+ All Categories
Home > Documents > In-situ synchrotron radiation studies of the deformation ...€¦ · martensitic phase...

In-situ synchrotron radiation studies of the deformation ...€¦ · martensitic phase...

Date post: 04-Apr-2021
Category:
Upload: others
View: 0 times
Download: 0 times
Share this document with a friend
8
In-situ synchrotron radiation studies of the deformation mechanisms in metastable austenitic TRIP/TWIP steels C Ullrich 1 , S. Martin 1 , C Schimpf 1 , A Stark 2 , N Schell 3 and D Rafaja 1 1 Institute of Materials Science, TU Bergakademie Freiberg, Gustav-Zeuner-Str. 5, 09599 Freiberg, Germany 2 Institute of Materials Research, Helmholtz-Zentrum Geesthacht, Max-Planck-Str. 1, 21502 Geesthacht, Germany 3 Structural Research on New Materials, Helmholtz-Zentrum Geesthacht Outstation at DESY, Hamburg, Germany E-mail: [email protected] Abstract. Many deformation mechanisms in metastable austenitic steels are governed by the stacking fault energy. In this study, microstructure changes in Cr-Mn-Ni steels with different Ni contents (3, 6 and 9 wt.%) and thus with different stacking fault energies (SFE = 7.5, 16.7 and 24.3 mJ/m²) are investigated using in-situ synchrotron X-ray diffraction under uniaxial compression. The martensitic phase transformation is tracked by the evolution of the phase composition. The changes in the defect densities (microstrain due to dislocations, stacking fault probabilities) are monitored and correlated with the lattice strain in austenite. The onset of - martensite formation was found to be coupled to an abundant stacking fault formation and to their dense arrangement. The ´-martensite forms mainly in the steels with low SFE; it was not found in the steel with SFE > 20 mJ/m². Increasing SFE reduces the stacking fault probability and shifts the onset of the stacking fault formation to higher deformations. Concurrently, higher SFE enhances the formation of dislocations and their slip activity. The interplay of the microstructure features and defects and their effect on the stress-strain curve are discussed. 1. Introduction High-alloy metastable austenitic steels, which show the TRIP (Transformation Induced Plasticity) or TWIP (Twinning Induced Plasticity) effect, exhibit outstanding properties, especially with regard to the strength, ductility and absorption of mechanical energy during plastic deformation [1-3]. Their deformation behavior is largely controlled by the stacking fault energy (SFE) of austenite, which depends on the chemical composition of the steel and on the deformation temperature [4-7]. For very low SFE (< 20 mJ/m²), the dissociation of partial dislocations, the formation of stacking faults and their bunching in deformation bands are observed. A high stacking fault density leads to a local arrangement of stacking faults in the deformation bands. This arrangement resembles a hexagonal crystal structure, which is identified as -martensite by diffraction methods [8-10]. At higher deformations, ´-martensite forms – preferentially inside of the deformation bands or at their intersections [9-11]. Increasing SFE postpones the stacking fault formation and the formation of ´- martensite to higher deformation stages. For SFE above 20 mJ/m², the formation of -martensite is gradually replaced by twinning. The twins are also produced via dissociation of partial dislocations and subsequent accumulation of stacking faults [8, 12-15], but the stacking fault density is higher in
Transcript
  • In-situ synchrotron radiation studies of the deformation

    mechanisms in metastable austenitic TRIP/TWIP steels

    C Ullrich1, S. Martin1, C Schimpf1, A Stark2, N Schell3 and D Rafaja1

    1 Institute of Materials Science, TU Bergakademie Freiberg, Gustav-Zeuner-Str. 5,

    09599 Freiberg, Germany 2 Institute of Materials Research, Helmholtz-Zentrum Geesthacht, Max-Planck-Str. 1,

    21502 Geesthacht, Germany 3 Structural Research on New Materials, Helmholtz-Zentrum Geesthacht Outstation at

    DESY, Hamburg, Germany

    E-mail: [email protected]

