high purity ultrafine-grained nickel processed by dynamic plastic deformation: microstructure and...

7
DOI: 10.1002/adem.201200034 High Purity Ultrafine-Grained Nickel Processed by Dynamic Plastic Deformation: Microstructure and Mechanical Properties** By Lukasz Farbaniec, Akrum Abdul-Latif, Jeno ˜ Gubicza and Guy Dirras* Metals and their alloys play an important role in many engineering fields such as construction and manufacturing of equipments and tools. Therefore, new materials and/or microstructure improvement of well-known materials (lead- ing to a combination of enhanced properties and long life cycle) are required. As for structural materials, a high strength combined with good ductility are properties of interest. Indeed, the grain refinement is an effective way to increase the strength of materials as expressed by the Hall–Petch relation- ship. [1,2] For this reason the ultrafine-grained (UFG, the average grain size varies between 100 and 1000 nm) and nanocrystalline (the average grain size is less than 100 nm) materials have attracted considerable attention in recent years since the pioneering work of Gleiter. [3–5] The increase in strength and hardness in metals can be achieved by grain subdivision during plastic deformation. For this reason a lot of severe plastic deformation (SPD) methods were developed in the last decades. [6,7] Among them, high pressure torsion (HPT) and equal channel angular pressing (ECAP) are the most frequently used SPD techniques. [8–19] In these proce- dures, an initial coarse-grained bulk material is subjected to large strain under high hydrostatic pressure yielding a uniform microstructure with fine grains without damage of the specimen. Lately, a new method referred to as dynamic plastic deformation (DPD) also showed its ability to achieve grain refinement in different materials. [20,21] Actually, it was found that under dynamic loading conditions the deformed microstructure is very similar to that processed by ECAP, including elongated or equiaxed grains, and subgrain structure, depending on the applied route of deformation. [23] DPD technique was recently used to produce UFG micro- structures in Al, Cu, or Cu–Ag alloy [20–23] in one step or in combination with powder metallurgy methods. [24–26] These studies revealed that strain rate, deformation temperature and stacking fault energy (SFE) play a significant role in dislocation patterning which subsequently results in the formation of new grain boundaries during further strain- ing. [27–31] In this context, the combined effect of the strain rate and the deformation temperature on the evolution of the COMMUNICATION [*] Prof. G. Dirras, L. Farbaniec Universite´Paris 13, Sorbonne Paris Cite´ LSPM-CNRS, 99, Avenue J. B. Cle´ment, 93430 Villetaneuse, France E-mail: [email protected] Prof. A. Abdul-Latif Laboratoire d’Inge´nierie des Syste`mes Me´caniques et des Mate´riaux 3, Rue Fernand Hainaut, 93407 St. Ouen Cedex, France Dr. J. Gubicza Department of Materials Physics, Eo ¨tvo ¨s Lora´nd University Budapest, P.O. Box 32, H-1518, Hungary [**] The authors like to thank Mr. Thierry Chauveau for fruitful discussions and texture analysis. This work was supported by the French National Research Agency, ANR, via the ‘‘MIMIC’’ Project (ANR-09-BLAN-0010), the Hungarian Scientific Research Fund, OTKA, Grant No. K-81360 and partly by the French-Hungarian bilateral project (PHC Balaton, Hungarian Grant No. TeT_10-1-2011-0737). Bulk ultrafine-grained samples are processed by dynamic plastic deformation at an average strain rate of 3.3 10 2 s 1 from bulk coarse-grained nickel with purity higher than 98.4 wt.%. The obtained microstructure is investigated by electron backscattering diffraction, transmission electron microscopy and X-ray line profile analysis. After dynamic deformation the microstructure evolves into sub- micron-size lamellar and subgrain structures. Evaluation of average grain size shows a heterogeneous microstructure along both the diameter and the thickness of the sample. X-ray line profile analysis reveals high dislocation density of about 13 2 10 14 m 2 in the impacted material. The mechanical properties are investigated by means of uniaxial quasi-static compression tests conducted at room temperature. The stress–strain behavior of the impacted Ni depends on the location in the impacted disk and on the orientation of the compression axis relative to the impact direction. ADVANCED ENGINEERING MATERIALS 2012, 14, No. 11 ß 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim wileyonlinelibrary.com 1027

