8
Determination of dislocation density by electron backscatter diffraction and X-ray line prole analysis in ferrous lath martensite Tibor Berecz a, , Péter Jenei b , András Csóré a , János Lábár c , Jenő Gubicza b , Péter János Szabó a a Department of Materials Science and Engineering, Budapest University of Technology and Economics, 1111 Budapest, Bertalan Lajos utca 7, Hungary b Department of Materials Physics, Eötvös Loránd University, 1117 Budapest, Pázmány Péter sétány 1/a, Hungary c Institute for Technical Physics and Materials Science, Centre for Energy Research, Hungarian Academy of Sciences, 1121 Budapest, Konkoly Thege Miklós út 2933, Hungary abstract article info Article history: Received 29 September 2015 Accepted 16 November 2015 Available online 17 November 2015 The microstructure and the dislocation density in as-quenched ferrous lath martensite were studied by different methods. The blocks, packets and variants formed due to martensitic transformation were identied and their sizes were determined by electron backscatter diffraction (EBSD). Concomitant transmission electron microsco- py (TEM) investigation revealed that the laths contain subgrains with the size between 50 and 100 nm. A novel evaluation procedure of EBSD images was elaborated for the determination of the density and the space distribu- tion of geometrically necessary dislocations from the misorientation distribution. The total dislocation density obtained by X-ray diffraction line prole analysis was in good agreement with the value determined by EBSD, indicating that the majority of dislocations formed due to martensitic transformation during quenching are geometrically necessary dislocations. © 2016 Elsevier Inc. All rights reserved. Keywords: Martensite Electron back scatter diffraction Misorientation X-ray diffraction Dislocation density 1. Introduction The martensitic transformation is a diffusionless process, resulting in a martensite phase from an austenite parent phase. The ultra-high strength steels usually have martensitic microstructures, therefore the martensite in steels (ferrous martensite) is of practical importance. Martensitic transformation occurs during quenching when the face- centered cubic (fcc) austenite transforms into body-centered cubic (bcc) or body-centered tetragonal (bct) ferrous martensite. Therefore, the martensite in steels is actually a metastable ferritic phase supersat- urated with carbon and it includes a high density of crystallographic defects, especially dislocations [13]. Ferrous martensite can appear in several forms, such as lath, lentic- ular and plate, depending mainly on the chemical composition of steel. Among these martensite structures the lath martensite has high industrial signicance because of its high strength and toughness. Lath martensite can appear usually in the technologically more important low (and medium) carbon, low cost and low alloyed steels and consists of stacks of parallel, parquet- or lath-shape crystals [24]. The lath martensite morphology exhibits a characteristic multilevel microstructure. A parent austenite grain consists of several packets (the group of laths with the same habit plane). Each packet is divided into parallel blocks and a block is further subdivided into laths. The martensite lath is a single crystal of martensite which contains lattice defects, such as dislocations, with high density. The size of individual martensite laths is very small (typically few micrometers), therefore they cannot be seen clearly by optical microscope. The blocks and packets are separated by high-angle grain boundaries with the misori- entation larger than 15°, while between the laths the misorientation angle is smaller (~10°) [1,57]. The strength and toughness of lath martensitic steels strongly depend on the packet and block sizes, as well as the size, shape and ar- rangement of laths. The high strength and toughness of lath martensitic steels is mainly due to the high-angle boundaries between the blocks and packets which hinder dislocation movements [3,7,8]. Among the alloying elements carbon has the most important effect on the micro- structure and strength of lath martensite in low carbon steels. It was found that higher C content results in a ner packet size, and the ner packet size yields a higher ow stress [3,5,6]. When austenite transforms into martensite, particular crystallo- graphic orientation relationships (ORs) are maintained between them. There are three different orientation relationships observed between the parent austenite and the inherited martensite in ferrous al- loys. These are the following: KurdjumovSachs (KS), NishiyamaWassermann (NW) and GreningerTroiano (GT) ORs [2,4]. It should be noted that NW and GT ORs are close to the KS OR with deviations of 5.3° and 2.7°, respectively [9]. In low carbon low alloyed steels the KS OR is kept between the parent and inherited phases according to the above referred studies and our experiences. The microstructure developed during martensitic transformation in steels has been investigated in numerous works [1,2,6,10,11]. In the Materials Characterization 113 (2016) 117124 Corresponding author. E-mail addresses: [email protected] (T. Berecz), [email protected] (P. Jenei), [email protected] (A. Csóré), [email protected] (J. Lábár), [email protected] (J. Gubicza), [email protected] (P.J. Szabó). http://dx.doi.org/10.1016/j.matchar.2015.11.014 1044-5803/© 2016 Elsevier Inc. All rights reserved. Contents lists available at ScienceDirect Materials Characterization journal homepage: www.elsevier.com/locate/matchar

Determination of dislocation density by electron

  • Upload
    others

  • View
    10

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Determination of dislocation density by electron

Materials Characterization 113 (2016) 117–124

Contents lists available at ScienceDirect

Materials Characterization

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

Determination of dislocation density by electron backscatter diffractionand X-ray line profile analysis in ferrous lath martensite

Tibor Berecz a,⁎, Péter Jenei b, András Csóré a, János Lábár c, Jenő Gubicza b, Péter János Szabó a

a Department of Materials Science and Engineering, Budapest University of Technology and Economics, 1111 Budapest, Bertalan Lajos utca 7, Hungaryb Department of Materials Physics, Eötvös Loránd University, 1117 Budapest, Pázmány Péter sétány 1/a, Hungaryc Institute for Technical Physics and Materials Science, Centre for Energy Research, Hungarian Academy of Sciences, 1121 Budapest, Konkoly Thege Miklós út 29–33, Hungary

