+ All Categories
Home > Documents > Dislocation densities and prevailing slip-system types determined by X-ray line profile analysis in...

Dislocation densities and prevailing slip-system types determined by X-ray line profile analysis in...

Date post: 30-Nov-2016
Category:
Upload: tamas
View: 214 times
Download: 1 times
Share this document with a friend
8
research papers J. Appl. Cryst. (2013). 46, 55–62 doi:10.1107/S0021889812046705 55 Journal of Applied Crystallography ISSN 0021-8898 Received 10 April 2012 Accepted 12 November 2012 # 2013 International Union of Crystallography Printed in Singapore – all rights reserved Dislocation densities and prevailing slip-system types determined by X-ray line profile analysis in a textured AZ31 magnesium alloy deformed at different temperatures Bertalan Jo ´ni, a Talal Al-Samman, b Sandip Ghosh Chowdhury, c Ga ´bor Csisza ´r a and Tama ´s Unga ´r a * a Department of Materials Physics, Eo ¨ tvo ¨ s University Budapest, Hungary, b Department of Metal Physics and Physical Metallurgy, RWTH Aachen University, Germany, and c CSIR–National Metallurgical Laboratory, Jamshedpur 831007, India. Correspondence e-mail: [email protected] Tension experiments were carried out at room temperature, 473 K and 673 K on AZ31-type extruded magnesium alloy samples. The tensile deformation has almost no effect on the typical extrusion texture at any of the investigated temperatures. X-ray diffraction patterns provided by a high-angular-resolution diffractometer were analyzed for the dislocation density and slip activity after deformation to fracture. The diffraction peaks were sorted into two groups corresponding either to the major or to the random texture components in the specimen. The two groups of reflections were evaluated simultaneously as if the two texture components were two different phases. The dislocation densities in the major texture components are found to be always larger than those in the randomly oriented grain populations. The overwhelming fraction of dislocations prevailing in the samples is found to be of hai type, with a smaller fraction of hc + ai-type dislocations. The fraction of hci-type dislocations is always obtained to be zero within experimental error. 1. Introduction Mg alloys have been the subject of intense research during recent years, owing to their excellent properties, such as low density and high specific strength, and because of the increasing interest in weight savings for automotive and aerospace applications. In contrast to cubic crystals, particu- larly face-centered-cubic crystals, there are a wide variety of slip and twinning systems in hexagonal crystals, the activation of which depends strongly on the c/a ratio (Jones & Hutch- inson, 1981). Magnesium has a c/a ratio of 1.624, very close to the ideal ratio of 1.633. The onset of plastic deformation is usually characterized by the yield stress that is related to the beginning of a large scale dislocation motion. Such dislocation motion is driven by a resolved shear stress and takes place preferentially on close-packed crystallographic planes and almost exclusively in close-packed directions. This applies in magnesium to basal slip in the closest-packed plane (0001) along the closest-packed direction h11 20i (hai slip). Less densely packed slip planes and larger slip vectors, e.g. h11 23i (hc + ai slip), are more difficult to activate because the critical resolved shear stress of this slip system is considerably higher than that of basal slip. However, at higher temperatures the hc + ai slip systems can also be activated thermally (Agnew & Duygulu, 2005; Lukac & Mathis, 2002; Mathis, Nyilas et al. , 2004; Mathis et al. , 2005; Obara et al., 1973; Wang et al., 2010). This behavior is in correlation with the limited ductility and poor formability of Mg at around room temperature (Yoo, 1981). The activation of slip and twinning systems during deformation of a polycrystal dictates its texture and micro- structure evolution and thus determines its deformation behavior. The determination of the dislocation density and slip activity in different texture components was attempted by Klimanek (1994) and Klimanek et al. (1996). Hot- and cold-rolled commercial-purity titanium plates were piled together to form thick metal pieces. Plane surfaces in correlation with the texture were explored by cutting and polishing, and line profile analysis was carried out in order to determine the dislocation density in specific texture compo- nents. In the present work a novel procedure is presented with two purposes. On the one hand, for a better understanding of the plastic deformation of magnesium alloys at different temperatures, the method of X-ray line profile analysis is applied to obtain qualitative and quantitative microstructure information about dislocation densities and slip activity as a function of the temperature of deformation. On the other hand, the dislocation density and slip activity in different texture components is determined on the same sample piece without sectioning for the different reflection planes. It is also shown that the parallel-beam geometry with high angular
Transcript

research papers

J. Appl. Cryst. (2013). 46, 55–62 doi:10.1107/S0021889812046705 55

Journal of

AppliedCrystallography

ISSN 0021-8898

Received 10 April 2012

Accepted 12 November 2012

# 2013 International Union of Crystallography

Printed in Singapore – all rights reserved

Dislocation densities and prevailing slip-systemtypes determined by X-ray line profile analysis ina textured AZ31 magnesium alloy deformed atdifferent temperatures

Bertalan Joni,a Talal Al-Samman,b Sandip Ghosh Chowdhury,c Gabor Csiszara and

Tamas Ungara*

aDepartment of Materials Physics, Eotvos University Budapest, Hungary, bDepartment of Metal

