+ All Categories
Home > Documents > Determination of composition, residual stress and stacking fault depth profiles in expanded...

Determination of composition, residual stress and stacking fault depth profiles in expanded...

Date post: 20-Dec-2016
Category:
Upload: maj
View: 212 times
Download: 0 times
Share this document with a friend
6
Determination of composition, residual stress and stacking fault depth proles in expanded austenite with energy-dispersive diffraction S. Jegou a, , 1 , T.L. Christiansen a , M. Klaus b , Ch. Genzel b , M.A.J. Somers a a Technical University of Denmark, Dept. Mechanical Engineerings, Kgs. Lyngby, Denmark b Helmholtz-Zentrum Berlin für Materialien und Energie GmbH, Berlin, Germany abstract article info Available online 9 June 2012 Keywords: Surface engineering Residual stress Energy-dispersive diffraction Reconstruction prole A methodology is proposed combining the scattering vector method with energy dispersive diffraction for the non-destructive determination of stress- and composition-depth proles. The advantage of the present method is a relatively short measurement time and avoidance of tedious sublayer removal; the disadvantage as compared to destructive methods is that depth proles can only be obtained for depth shallower than half the layer thickness. The proposed method is applied to an expanded austenite layer on stainless steel and al- lows the separation of stress, composition and stacking fault density gradients. © 2012 Elsevier B.V. All rights reserved. 1. Introduction Residual stresses are widely and deliberately introduced within the near surface region of materials to locally modify the mechanical properties and enhance the component performance with respect to wear and/or fatigue. Surface engineering associated with tailoring of the surface properties and residual stress can be achieved by thermal, chemical or mechanical treatment [1] and yields a functionally graded material that changes its properties from surface to interior. The quantication of residual stress-depth proles to investigate the ef- fect of the surface engineering treatment can be performed by X-ray diffraction analysis [2]. This technique relies on the determination of hkl specic lattice strains for various orientations of the scattering vector with respect to the sample surface normal combined with an appropriate grain-interaction model [3]. Numerous factors affect the so-called X-ray diffraction stress analysis, e.g. grain size, triaxiality of the stress state and preferred orientation. The evaluation of stress-depth proles in functionally graded materials can be inuenced by the stress gradient itself, as well as by other gradients. Steep residual stress gradients can lead to the so-called ghost stresses, i.e. systematic errors inherent to the applied measurement and/or evaluation procedure, if no precautions are taken. When superimposition of composition and stress gradients occurs, such as for a composition-induced stress gradient, stress evaluation over the information depth also depends on composition, because the reference spacing is composition dependent. This can lead to dra- matic ghost stresses if not taken into account during data acquisition and evaluation [4,5]. Among the various techniques developed for non-destructive depth resolved stress determination [3,69], energy-dispersive dif- fraction methods, using white radiation, give some advantages associ- ated with multiple reections recorded in one energy spectrum and deeper information depths [1013]. Stress-induced errors can effec- tively be avoided combining a modied multi-wavelength approach with the sin 2 ψ method or the scattering vector method [14]. In [15] it was shown that the energy-dispersive method can be applied even to the detection of very steep in-plane residual stress gradients in surface treated hard coatings, if the information depth is adapted to the steepness of the gradient. However, for a composition- induced (self-induced) stress gradient, the optimisation proceduredeveloped for the scattering vector method cannot be applied straightforwardly, because the lattice spacing in the strain-free direc- tion varies with the information depth. Instead a sin 2 ψ-based ap- proach should be considered, where sin 2 ψ dependencies at pre- chosen information depths are evaluated by interpolation among the experimental data. The reference lattice parameter for the appro- priate information depth follows from interpolation among the data in the strain free direction or from independent spectroscopic analy- sis and knowledge of the relation between lattice parameter and composition. This work deals with the evaluation of residual stress by means of non-destructive energy-dispersive diffraction under the inuence of steep stress- and composition gradients. Steep superimposed mul- tigradients arise after low temperature thermochemical surface treat- ments of stainless steel [16]. Such treatments (nitriding, carburising or nitrocarburising) give rise to the formation of a surface zone of Thin Solid Films 530 (2013) 7176 Corresponding author at: Arts & Métiers ParisTech, MécaSurf Laboratory, 2 cours des Arts et Métiers, 13617 Aix-en-Provence, France. Tel.: + 33 442938174; fax: + 33 442938114. E-mail addresses: [email protected] (S. Jegou), [email protected] (T.L. Christiansen), [email protected] (M. Klaus), [email protected] (C. Genzel), [email protected] (M.A.J. Somers). 1 Now with: Arts & Métiers ParisTech, MécaSurf Laboratory, Aix-en-Provence, France. 0040-6090/$ see front matter © 2012 Elsevier B.V. All rights reserved. doi:10.1016/j.tsf.2012.06.029 Contents lists available at SciVerse ScienceDirect Thin Solid Films journal homepage: www.elsevier.com/locate/tsf
Transcript