⁎ Corresponding author.E-mail addresses: [email protected] (T. Berecz), jene

[email protected] (A. Csóré), [email protected] (J. Lábá(J. Gubicza), [email protected] (P.J. Szabó).

http://dx.doi.org/10.1016/j.matchar.2015.11.0141044-5803/© 2016 Elsevier Inc. All rights reserved.

a b s t r a c t

a r t i c l e i n f o

Article history:Received 29 September 2015Accepted 16 November 2015Available online 17 November 2015

The microstructure and the dislocation density in as-quenched ferrous lath martensite were studied by differentmethods. The blocks, packets and variants formed due to martensitic transformation were identified and theirsizes were determined by electron backscatter diffraction (EBSD). Concomitant transmission electron microsco-py (TEM) investigation revealed that the laths contain subgrains with the size between 50 and 100 nm. A novelevaluation procedure of EBSD images was elaborated for the determination of the density and the space distribu-tion of geometrically necessary dislocations from the misorientation distribution. The total dislocation densityobtained by X-ray diffraction line profile analysis was in good agreement with the value determined by EBSD,indicating that the majority of dislocations formed due to martensitic transformation during quenching aregeometrically necessary dislocations.

© 2016 Elsevier Inc. All rights reserved.

Keywords:MartensiteElectron back scatter diffractionMisorientationX-ray diffractionDislocation density

1. Introduction

Themartensitic transformation is a diffusionless process, resulting ina martensite phase from an austenite parent phase. The ultra-highstrength steels usually have martensitic microstructures, therefore themartensite in steels (ferrous martensite) is of practical importance.Martensitic transformation occurs during quenching when the face-centered cubic (fcc) austenite transforms into body-centered cubic(bcc) or body-centered tetragonal (bct) ferrous martensite. Therefore,the martensite in steels is actually a metastable ferritic phase supersat-urated with carbon and it includes a high density of crystallographicdefects, especially dislocations [1–3].

Ferrous martensite can appear in several forms, such as lath, lentic-ular and plate, depending mainly on the chemical composition ofsteel. Among these martensite structures the lath martensite has highindustrial significance because of its high strength and toughness. Lathmartensite can appear usually in the technologically more importantlow (andmedium) carbon, low cost and low alloyed steels and consistsof stacks of parallel, parquet- or lath-shape crystals [2–4].

The lath martensite morphology exhibits a characteristic multilevelmicrostructure. A parent austenite grain consists of several packets(the group of laths with the same habit plane). Each packet is dividedinto parallel blocks and a block is further subdivided into laths. The

[email protected] (P. Jenei),r), [email protected]

martensite lath is a single crystal of martensite which contains latticedefects, such as dislocations, with high density. The size of individualmartensite laths is very small (typically few micrometers), thereforethey cannot be seen clearly by optical microscope. The blocks andpackets are separated by high-angle grain boundaries with the misori-entation larger than 15°, while between the laths the misorientationangle is smaller (~10°) [1,5–7].

The strength and toughness of lath martensitic steels stronglydepend on the packet and block sizes, as well as the size, shape and ar-rangement of laths. The high strength and toughness of lath martensiticsteels is mainly due to the high-angle boundaries between the blocksand packets which hinder dislocation movements [3,7,8]. Among thealloying elements carbon has the most important effect on the micro-structure and strength of lath martensite in low carbon steels. It wasfound that higher C content results in a finer packet size, and the finerpacket size yields a higher flow stress [3,5,6].

When austenite transforms into martensite, particular crystallo-graphic orientation relationships (ORs) are maintained betweenthem. There are three different orientation relationships observedbetween the parent austenite and the inheritedmartensite in ferrous al-loys. These are the following: Kurdjumov–Sachs (K–S), Nishiyama–Wassermann (N–W) and Greninger–Troiano (G–T) ORs [2,4]. It shouldbe noted that N–Wand G–T ORs are close to the K–S ORwith deviationsof 5.3° and 2.7°, respectively [9]. In low carbon low alloyed steels the K–S OR is kept between the parent and inherited phases according to theabove referred studies and our experiences.

The microstructure developed during martensitic transformation insteels has been investigated in numerous works [1,2,6,10,11]. In the

Page 2: Determination of dislocation density by electron

118 T. Berecz et al. / Materials Characterization 113 (2016) 117–124

recent years, several papers (e.g. [1,2,6]) studied the crystallographicand morphological characteristics of ferrous lath martensite crystalsby automated electron backscattering diffraction (EBSD) owing to theeasier access to a high-resolution scanning electron microscope (SEM)equipped with a field emission type gun (FE-SEM) and due to thenovel sample preparation techniques, such as ion-milling for EBSD sur-face treatment [10]. The EBSD imageswere analyzed in order to identifythe crystallographic variants and study the OR between the austeniteand martensite phases [1,2,6,11]. These investigations also reveal thatthe FE-SEM/EBSDmethod is an effective tool for crystallographic analy-sis in ferrous lath martensite.

The purpose of the present work is to study the dislocation densi-ty in lath martensitic microstructure. The dislocation density is de-termined in individual laths and blocks by automated electronbackscatter diffraction (EBSD) and these results are compared withthe dislocation density measured by X-ray line profile analysis(XLPA). As EBSD and XLPA measure the geometrically necessaryand the total dislocation densities, respectively, the comparison ofthe values determined by the two methods may yield the fractionsof geometrically necessary and statistically stored dislocations inthe microstructure.