    Abstract. Many deformation mechanisms in metastable austenitic steels are governed by the

    stacking fault energy. In this study, microstructure changes in Cr-Mn-Ni steels with different

    Ni contents (3, 6 and 9 wt.%) and thus with different stacking fault energies (SFE = 7.5, 16.7

    and 24.3 mJ/m²) are investigated using in-situ synchrotron X-ray diffraction under uniaxial

    compression. The martensitic phase transformation is tracked by the evolution of the phase

    composition. The changes in the defect densities (microstrain due to dislocations, stacking fault

    probabilities) are monitored and correlated with the lattice strain in austenite. The onset of -

    martensite formation was found to be coupled to an abundant stacking fault formation and to

    their dense arrangement. The ´-martensite forms mainly in the steels with low SFE; it was not

    found in the steel with SFE > 20 mJ/m². Increasing SFE reduces the stacking fault probability

    and shifts the onset of the stacking fault formation to higher deformations. Concurrently,

    higher SFE enhances the formation of dislocations and their slip activity. The interplay of the

    microstructure features and defects and their effect on the stress-strain curve are discussed.

    1. Introduction

    High-alloy metastable austenitic steels, which show the TRIP (Transformation Induced Plasticity) or

    TWIP (Twinning Induced Plasticity) effect, exhibit outstanding properties, especially with regard to

    the strength, ductility and absorption of mechanical energy during plastic deformation [1-3]. Their

    deformation behavior is largely controlled by the stacking fault energy (SFE) of austenite, which

    depends on the chemical composition of the steel and on the deformation temperature [4-7]. For very

    low SFE (< 20 mJ/m²), the dissociation of partial dislocations, the formation of stacking faults and

    their bunching in deformation bands are observed. A high stacking fault density leads to a local

    arrangement of stacking faults in the deformation bands. This arrangement resembles a hexagonal

    crystal structure, which is identified as -martensite by diffraction methods [8-10]. At higher

    deformations, ´-martensite forms – preferentially inside of the deformation bands or at their

    intersections [9-11]. Increasing SFE postpones the stacking fault formation and the formation of ´-

    martensite to higher deformation stages. For SFE above 20 mJ/m², the formation of -martensite is

    gradually replaced by twinning. The twins are also produced via dissociation of partial dislocations

    and subsequent accumulation of stacking faults [8, 12-15], but the stacking fault density is higher in

  • twinned austenite, where the stacking faults appear on every close-packed plane, than in -martensite

    accommodating the stacking faults on every second close-packed plane. For SFE > 40 mJ/m², the

    deformation is mostly accomplished through the slip of perfect dislocations, because the faulting and

    the martensitic transformation are suppressed and postponed in this SFE range [6].

    The aim of the present study is to analyze the microstructure evolution, the corresponding

    deformation mechanisms and the interaction between different types of microstructure defects and the

    martensitic phase transformation plus their effects on the mechanical behavior during deformation of

    three metastable austenitic Cr-Mn-Ni-steels with varying compositions and SFEs.

    2. Experimental

    Three fine-grained metastable austenitic 16Cr-7Mn-xNi steels were studied. Their composition varied

    in the Ni content, which amounted to 3, 6 and 9 wt.%. The SFE of these steels is known from in-situ

    XRD bending tests to be 7.5, 16.7 and 24.3 mJ/m², respectively [16, 17]. The samples with 3 and 9

    wt.% Ni were produced by powder metallurgy using field assisted sintering technique, whereas the

    16Cr-7Mn-6Ni sample was taken from a hot-rolled bar. In-situ compression tests at room temperature

    were done using a deformation dilatometer DIL 805A/D (TA Instruments, formerly Bähr), which is

    installed at the beamline P07, PETRA III at DESY synchrotron radiation facility. The cylindrical

    compression samples had a diameter of 4 mm, therefore the maximum stress applied with the 20 kN

    maximum load device was 1590 MPa. The samples were compressed stepwise with a short holding (5

    s) needed for recording of the diffraction patterns. The high-energy synchrotron radiation (100 keV,