Physics and Physical Metallurgy, RWTH Aachen University, Germany, and cCSIR–National

Metallurgical Laboratory, Jamshedpur 831007, India. Correspondence e-mail:

[email protected]

Tension experiments were carried out at room temperature, 473 K and 673 K on

AZ31-type extruded magnesium alloy samples. The tensile deformation has

almost no effect on the typical extrusion texture at any of the investigated

temperatures. X-ray diffraction patterns provided by a high-angular-resolution

diffractometer were analyzed for the dislocation density and slip activity after

deformation to fracture. The diffraction peaks were sorted into two groups

corresponding either to the major or to the random texture components in the

specimen. The two groups of reflections were evaluated simultaneously as if the

two texture components were two different phases. The dislocation densities in

the major texture components are found to be always larger than those in the

randomly oriented grain populations. The overwhelming fraction of dislocations

prevailing in the samples is found to be of hai type, with a smaller fraction of

hc + ai-type dislocations. The fraction of hci-type dislocations is always obtained

to be zero within experimental error.

1. IntroductionMg alloys have been the subject of intense research during

recent years, owing to their excellent properties, such as low

density and high specific strength, and because of the

increasing interest in weight savings for automotive and

aerospace applications. In contrast to cubic crystals, particu-

larly face-centered-cubic crystals, there are a wide variety of

slip and twinning systems in hexagonal crystals, the activation

of which depends strongly on the c/a ratio (Jones & Hutch-

inson, 1981). Magnesium has a c/a ratio of 1.624, very close to

the ideal ratio of 1.633. The onset of plastic deformation is

usually characterized by the yield stress that is related to the

beginning of a large scale dislocation motion. Such dislocation

motion is driven by a resolved shear stress and takes place

preferentially on close-packed crystallographic planes and

almost exclusively in close-packed directions. This applies in

magnesium to basal slip in the closest-packed plane (0001)

along the closest-packed direction h1120i (hai slip). Less

densely packed slip planes and larger slip vectors, e.g. h1123i

(hc + ai slip), are more difficult to activate because the critical

resolved shear stress of this slip system is considerably higher

than that of basal slip. However, at higher temperatures the

hc + ai slip systems can also be activated thermally (Agnew &

Duygulu, 2005; Lukac & Mathis, 2002; Mathis, Nyilas et al.,

2004; Mathis et al., 2005; Obara et al., 1973; Wang et al., 2010).

This behavior is in correlation with the limited ductility and

poor formability of Mg at around room temperature (Yoo,

1981). The activation of slip and twinning systems during

deformation of a polycrystal dictates its texture and micro-

structure evolution and thus determines its deformation

behavior. The determination of the dislocation density and

slip activity in different texture components was attempted

by Klimanek (1994) and Klimanek et al. (1996). Hot- and

cold-rolled commercial-purity titanium plates were piled

together to form thick metal pieces. Plane surfaces in

correlation with the texture were explored by cutting and

polishing, and line profile analysis was carried out in order to

determine the dislocation density in specific texture compo-

nents.

In the present work a novel procedure is presented with two

purposes. On the one hand, for a better understanding of the

plastic deformation of magnesium alloys at different

temperatures, the method of X-ray line profile analysis is

applied to obtain qualitative and quantitative microstructure

information about dislocation densities and slip activity as a

function of the temperature of deformation. On the other

hand, the dislocation density and slip activity in different

texture components is determined on the same sample piece

without sectioning for the different reflection planes. It is also

shown that the parallel-beam geometry with high angular

resolution can be usefully exploited for exploring the micro-

structure in specific texture components.

2. Experimental

2.1. Samples

The material investigated was commercially extruded

magnesium alloy AZ31 with the chemical composition (in

wt%) 2.90 Al, 0.84 Zn, 0.33 Mn, 0.02 Si, 0.004 Fe, 0.001 Cu,

0.001 Ni, Mg (balance). Round tension specimens with 6 mm

diameter and 37 mm gauge length were machined from the

extruded material with the extrusion direction oriented

parallel to the tension direction. The tension tests were

conducted at room temperature (RT), 473 K and 673 K, at a

constant strain rate of 10�2 s�1. The samples were strained to

failure, which took place shortly after necking. The samples

deformed at elevated temperatures were immediately quen-

ched in water after completion of the tests. Stress–strain

curves corresponding to the three temperatures are shown in

Fig. 1. For the purpose of the X-ray diffraction measurements

specimens were cut from the central part of the tensile

deformed samples with one surface perpendicular and one

parallel to the direction of extrusion, denoted as ND and ED

(normal and extruded directions), respectively. Specimen

surfaces for X-ray diffraction measurements were prepared by

conventional grinding and diamond polishing and finishing

with a colloidal silica solution.