2. Experimental material and methods

2.1. Sample preparation and experimental methods

The studied sample was a common, low carbon low alloyed steelwith the composition presented in Table 1. The specimen was annealedat 1100 °C for 15 min in order to achieve a stress-free austenite struc-ture. Then, the sample was quenched in ice cold (0 °C) water to obtaina full lath martensite structure.

Lath martensite forms in low- andmedium carbon steels with a car-bon content less than 0.6% [1,3]. It has been shown that there is a K–SORbetween the austenite and lath martensite crystallites in the low- andmedium carbon low alloyed steels [1,5,6,12,13]. Suikkanen et al. [4]found in 0.2C–2.0Mn–1.5Si–0.6Cr steel that the OR between austeniteand martensite is closer to a N–W OR than a K–S OR. Morito et al. [11]showed in low carbon low alloyed steels that the OR follows a near K–S OR that deviates from the exact K–S OR by 2–2.5° towards a N–WOR. However, Kitahara et al. [2] revealed the existence of a N–W OR inFe–29.6 wt.% Ni alloy. Therefore, due to the chemical compositiongiven in Table 1 our studied sample is expected to have a lathmartensitestructure with a K–S OR. The K–S orientation relationship between aus-tenite (γ) and martensite (α) is expressed as:

111ð Þγ== 011ð Þα and 101h i

γ== 111h i

α:

In the case of a K–S OR due to symmetry reasons there are 24 equiv-alent crystallographic variants (laths) in martensite, which can formfrom an austenite grain (also called prior austenite). A block consistsof laths with small misorientations. A packet is composed of blocksthat consist of martensite laths with different variants, which had thesame {111}γ planes in the prior austenite. Themaximumnumber of var-iants in a packet is six. There may be four crystallographically differentpackets in a prior austenite crystallite, because there are four different{111}γ planes [1,6].

There are ten possible misorientation angles between the differentvariants in the K–S OR, namely: 10.53°, 14.88°, 20.61°, 21.06°, 47.11°,

Table 1Chemical composition of the studied sample in wt.%.

Fe C Si Mn Ni

Balanced 0.147 0.383 1.177 3

49.47°, 50.51°, 51.73°, 57.21° and 60.00°. It can be seen that the mini-mum misorientation angle between the adjacent martensite variants(laths) is 10.53° according to the K–S OR. Therefore, themisorientationsat packet and block boundaries are higher than 10.53° [1,6]. The variantsin the present material were identified from pole figures obtained byEBSD using the method described in [1].

As it was mentioned above, the lath martensitic microstructurescontain a high density of lattice defects, especially dislocations,which makes EBSD analysis particularly difficult by traditional sam-ple preparation techniques, such as mechanical or electro-polishingmethods. Therefore, in order to obtain high quality EBSD imagesthe mechanically polished surfaces of the samples were ion milledfor 30 min by a 10 keV Ar-ion beam at an incidence angle of ~7°[14]. The EBSD analysis was performed by a FEI Quanta 3D dualbeam SEM, equipped with an EDAX type EBSD system and an OIMAnalysis 5.3 software configuration. The EBSD images were taken atvarious scanning step sizes of 500, 200, 80 and 35 nm.

The lattice defect structure in thematerial was also studied by X-rayline profile analysis (XLPA). For X-ray examination the specimen wasfirst mechanically polished by 1200-grit SiC paper and then by 1 μmALOX suspension. Subsequently, the sample was electropolished in anA8 electrolytic solution (made by Struers) at room temperature and afixed voltage of 10 V for 60 s. The X-ray line profiles were measuredby a high-resolution diffractometer using CoKα1 (λ= 0.17889 nm) ra-diation. The line profiles were evaluated by the Convolutional MultipleWhole Profile (CMWP) analysis [15]. In this method, the experimentalpattern is fitted by the convolution of the instrumental pattern andthe theoretical size and strain line profiles. The theoretical profilefunctions used in this fitting procedure are calculated on the basisof a model of the microstructure, where the crystallites have spheri-cal shape and log-normal size distribution, and the lattice strains areassumed to be caused by dislocations. In the present study, the area-weighted mean crystallite size, the average dislocation density and aparameter characterizing the edge/screw character of dislocationswere determined by the XLPAmethod. According to previous studies[15–18] the mean crystallite size (i.e. the average size of coherentlyscattering domains) gives the average subgrain-size in metallicmaterials which is usually smaller then the grain size obtained bymicroscopic methods.

Themicrostructurewas also investigated by a JEOL-200CX transmis-sion electronmicroscope (TEM) operating at 200 kV. The TEM foils wereprepared by mechanical polishing to the thickness of 100 μm, then jetpolish was applied until perforation of the foil using A8 electrolytic so-lution at 15 °C and 50 V.

2.2. Methodology of the dislocation density determination from EBSDmisorientations

A novel software was developed for the determination of the dislo-cation density from the misorientations measured between the neigh-boring pixels in EBSD images. In the present EBSD experiments thepixels were arranged into a square point grid. Each measurementgives an output file which contains the Euler angles and the Cartesiancoordinates of the pixels in the reference system fixed to the sample.From this output file the so-called Nye dislocation density tensor is cal-culated by our software written in programming language C#. This pro-gram applies the procedure proposed in refs. [19,20] which is describedbriefly in the following paragraphs.