Thin Solid Films 530 (2013) 71–76

Contents lists available at SciVerse ScienceDirect

Thin Solid Films

j ourna l homepage: www.e lsev ie r .com/ locate / ts f

Determination of composition, residual stress and stacking fault depth profiles inexpanded austenite with energy-dispersive diffraction

S. Jegou a,⁎,1, T.L. Christiansen a, M. Klaus b, Ch. Genzel b, M.A.J. Somers a

a Technical University of Denmark, Dept. Mechanical Engineerings, Kgs. Lyngby, Denmarkb Helmholtz-Zentrum Berlin für Materialien und Energie GmbH, Berlin, Germany

⁎ Corresponding author at: Arts & Métiers ParisTechdes Arts et Métiers, 13617 Aix-en-Provence, France. Te442938114.

E-mail addresses: [email protected] (S. Jego(T.L. Christiansen), [email protected] (M. Klaus(C. Genzel), [email protected] (M.A.J. Somers).

1 Now with: Arts & Métiers ParisTech, MécaSurfFrance.

0040-6090/$ – see front matter © 2012 Elsevier B.V. Alldoi:10.1016/j.tsf.2012.06.029

a b s t r a c t

a r t i c l e i n f o

Available online 9 June 2012

Keywords:Surface engineeringResidual stressEnergy-dispersive diffractionReconstruction profile

A methodology is proposed combining the scattering vector method with energy dispersive diffraction forthe non-destructive determination of stress- and composition-depth profiles. The advantage of the presentmethod is a relatively short measurement time and avoidance of tedious sublayer removal; the disadvantageas compared to destructive methods is that depth profiles can only be obtained for depth shallower than halfthe layer thickness. The proposed method is applied to an expanded austenite layer on stainless steel and al-lows the separation of stress, composition and stacking fault density gradients.

© 2012 Elsevier B.V. All rights reserved.

1. Introduction

Residual stresses are widely and deliberately introduced withinthe near surface region of materials to locally modify the mechanicalproperties and enhance the component performance with respect towear and/or fatigue. Surface engineering associated with tailoring ofthe surface properties and residual stress can be achieved by thermal,chemical or mechanical treatment [1] and yields a functionally gradedmaterial that changes its properties from surface to interior. Thequantification of residual stress-depth profiles to investigate the ef-fect of the surface engineering treatment can be performed by X-raydiffraction analysis [2]. This technique relies on the determination ofhkl specific lattice strains for various orientations of the scatteringvector with respect to the sample surface normal combined with anappropriate grain-interaction model [3]. Numerous factors affect theso-called X-ray diffraction stress analysis, e.g. grain size, triaxialityof the stress state and preferred orientation. The evaluation ofstress-depth profiles in functionally graded materials can beinfluenced by the stress gradient itself, as well as by other gradients.Steep residual stress gradients can lead to the so-called ghost stresses,i.e. systematic errors inherent to the applied measurement and/orevaluation procedure, if no precautions are taken.

When superimposition of composition and stress gradients occurs,such as for a composition-induced stress gradient, stress evaluation

, MécaSurf Laboratory, 2 coursl.: +33 442938174; fax: +33

u), [email protected]), [email protected]

Laboratory, Aix-en-Provence,

rights reserved.

over the information depth also depends on composition, becausethe reference spacing is composition dependent. This can lead to dra-matic ghost stresses if not taken into account during data acquisitionand evaluation [4,5].

Among the various techniques developed for non-destructivedepth resolved stress determination [3,6–9], energy-dispersive dif-fractionmethods, using white radiation, give some advantages associ-ated with multiple reflections recorded in one energy spectrum anddeeper information depths [10–13]. Stress-induced errors can effec-tively be avoided combining a modified multi-wavelength approachwith the sin 2ψ method or the scattering vector method [14]. In [15]it was shown that the energy-dispersive method can be appliedeven to the detection of very steep in-plane residual stress gradientsin surface treated hard coatings, if the information depth is adaptedto the steepness of the gradient. However, for a composition-induced (self-induced) stress gradient, the ‘optimisation procedure’developed for the scattering vector method cannot be appliedstraightforwardly, because the lattice spacing in the strain-free direc-tion varies with the information depth. Instead a sin 2ψ-based ap-proach should be considered, where sin 2ψ dependencies at pre-chosen information depths are evaluated by interpolation amongthe experimental data. The reference lattice parameter for the appro-priate information depth follows from interpolation among the datain the strain free direction or from independent spectroscopic analy-sis and knowledge of the relation between lattice parameter andcomposition.