2.2. X-ray diffraction experiments

2.2.1. Texture measurements. For X-ray texture measure-

ments a set of six incomplete pole figures, {1010}, {0002},

{1011}, {1012}, {1020} and {1013}, were measured between 5

and 75� in the back-reflection mode, using Co K� radiation at

35 kV and 28 mA. The measured pole figure data were used to

calculate the orientation distribution function (ODF) using

the MTEX toolbox (Hielscher & Schaeben, 2008). The {0001}

and {0110} pole figures, and ODF sections at ’2 = 0 and 30� for

the specimens deformed at RT, 473 K and 673 K, are shown in

Figs. 2 and 3, respectively. The volume fractions of the fiber

textures h1010i || ED shown in Figs. 2 and 3 were calculated

using de la Vallee Poussin as the kernel function with a half-

width value of 5�. The allowed angular spread was 25�. In

addition, the volume fractions of the local texture maxima in

the ODFs were calculated using the same angular spread of

25�. The measured surface (TD–ED) was parallel to the

tension axis. We note that the texture strength increased at

473 K and decreased at 673 K. The most probable reason for

this is that at 473 K prismatic hai slip is dominant, which helps

to maintain the strong initial [0110] texture. At 673 K, pyra-

midal hc + ai slip becomes thermally activated and its opera-

tion provides a multiple slip condition that is known to weaken

the texture. More details about the texture measurements and

results are given by Al-Samman et al. (2010).

2.2.2. X-ray diffraction for line profile analysis. The X-ray

diffraction measurements were carried out using a high-reso-

lution diffractometer dedicated to line profile analysis with a

plane Ge(220) primary monochromator operated at a Cu K�fine-focus rotating copper anode (Nonius, FR-591) at 45 kV

and 80 mA (Ungar et al., 1998). The distance between the

research papers

56 Bertalan Joni et al. � Dislocation densities and slip-system types J. Appl. Cryst. (2013). 46, 55–62

Figure 1The true-stress–true-strain curves of tensile deformation measured atdifferent temperatures.

Figure 2The {0001} and {0110} pole figures for the specimens tensile deformed atdifferent temperatures up to fracture.

source and the monochromator is 240 mm and a slit of about

160 mm is positioned before the monochromator, at a distance

of 200 mm from the X-ray source. At this distance the K�1 and

K�2 components are received by the Ge(220) crystal at a large

enough separation to allow for cutting off the K�2 component

by means of the 160 mm-wide slit. The Cu K�1 beam has a size

of about 0.2 � 3 mm on the specimen surface. The scattered

radiation is registered by three flat imaging plates (IPs) with a

linear spatial resolution of 50 mm. The first two and the third

IP are placed at distances of 500 and 300 mm from the

specimen, respectively. The first two and the third IP cover the

2� angular ranges from 30 to 52, 56 to 94 and 98 to 127�,

respectively. Since the first two IPs are at a larger distance

from the specimen they allow for a better angular resolution in

the case of the narrower peaks in the smaller angular range. In

the higher angular range the peaks are broader by nature;

therefore the somewhat smaller angular resolution in this case

provides the same relative angular resolution in the entire

angular range measured. At the same time, the shorter

specimen-to-IP distance in the higher angular range provides

better counting statistics for the considerably broader and

weaker peaks. The distances between the specimen and

detector are selected such that the instrumental effect is

always less than 10% of the physical broadening. The

diffraction geometry is of parallel-beam type, and therefore

the specimen does not have to be moved while the angular

resolution is sufficiently good over the entire angular range of

measurement. The arrangement of the diffraction geometry is

shown in Fig. 4. The specimen surface is normal and parallel to

the z and y axes, respectively, and the incident beam makes an

angle ! with the specimen surface, which is constant during

the measurement of one particular diffraction pattern. Here

! ’ 15� was used. The instrumental effect in the double-

crystal diffractometer is considerably smaller than the line

broadening corresponding to the NIST SRM-660a LaB6

standard material, as was shown quantitatively in Fig. 4 of

Gemes et al. (2010). The beam size on the specimen in the

plane of incidence is 200 mm, the distance between the

specimen and detector is 500 mm in the lower and 300 mm in

the higher angular range, and the pixel size of the IP detector

is 50 mm. The beam divergence is 5� 10�4 rad. On the basis of

these values the FWHM of the instrumental broadening is

between 2 � 10�3 and 5 � 10�3 nm�1 (Gemes et al., 2010),

corresponding to a dislocation density of less than about

1012 m�2.

The diffraction patterns were obtained by integrating the

intensity distributions along the corresponding Debye–

Scherrer arcs on the IPs. Only the central quarter, about

25 mm-high regions of the images, were used. With the

geometrical values of the beam the breadths of the peaks

research papers

J. Appl. Cryst. (2013). 46, 55–62 Bertalan Joni et al. � Dislocation densities and slip-system types 57

Figure 4A schematic representation of the arrangement of the diffractiongeometry. ! is the angle between the incoming beam direction and thespecimen surface, �hkl is the Bragg angle of the hkl reflection, ehkl is thenormal unit vector of the hkl planes, and � � ! is the angle between ehkl

and the normal of the measured specimen surface, z. The y coordinate lieson the specimen surface. The dashed cone indicates schematically theangular range, 2��hkl, within which the hkl reflecting planes correspondto the same texture component. The 2��hkl values are correlated to thevalue of m.r.d.