First, the Bunge orientation tensor [21] denoted by G is determinedfor each pixel from the Euler angles φ1, φ2, and ϕ as:

G ¼cosφ1 cosφ2− sinφ1 sinφ2 cosϕ sinφ1 cosφ2 þ cosφ1 sinφ2 cosϕ sinφ2 sinϕ

− cosφ1 sinφ2− sinφ1 cosφ2 cosϕ − sinφ1 sinφ2 þ cosφ1 cosφ2 cosϕ cosφ2− sinϕsinφ1 sinϕ − cosφ1 sinϕ cosϕ

24

35:

ð1Þ

Page 3: Determination of dislocation density by electron

Fig. 1. Sketch of the square grid used in the evaluation of EBSD images. Black solid circlesrepresent the measuring points which have neighbors at both their right side and belowthem. Red solid circles do not have all necessary neighbors, hence these are omitted inthe evaluation procedure. (For interpretation of the references to color in this figure leg-end, the reader is referred to the web version of this article.)

119T. Berecz et al. / Materials Characterization 113 (2016) 117–124

Then, themisorientationmatrix for a pixel and one of its neighbors isobtained from the Bunge orientation tensors of the two pixels denoted

by G and G0, respectively, using the following equation:

M¼ GC G0C0� �−1

¼ g g 0−1

; ð2Þ

where C and C 0 are symmetry matrices, g ¼ GC and g 0 ¼ G0C 0 . Forcubic crystals there are 24 symmetrymatrices, yielding 576misorienta-tion matrices for two neighboring pixels. From these matrices only thatone is used in the analysis for which the misorientation angle isminimal. The misorientation angle can be easily obtained from thequaternionic representation of matrices [19]. Quaternions are mathe-matical objects consisting of one scalar and a three dimensional vectorpart: q = (q0, q

!). In this representation, q0 and the three componentsof q! can be obtained from tensor g as:

q0 ¼ 12

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiXlgll þ 1

q; qi ¼ −

εijk gjk4q0

; ð3Þ

where εijk is the Levi–Civita symbol and gjk are the components of theBunge orientation tensor. Using the quaternions of the two neighboringpixels another quaternion, Δq, can be obtained as.

q q’ð Þ−1 ¼ Δq; ð4Þ

which gives the misorientation angle θ between the two pixels as[19]:

θ ¼ 2 arccos Δq0ð Þ; ð5Þ

where Δq0 is the scalar part of the Δqmisorientation quaternion. Asmentioned above between two points there are 576misorientationma-trices (or in quaternionic representation: 576 Δq quantities). For thefurther calculations we selected the quaternion Δq which belongs tothe minimal misorientation angle. Then, this quantity was used for thedetermination of the Nye's dislocations tensor:

α ¼α11 α12 α13α21 α22 α23α31 α32 α33

24

35; ð6Þ

where the components of the tensor can be obtained according tothe following formula [19]:

αlk ¼ Δqk2arccos Δq0ð Þffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

1−Δq20q 1

Δxl: ð7Þ

In Eq. (7) the step size (distance between the pixels) in the differentdirections are denoted by Δxl where indices l = 1 and 2 correspond tosteps in the plane of the EBSD image (see Fig. 1) while l = 3 is relatedto steps perpendicular to this plane. Δqk is the k-th component of thevectorial part of quaternion Δq. It is noted that in the calculation of αlk

the quaternion component Δqk obtained in the xl direction must beused in Eq. (7). Since the measuring points are arranged in a squaregrid, Δx1 = Δx2. As local lattice orientations are determined only inthe plane of investigation (along x1 and x2) but not perpendicular to it(along x3), some components in matrix α cannot be determined. Thus,Nye's dislocation tensor is an incomplete matrix, and therefore only alocal apparent dislocation density can be calculated from the availabletensor components as:

ρ ¼ 1b

α12j j þ α13j j þ α21j j þ α23j j þ α33j jð Þ; ð8Þ

where b is the magnitude of the Burgers-vector. It follows fromEq. (7) that only those pixels were incorporated in the evaluationwhich have neighbors at both right and below this point in the squaregrid as shown in Fig. 1. For calculating the total apparent dislocationdensity for the investigated area, the values obtained for the individualpixels are averaged. The input of our software is a “.txt” file prepared bythe electron microscope which contains the Euler angles for all mea-sured pixels. The output file contains the elements of the Nye-tensorand the dislocation density and in addition the dislocation densitymap in the investigated area can be generated.

3. Results

Fig. 2a shows an example for the EBSD orientation maps taken onthe as-quenched microstructure with the step size of 500 nm. The crys-tallographic orientations of the martensite crystallites are indicated bydifferent colors. According to the K–S OR between the original austeniteand the formed martensite phases, the parent austenite grains wereidentified and their boundaries are indicated by black lines in Fig. 2a.The size of the parent austenite grains varies between 20 and 200 μm.A former austenite grain from the upper right corner of Fig. 2a was se-lected to visualize the orientations of the as-formedmartensite variants.The identification of the variants was carried out with the help of polefigures. Fig. 2c shows the theoretical {001} pole figure with the 24mar-tensite variants. Fig. 2d is an experimental {001} pole figure obtainedfrom the grain shown in Fig. 2b. The different variants around the{001} pole are indicated by various colors. The same colors are used inFig. 2b for mapping the space distribution of these variants. The thick-ness of martensite laths was a few micrometers.