    = 0.1235 Å) allowed to conduct the diffraction experiments in transmission geometry. The patterns

    were recorded using a 2D detector (PerkinElmer). In this setup, the angle between the diffraction

    vector and the uniaxial load direction changes with the azimuthal angle (along the Debye rings)

    between a half of the diffraction angle and 90°, thus the dependence of the line positions on the

    azimuthal angle can be used for the analysis of the lattice deformation upon deformation. The electron

    channeling contrast imaging (ECCI) and electron back-scatter diffraction (EBSD) on deformed

    samples were done post mortem in a scanning electron microscope Zeiss LEO 1530 FEGSEM at 20

    kV acceleration voltage using the HKL Channel 5 EBSD software and a Nordlys detector.

    Figure 1. True stress-true strain curves

    measured for the metastable austenitic steel

    samples under study.

    Figure 2. Diffraction patterns of sample 16Cr-

    7Mn-6Ni after 21% compression obtained from the

    integration of 5° sections in a single 2D diffraction

    pattern. The angles in the legend refer to the

    azimuthal position of the respective section with

    respect to the load direction. Measured intensities

    are plotted by dots, refined intensities by lines. The

    peaks are labelled by diffraction indices.

  • 3. Data analysis

    The 2D diffraction images were integrated within azimuthal sections having the width of 5°, and

    analyzed using the Rietveld refinement with the software MAUD [18]. The Rietveld analysis revealed

    the phase fractions of fcc austenite (space group 𝐹𝑚3̅𝑚), hcp -martensite (𝑃63/𝑚𝑚𝑐) and bcc ´-martensite (𝐼𝑚3̅𝑚), and the macroscopic lattice deformation, the stacking fault probability and the microstrain in austenite. The austenite lattice parameters for the undeformed steels were determined to

    be 3.5954(7) Å for 16Cr-7Mn-3Ni, 3.5930(6) Å for 16Cr-7Mn-6Ni and 3.5923(7) Å for 16Cr-7Mn-9Ni, respectively. The macroscopic lattice deformation was described using the Moment Pole Stress

    Model [19], which includes the anisotropy of the elastic constants and allows therefore the description

    of the anisotropic line shift due to strain. Elastic constants of austenite were taken from Refs. [20] and

    [21]. For the description of the anisotropic line broadening, the Popa model [22] was applied. It was

    assumed that the refined microstrain, which corresponds to the variation of lattice plane distances, is

    caused by dislocations and thus proportional to the square root of the dislocation density [23, 24]. The

    stacking fault probability was concluded from the line shift and from the line broadening according to

    the Warren model [25]. Preferred orientation of crystallites during deformation was described by

    orientation distribution functions, which were refined using the E-WIMV method that is incorporated

    in the MAUD routine. For the description of the texture, a model developed by Wenk et al. [26]

    assuming a fiber texture was employed.

    4. Results and discussion

    The true stress-strain curves obtained from mechanical experiments (figure 1) show differently

    pronounced strain hardening in the austenitic steels with different chemical compositions. The lowest

    stress level was observed in the steel with 9 wt.% Ni, the highest stress level in the steels with less Ni

    and thus with lower SFE. The stress-strain curve of the steel 16Cr-7Mn-3Ni has a distinct sigmoidal

    shape, indicating the formation of a high amount of bcc ’-martensite. The steel 16Cr-7Mn-6Ni

    possesses the highest yield strength, probably due to an incompletely recrystallized microstructure. As

    the experiments were conducted in force-controlled mode, the maximum technical stress was equal for

    all samples, but the deformations achieved at the respective stress level differ. The smallest sample

    deformation was observed in sample 16Cr-7Mn-3Ni, the largest one in 16Cr-7Mn-9Ni. This strength

    and hardening behavior can be expected, because a lower SFE and a larger extent of martensitic

    transformation in steels with a lower Ni content increase the stress level and reduce the deformability

    [1, 2]. Still, the mechanical behaviors of samples 16Cr-7Mn-3Ni and 16Cr-7Mn-6Ni differ less than

    expected. The possible reasons for this similarity will be discussed in detail later, as they are also

    reflected in the results of in situ synchrotron radiation diffraction.