Upload: independent

Post on 11-Nov-2023

0 views

Category:

Documents


0 download

TRANSCRIPT

CO

DOI: 10.1002/adem.201200034

MM

UNIC

AT

High Purity Ultrafine-Grained Nickel Processed byDynamic Plastic Deformation: Microstructure andMechanical Properties**

ION

By Lukasz Farbaniec, Akrum Abdul-Latif, Jeno Gubicza and Guy Dirras*

Bulk ultrafine-grained samples are processed by dynamic plastic deformation at an average strain rateof 3.3� 102 s�1 from bulk coarse-grained nickel with purity higher than 98.4 wt.%. The obtainedmicrostructure is investigated by electron backscattering diffraction, transmission electron microscopyand X-ray line profile analysis. After dynamic deformation the microstructure evolves into sub-micron-size lamellar and subgrain structures. Evaluation of average grain size shows a heterogeneousmicrostructure along both the diameter and the thickness of the sample. X-ray line profile analysisreveals high dislocation density of about 13� 2� 1014 m�2 in the impacted material. The mechanicalproperties are investigated by means of uniaxial quasi-static compression tests conducted at roomtemperature. The stress–strain behavior of the impacted Ni depends on the location in the impacted diskand on the orientation of the compression axis relative to the impact direction.

Metals and their alloys play an important role in many

engineering fields such as construction and manufacturing of

equipments and tools. Therefore, new materials and/or

microstructure improvement of well-known materials (lead-

ing to a combination of enhanced properties and long life

cycle) are required. As for structural materials, a high strength

combined with good ductility are properties of interest.

Indeed, the grain refinement is an effective way to increase the

strength of materials as expressed by the Hall–Petch relation-

ship.[1,2] For this reason the ultrafine-grained (UFG, the

[*] Prof. G. Dirras, L. FarbaniecUniversite Paris 13, Sorbonne Paris CiteLSPM-CNRS, 99, Avenue J. B. Clement, 93430 Villetaneuse,FranceE-mail: [email protected]

Prof. A. Abdul-LatifLaboratoire d’Ingenierie des Systemes Mecaniques et des Materiaux3, Rue Fernand Hainaut, 93407 St. Ouen Cedex, France

Dr. J. GubiczaDepartment of Materials Physics, Eotvos Lorand UniversityBudapest, P.O. Box 32, H-1518, Hungary

[**] The authors like to thank Mr. Thierry Chauveau for fruitfuldiscussions and texture analysis. This work was supportedby the French National Research Agency, ANR, via the‘‘MIMIC’’ Project (ANR-09-BLAN-0010), the HungarianScientific Research Fund, OTKA, Grant No. K-81360 andpartly by the French-Hungarian bilateral project (PHC Balaton,Hungarian Grant No. TeT_10-1-2011-0737).

ADVANCED ENGINEERING MATERIALS 2012, 14, No. 11 � 2012 WILEY-VCH Ve

average grain size varies between 100 and 1000 nm) and

nanocrystalline (the average grain size is less than 100 nm)

materials have attracted considerable attention in recent years

since the pioneering work of Gleiter.[3–5] The increase in

strength and hardness in metals can be achieved by grain

subdivision during plastic deformation. For this reason a lot

of severe plastic deformation (SPD) methods were developed

in the last decades.[6,7] Among them, high pressure torsion

(HPT) and equal channel angular pressing (ECAP) are the

most frequently used SPD techniques.[8–19] In these proce-

dures, an initial coarse-grained bulk material is subjected to

large strain under high hydrostatic pressure yielding a

uniform microstructure with fine grains without damage of

the specimen. Lately, a new method referred to as dynamic

plastic deformation (DPD) also showed its ability to achieve

grain refinement in different materials.[20,21] Actually, it was

found that under dynamic loading conditions the deformed

microstructure is very similar to that processed by ECAP,

including elongated or equiaxed grains, and subgrain

structure, depending on the applied route of deformation.[23]