This work deals with the evaluation of residual stress by means ofnon-destructive energy-dispersive diffraction under the influence ofsteep stress- and composition gradients. Steep superimposed mul-tigradients arise after low temperature thermochemical surface treat-ments of stainless steel [16]. Such treatments (nitriding, carburisingor nitrocarburising) give rise to the formation of a surface zone of

72 S. Jegou et al. / Thin Solid Films 530 (2013) 71–76

so-called expanded austenite which essentially is a solid solution ofcolossal amounts of interstitials (carbon and/or nitrogen) in the aus-tenite lattice. This results in biaxial compressive residual stresses ofseveral GPa's that find their origin in the lattice misfit between theexpanded austenite “case” and the untreated core [16,17].

2. Non destructive depth profiling with energy-dispersive X-raystress analysis

X-ray stress analysis is based on the lattice strain measurement ε-φψ

hkl experienced by a set of lattice planes {hkl} in a given directiondefined by the azimuth, φ, and inclination, ψ, with respect to the sam-ple surface normal (Fig. 1):

εhklφψ ¼ dhklφψ

dhklo

−1 ð1Þ

where dohkl is the unstrained lattice spacing.

In energy-dispersive diffraction using a white beam, measure-ments are carried out for fixed and predetermined diffraction andscattering angles. The Bragg equation then takes the following form:

dhkl ¼ hc2 sinθ

1Ehkl

ð2Þ

where 2θ is the scattering angle, h is Planck's constant, c is the veloc-ity of light and Ehkl is the energy for which diffraction of the hkl latticeplanes occurs.

Introducing Eq. (2) in Eq. (1) gives the lattice strain εφψhkl in themeasuring direction defined by φ and ψ as:

εhklφψ ¼ Ehklo

Ehklφψ

−1 ð3Þ

where Eohkl corresponds to the unstrained lattice spacing do

hkl.For sur-face engineered quasi-isotropic polycrystalline materials usually astate of rotationally symmetric biaxial stress (σ13=σ23=σ33=0and σ11=σ22=σ//) can be assumed, leading to:

εhklψ zð Þ ¼ 2Shkl1 σ == zð Þ þ 12Shkl2 σ == zð Þ sin2ψ ð4Þ

where S1hkl and 1/2S2hkl are diffraction elastic constants, depending onthe crystal orientation hkl and elastic interaction among the crystals.The lattice spacing, ⟨dψhkl⟩ (or equivalently the energy at which diffrac-tion occurs) determined in an X-ray diffraction experiment for a

Fig. 1. Diffraction geometries in X-ray stress analysis from [17]. η denotes the rotationof the sample around the scattering vector g

→φψ for a fixed measuring direction (φ,ψ)

with respect to the sample system P. PB and SB denote primary and secondary(diffracted) beam.

sample (or layer) of thickness, t, is the diffracted intensity-weightedaverage over depth, z, i.e.:

dhklψ

D E¼ ∫t

0dhklψ zð Þ exp −μ Eð Þkzf gdz∫t0 exp −μ Eð Þkzf gdz

ð5aÞ

where, for measurement in reflection geometry (as practised in thepresent work)

k ¼ 2 sinθ cosψsin2θ− sin2ψþ cos2θ sin2ψ sin2η

ð5bÞ

describes the diffraction geometry, μ(E) is the linear absorption coef-ficient which, for a homogeneous layer, depends on the photon ener-gy and η denotes the rotation angle around the scattering vector, g

→φψ

(Fig. 1). For completeness it is mentioned that μ(E) depends on com-position. This second order effect is not considered here.2 Hence, it isobtained for the lattice strain, averaged over the diffracting volume,⟨εψhkl⟩ :

εhklψ

D E¼ ∫t

0dhklψ zð Þ exp −μ Eð Þkzf gdz

∫t0d

hklo zð Þ exp −μ Eð Þkzf gdz

−1 ð6Þ

Note that these equations are only valid for the case where thestudied layers are well within the gauge volume. From Eq.(6) it is ob-served that the lattice strain evaluated for experimental lattice spac-ings has to be evaluated from strained and unconstrained latticespacings weighted over the same depth range. This lattice strain canbe assigned to the information depth, τ:

τ Eð Þ ¼ zh i ¼ ∫t0z⋅ exp −μ Eð Þkzf gdz∫t0 exp −μ Eð Þkzf gdz

¼ 1μ Eð Þkþ t

exp −μ Eð Þktf gexp −μ Eð Þktf g−1

ð7Þ

Note that the information depth in a layer is maximally t/2 for thecase where the layer can be considered infinitely thin as compared tothe penetration of the X-rays.3 For an infinitely thick layer the infor-mation depth equals 1/[μ(E)k], which for the present case amountsto 27 μm. It is important to realise that, in general, ⟨dψhkl⟩ and ⟨do

hkl⟩

are not experimentally determined at the same information depth,because the strain-free lattice spacing applies only for one specificvalue for ψ (and thus τ), the so-called strain-free direction, ψo, de-

fined by sin2ψo ¼ − 2Shkl112S

hkl2

(as obtained by equating Eq. (4) to zero).

Consequently, application of Eq. (6) requires that a value for ⟨dohkl⟩

at τψ is obtained by interpolation among the experimentally deter-mined strain-free lattice spacing-depth profile ⟨do

hkl(z)⟩.In the case of stress-depth profiling, various methods have been

developed, based on either successive layer removal (destructivemethods) or assigning the evaluated data to a depth below the sur-face (non-destructive methods). According to Eqs. (7) and (5b) for afixed value of θ the information depth can be varied by variation ofthe angles ψ and η or, for energy dispersive analysis, by selecting an-other energy E where diffraction occurs. In the present work the scat-tering vector method (varying η and ψ) and the multi-wavelengthmethod (varying E and ψ) are combined for non-destructive depthprofiling of the composition, stress and stacking fault probability inlow temperature hardened stainless steel.

2 For the present case where a layer of expanded austenite on stainless steel is con-sidered, the error in assuming μ(E) independent of depth (i.e. a homogeneous layer)lies in the range 3.94 to 4.39% for a composition ranging from yN=0.30 to yN=0.50(cf. Fig. 4).

3 For ‘Real space’ method, the measuring depths are not limited to t/2 since thegauge volume is used to define the observed volume.

36 38 40 42 44 46 48 50 52 540

2

4

6

8

10

12

14x103

Energy, keV

Inte

nsity

, cou

nts/

s

200N

111N

111

200

Fig. 2. (Smoothed) energy spectrum of nitrided 316 L stainless steel. 2θ=8°, φ=0°,ψ=18.43°, η=86.9°, counting time: 300 s. Expanded austenite and austenite are den-oted by γN and γ respectively.

4 It is noted that if texture gradients contribute to the asymmetry of the line profiles,an error is introduced by assigning the peak position to the centroid position.

73S. Jegou et al. / Thin Solid Films 530 (2013) 71–76

3. Experimental details

3.1. Nitriding

A disc of AISI 316 L stainless steel with diameter 10 mm was gro-und and polished to a mirror like finish. The final thickness of the discwas 2.18 mm. The disc was austenitised in flowing hydrogen at atemperature of 1355 K in order to obtain a fully recrystallized austen-itic structure in the sample. Nitriding was performed in a Netzsch 449thermo-balance equipped with electronic mass-flow controllers foraccurate gas control. The sample was activated in a gaseous atmo-sphere and subsequently nitrided at 440 °C for 14 h in a gas mixtureconsisting of 100 ml/min NH3+5 ml/min N2 (N2 was led throughthe measurement compartment to protect the electronics of the ap-paratus from interaction with NH3). This treatment yielded a zoneof expanded austenite with a thickness of about 16 μm.

3.2. X-ray diffraction

The energy dispersive diffraction experiments were performed atthe materials science beamline EDDI@BESSY II [18]. The white syn-chrotron beam with usable photon energies between about 8 keVand 120 keV is provided by a superconducting 7 T multipole wiggler.The primary beam cross-section was 0.3×0.3 mm², the equatorial di-vergence Δ2θ in the diffracted beam was limited by a double slit sys-tem with an aperture of 0.03×5 mm² to values smaller than 0.01°.Hence, the gauge volume (which is the part of the sampling volumedefined by the beam limiting slits, that immerses in the sample)takes a rather complex geometrical shape [19]. In the case of steel,however, the limiting factor for the information depth is not givenby this ‘geometrical’ gauge volume but due to the high absorptionby the 1/e information depth τ(Ε) = 1/[μ(Ε)⋅k] (cf. Eq. (7)). Referencemeasurements were carried out under identical experimental condi-tions on stress-free powder to ensure that geometrically inducedline shifts were smaller than ΔE=10 eV and therefore, had not tobe taken into account in data evaluation. A measuring time of 300 swas chosen for recording the diffraction patterns in order to achievegood counting statistics. For data acquisition a solid state germaniumdetector (Canberra model GL0110) was used.