    Figure 3. Evolution of the phase fractions as obtained from the Rietveld refinement. a) all phases,

    b) -martensite.

  • Figure 2 shows exemplarily the diffraction patterns that were extracted from the azimuthal sections

    of the 2D scans recorded for the sample 16Cr-7Mn-6Ni at 21 % compression together with the

    respective Rietveld fits. The evolution of the phase fractions under compressive load, which was

    obtained from the Rietveld refinement, is summarized in figure 3. The initial microstructures of all

    three steels were fully austenitic. In the steels with 3 and 6 wt.% Ni, the formation of hcp -martensite

    starts very soon after the onset of the plastic deformation. The diffraction lines from -martensite were

    detected already at 0.8 % compression for both compositions. The maximum amount of -martensite

    (19 vol.%) was attained at ca. 17 % compression. Beyond this maximum, a slight reduction of the -

    martensite fraction was observed for both steels, as -martensite transforms partially into ´-

    martensite. The formation of the bcc ´-martensite started between 2 % and 3 % compression. The

    maximum content of ´-martensite amounted 40 vol.% for 16Cr-7Mn-3Ni (at 21 % compression) and

    32 vol.% for 16Cr-7Mn-6Ni (at 23 % compression). The 16Cr-7Mn-9Ni steel revealed the first -

    martensite peaks at 5 % compression, the maximum fraction of 12 vol.% was reached at the maximum

    compression of 33 %. In this sample, the bcc ´-martensite did not form in the investigated

    deformation range, as the high Ni content stabilizes austenite, increases its SFE and lowers the driving

    force for the (’) martensitic transformation.

    The effect of the Ni content on the SFE can also be seen on different stacking fault probabilities

    (SFP) in individual samples (figure 4). In sample 16Cr-7Mn-3Ni, an enormous increase of SFP was

    observed beyond 0.7 % compression. This means that the plastic deformation of this steel is

    dominated by the stacking fault formation. For 16Cr-7Mn-6Ni, a similar course is observed, with a

    comparable, i.e., early start of the SFP increase. The final values of SFP are even higher for 16Cr-

    7Mn-6Ni than for 16Cr-7Mn-3Ni. This behavior is probably again due to the incompletely

    recrystallized initial microstructure, which provides more nucleation sites for stacking fault formation

    and also creates higher levels of internal stress, which is necessary to dissociate partial dislocations

    [16, 17]. Additionally, a higher SFP was observed in 16Cr-7Mn-6Ni than in 16Cr-7Mn-3Ni already in

    the non-deformed state. In steel 16Cr-7Mn-9Ni, almost no stacking faults are present prior to the 3 %

    compression. The deformation in this stage runs exclusively via slip of perfect dislocations. The

    critical stress for wide dissociation of partial dislocations is reached first at higher deformations (>

    3%), after a certain amount of work hardening due to the dislocations is achieved. As the microstrain

    in sample 16Cr-7Mn-9Ni is relatively low (figure 5), these dislocations form probably dislocation

    structures with partially compensated strain fields [23]. Although the plastic deformation of low-SFE

    alloys is dominated by stacking faults, at least initial plastic deformation by perfect dislocation slip is

    generally necessary for the initiation of the stacking fault formation.

    Figure 4. Stacking fault probability for isolated

    intrinsic stacking faults in austenite obtained by

    using the Warren model [25].