DPD technique was recently used to produce UFG micro-

structures in Al, Cu, or Cu–Ag alloy[20–23] in one step or in

combination with powder metallurgy methods.[24–26] These

studies revealed that strain rate, deformation temperature

and stacking fault energy (SFE) play a significant role in

dislocation patterning which subsequently results in the

formation of new grain boundaries during further strain-

ing.[27–31] In this context, the combined effect of the strain

rate and the deformation temperature on the evolution of the

rlag GmbH & Co. KGaA, Weinheim wileyonlinelibrary.com 1027

COM

MUNIC

ATIO

N

L. Farbaniec et al./High Purity Ultrafine-Grained Nickel

microstructure during plastic deformation can be expressed

by the so-called Zener–Hollomon parameter:[32]

Z ¼ _"� expQ

RT

� �

where Q is the activation energy of the mechanism controlling

the rate of deformation, T is the absolute temperature, and

R is the universal gas constant. High strain rate and low

deformation temperature (corresponding to high Z-value)

effectively suppress annihilation of dislocations, resulting

in grain refinement by formation of new boundaries from

dislocations. Contrariwise, deformation conditions corre-

sponding to lower Z values lead to smaller dislocation

density, and therefore to lower amount of grain boundaries

and larger grains.[33] At a given temperature the lower the

SFE, the more difficult the annihilation of dislocations and

the larger the probability of the occurrence of twinning. For

materials with medium or high SFE, the deformation at

ambient temperatures occurs mainly by the motion of full

lattice dislocations. When the deformation temperature is low

(e.g., at cryogenic temperatures) the full lattice dislocation

activity may be suppressed and deformation by twinning

plays a key role.[21] In our previous work,[25] we proposed a

DPD method for processing UFG microstructures in coarse-

grained high SFE materials, such as pure Al. The objective of

the present work is to evaluate the potential of this DPD

technique for grain refinement in commercially pure Ni

with lower SFE than for Al. The influence of as-processed

microstructure on subsequent mechanical behavior is inves-

tigated under quasi-static compression conditions.

1. Experimental Procedure

The initial material was a bulk coarse-grained nickel with

an average grain size of about 25mm. According to the

chemical composition given in Table 1 the purity of Ni was

higher than 98.4%. Test specimens having a diameter of

9.5mm and a length of 16mm were machined from the

rod shaped as-received material supplied by Goodfellow

(www.goodfellow.com). Before DPD tests, the samples were

heat-treated in an electrical air circulating furnace (Carbolite

HRF 7/45 type) at 7508C for 1 h and then air cooled in order to

homogenize the microstructure.

DPD tests were conducted by using a 6-m-height dynamic

drop mass bench test machine.[24] An object with a mass of

45 kg fell almost freely from the drop height of 5.5m and hit

the upper surface of the specimen. At the moment of the

Table 1. Chemical composition of the studied nickel material.

Ni Cu Fe Mg

Composition [mg � g�1] Base <2500 <4000 <200

1028 http://www.aem-journal.com � 2012 WILEY-VCH Verlag GmbH & C

impact the kinetic energy of themasswas approximately 2.5 kJ

that was absorbed by the plastic deformation of the specimen.

The corresponding initial maximum impact velocity was

10m � s�1 that yielded an average strain rate of 3.3� 102 s�1. A

dynamic load cell of 20 tons measured the applied axial force

on specimen. The M5L series Bullier International laser

displacement sensor, 5000g accelerometer and laser beam

displacement transducer measured the displacement of the

impacting object. The force and displacement as a function of

time were recorded under control of a rapid acquisition chain

(2.5MHz). A synchronization signal is triggered by a signal

from two photocells.

A Zeiss Supra 40VP FEG scanning electron microscope

(SEM) was used for electron backscattering diffraction (EBSD)

investigations of the microstructure of the initial and the

impacted samples. EBSD maps were recorded with a step

size of 0.1mm between neighboring measurement positions.

The average grain size and the fraction of low and high

angle grain boundaries (LAGB and HAGB, respectively)

were determined utilizing OIM software v.4 from TexSem

Laboratories (TSL). Mechanical polishing followed by electro-

polishing was used to prepare the specimens for microstruc-

tural examination. The electropolishing step was performed in

an A2 electrolytic solution at room temperature (RT) using a

Tenupol1 electropolishing apparatus from Struers.