Figure 3ODF sections at ’2 = 0 and 30� for the specimens deformed at RT (a),473 K (b) and 673 K (c). The maxima corresponding to the major texturecomponent are at ’1 = ’2 = 0� and � = 31.8�, �= 73.37� and � = 73.41� forthe specimens deformed at RT, 473 K and 673 K, respectively.

obtained from this region are affected less than 5 �

10�4 nm�1, which is less than 5% in the case of the arcs with

the strongest curvature. This geometrical error becomes

exactly zero when the arcs straighten. Since line profile

analysis depends neither on the absolute nor on the relative

intensities of diffraction peaks, all Debye–Scherrer arcs were

integrated within the same rectangular region in all the

imaging plates used. In order to obtain the precise positions of

the diffraction peaks, diffraction patterns were also measured

in the conventional Bragg–Brentano geometry with a Philips

X’pert diffractometer.

3. Evaluation of the X-ray diffraction experiments

3.1. Identifying the diffraction peaks corresponding to thedifferent texture components

The texture measurements show that the specimens have

well developed fiber textures. The different reflections in the

diffraction patterns correspond to different grain populations.

These different grain populations may correspond to different

texture components and can have different substructures in

terms of Burgers vector types and dislocation densities.

Therefore the groups of reflections belonging to different

texture components have to be evaluated separately. The

grouping of reflections is done as follows.

We define the Cartesian coordinate system of the specimen

by the z axis normal to the specimen surface and the incoming

beam is in the yz plane. When the c and one of the a axes are

parallel to the z and y axes, respectively, the normal vector of

the (hkl) plane is

ehkl ¼ dhkl

2hþ k

31=2a;

k

a;

l

c

� �; ð1Þ

where dhkl is the separation of the (hkl) planes and a and c are

the lattice parameters. The pole figures define the orientation

of the hexagonal unit cell in the specimen. Denoting the

standard three-dimensional rotation matrix by R’1,�,’2, where

’1, �, ’2 are the Eulerian angles of rotation as defined by

Bunge (1996), the normal vectors of the (hkl) planes corre-

sponding to a specific texture are

e0hkl ¼ R’1;�;’2 ehkl: ð2Þ

With this notation, and in accordance with Fig. 4, the

diffraction angle of an (hkl) plane relative to the incoming

beam direction is

�hkl ¼ 90� � arccosðe0y cos!þ e0z sin!Þ: ð3Þ

Taking into account the multiple-of-random-distribution

(m.r.d.) values of the textures, an hkl reflection in the

measured pattern corresponds to the major texture compo-

nent if its �hkl value is within �hkl = �B � 25�, where �B is the

Bragg angle of the hkl reflection. The reflections that do not

satisfy this criterion are considered to correspond to the

random texture component (or grains). Though the reflections

are sorted into two groups, related to either the major (Ma) or

the random (R) texture components, the entire pattern is

evaluated simultaneously as if the two texture components

were two ‘phases’ in the specimen.

A typical diffraction pattern of the specimen deformed at

RT and measured on the ND surface is shown in Fig. 5(a).

Enlarged sections of the patterns obtained on the ND surfaces

of the specimens deformed at RT, 473 and 673 K, with loga-

rithmic intensity scales and shifted relative to each other, are

shown in Fig. 5(b). The quality of the fits is given by the

goodness-of-fit (GoF) values defined in Table 1 on p. 153 of

Langford & Louer (1996). The GoF values of the measured

patterns are given in Table 1 in the present work.

research papers

58 Bertalan Joni et al. � Dislocation densities and slip-system types J. Appl. Cryst. (2013). 46, 55–62

Figure 5(a) Measured (open circles) and fitted (line; red in the electronic versionof the journal) patterns of the specimen deformed at RT and measured onthe ND surface. The difference between the measured and fitted data isshown at the bottom of the figure. (b) Enlarged sections of the measured(open circles) and fitted (red lines) patterns, with logarithmic intensityscales, of the specimens deformed at RT, 473 K and 673 K obtained on theND surfaces. The pattern sections are shifted vertically for cleardistinction. ‘Ma’ or ‘R’ at the hkl values indicate that the reflectioncorresponds to the major or random texture component, respectively.[The Ma/R labels are never shown redundantly. In (b) about 80% of thedata points have been omitted for the purpose of clarity.]

It is noted here that, though in the present investigation we

used only one single angle of incidence, !, it is easy to apply

two or three different ! values and obtain two or three

diffraction patterns, each corresponding to a different ! value.

By collecting these different patterns the number of reflec-

tions, and thus the number of peak profiles corresponding to

the different texture components, can be increased.

