysis, Adv. X-ray Anal. 44, 128–133 (2001) V. Hauk: In: Residual Stress and Stress Relaxation, ed. by E. Kula, V. Weiss (Plenum, New York 1982) ASTM: Standard test method for determining the effective elastic parameter for x-ray diffraction measurements of residual stress, E1426. In: Annual Book of ASTM Standards (ASTM, West Cohshohocken 2007) P. Dawson, D. Boyce, S. MacEwen, R. Rogge: On the influence of crystal elastic moduli on computed lattice strains in AA-5182 following plastic straining, Mater. Sci. Eng. A 313, 123–144 (2001) M.R. Daymond: The determination of a continuum mechanics equivalent elastic strain from the analysis of multiple diffraction peaks, J. Appl. Phys. 96(2), 4263–4272 (2004) M. v. Leeuwen, J.-D. Kamminga, E.J. Mittemeijer: Diffraction stress analysis of thin films: Modeling and experimental evaluation of elastic constants and grain interaction, J. Appl. Phys. 86, 1904–1914 (1999) F. Witt, R.W. Vook: Thermally induced strains in cubic metal films, J. Appl. Phys. 39, 2773–2776 (1968) A. Segmüller, M. Murakami: X-ray diffraction analysis of strains and stresses in thin films, Treatise Mater. Sci. Technol. 27, 143–197 (1988) I.C. Noyan, T.C. Huang, B.R. York: Residual stress/strain analysis in thin films by x-ray diffraction, Crit. Rev. Solid State Mater. Sci. 20(2), 125–177 (1995) ASTM: Standard test method for verifying the alignment of x-ray diffraction instrumentation for residual stress measurement, E915. In: Annual Book of ASTM Standards (ASTM, West Cohshohocken 2007) R.H. Marion: X-Ray Stress Analysis of Plastically Deformed Metals, PhD Thesis (Northwestern University, Evanston 1973) J. Jo, R.W. Hendricks: Diffractometer misalignment errors in x-ray residual stress measurements, J. Appl. Cryst. 24, 878–887 (1991) T.A. Lohkamp, R.A. Winholtz: Assessing the validity of diffraction stress data with the goodness-of-fit statistic, Adv. X-ray Anal. 39, 281–289 (1997) U.F. Kocks, C.N. Tome, H.-R. Wenk: Texture and Anisotropy (Cambridge Univ. Press, Cambridge 1998) S. Matthies, H.G. Priesmeyer, M.R. Daymond: On the diffractive determination of single-crystal elastic constants using polycrystalline samples, J. Appl. Cryst. 34, 585–601 (2001) References 820 Part C Noncontact Methods 28.52 28.53 28.54 Part C 28 28.55 28.56 28.57 R. Barabash, G.E. Ice, B.C. Larson, G.M. Pharr, K.-S. Chung, W. Yang: White microbeam diffraction from distorted crystals, Appl. Phys. Lett. 79(2), 749–51 (2001) B. Jakobsen, H.F. Poulsen, U. Lienert, J. Almer, S.D. Shastri, H.O. Sorensen, C. Gundlach, W. Pantleon: Formation and subdivision of deformation structures during plastic deformation, Science 312(2), 889–892 (2006) N. Tamura, R.S. Celestre, A.A. MacDowell, H.A. Padmore, R. Spolenak, B.C. Valek, N.M. Chang, A. Manceau, J.R. Patel: Submicron x-ray diffraction and its applications to problems in materials and environmental science, Rev. Sci. Instrum. 73(2), 1369–1372 (2002) J. Baruchel, P. Cloetens, J. Hartwig, W. Ludwig, L. Mancini, P. Pernot, M. Schlenker: Phase imaging using highly coherent X-rays: radiography, tomography, diffraction topography, J. Synchrotron Radiat. 7, 196–201 (2000) R. Sinclair, M. Preuss, P.J. Withers: Imaging and strain mapping fibre by fibre in the vicinity of a fatigue crack in a Ti/SiC fibre composite, Mater. Sci. Technol. 21(2), 27–34 (2005) A. Pyzalla, B. Camin, T. Buslaps, M. Di Michiel, H. Kaminski, A. Kottar, A. Pernack, W. Reimers: Simultaneous tomography and diffraction analysis of creep damage, Science 308(2), 92–95 (2005) 28.58 28.59 28.60 28.61 28.62 28.63 28.64 28.65 28.66 M. Pick, R. Lovell, A.H. Windle: Detection of elastic strain in an amorphous polymer by X-ray scattering, Nature 281, 658–659 (1979) H.F. Poulsen: Measuring strain distributions in amorphous materials, Nat. Mater. 1, 33–36 (2005) T.C. Hufnagel, R.T. Ott, J. Almer: Structural aspects of elastic deformation of a metallic glass, Phys. Rev. B 73(6), 64204 (2006) J.D. Almer, S.R. Stock: Micromechanical response of mineral and collagen phases in bone, J. Struct. Biol. 157, 365–370 (2007) H.S. Gupta, P. Messmer, P. Roschger, S. Bernstorff, K. Klaushofer, P. Fratzl: Synchrotron diffraction study of deformation mechanisms in mineralized tendon, Phys. Rev. Lett. 93(15), 158101 (2004) T. Kallaste, J. Nemliher: Apatite varieties in extant and fossil vertebrate mineralized tissues, J. Appl. Cryst. 38, 587–594 (2005) T. Pirling: Stress determination with high lateral resolution using neutron diffraction, Thermec’2003, Vol. 1–5 (2003) pp. 3975–3980 Y.D. Wang, R.L. Peng, R.L. McGreevy: A novel method for constructing the mean field of grainorientation-dependent residual stress, Philos. Mag. Lett. 81(2), 153–163 (2001) H. Behnken: Strain-function method for the direct evaluation of intergranular strains and stresses, Phys. Stat. Sol. (a) 177(2), 401–418 (2000)
© Copyright 2025 Paperzz