Transmission electron microscopy (TEM) investigations

were carried out by a JOEL-2010 electron microscope

operating at 200 kV. Thin foils were prepared using a GatanTM

precision ion polishing system (PIPS) with a current of 0.5mA

and a tension of 15 kV. Complementary microstructural

investigations were carried out on the impacted specimen

by X-ray line profile analysis (XLPA). The X-ray line profiles

were measured on the cross-section perpendicular to the

impacted surface by a high-resolution rotating anode

diffractometer (Nonius, FR 591) using Cu Ka1 radiation

(l¼ 0.15406 nm). Two-dimensional imaging plates detected

the Debye–Scherrer diffraction rings. The line profiles were

determined as the intensity distribution perpendicular to the

rings obtained by integrating the two dimensional distribu-

tion of the intensity along the rings. The line profiles were

evaluated by the convolutional multiple whole profile

(CMWP) analysis.[34,35] In this method, the diffraction pattern

is fitted by the sum of a background spline and the

convolution of the instrumental pattern and the theoretical

line profiles related to the crystallite size, dislocations and

twin faults. The mean crystallite size, the dislocation density

(r) and the twin boundary frequency (b) were obtained from

Elements

Mn Si Ti C S

0 <3500 <1500 <1000 <1500 <100

o. KGaA, Weinheim ADVANCED ENGINEERING MATERIALS 2012, 14, No. 11

COM

MUNIC

ATIO

N

L. Farbaniec et al./High Purity Ultrafine-Grained Nickel

Fig. 1. Reference coordinate system attached to the impacted sample. The squaresnumbered as 1 and 2 indicate the locations of the specimens cut from the center and theedge of the impacted sample for mechanical testing.

the fitting. The latter quantity is defined as the relative fraction

of twin boundaries among {111} lattice planes.

X-ray diffraction (XRD) was used to characterize the

crystallographic texture after impact test. A macroscopic

reference coordinate system (X1, X2, X3) was attached

to the sample, where X3 is the normal direction to the

impacted surface while X1 and X2 directions are lying on

the impacted surface (Figure 1). The XRD measurements

Fig. 2. (a) The shape and the dimensions of the sample before and after DPD. Grain boundar(d) regions.

ADVANCED ENGINEERING MATERIALS 2012, 14, No. 11 � 2012 WILEY-VCH Verl

were performed using an InelTM four-circles goniometer

(Bragg-Brentano geometry) with a point focus X-ray source

having a Co Ka radiation (l¼ 0.17902 nm). Experimental

pole figures were determined for the selected {111} and {220}

reflections. After corrections (background, defocusing, and

normalization), the orientation distribution functions (ODF)

were obtained using LabotexTM software from Labosoft

(www.labotex.com).

The mechanical properties of the impacted specimen

were investigated by uniaxial compression tests at RT and a

strain rate of 2� 10�3 s�1 using a 5 kN Deben MICROTEST

tensile/compression/bending stage (Deben UK Ltd.). The

stage has a linear scale with a resistive extensometer for

accurate elongation measurement. The speed of motor was

measured by means of an optical encoder and acquisition

software registered the force-extension data. Prismatic

samples with dimensions of 2mm� 2mm� 3.2mm were

cut radially at two locations in the middle of the impacted

specimen (Figure 1). The directions of the compression axes

were either parallel or perpendicular to the initial impact

direction.

2. Results

2.1. EBSD and TEM Investigations

Figure 2a shows the samples before and after DPD

illustrating the changes in the dimensions due to impact

deformation. The diameter of the sample after impact is twice

as large as in the initial state, while the height was reduced to

about one fourth of the initial value that corresponds to a true

strain of about 1.4. Such deformation can cause significant

y maps obtained by EBSD (b) before DPD and after DPD in the peripheral (c) and center

ag GmbH & Co. KGaA, Weinheim http://www.aem-journal.com 1029

COM

MUNIC

ATIO

N

L. Farbaniec et al./High Purity Ultrafine-Grained Nickel

Fig. 3. TEM images obtained at the periphery with heterogeneous microstructure (a)and lamellae in the center (b) of the DPD-processed sample. (c) Grain-size distributionsfor the peripheral and center regions evaluated from the TEM micrographs.

changes in the microstructure. Indeed, Figure 2b–d present

the grain boundary maps obtained by EBSD for the specimens

before and after DPD. The heat-treatment in processing of

the initial sample resulted in a well-recrystallized micro-

structure as revealed by the uniform equiaxed grain structure

with a high fraction of HAGBs (about 98%) in Figure 2b.