Fig. 3 shows the total misorientation distribution for the angle rangebetween 0 and 60° determined by EBSD with the step size of 500 nm.The K–S OR predicts theoretical orientation relationships between thedifferent variants which are indicated by arrows in Fig. 3. These misori-entations correspond to the crystallographically determined orientationdifference between the neighboringmartensite laths related to differentvariants. The exact values of these misorientation angles are listed inRef. [1]. Former investigations revealed that the experimentally

Page 4: Determination of dislocation density by electron

Fig. 2. (a) EBSD orientationmap for the studied lath martensite. The former austenite grain boundaries are indicated by black lines. (b) A grain from the upper right corner of (a) inwhichthe variants are indicated by different colors. (c) and (d) show the theoretical and experimental {001} pole figures for the 24martensite variants, respectively. The color code in (b) can beidentified from the pole figure in (d). (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

Fig. 3. The total misorientation distribution for the angle range between 0 and 60° deter-mined by EBSDwith the step size of 500 nm. The exact misorientations expected from theideal K–S orientation relationship are indicated by arrows. The inset shows the bottompart of the figure with a higher magnification.

120 T. Berecz et al. / Materials Characterization 113 (2016) 117–124

determined misorientations may slightly deviate from the angle valuescalculated theoretically from the K–S OR [11]. Either this deviation isconsidered or not, the misorientations smaller than 5° are certainlynot related to K–S OR angles in Fig. 3. Therefore, the misorientationslower than 5° were used in the EBSD images for the determination ofthe dislocation density.

There is an uncertainty in the determination of the Euler angles fromthe Kikuchi patterns obtained by EBSD which results in an error in thecalculation of the dislocation density from the misorientations. Inorder to estimate the error in the misorientation angles, EBSD experi-ments were performed on an undeformed single crystalline siliconwafer. For this material the difference between the crystallographic ori-entations of the neighboring EBSD image pixels was expected to bepractically zero. Fig. 4 shows themisorientation distribution for a Si sin-gle crystal obtained with the EBSD step size of 35 nm. The measuredmaximum misorientation angle was 0.6° and the median of the distri-bution was about 0.25°. This means that a small but considerable mis-orientation can be observed even if the material is practically defect-free. The same misorientation distributions were obtained for Si withother EBSD step sizes. Since this misorientation angle distribution prac-tically ends at 0.5°, this value was selected as a lower limit of the

Page 5: Determination of dislocation density by electron

Fig. 4. The misorientation distribution for a single crystalline Si measured with the EBSDstep size of 35 nm.

Fig. 6. EBSD orientation map (a) and the corresponding dislocation density map(b) obtained with the step size of 35 nm. In (b) the grain boundaries with misorientationangles larger than 5° are indicated by thin blue lines, and the higher the dislocation den-sity, the darker the gray contrast. (For interpretation of the references to color in thisfigurelegend, the reader is referred to the web version of this article.)

121T. Berecz et al. / Materials Characterization 113 (2016) 117–124

misorientations used for the calculation of the dislocation density in lathmartensite. Thismeans that if bothmisorientations between a pixel andits two neighbors (at the right side and below the studied pixel) aresmaller than 0.5°, the dislocation density is taken as zero in that pixel.If any of the two misorientations is higher than 5°, the pixel is omittedfrom the dislocation density calculation. For all other pixels the disloca-tion density was calculated from the misorientations, as described inSection 2.2.

Fig. 5 shows the misorientation distribution density functions up tothe angle of 5° for a grain in the studied lath martensite with differentEBSD step sizes. The streaked area corresponds to the omitted angle re-gime below 0.5°. The fraction of small misorientations increases withdecreasing EBSD step size. This trend suggests that the grains arefragmented into very small parts with small misorientations. Thesesmall changes in the orientation cannot be observed if the EBSD stepsize is large. Fig. 6a and b shows an EBSD orientation map and the cor-responding dislocation density map, respectively, obtained with thestep size of 35 nm. In Fig. 6b the grain boundaries with misorientationslarger than 5° are indicated by thin blue lines, and the higher the dislo-cation density, the darker the gray contrast.

Fig. 7 shows the average dislocation density determined for thewhole studied area as a function of the step size. The error was calculat-ed from theuncertainty of themisorientation determination,whichwasestimated by the median of the misorientation distribution (0.25°) ob-tained on single crystalline Si. The error in the dislocation density was

Fig. 5. Themisorientation distribution density functions up to the angle of 5° for a grain inthe studied lath martensite with different EBSD step sizes.

obtained by shifting the measured misorientation angles by 0.25°,cutting the two ends of the distribution at 0.5° and 5°, and thenrecalculating the dislocation density. Fig. 7 reveals that the smaller thepixel size, the higher the calculated dislocation density. This trend sug-gests that the microstructure is finer than the largest three step sizes

Fig. 7. The dislocation density determined by EBSD as a function of the scan step size.

Page 6: Determination of dislocation density by electron

Fig. 8. Bright (a) and dark (b) field TEM images showing the subgrain structure inside the laths.

122 T. Berecz et al. / Materials Characterization 113 (2016) 117–124

used in EBSD (80, 200 and 500 nm). In order to reveal the subgrainstructure, TEM was performed. Fig. 8 shows TEM bright and dark fieldimages which indicate that the laths contain subgrains with the sizevarying between 50 and 100 nm. There is a considerable contrast varia-tion inside the subgrains in the dark-field TEM image in Fig. 8b, suggest-ing lattice strains which are most probably caused by dislocations. Inaddition, the boundaries of subgrains usually consist of dislocations,therefore in the present EBSD experiments the smallest achievablestep size (35 nm) provides themost reliable values of the geometricallynecessary dislocation density (~30 × 1014 m−2). The dislocation densi-ties in the different variants can only be determined for the large stepsizes (200 and 500 nm), since for smaller step sizes the variants cannotbe identified in the pole figures due to the smaller investigated area. Theidentification of variants is based on the comparison of the experimen-tal and theoretical pole figures. This procedure can only be carried out ifa sufficient number of different variants appears on the EBSD image.Since the investigated area was less than 40 μm × 40 μm for lowerstep sizes (35 and 80 nm), in these cases the evaluation of varianttypes was not possible. According to the present investigations, thereis no significant difference between the dislocation densities deter-mined in the various variants.