The 111 and 200 diffraction lines of the expanded austenite phasewere investigated. The ⟨dψ

hkl⟩ versus the information depth τ wereevaluated for a constant diffraction angle 2θ=8° and azimuthφ=0° at nine different ψ angles, ranging from 18.43° to 71.57°. Foreach value of ψ, 18 values of the rotation angle η about the scatteringvector were selected. The range of rotation angles, η, depends on theinclination angle, ψ; the larger ψ the smaller is the range of η values.According to Fig. 1 the orientation η=90° corresponds to theΨ−mode of X-ray stress analysis and, therefore, to the largest infor-mation depth. The case η=0° corresponds to the Ω−mode (ψb0)and can only be realised for θ>ψ. For the small Bragg angles betweenabout 4° and 10° used in ED diffraction, this condition is usually notfulfilled. Here, the minimum rotation angle ηwhich would correspondto an inclination angle α between the incoming beam and the samplesurface, is given byηmin ¼ arcsin

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffisin2 ψ− sin2 θ

q= sinψ cosθð Þ

� �[20]. The dif-

fraction elastic constants for the analysed reflections were calculatedusing the Eshelby/Kröner model with the single crystal constants ofFe-18Cr-12Ni [21].

4. Results and interpretation

An energy spectrum of the nitrided sample is given in Fig. 2. Bothexpanded austenite and the unaffected austenite in the substrate areobserved; the 111 line profiles are observed at 40 and 42.8 keV forexpanded austenite and austenite, respectively. The difference in en-ergy position between the peaks for expanded austenite (case) and

austenite (core) is attributed to the combined effect of residual stress,composition and stacking fault energy (see [22]). Asymmetry of thepeaks obtained for expanded austenite is ascribed to the presence ofgradients in stress, composition and stacking fault probability,where those in stress and composition will dominate [22]. In addi-tion, also texture gradients could contribute to the asymmetry. Suchtexture gradients originate from grain rotation caused by plastic de-formation in the expanded austenite case during growth [23].Recognising that the line profiles are an average over a depth rangeand, thus, a certain part of the profile, the centroid position of theγN line profile was taken as the peak position and assigned to the in-formation depth (cf. Eq. (7)).4

The lattice spacing ⟨dψhkl⟩ as measured for nine different ψ angles

and 18 η angles per value of ψ is given in Fig. 3a as a function of thecorresponding information depth τ. Clearly, the variation in τ realisedby rotation about the diffraction vector is largest for the lowest valueof ψand is ascribed to larger range of η values for smallerψ. The data inFig. 3a is assigned to relatively shallow information depths, smallerthan t/2=8 μm (see comment below Eq. (7)). From interpolationamong the data in Fig. 3a it is possible to evaluate a ⟨dψ

111⟩ vs sin 2ψdependence at a chosen value for τ. An example is given in Fig. 3bat a depth of 4 μm (indicated by the vertical dashed line in Fig. 3a).Within experimental accuracy a straight line is obtained, suggestingno influence of steep stress/composition gradients on the evaluationmethod.

The dependence of the strain-free lattice spacing on informa-tion depth was obtained by linear interpolation in ⟨dψ

111⟩ vs sin 2ψand ⟨dψ

200⟩ vs sin 2ψ data for the respective strain-free directions,i.e. sin 2ψo=0.324 and 0.48 for 111 and 200, respectively. The re-sults are given in Fig. 4 and converted to the nitrogen content in ex-panded austenite by applying the relation determined in [24]. Anon-negligible difference in lattice parameter/composition is ob-served for the two selected hkl. The lattice parameter/nitrogencontent determined from the 200 line profile is systematicallyhighest. This discrepancy can be understood from the introductionof stacking faults associated with the partial plastic accommoda-tion of the colossal, chemically induced, stresses introduced in ex-panded austenite during growth of this zone into the austenitesubstrate. The presence of stacking faults on the peak positions of111 and 200 line profiles is antagonistic: for 200 a shift towardslower energy (higher lattice spacing) and for 111 a shift towardshigher energy (lower lattice spacing) would occur, which is inagreement with the observed discrepancy. Adopting the Warrenequation for the relation between the peak shift as a consequence

0 1 2 3 4 5 6 7 80.218

0.220

0.222

0.224

0.226

0.228

, m

<d

111 >

, nm

0.10.20.30.40.50.60.70.80.9

sin2

0.0 0.2 0.4 0.6 0.8 1.00.218

0.220

0.222

0.224

0.226

sin2

<d

111 >

, nm

= 4 m

a

b

Fig. 3. a) As measured lattice spacing-depth, ⟨dψ111⟩ vs τ profiles of expanded austenitemeasured with the scattering vector method applied to a nitrided 316 L stainless steel.b) A reconstructed sin 2ψ plot at chosen information depth, τ, of 4 μm is also given.