Among HAGBs, the fraction of grain boundaries with a

misorientation of 608 (S3 boundaries) was particularly high

(50%). Most of these boundaries appear to be twin boundaries.

The arithmetic and the area-weighted mean grain sizes are

about 9 and 25mm, respectively. Figure 2c and d show

cross-sectional (the plane containing X2 and X3 axes) view of

the microstructures in the periphery and the middle of the

impacted sample, respectively. Figure 2c illustrates that in

the periphery of the impacted disk the pristine grains were

divided into subgrains bounded by LAGBs with a misor-

ientation angle less than 158 which were most probably

developed from dislocations formed during DPD. Smaller

lamellar grains characterize the microstructure in the middle

of the impacted disc (Figure 2d). The lamellae are lying in

direction X2. The grains are delineated by HAGBs and their

interiors contain a large density of LAGBs. The grain size

measured by EBSD and averaged for the entire specimen is

about 1094� 531 nm, however, smaller grains are observed in

the center of the sample (Figure 2d).

Cross-sectional TEM observations for both the periphery

and the middle of the impacted sample are presented in

Figure 3a and b, respectively. In both locations, a significant

grain size reduction occurred due to DPD-processing. Both

cellular and lamellar grains can be observed whose interiors

are populated by dislocations with a high density. Lamellae

dominate in the middle of the sample, while both cell blocks

and lamellae appear in the periphery. These observations are

in line with EBSD investigations. According to the litera-

ture,[28–30] most probably the new grain and subgrain

boundaries formed during DPD. Figure 3c shows the

grain-size distributions in the middle and the periphery of

the DPD-processed sample as obtained from the TEM

micrographs. Both distributions can be well fitted by

lognormal functions, and only a slight difference in the

mean grain sizes was observed. Actually, the computed

average grain sizes (based on the measurement of more than

500 grains) were about 220 and 230 nm in the middle and the

periphery of the sample, respectively.

2.2. XLPA and XRD Characterizations

Further microstructure investigation was carried out at the

middle and half the radius of the impacted sample (in the

lamellae zone) by XLPA. The fitting of the diffraction pattern

in logarithmic intensity scale is shown in Figure 4. The open

circles and the solid line represent the measured data and the

fitted curves, respectively. The dislocation density was very

large (13� 2� 1014m�2) as determined by XLPA while

significant twinning was not observed. The mean crystallite

size was 110� 10 nm that is listed in Table 2 together with the

grain sizes determined by TEMand EBSD techniques. Figure 5

1030 http://www.aem-journal.com � 2012 WILEY-VCH Verlag GmbH & C

presents the complete pole figures recalculated from the

orientation distribution functions in the plane containing axes

X1 and X2 at the center (location 1, Figure 1) and the edge

(location 2, Figure 1) of the DPD-processed specimen. The

analysis revealed some preferential orientations including a

strong h220i fiber texture parallel to the impact direction (axis

o. KGaA, Weinheim ADVANCED ENGINEERING MATERIALS 2012, 14, No. 11

COM

MUNIC

ATIO

N

L. Farbaniec et al./High Purity Ultrafine-Grained Nickel

Fig. 4. The measured (open circle) and the fitted (solid line) X-ray diffraction patterns inlogarithmic intensity scale.

Table 2. The average crystallite size obtained by XLPA, TEM, and EBSD.

XLPA TEM EBSD

Crystallite size [nm] 110� 10 230� 9 1094� 531

X3). The volume fraction of the main fiber texture was

estimated as 46% in the center and 38% at the periphery. The

intensity of this texture component varies along the diameter

Fig. 5. Recalculated {111} and {220} pole figures in the (X1, X2) plane (a) at the center a

ADVANCED ENGINEERING MATERIALS 2012, 14, No. 11 � 2012 WILEY-VCH Verl

and the thickness of the disk, in accordance with the changes

of the microstructure.