    Figure 5. Evolution of the squared microstrain

    〈ε1002 〉 in austenite.

  • The squared microstrain plotted in figure 5 is as a measure of the dislocation density [23].

    However, as already mentioned above, the microstrain values are possibly affected by the correlations

    in the dislocation positions. For the steel with the lowest SFE (16Cr-7Mn-3Ni), an initial increase in

    squared microstrain up to ca. 7 % compression is observed. Afterwards the increase is lowered, and

    the microstrain almost stagnates between 12 and 16 % compression. This indicates that the dislocation

    slip is a less important mechanism of the plastic deformation in this compression range, as the

    formation of stacking faults is dominant (cf. figure 4). For 16Cr-7Mn-6Ni, the microstrain increases

    during the whole deformation process. High initial microstrain indicates again that the sample was not

    completely recrystallized, as it had a high dislocation density in the non-deformed state. The high

    dislocation density in the starting state may also account for the early onset of the martensitic

    transformation that occurs at a similar deformation like in the steel 16Cr-7Mn-3Ni (figure 3) despite

    its lower SFE. Nevertheless, the microstrain in 16Cr-7Mn-6Ni attains a clearly higher level than in

    sample 16Cr-7Mn-3Ni, and shows a constant increase, what indicates an enhanced dislocation slip

    activity. The steel containing 9 wt.% Ni has nearly the same initial microstrain as steel 16Cr-7Mn-3Ni,

    and shows a permanent increase of the microstrain. At highest deformations, the maximum

    microstrain in sample 16Cr-7Mn-9Ni is comparable with the maximum microstrain in sample 16Cr-

    7Mn-6Ni. High SFE in sample 16Cr-7Mn-9Ni retards the stacking fault formation and promotes the

    dislocation slip activity of undissociated perfect dislocations instead. New dislocations are formed

    during the whole deformation process, even in the deformation range, in which the stacking faults

    form. By comparing the microstrain evolution in these three steels, it is obvious that with decreasing

    SFE, the portion of dislocation slip for plastic deformation decreases, because other deformation

    mechanisms such as stacking faults and phase transformation excel them.

    Other outputs of the in situ diffraction measurement under deformation were the lattice stress in

    austenite and its dependence on the macroscopic deformation (figure 6a) as a complement to the

    stress-strain curve from figure 1. The lattice stress was obtained from the anisotropic shift of the

    diffraction lines measured at different angles between the diffraction vector and the compression

    direction [26, 27]. In the elastic range (approx. below 0.5 % compression), the mechanical stress and

    the lattice stress in austenite agree quite well. After the plastic deformation starts, the mechanical

    stress becomes higher than the lattice stress in austenite. Whereas the mechanical stress is a sum of

    stresses needed for both, elastic and plastic deformations of all phases in the sample, the lattice stress

    comprises only the elastic component in the austenite. Thus, the difference between both stresses

    (figure 6b) is a measure of the driving force for the plastic deformation. The stress difference increases

    Figure 6. (a) Lattice stress in

    austenite (in the compression

    direction) obtained from the shift of

    the diffraction lines and (b) the

    difference between mechanical true

    stress (cf. fig. 1) and the lattice

    stress.

  • during the compression, which confirms the expected strain hardening of the samples due to the

    increasing density of microstructure defects, intense interaction between their strain fields and

    deformation induced phase transformations.

    The extents of individual hardening mechanisms follow from the comparison of the stress

    difference (figure 6b) with the phase fractions (figure 3), stacking fault probability (figure 4) and

    squared microstrain (figure 5). The steep increase of the stress difference that was observed in samples

    16Cr-7Mn-3Ni and 16Cr-7Mn-6Ni at low compressions is mainly caused by the formation of stacking

    faults and -martensite. As -martensite originates from faulted austenite with dense packed stacking

    faults [8], the formation of stacking faults and -martensite are coupled phenomena that have a similar

    effect on the stress-strain curve. The main contribution to the increase of the stress difference and thus

    to the strain hardening in sample 16Cr-7Mn-3Ni at the deformations > 9 % comes from the formation

    of ´-martensite. The formation of dislocations plays a minor role. This model is also applicable for

    sample 16Cr-7Mn-9Ni, in which the increase of the stress difference is delayed in the same manner as

    the formation of the stacking faults and -martensite.