74 S. Jegou et al. / Thin Solid Films 530 (2013) 71–76

of stacking faults in angle dispersed X-ray diffraction [25]:

Δ 2θhkl� �

¼ 0:2756⋅α⋅Ghkl⋅ tanθhkl ð8aÞ

where Δ(2θhkl) is given in radians and α is the stacking fault proba-bility, G111=1/4 and G200=−1/2. Equivalently, for the change in

0 2 4 6 80.375

0.380

0.385

0.390

0.395

0.400

0.2

0.3

0.4

0.5

0.6

, m

<a 0

>, n

m

111

200

corrected

<y N

>

Fig. 4. Lattice parameter, ⟨ao⟩, and associated composition-depth profiles, ⟨yN⟩, evaluatedfor 111 and 200 diffraction lines of expanded austenite from the sin 2ψ plots at chosenpenetration depths τ and reconstructed from lattice-spacing-depth profiles ⟨dψ

111⟩ mea-sured from the scattering vector method applied to a nitrided 316 L stainless steel. Thecorrected ⟨ao⟩ profile was evaluated by applying Eq. (8a) and (8b), respectively.

lattice spacing :

Δ dhkl� �dhkl

¼ −0:1378⋅α⋅Ghkl ð8bÞ

Assuming that the same composition, and thus strain-free latticeparameter should be obtained from 111 to 200 line profiles, the stac-king fault probability and strain free lattice spacing can be deter-mined, provided that the same stacking fault probabilities prevailfor 111 and 200 as measured in the strain-free direction. In terms ofthe strain-free lattice parameter ao, it is obtained from Eq. (8b) :

a111o −a200o

ao¼ −0:1378⋅α⋅ 1

4−1

2

� �ð8cÞ

with ahkl the lattice parameter following from the as measured hkl lat-tice spacing.

The depth dependencies of the strain-free lattice parameter a0 andthe stacking fault probability α for the depth range τ=4–6 μm(wherethey can be determined), are given in Figs. 4 and 5, respectively.

Using the strain-free lattice parameter from Fig. 4, residual stress-es can be evaluated for 111 and 200 reflections from interpolatedgraphs as Fig. 3b. Fig. 6 shows the resulting residual stress depth pro-files σ//(τ) for both diffraction lines 111 and 200 of expanded austen-ite after reconstruction of sin 2ψ plots at predefined depths τ. It has tobe noted that the error bars are given as twice the standard deviation,which is obtained from the determination procedure of the scatteringpeak position and propagated through all calculations.

5. Discussion

Separation of stress-, composition- and stacking fault probabilitydepth profiles from X-ray stress analysis of nitrided 316 L stainlesssteel was investigated through an energy-dispersive method. Thescattering-vector method was applied for its fast depth profiling reli-ability and combined with a procedure for reconstruction of sin 2ψplots at chosen information depths τ. It is noted explicitly that theprofiles shown in Figs. 2–6 do not reflect the actual depth profiles.By straightforward calculus it can be shown that an actual depth pro-file, i.e. vs depth z, is only (in a certain depth range) identical to thatfor the profile vs τ, if the profile depends linearly on depth [26]. In thisrespect it is important to realise that the lattice spacing data used forthe evaluation are diffracted intensity weighted lattice spacings (cf.Fig. 3a) and that the larger information depth to which the observedlattice spacing (or stress, composition, stacking fault probability) isassigned, the less will the average value reflect the actual lattice spac-ing (or stress, composition, stacking fault probability) at this depth.

0 2 4 6 8 100.00

0.02

0.04

0.06

0.08

0.10

0.12

0.14

0.2

0.3

0.4

0.5

0.6

, m

<yN>

<y N

>

Fig. 5. Stacking fault probability, α, vs information depth, τ, determined from the an-tagonistic shifts of the 111 and 200 Bragg reflections. The corresponding nitrogen con-tent, ⟨yN⟩, is given too and identical to the drawn line in Fig. 4.