3.2. Determination of the substructure parameters in termsof dislocation densities, slip-system types, twin-boundarydensities and subgrain size

The evaluation of diffraction patterns is done by assuming

that the different physical effects are superimposed by

convolution (Warren, 1959). The measured, IM(2�) diffraction

patterns are matched by the theoretically calculated and

convoluted profile functions of the effects of size, distortion

and planar defects as well as instrumental effects (IShkl, ID

hkl , ISFhkl

and IINSThkl , respectively):

IMð2�Þ ¼

Phkl

IShkl � ID

hkl � ISFhkl � IINST

hkl þ IBG; ð4Þ

whereP

hkl IShkl � ID

hkl � ISFhkl � IINST

hkl is the theoretically calcu-

lated pattern and IBG is the background determined separately

(Ribarik et al., 2004). Assuming an equiaxed shape and

lognormal size distribution the size profile is (Ungar et al.,

2001)

ISðsÞ ¼

Z1

0

xsin2ðx�sÞ

ð�sÞ2 erfc

lnðx=mÞ

21=2�

� �dx; ð5Þ

where s = (2cos�/�)��, erfc is the complementary error

function, and m and � are the median and variance of the

lognormal size-distribution density function. The Fourier

transform of the strain profile is (Warren, 1959)

FTðIDhklÞ ¼ expð�2�2L2g2

h"2g;LiÞ; ð6Þ

where g is the absolute value of the diffraction vector and L is

the Fourier variable. The mean-square strain in dislocated

crystals is given by

h"2g;Li ffi

�Cb2

4�f ðÞ; ð7Þ

where � and b are the density and the magnitude of the

Burgers vector of dislocations and C is the dislocation contrast

factor. f() is the Wilkens function, where = L/Re and Re is

the effective outer cutoff radius of dislocations (Wilkens,

1970). In a polycrystal, or if all possible slip systems are

equally populated, the dislocation contrast factors, C, can be

averaged over the permutations of the hkl indices and the

mean-square strain. If, however, more Burgers vectors can be

present, especially with different absolute values and espe-

cially in hexagonal crystals, the averaging has to be extended

over the Burgers vectors, as indicated by the bar in equation

(8) (Ungar et al., 2007).

In Mg there are a large number of slip-plane families with

three different Burgers vector types (Jones & Hutchison,

1981). The average value of Cb2 can be written as (Ungar et al.,

2007)

Cb2 ¼PNi¼1

fiCðiÞ

b2i ; ð8Þ

where N is the number of slip-plane families, CðiÞ

and bi are the

average dislocation contrast factors and the Burgers vector

magnitudes corresponding to the ith slip-plane family, and fi

are the fractions by which the particular slip-plane families

contribute to the broadening of a specific reflection. In the

case of hexagonal crystals N is equal to 11 (Jones & Hutchison,

1981). In a texture-free polycrystal or if all possible Burgers

vectors are activated in a particular slip system the dislocation

contrast factors can be averaged over the permutations of the

corresponding hkl indices (Ungar & Tichy, 1999). The average

contrast factors for a particular slip-plane family in hexagonal

crystals can be given as (Ungar et al., 1999)

CðiÞ

hk:l ¼ CðiÞ

hk:0½1þ qðiÞ1 xþ q

ðiÞ2 x2; ð9Þ

where x = (2/3)(l/ga)2, qðiÞ1 and q

ðiÞ2 are parameters depending

on the ith slip-plane family and the elastic properties of the

material, CðiÞ

hk:0 is the average contrast factor of the ith slip-

plane family corresponding to the hk.0-type reflections, and a

is the lattice constant in the basal plane. The qðiÞ1 and q

ðiÞ2

parameters and the values of CðiÞ

hk:0 have been evaluated

numerically and compiled for a large number of hexagonal

crystals and compounds by Ungar et al. (1999). The measured

values of q1 and q2, denoted as qðmÞ1 and q

ðmÞ2 , are provided by

the extended convolutional whole profile (eCMWP) evalua-

tion procedure (Balogh et al., 2009). The X-ray measurements

provide only two independent parameters: qðmÞ1 and q

ðmÞ2 . Since

the number of slip-plane families is about 11, it is not possible

to determine the fraction, fi, of each slip-plane family but only

the fractions of slip-plane family types, i.e. the fractions of hai,

hc + ai and hci family types, hhai, hhc+ai and hhci, respectively.

research papers

J. Appl. Cryst. (2013). 46, 55–62 Bertalan Joni et al. � Dislocation densities and slip-system types 59

Table 1The q1 and q2 parameter values in the contrast factors, as in equation (9),the dislocation densities, �, and the fractions of the prevailing dislocationtypes, hhai, hhci and hhc+ai.

The numerical quality of the fit is given by the GoF parameters for thepatterns measured on the ED and ND surfaces. The GoF values are listed onlyonce for each pattern since they correspond to the pattern itself and not to thetexture components. Values in parentheses are standard uncertainties on theleast significant digits.

Major texture component.

T(K) q1 q2

�(1014 m�2)

hhai(%)

hhci(%)

hhc+ai

(%)GoF(ED/ND)

RT 0.4 (2) �0.4 (2) 3 (1) 0.70 (15) 0.0 (1) 0.3 (1) 0.122/0.118473 0.7 (2) �0.3 (2) 1.9 (7) 0.6 (1) 0.1 (1) 0.3 (1) 0.057/0.053673 0.9 (3) �0.4 (2) 0.46 (10) 0.5 (1) 0.1 (1) 0.4 (1) 0.112/0.108

Other grains.