    In steel 16Cr-7Mn-9Ni, the stacking fault probabilities are much lower than in steels with lower Ni

    contents, despite the low yield strength. As the plasticity of the steel with 9 wt.% Ni can be explained

    neither by high dislocation density nor by formation of random stacking faults nor by the -martensite

    formation, the only possible deformation mechanisms are the absence of ’-martensite and the

    presence of twins. The presence of twins affects both, the broadening and asymmetry of X-ray

    diffraction lines, but the X-ray diffraction is generally less sensitive to the presence of twins than to

    the presence of stacking faults and dislocations. Therefore, the existence of twins in deformed samples

    was examined by ECCI and EBSD in SEM.

    Representative ECCI microstructure images are shown in Figure 7. For the 16Cr-7Mn-3Ni,

    deformation bands generally occur on several slip planes, providing nucleation sites for the ´-

    a) 16Cr-7Mn-3Ni

    b) 16Cr-7Mn-6Ni

    c) 16Cr-7Mn-9Ni

    Figure 7. ECCI micrographs of the samples after

    deformation. a) Deformation bands (marked by

    straight lines), stacking faults (arrows) and ´-

    martensite (circles) in 16Cr-7Mn-3Ni. b) Very

    high defect density with dislocations (blue

    arrows), dense deformation bands (lines) and ´-

    martensite (circles) in 16Cr-7Mn-6Ni. c)

    Dislocation structures (blue arrows) and

    deformation bands (lines) identified as -

    martensite and twins in 16Cr-7Mn-9Ni.

  • martensite formation. Additionally, nucleation sites can also be found inside deformation bands

    consisting of -martensite. In the remaining austenite matrix, a high concentration of stacking faults

    can be detected, but the dislocation slip and accordingly the formation of dislocation structures is less

    noticeable. The microstructure of the steel 16Cr-7Mn-6Ni also contains a high stacking fault density in

    austenite, many deformation bands on several slip systems and nuclei of ´-martensite. Additionally, the pronounced contrasts from dislocations evidence a highly partitioned microstructure and verify the

    high dislocation density found by the XRD analysis. The ECCI image of the steel 16Cr-7Mn-9Ni

    clearly reveals the dominant role of the dislocation slip in the deformation of this sample. Some areas

    are free of typical deformation bands, while tangled dislocations and dislocation walls are visible. The

    EBSD measurements revealed considerable misorientations inside of grains, which can be interpreted

    as local lattice rotations due to presence of dislocation structures. The interaction between deformation

    bands and dislocations results in a curvature of the deformation bands. Furthermore, the results of

    EBSD demonstrate that the deformation bands consist partly of twins and partly of -martensite. The

    ´-martensite was not found using EBSD, which is also in a good agreement with the result of the XRD phase analysis.

    5. Conclusions

    The microstructure evolution in metastable austenitic steels with different stacking fault energies was

    studied using in-situ synchrotron XRD under compressive load. The formation of dislocations,

    stacking faults and twins, and the martensitic transformation of austenite to -martensite and ´-

    martensite were correlated with the stress-strain curves obtained from the mechanical experiments and

    with the elastic lattice deformation obtained from the diffraction experiments. In all steels under study,

    the plastic deformation starts with the production of perfect dislocations. The significance of

    individual deformation mechanisms at higher deformations depends strongly on the stacking fault

    energy, which controls the onset of the rapid dissociation of partial dislocations and the stacking fault

    formation. Especially in steels with higher stacking fault energies, some critical stress must be reached

    to be able to activate the partial dislocation dissociation and the stacking fault formation. The critical

    stress can be produced by interacting microstructure defects, in particular by interaction of perfect

    dislocations. With increasing Ni content and stacking fault energy, the stacking fault formation and the

    martensitic phase transformations are retarded, the formation of ´-martensite is suppressed, and the importance of the dislocation slip increases. A higher initial defect density, as present in sample 16Cr-