0 2 4 6 8 10, m

//, M

Pa

111

200

Fig. 6. Residual stress-depth profiles σ// evaluated from 111 to 200 diffraction lines ofexpanded austenite from the sin 2ψ plots at chosen information depths, τ,reconstructed from lattice-spacing-depth profiles ⟨dψ

hkl⟩ measured from the scatteringvector method applied to a nitrided 316 L stainless steel.

5 Note that nitrogen in austenite leads to strengthening and, thus, an increase of theyield stress.

75S. Jegou et al. / Thin Solid Films 530 (2013) 71–76

The actual depth profiles could be obtained by reconstructing the ac-tual lattice spacing profiles from those in Fig. 3a by assuming a poly-nomial description of the actual profile (for the case of layer/substratesystems see [5]).

5.1. Composition

Clearly, the nitrogen content decreases with depth, as anticipatedfor a growth process largely governed by solid state diffusion of nitro-gen through the developed layer. Presuming local equilibrium at thesurface in an unconstrained condition, the nitrogen content at thesurface would be yN

s =0.61 [24]. This predicted value comparesfavourably with the present results, as suggested by extrapolation ofthe current experimental data towards the surface. Actually, a dis-crepancy would be expected between experimental and predictedsurface concentrations as a consequence of the huge compressivestresses, which affect the thermodynamics of the system, such thatthe solubility is reduced as compared to the unconstrained condi-tion [27]. In this respect it should also be mentioned that the adoptedelastic constants were assumed to be identical to those for austenite.For the nitrogen content close to the unnitrided stainless steel sub-strate (at a depth of 16 μm) a content corresponding to a ratio Cr:N=1:1 (i.e. yN=0.17) is expected at the transition from expandedaustenite to “substrate”. It is not possible to make an accurate extrap-olation of the few present data to this information depth, but a firstestimate provides a value of 0.21, which is in fair agreement withthe predicted value, taking the uncertainty into account.

5.2. Stacking fault probability

In this discussion it has to be remarked that the Warren methodfor the determination of stacking fault probabilities only applies forrelatively low stacking fault probabilities and that the method byVelterop et al. [28] should be preferred for probabilities of values de-termined. Seen in this light the current evaluation should be consid-ered a first attempt. Interpreting stacking faults in terms of theVelterop-model would require additional measurements.

The stacking fault probability increases with depth from a smallvalue for shallow depths to a higher value closer to the “substrate”and is opposite to the depth dependence of strain-free lattice param-eter. The origin of plastic deformation in expanded austenite is plasticaccommodation of the composition-induced stresses in the growingexpanded austenite zone, which exceeds the yield stress. It has beendemonstrated convincingly that grain rotation and texture changesoccur as a consequence of the plastic accommodation [23,29]. If thestacking faults are caused by this plastic deformation, it would be

expected that the stacking fault probability is highest at the surfaceand decreases with depth. Provided that the evaluation procedure iscorrect and the assumptions made are justified, the present datashow the opposite trend (within the small information depth rangewhere the analysis can be made). This could be understood as follows.The part of the expanded austenite zone that grows into the “sub-strate” can accommodate the compositionally induced lattice misfitlargely elastically.5 Ahead of the growing expanded austenite thecompensating tensile stress in the unnitrided (and not strengthened)austenite leads to deformation martensite and plastic deformation.Upon continued nitriding this structure is transformed into expandedaustenite but it could be that the stacking faults remain. For the partclosest to the surface, in an early stage of nitriding, such stresses inthe unnitrided austenite can be (partly) relaxed by the adjacent sur-face and inhomogeneous thickness of the expanded austenite zone.

5.3. Residual stress

The stress data for 111 and 200 show a trend of a decreasing resid-ual compressive stress with increasing information depth. In this re-spect it has to be noted that the stress values at the largestaccessible information depths for both 111 and 200 are very sensitivefor small variations in the data, because the number of data contribut-ing to the dψ vs sin 2ψ graph decreases with increasing τ (cf. Fig. 3 for111). These data should therefore be considered as less reliable.

The occurrence of residual stresses of the magnitude observed inFig. 6 is consistent with the previous work [16], where depth profilesof stress and composition were reconstructed from measurementsafter (destructive) successive layer removals. The present methodhas the advantage that significantly shorter measurement times arenecessary and no tedious removal of thin (sub)layers from the sampleis necessary, which affects the stress state. On the other hand the ac-tual stress profile is not obtained, only a diffracted intensity weightedaverage.