T(K) q1 q2

�(1014 m�2)

hhai(%)

hhci(%)

hhc+ai

(%)GoF(ED/ND)

RT 1.0 (4) �0.7 (3) 1.8 (10) 0.75 (15) 0.0 (1) 0.25 (10) –473 1.0 (4) �0.7 (3) 1.3 (7) 0.70 (15) 0.0 (1) 0.3 (1) –673 0.5 (2) �0.7 (3) 0.4 (1) 0.6 (1) 0.0 (1) 0.4 (1) –

The details of this evaluation procedure are given by Ungar et

al. (2007).

The effect of twinning on the diffraction patterns in hexa-

gonal close-packed materials has been worked out for the

{10.1}h10.2i and {11.2}h11.3i compressive and {10.2}h10.1i and

{11.1}h11.6i tensile twin planes (Balogh et al., 2009). It was

shown that the profile functions for twinning are the sum of

symmetrical and antisymmetrical Lorentz functions (Balogh et

al., 2009). The antisymmetrical Lorentz function is the

consequence of interference between two overlapping subre-

flections in reciprocal space, which correspond to the parent

and the twin crystal. The symmetrical and antisymmetrical

parts of the subreflections are strictly correlated and can be

characterized by a breadth value, B, and an antisymmetry

parameter, A (Balogh et al., 2009). The B and A values for the

symmetrical and antisymmetrical Lorentzian profile functions

have been parameterized as a function of fault density for the

different stacking faults and twins and are available at http://

metal.elte.hu/~levente/stacking.

4. Results and discussion

4.1. Mechanical response, texture evolution and deformationmechanisms

The mechanical response of the deformed specimens under

tension at different temperatures, i.e. RT, 473 K and 673 K, at

a constant strain rate of 10�2 s�1 up to failure is shown in Fig. 1.

The flow curves clearly show that the strength and elongation

to failure strongly depend on the temperature of deformation.

At RT the constraints of the limited number of available slip

systems cause a considerable initial hardening and a high

ultimate tensile stress of about 255 MPa. For the current

deformation geometry and starting texture (typical h1010i ||

ED extrusion texture), the resolved shear stress on the basal

planes was close to zero, and the only possible deformation

mechanisms were prismatic slip and {1012}-compression

twinning. From the texture evolution in Fig. 2, it is evident that

the strong extrusion texture was retained during tensile

deformation at RT up to " ’ 0.12, the failure strain. This is

usually an indication of a predominant prismatic slip activity

with the resulting slip direction lying parallel to the loading

axis. In such a case, accommodation of plastic strain is

maximum in the basal plane and equal to zero along the c axis,

which means that the orientation of individual grains during

straining remains virtually unchanged (Agnew & Duygulu,

2005). As mentioned above, compression twinning is a

possible deformation mechanism in the current condition,

despite its high activation stress. However, since its volume

fraction is usually quite low, it is not likely that it would affect

the texture evolution and mechanical response in a quantita-

tive manner. At higher temperatures, i.e. at 473 and 673 K,

additional slip systems, such as pyramidal hc + ai, become

thermally activated and can thus contribute to plastic defor-

mation, which is one reason for the ductility enhancement

observed in the flow curves in Fig. 1. However, the corre-

sponding texture evolution in Fig. 2 indicates that prismatic

slip remained the dominant mechanism for strain accom-

modation, simply because the main strain component along

the ED does not require a c-axis accommodation mechanism.

The other reason for the delay of failure is dynamic recrys-

tallization during the deformation process. The calculated

volume fraction of the fiber texture with the h1010i || ED fiber

orientation (see also x2.2.1) was 0.697 at RT, 0.726 at 473 K

and 0.628 at 673 K. The calculated volume fraction of local

maxima (part of the fiber texture) was 0.314 at RT, 0.321 at

473 K and 0.279 at 673 K.

4.2. Dislocation densities and slip-system activity

The values of the q1 and q2 parameters, as defined in

equation (9), the dislocation densities, �, and the fractions of

the slip-system types, hhai, hhci and hhc+ai, corresponding to the

different temperatures of deformation are listed Table 1. All

parameter values are evaluated separately for the main

texture component and for the grain population that does not

belong to the main texture component. The latter are denoted

in the table as ‘other grains’. The values in Table 1 indicate

that (i) the dislocation density decreases strongly with the

temperature of deformation and (ii) in the main texture

component the dislocation density is larger than that in the

other grains when the deformation is performed at RT or

473 K. After deformation at 673 K the dislocation densities

are very small and almost equal within the experimental error.

The results indicate that the major part of deformation and

work hardening is carried within the grain population corre-

sponding to the main texture component. This suggests that

the grain population not within the main texture component,

i.e. the ‘other grains’, most probably acts for balancing strain

incompatibility produced during the deformation of the grain

population inside the major texture component. More specific

conclusions would require numerical polycrystal plasticity

simulations.

The slip-system types are determined by carrying out the

analysis described briefly in x3.2 and equations (8) and (9).