Fig. 9. X-ray diffractogram evaluated by the CMWP fitting method. The open circles andthe solid line represent the measured and the fitted X-ray diffraction patterns,respectively.

The total dislocation density was determined by XLPA using theCMWP fitting procedure. The CMWP fitting for the studied sample isshown in Fig. 9. The open circles and the solid line represent the mea-sured data and the fitted curves, respectively. The parameters of themi-crostructure obtained by XLPA are listed in Table 2. The area-weightedmean crystallite size (bxNarea) is about 71 nm, which is in reasonableagreement with the average size determined for a limited number ofsubgrains by TEM. XLPA gives ~35 × 1014 m−2 for the dislocation den-sity (ρ) which agrees well with the value obtained from EBSD(~30 × 1014 m−2). It should be noted that the latter procedure givesonly the density of the geometrically necessary dislocations, whileXLPA provides the total dislocation density (including both geometri-cally necessary and statistically stored dislocations). Therefore, it canbe concluded that in this as-quenched lath martensite the majority ofdislocations are geometrically necessary. It is also noted that thedislocation density obtained for the present material by XLPA is ingood agreement with the value determined for another as-quenchedlath martensite by other authors using X-ray diffraction broadeninganalysis [22]. XLPA also gives a parameter describing the edge/screwcharacter of dislocations. The experimental value of parameter q is1.9 ± 0.1 which agrees within the experimental error with the arith-metic average of the theoretical values calculated for pure edge andscrew dislocations, 1.28 and 2.67, respectively. This observation sug-gests that the dislocation structure has a half edge–half screw characterwhich is in good agreementwith recently published results obtained onanother lath martensite steel [23].

4. Discussion

The Kikuchi patterns detected in EBSD experiments can be used forthe determination of the internal stress distribution [24,25] and thedensity of geometrically necessary dislocations (GNDs) [19,26]. The ro-tation gradients obtained from the difference between the Kikuchi pat-terns of the neighboring image pixels are used for the calculation ofGND density according to the Nye's theory [19]. The effectiveness ofthis method is limited by the fact that the misorientation angle canbe resolved only above a limit, which is usually between 0.5 and 1.5°

Table 2The area-weightedmean crystallite size (bxNarea), the dislocation density (ρ) and param-eter q describing the edge/screw character of dislocations.

bxNarea [nm] ρ [1014 m−2] q

71 ± 8 35 ± 4 1.9 ± 0.1

Page 7: Determination of dislocation density by electron

123T. Berecz et al. / Materials Characterization 113 (2016) 117–124

[19,27]. This is caused by the uncertainty of the band detection inKikuchi patterns which also results in an error in the obtained disloca-tion density values. A better angular resolution (0.03–0.1°) can beachieved by cross-correlation-based EBSD measurements [26,28,29].The latter method has been successfully applied for the determinationof GNDs in Cu [26] and Ti-alloys [28,30]. If the method based on theNye's theory is applied for severely deformedmaterials with high dislo-cation density, the small required EBSD pixel size and the relativelylarge uncertainty in the disorientation angle may yield a large error inthe dislocation density [19]. In the present study, this error was deter-mined from the uncertainty of the misorientation angle measured ona defect-free Siwafer. Although, in this sample there are no considerablemisorientations, non-zero disorientation angles were measured. Thesemisorientation angles have a median value of about 0.25° (see Fig. 4)which was taken as the error of the disorientation angles determinedexperimentally in the lath martensite. Moreover, it was shown thatthis error is independent of the step size of the EBSD investigation.The uncertainty of the misorientation angle determines the error ofthe dislocation density. This uncertainty can be reduced by increasingthe resolution of the EBSD detector. In our experiments the camera(EDAX “Hikari”) recorded the Kikuchi patterns in 120 × 120 pixels. Inorder to evaluate the effect of camera resolution on the misorientationangle error, the EBSD measurements were repeated on the Si waferusing higher detector resolution (240 × 240 pixels). In this case, themeasured maximum misorientation angle was 0.4°, the median of thedistribution was under 0.15° and it was also independent from thestep size. The lower error can be explained by the fact that for a highercamera resolution the EBSD evaluation software can fit the theoreticalKikuchi lines to a higher number of experimental data points. Furtherincrement in the camera resolution most probably results in further re-duction of the error. It should be noted, however, that the better resolu-tion yields longer EBSD measuring time.

There are twomain sources of thehigh dislocation density in the lathmartensites. The first is the plastic accommodation of transformationstrains in the martensite [32]. The volume change during transforma-tion was found to be the most dominant factor influencing the disloca-tion density in the lath martensite. It should be noted that the austenitematrix is also deformed during transformation, yielding dislocationmultiplication in this phase too. These dislocations are inherited into amartensite phase [32]. The second origin of the high dislocation densityin ferrous lath martensite is themisfit between the parent (fcc) and theproduct (bcc) lattices [31]. The misfit strains at phase boundaries canrelax by the formation of dislocations. It was found that dislocation den-sity in lathmartensite depends on the temperature ofmartensitic trans-formation, the cooling rate and the chemical composition [32]. Thelower transformation temperature and the higher cooling rate resultin a higher dislocation density due to the suppressed dislocation annihi-lation. It has also been shown that as the carbon content in the lathmartensite increases, the dislocation density increases [32].