    7Mn-6Ni, brings about higher defect densities (stacking faults and dislocations) and an earlier onset of

    stacking fault formation and phase transformation during deformation, mainly due to a higher level of

    lattice stress in the austenite.

    Acknowledgements

    This work was funded by the German Research Foundation (DFG) as part of a research project within

    the Collaborative Research Centre SFB 799. We would like to thank Dr. S. Decker for sintering

    samples by field assisted sintering technology and Mr. R. Prang for the SEM sample preparation.

    References

    [1] Jahn A, Kovalev A, Weiß A, Wolf S, Krüger L and Scheller P R 2011 steel res.int. 82 39 [2] Krüger L, Wolf S, Martin U, Martin S, Scheller P R, Jahn A and Weiß A 2010 15th

    International Conference on the Strength of Materials (ICSMA-15) J. Phys., 240 1

    [3] Frommeyer G, Brüx U and Neumann P 2003 ISIJ Int. 43 438 [4] Rémy L 1977 Acta metal. 25 173 [5] Rémy L 1977 Metall. Trans. A 8A 253 [6] Martin S, Wolf S, Martin U, Krüger L and Rafaja D 2016 Metall. Mater. Trans. A A47 49 [7] Martin S, Fabrichnaya O and Rafaja D 2015 Mater. Lett. 159 484

  • [8] Martin S, Ullrich C, Simek D, Martin U and Rafaja D 2011 J. Appl. Cryst. 44 779 [9] Martin S, Ullrich C and Rafaja D 2015 Mater. Today: Proceedings 2S 643 [10] Bracke L, Kestens L and Penning J 2007 Scripta Mater. 57 385 [11] Olson G B and Cohen M 1972 J. Less Common Metals 28 107 [12] Ullrich C, Eckner R, Krüger L, Martin S, Klemm V and Rafaja D 2016 Mater. Sci. Eng. A 649

    390

    [13] Shen Y, Li X, Sun Y, Wang L and Zuo L 2012 Mater. Sci. Eng. A 552 514 [14] Hamada A S, Karjalainen L P, Misra R D K and Talonen J 2013 Mater. Sci. Eng. A 559 336 [15] Bracke L, Kestens L and Penning J 2009 Scripta Mater. 61 220 [16] Rafaja D, Krbetschek C, Borisova D, Schreiber G and Klemm V 2013 Thin Solid Films 530 105 [17] Rafaja D, Krbetschek C, Ullrich C and Martin S 2014 J. Appl. Cryst. 47 936 [18] Lutterotti L, Matthies S and Wenk H R 1999 IUCr Commission on Powder

    DiffractionNewsletter 21 14

    [19] Matthies S and Vinel G W 1982 phys. stat. solidi 112 111 [20] Pierce D T, Nowag K, Montagne A, Jiménez J A, Wittig J E and Ghisleni R 2013 Mater. Sci.

    Eng. A 578 134

    [21] Ledbetter H M 1984 phys. stat. solidi A 86 89 [22] Popa N C 1998 J. Appl. Cryst. 31 176 [23] Wilkens M 1970 phys. stat. solidi A 2 359 [24] Ungár T and Borbély A 1996 Appl. Phys. Lett. 69 3173 [25] Warren B E 1969 X-Ray Diffraction New York: Dover [26] Wenk H R and Grigull S 2003 J. Appl. Cryst. 36 1040 [27] Merkel S, Liermann H P, Miyagi L and Wenk H R 2013 Acta mater.61 5144


Recommended