Here the q1 and q2 parameters in the dislocation contrast

factors in equation (9) play a crucial role. In Table 1 it can be

seen that, though these parameter values do change, the

research papers

60 Bertalan Joni et al. � Dislocation densities and slip-system types J. Appl. Cryst. (2013). 46, 55–62

Figure 6The physically possible range in which the q1 and q2 parameter values inequation (9) can change (the large rectangle), along with the range inwhich the measured q1 and q2 values are varying (small dashed rectangle).

variation of the slip-system types is relatively small. In order to

make this situation clear, in Fig. 6 we plot the physically

possible range in which the q1 and q2 parameter values can

change (the large rectangle), along with the range in which the

measured q1 and q2 values are varying (small dashed

rectangle). The figure shows that the measured q1 and q2

values vary in a range that is relatively small compared to the

range of possible values.

From the values in Table 1 it can be seen that the largest

fraction of prevailing dislocations are of hai type, the fraction

of hc + ai-type dislocations varies between 0.25 and 0.4, and

hci-type dislocations are either missing or only present with

fractions within the experimental error. The values also indi-

cate that with increasing temperature of deformation the

fraction of hai-type dislocations decreases, whereas that of the

hc + ai-type dislocations increases concomitantly. This is in

good correlation with earlier X-ray diffraction (Mathis et al.,

2005), acoustic emission (Mathis, Chmelik et al., 2004; Mathis

et al., 2011) and transmission electron microscopy (Agnew et

al., 2002; Jain et al., 2008; Mathis et al., 2005) results. The grain

size of the specimens was of the order of about 100 mm. The

subgrain size provided by the X-ray diffraction analysis is

larger than about 500 nm in all specimens after deformation at

the three temperatures investigated here. This means that

there is practically no size effect on X-ray line broadening.

The correlation between the substructure determined here

and the ultimate tensile stress, �UTS, given by the stress–strain

curves in Fig. 1 can be investigated using the Taylor (1934)

equation:

�UTS ¼ �o þ fMa�MTGb�1=2Ma þ fR�MTGb�1=2

R ; ð10Þ

where �o is the friction stress in the alloy, fMa and fR and �Ma

and �R are the volume fractions and dislocation densities of

the material corresponding to the major texture component

and to the grain population outside this, respectively, � is a

constant between zero and unity, G is the shear modulus, and

MT is the Taylor factor. The values of �o are taken from the

flow stress curves, also shown in Fig. 1, i.e. �o = 185 (15),

115 (15) and 40 (8) MPa for the specimens deformed at RT,

473 K and 673 K, respectively. The volume fractions of the

major and random texture components, fMa and fR, are

obtained from the numerical evaluation of the ODF sections

shown in Fig. 3: fMa = 0.7 and fR = 0.3, respectively. The shear

modulus is G = 17 GPa (Caceres et al., 2002). The Taylor factor

is determined by taking into account the fractions of the active

slip systems by using the slip-system phase diagram in Fig. 5 of

Caceres & Lukac (2008), based on the calculations of Graff et

al. (2007). The values of MT vary between about 2.5 and 4. In

the present case, taking into account the fractions of the hai-

and hc + ai-type slip systems, we use MT = 3.2. Since the

prevailing slip systems in the major and random texture

components are the same within experimental error, we use

the same Taylor factor for the two texture components. The

Burgers vector, b, is taken as the weighted average of Burgers

vectors of the basal and pyramidal slip-plane type in accor-

dance with the values in Table 1: b = 0.4083 nm. With these

parameter values the measured and calculated �UTS values,

�measured and �calculated, are plotted in Fig. 7. The measured and

calculated �UTS values in the figure are obtained with � = 0.25

and �o = 180 (15), 110 (15) and 30 (8) MPa for the specimens

deformed at RT, 473 K and 673 K, respectively.

5. Conclusions

(1) An algorithm is elaborated to sort diffraction peaks

produced by grain populations corresponding to particular

texture components. The algorithm allows the correlation of

groups of reflections in a diffraction pattern with texture

described either by pole figures or by orientation distribution

functions.

(2) Dislocation densities and the prevailing slip-system

types are determined by separating the reflections into

different groups, where each group consists of diffraction

peaks produced by grains corresponding to one particular

texture component or orientation. The different groups of

grains are treated as if they belong to different phases.

(3) We have shown that the dislocation densities in the

major texture components are always larger than in other

grain populations at all temperatures of deformation investi-

gated here.

(4) About a fraction of two-thirds of the prevailing dislo-

cations are of hai type after deformation at each temperature.

The fraction of hc + ai type increases from about one-third to

slightly larger values with the temperature of deformation, in

correlation with other observations.

(5) The ultimate tensile stress values measured at different

temperatures are in good correlation with the dislocation

densities when using Taylor’s equation with � = 0.25.

(6) Though the evaluation by using the eCMWP procedure

allows for twins, the average densities of the twin boundaries

are obtained to be zero within experimental error. This result

is in good correlation with the stress–strain curves in Fig. 1.