In recent years several studies have been published inwhich the dis-location density in ferrous lath martensite was determined by differentmethods. Morito et al. [32] measured 11 × 1014m−2 as the average dis-location density by TEM in an Fe–0.18C sample after quenching from1200 or 1100 °C. Takebayashi et al. [33] determined 63 × 1014 m−2

and 200 × 1014 m−2 for the dislocation density in as-quenched Fe–0.3C–0.49Si–1.98Mn steel by different X-ray diffraction methods. Conget al. [34] obtained 29 × 1014 m−2 as the maximum dislocation densityin a quenched Fe–0.1C steel sample by two types of slip deformation(TTSD) model. HajyAkbary et al. [22] applied different X-ray diffractionmethods for the determination of the dislocation density in a quenchedFe–0.3C–1.6Si–3.5Mn sample and the obtained values were between~30 × 1014 m−2 and ~85 × 1014 m−2. Shi et al. [23] has received~30 × 1014 m−2 for the dislocation density in Fe–0.22C–0.87Si–1.64Mn–0.024Ti–0.0015B–0.0025 N (wt.%) lath martensite steelusing neutron diffraction profile analysis. The dislocation density evalu-ated in the present study by both EBSD and XLPA methods

(~35 × 1014 m−2) has the same order of magnitude as the values re-ported in the literature.

5. Conclusions

The microstructure of an as-quenched lath martensite was investi-gated by EBSD and XLPA. The following results were obtained:

1. The blocks, packets and variants formed due to martensitic transfor-mation were identified from the pole figures obtained by EBSD. Thesize of the parent austenite grains varies between 20 and 200 μm.The thickness of martensite laths was a few micrometers.

2. Transmission electron microscopy revealed that the laths containsubgrains with the size varying between 50 and 100 nm. This is inreasonable agreement with the average crystallite size value(71 nm) determined by XLPA.

3. For the determination of the dislocation density the misorientationsin the EBSD images were evaluated. In this method the local disloca-tion density was obtained from the misorientations between theneighboring image pixels. Considering the crystallographic orienta-tion relationship between the variants formed during martensitictransformation, only the misorientation angles lower than 5° are re-lated to dislocations. In addition, EBSD experiments on a defect-freeSi crystal revealed that the practical lower limit for the determinationofmisorientation angles is about 0.5°. Therefore, only themisorienta-tions in the range of 0.5–5° were used in the evaluation of the dislo-cation density. No significant difference was obtained between theaverage dislocation densities in the different variants.

4. The dislocation density was determined as a function of the EBSDstep size. It was found that the dislocation density increasedwith de-creasing the step size: for the smallest (35 nm) and largest (500 nm)step sizes the dislocation densities were about 30 and 6 × 1014 m−2,respectively. This trend can be explained by the fine subgrain struc-ture in the laths, as revealed by TEM. Since the subgrain size isunder 100 nm, the more reliable dislocation density was obtainedwith the smallest step size (35 nm). This dislocation density(30 × 1014 m−2) is in good agreement with the value determinedby XLPA (35 × 1014m−2). As EBSD determines the dislocation densi-ty from the misorientations in the microstructure, it can be conclud-ed that the majority of dislocations are geometrically necessarydislocations in the present ferrous lath martensite.

Acknowledgements

This work was financially supported by the Hungarian Scientific Re-search Fund, OTKA, through Grant Nos. K-109021 and PD-101028.

References

[1] H. Kitahara, R. Ueji, N. Tsuji, Y. Minamino, Crystallographic features of lath martens-ite in low-carbon steel, Acta Mater. 54 (2006) 1279–1288.

[2] H. Kitahara, R. Ueji, M. Ueda, N. Tsuji, Y. Minamino, Crystallographic analysis of platemartensite in Fe–28.5 at.% Ni by FE-SEM/EBSD, Mater. Charact. 54 (2005) 378–386.

[3] G. Krauss, Martensite in steel: strength and structure, Mater. Sci. Eng. A-Struct. 273-275 (1999) 4057.

[4] P.P. Suikkanen, C. Cayron, A.J. DeArdo, L.P. Karjalainen, Crystallographic analysis ofmartensite in 0.2C–2.0Mn–1.5Si–0.6Cr steel using EBSD, J. Mater. Sci. Technol. 27(10) (2011) 920–930.

[5] T. Furuhara, S. Morito, T. Maki, Morphology, substructure and crystallography of lathmartensite in Fe–C alloys, J. Phys. IV France 112 (2003) 255–258.

[6] S. Morito, H. Tanaka, R. Konishi, T. Furuhara, T. Maki, The morphology and crystal-lography of lath martensite in Fe–C alloys, Acta Mater. 51 (2003) 1789–1799.

[7] S. Morito, H. Yoshida, T. Maki, X. Huang, Effect of block size on the strength of lathmartensite in low carbon steels, Mater. Sci. Eng. A-Struct. 438–440 (2006) 237–240.

[8] B. Sonderegger, S. Mitsche, H. Cerjak, Martensite laths in creep resistant martensitic9–12% Cr steels — calculation and measurement of misorientations, Mater. Charact.58 (2007) 874–882.

[9] H.K.D.H. Bhadeshia, Worked Examples in the Geometry of Crystals, 2nd editionInsitute of Materials, London, 2001.