6. Conclusion

The present work explores the use of the scattering vector methodfor non-destructive X-ray diffraction analysis and separation ofcomposition-, stress- and stacking fault probability gradients in afunctionally graded material. Based on the energy-dispersive method,complementary information can be deduced from recording variousdiffraction lines simultaneously such that from assuming an identicalstrain-free lattice parameter evaluated from these reflections, thecomposition and stacking fault probability can be estimated. The re-sults appear to be in agreement with earlier work on a similar system.The advantages of the present method are relatively short measure-ment time and a non-destructive analysis. The disadvantage is thatno actual profiles are obtained, but rather diffraction intensityweighted averages are assigned to the information depth. The latteris maximally half the layer thickness, so the deeper part of the layerremains “invisible”.

Acknowledgements

Financial support from the Danish Research Council for Technolo-gy and Production Sciences under grant no. 274-07-0344 is gratefullyacknowledged.

References

[1] P.J. Withers, H.K.D.H. Bhadeshia, Mater. Sci. Technol. 17 (2001) 355.[2] I..C. Noyan, J.B. Cohen, Residual Stress, Measurement by Diffraction and Interpre-

tation, Springer, New York, 1987.

76 S. Jegou et al. / Thin Solid Films 530 (2013) 71–76

[3] V. Hauk, Structural and Residual Stress Analysis by Nondestructive Methods,Elsevier, Amsterdam, 1997.

[4] M.A.J. Somers, E.J. Mittemeijer, Metall. Trans. A 21 (1990) 189.[5] T. Christiansen, M.A.J. Somers, Mater. Sci. Eng., A 424 (2006) 181.[6] Ch Genzel, Phys. Status Solidi A 146 (1994) 629.[7] A.J. Allen, M.T. Hutchings, C.G. Windsor, C. Andreani, Adv. Phys. 34 (1985) 445.[8] A. Pyzalla, J. Nondestruct. Eval. 19 (2000) 21.[9] Ch Genzel, in: E.J. Mittemeijer, P. Scardi (Eds.), Diffraction Analysis of the Micro-

structure of Materials, Springer Series in Materials Science, 68, 2004, p. 473.[10] H. Ruppersberg, Adv. X-Ray Anal. 37 (1994) 235.[11] Ch. Genzel, I.A. Denks, R. Coelho, D. Thomas, R. Mainz, D. Apel, M. Klaus, J. Strain,

Analysis 46 (2011) 615.[12] C. Stock, Ch. Genzel, W. Reimers, Mater. Sci. Forum 404–407 (2002) 13.[13] M. Klaus, Ch. Genzel, H. Holzschuh, Thin Solid Films 517 (2008) 1172.[14] Ch. Genzel, C. Stock, W. Reimers, Mater. Sci. Eng., A 372 (2004) 28.[15] M. Klaus, W. Reimers, Ch. Genzel, Powder Diffr. (Suppl. 24) (2009) S82.[16] T.L. Christiansen, M.A.J. Somers, Int. J. Mater. Res. 100 (2009) 10.[17] T.L. Christiansen, M.A.J. Somers, Metall. Mater. Trans. A 40 (2009) 1791.

[18] Ch. Genzel, I.A. Denks, J. Gibmeier, M. Klaus, G. Wagner, Nucl. Instrum. MethodsPhys. Res., Sect. A 578 (2007) 23.

[19] Ch. Genzel, S. Krahmer, M. Klaus, I.A. Denks, J. Appl. Crystallogr. 44 (2011) 1.[20] Ch. Genzel, J. Appl. Crystallogr. 32 (1999) 770.[21] H.M. Ledbetter, Br. J. Nondestruct. Test. (1981) 286.[22] T.L. Christiansen, T.S. Hummelshøj, M.A.J. Somers, Surf. Eng. 26 (2010) 242.[23] J.C. Stinville, P. Villechaise, C. Templier, J.P. Rivière, M. Drouet, Acta Mater. 58

(2010) 2814.[24] Th. Christiansen, M.A.J. Somers, Metall. Mater. Trans. A 37 (2006) 675.[25] B.E. Warren, X-ray Diffraction, Dover Publications Inc., 1990[26] T. Christiansen Ph.D thesis, Low temperature surface hardening of stainless steel”,

Technical University of Denmark 2004.[27] T. Christiansen, K.V. Dahl, M.A.J. Somers, Mater. Sci. Technol. 24 (2008) 159.[28] L. Velterop, R. Delhez, Th.H. de Keijser, E.J. Mittemeijer, D. Reefman, J. Appl.

Crystallogr. 33 (2000) 296.[29] C. Templier, J.C. Stinville, P. Villechaise, P.O. Renault, G. Abrasonis, J.P. Rivière, A.

Martinavičius, M. Drouet, Surf. Coat. Technol. 204 (2010) 2551.


Recommended