(7) By collecting diffraction patterns corresponding to more

than one ! value the number of Bragg peaks and thus the

research papers

J. Appl. Cryst. (2013). 46, 55–62 Bertalan Joni et al. � Dislocation densities and slip-system types 61

Figure 7The measured and calculated �UTS values, �measured and �calculated, asevaluated using equation (10). The thick vertical line indicates the error.

number of peak profiles corresponding to each of the different

texture components can be increased.

TU is grateful to the Hungarian National Science Founda-

tion (OTKA; grant Nos. 71594 and 67692) for the support of

this work. The European Union and European Social Fund

have provided financial support to this project under grant

agreement No. TAMOP 4.2.1./B-09/1/KMR-2010-0003 (to

TU).

References

Agnew, S. R. & Duygulu, O. (2005). Int. J. Plast. 21, 1161–1193.Agnew, S. R., Horton, J. A. & Yoo, M. H. (2002). Metall. Mater. Trans.

A, 33, 851–858.Al-Samman, T., Li, X. & Chowdhury, S. G. (2010). Mater. Sci. Eng. A,

527, 3450–3463.Balogh, L., Tichy, G. & Ungar, T. (2009). J. Appl. Cryst. 42, 580–591.Bunge, H. J. (1996). Textures Microstruct. 25, 71–108.Caceres, C. H., Griffiths, J. R., Davidson, C. J. & Newton, C. L. (2002).

Mater. Sci. Eng. A, 325, 344–355.Caceres, C. H. & Lukac, P. (2008). Philos. Mag. 88, 977–989.Gemes, Gy., Balogh, L. & Ungar, T. (2010). Metall. Mater. (Kovove

Mater.), 48, 33–39.Graff, S., Brocks, W. & Steglich, D. (2007). Int. J. Plast. 23, 1957–1978.Hielscher, R. & Schaeben, H. (2008). J. Appl. Cryst. 41, 1024–1037.Jain, A., Duygulu, O., Brown, D. W., Tome, C. N. & Agnew, S. R.

(2008). Mater. Sci. Eng. A, 486, 545–555.Jones, I. P. & Hutchinson, W. B. (1981). Acta Metall. 29, 951–968.Klimanek, P. (1994). Mater. Sci. Forum, 157–162, 1119–1130.Klimanek, P., Weidner, A., Esling, C. & Philippe, M.-J. (1996).

Proceedings of the 11th International Conference on Textures of

Materials, edited by Z. Liang, L. Zuo & Y. Chu, pp. 1443–1448.Beijing: International Academic Publishers,

Langford, J. I. & Louer, D. (1996). Rep. Prog. Phys. 59, 131–234.Lukac, P. & Mathis, K. (2002). Kovove Mater. 40, 281–289.Mathis, K., Capek, J., Zdrazilova, Z. & Trojanova, Z. (2011). Mater.

Sci. Eng. A, 528, 5904–5907.Mathis, K., Chmelik, F., Trojanova, Z., Lukac, P. & Lendvai, J. (2004).

Mater. Sci. Eng. A, 387–389, 331–335.Mathis, K., Gubicza, J. & Nam, N. H. (2005). J. Alloys Compd. 394,

194–199.Mathis, K., Nyilas, K., Axt, A., Dragomir-Cernatescu, I., Ungar, T. &

Lukac, P. (2004). Acta Mater. 52, 2889–2894.Obara, T., Yoshinaga, H. & Morozumi, S. (1973). Acta Metall. 21, 845–

853.Ribarik, G., Gubicza, J. & Ungar, T. (2004). Mater. Sci. Eng. A, 387–

389, 343–347.Taylor, G. I. (1934). Proc. R. Soc. London Ser. A, 145, 362–387.Ungar, T., Castelnau, O., Ribarik, G., Drakopoulos, M., Bechade, J. L.,

Chauveau, T., Snigirev, A., Snigireva, I., Schroer, C. & Bacroix, B.(2007). Acta Mater. 55, 1117–1127.

Ungar, T., Dragomir, I., Revesz, A. & Borbely, A. (1999). J. Appl.Cryst. 32, 992–1002.

Ungar, T., Gubicza, J., Ribarik, G. & Borbely, A. (2001). J. Appl.Cryst. 34, 298–310.

Ungar, T., Ott, S., Sanders, P. G., Borbely, A. & Weertman, J. R.(1998). Acta Mater. 46, 3693–3699.

Ungar, T. & Tichy, G. (1999). Phys. Status Solidi (a), 147, 425–434.Wang, H., Raeisinia, B., Wu, P. D., Agnew, S. R. & Tome, C. N. (2010).

Int. J. Solid Struct. 21, 2905–2917.Warren, B. E. (1959). Prog. Met. Phys. 8, 147–202.Wilkens, M. (1970). Fundamental Aspects of Dislocation Theory,

edited by S. A. Simmons, R. de Wit & R. Bullough, Vol. II, NationalBureau of Standards Special Publication No. 317, p. 1195.Washington, DC: US Government Printing Office.

Yoo, M. H. (1981). Metall. Trans. A, 12, 409–418.

research papers

62 Bertalan Joni et al. � Dislocation densities and slip-system types J. Appl. Cryst. (2013). 46, 55–62


Recommended