[10] Z. Dankházi, S.Z. Kalácska, A. Baris, G. Varga, Z.S. Radi, K. Havancsák, EBSD samplepreparation: high energy Ar ion milling, Mater. Sci. Forum 812 (2015) 309–314.

Page 8: Determination of dislocation density by electron

124 T. Berecz et al. / Materials Characterization 113 (2016) 117–124

[11] S. Morito, X. Huang, T. Furuhara, T. Maki, N. Hansen, Themorphology and crystallog-raphy of lath martensite in alloy steels, Acta Mater. 54 (2006) 5323–5331.

[12] S. Morito, Y. Edamatsu, K. Ichinotani, T. Ohba, T. Hayashi, Y. Adachi, T. Furuhara, G.Miyamoto, N. Takayama, Quantitative analysis of three-dimensional morphologyof martensite packets and blocks in iron–carbon–manganese steels, J. AlloysCompd. 577S (2013) S587–S592.

[13] G. Miyamoto, N. Iwata, N. Takayama, T. Furuhara, Quantitative analysis of variant se-lection in ausformed lath martensite, Acta Mater. 60 (2012) 1139–1148.

[14] T. Berecz, S.Z. Kalácska, G. Varga, Z. Dankházi, K. Havancsák, Effect of high energy Ar-ion milling on surface of quenched low-carbon low-alloyed steel, Mater. Sci. Forum812 (2015) 285–290.

[15] G. Ribárik, J. Gubicza, T. Ungár, Correlation between strength and microstructure ofball-milled Al-Mg alloys determined by X-ray diffraction, Mater. Sci. Eng. A-Struct.387-389 (2004) 343–347.

[16] P. Scardi, M. Leoni, Whole powder pattern modelling, Acta Crystallogr. A 58 (2002)190–200.

[17] T. Ungár, J. Gubicza, G. Ribárik, A. Borbély, Crystallite size distribution and disloca-tion structure determined by diffraction profile analysis: principles and practicalapplication to cubic and hexagonal crystals, J. Appl. Crystallogr. 34 (2001) 298–310.

[18] T. Ungár, G. Tichy, J. Gubicza, R.J. Hellmig, Correlation between subgrains and coher-ently scattering domains, Powder Diffract. 20 (4) (2005) 366–375.

[19] W. He, W. Ma, W. Pantleon, Microstructure of individual grains in cold-rolled alu-minium from orientation inhomogeneities resolved by electron backscattering dif-fraction, Mater. Sci. Eng. A-Struct. 494 (2008) 21–27.

[20] W. Pantleon, Resolving the geometrically necessary dislocation content by conven-tional electron backscattering diffraction, Scr. Mater. 58 (2008) 994–997.

[21] H.-J. Bunge, Texture Analysis in Materials Science: Mathematical Methods,Butterworths, London, 1982.

[22] F. HajyAkbary, J. Sietsma, A.J. Böttger, M.J. Santofimia, An improved X-ray diffractionanalysis method to characterize dislocation density in lath martensitic structures,Mater. Sci. Eng. A-Struct. 639 (2015) 208–218.

[23] Z.M. Shi, W. Gong, Y. Tomota, S. Harjo, J. Li, B. Chi, J. Pu, Study of tempering behaviorof lath martensite using in situ neutron diffraction, Mater. Charact. 107 (2015)29–32.

[24] J. Jiang, T.B. Britton, A.J. Wilkinson, Mapping type III intragranular residual stress dis-tributions in deformed copper polycrystals, Acta Mater. 61 (2013) 5895–5904.

[25] A.J. Wilkinson, E. Tarleton, A. Vilalta-Clemente, J. Jiang, T.B. Britton, D.M. Collins,Measurement of probability distributions for internal stresses in dislocated crystals,Appl. Phys. Lett. 105 (2014) 181907.

[26] J. Jiang, T.B. Britton, A.J. Wilkinson, Evolution of dislocation density distributions incopper during tensile deformation, Acta Mater. 61 (2013) 7227–7239.

[27] B.L. Adams, J. Kacher, EBSD-based microscopy: resolution of dislocation density,CMC-Comput. Mater. Con. 14 (3) (2009) 185–196.

[28] P.D. Littlewood, T.B. Britton, A.J.Wilkinson, Geometrically necessary dislocation den-sity distributions in Ti–6Al–4V deformed in tension, Acta Mater. 59 (2011)6489–6500.

[29] C. Maurice, J.H. Driver, R. Fortunier, On solving the orientation gradient dependencyof high angular resolution EBSD, Ultramicroscopy 113 (2012) 171–181.

[30] P.D. Littlewood, A.J.Wilkinson, Geometrically necessary dislocation density distribu-tions in cyclically deformed Ti–6Al–4V, Acta Mater. 60 (2012) 5516–5525.

[31] B.P.J. Sandvik, C.M. Wayman, Characteristics of lath martensite: part II. The martens-ite–austenite interface, Metall. Trans. A 14 (1983) 823–834.

[32] S. Morito, J. Nishikawa, T. Maki, Dislocation density within lath martensite in Fe–Cand Fe–Ni alloys, ISIJ Int. 43 (9) (2003) 1475–1477.

[33] S. Takebayashi, T. Kunieda, N. Yoshinaga, K. Ushioda, S. Ogata, Comparison of thedislocation density in martensitic steels evaluated by some X-ray diffractionmethods, ISIJ Int. 50 (6) (2010) 875–882.

[34] Z. Cong, Y. Murata, Y. Tsukada, T. Koyama, Prediction of the maximum dislocationdensity in lath martensitic steel by elasto-plastic phase-field method, Mater.Trans. 53 (9) (2012) 1598–1603.