2.3. Mechanical Properties

Figure 6 shows the true stress–true plastic strain compres-

sion curves. The compressive yield strength of the initial

material was about 90MPa and the corresponding stress–

strain curve exhibits a strain hardening behavior for the full

range of the investigated strain. After DPD, two types of

mechanical tests were performed: the compression axis was

oriented either perpendicular or parallel to the lamellar

boundaries (LBs) formed during the impact test. If the LBs are

oriented horizontally in the tested samples (Figure 6), the

loading axis was aligned vertically or horizontally in the first

or the second case, therefore these specimens are denoted as V

and H, respectively. In addition, specimens were cut from the

center and the edge with the notations of 1 and 2, respectively

(Figure 1). Therefore, for example the specimen cut from the

center and compressed perpendicularly to the LBs is denoted

as V1. Several observations can be made on the basis of the

mechanical tests: (i) the compressive yield strength increased

by a factor of 5–6 due to the impact. Depending on the location

and the orientation of the specimen the saturation flow stress

varies between 645 and 780MPa; (ii) in general, H specimen

exhibited higher maximum stress than V specimen from the

same location; (iii) for the same orientation of the compression

axis relative to LBs the specimen cut from the location 2

showed higher maximum stress than the sample obtained

from the location 1; (iv) for all DPD-processed samples after a

short hardening stage (up to a plastic strain of about 2%) a

stress plateau or even work softening was observed; (v) for

nd (b) the edge.

ag GmbH & Co. KGaA

sample H1 a resumed work hardening was

observed for strains larger than 5%, which

was not the case for the sample H2.

3. Discussion

Our results showed that intermediate

strain rate impact yielded a microstructure

composed of subgrains delineated by LAGBs

and lamellar grains bounded by HAGBs.

Indeed, EBSD and TEM investigations

revealed significant difference between dis-

location patterns in the middle and the

periphery of the DPD-processed sample,

leading to a heterogeneous microstructure

and texture distribution (Figure 2 and 3).

These results are in line with previous work

on high purity and commercially pure

Al.[25,26] It is observed here that the initial

grain boundary structure, even though dis-

torted, is still visible on the periphery of the

sample (Figure 2c). The amount of LAGBs

observed in that region is also lower than in

the middle of the sample. In general, a

, Weinheim http://www.aem-journal.com 1031

COM

MUNIC

ATIO

N

L. Farbaniec et al./High Purity Ultrafine-Grained Nickel

Fig. 6. True stress–true plastic strain obtained by compression. If the lamellar boundaries are orientedhorizontally in the tested samples, the specimens with vertical or horizontal compression axis are denotedas V or H, respectively. In addition, the specimens cut from the center or the periphery are denoted as 1 or 2,respectively.

non-uniform grain-refinement occurs due to DPD as sug-

gested by EBSD analysis (Figure 2c and d). However, TEM

investigations revealed that the difference in the grain sizes

and grain-size distributions along the diameter of the

impacted disk is small. This suggests that the misorientation

angle between neighboring grains/subgrains is very small at

the periphery of the sample and difficult to access by EBSD

technique. Moreover, the actual step size (0.1mm) used for

EBSD analysis also does not take into account the very

small grains detected by TEM. In addition, the crystallite

size measured by XLPA is the lowest compared to grain

sizes obtained by either EBSD or TEM (Table 2). The

discrepancies between TEM and XRD measurement have

been reported[25,38] in the case of metallic materials deformed

by SPD routes. As an example, the grain size measured

by TEM is about three times larger than the crystallite

size measured by XPLA. As was discussed by Saada and

Dirras,[39] this discrepancy can be related to the fact that X-rays

have better sampling statistics than TEM. In addition,

dislocation walls such as dipolar walls do not give contrast

difference in TEM while they break coherency of X-ray

scattering. Finally, the size of scattering domains is equivalent

to the grain size in materials without a subgrain structure

within individual grains. This is not the case in SPD-processed

materials. TEM investigations (Figure 3) showed high

dislocation density within grains contrary to the case of pure

Al deformed under the same conditions, where the subgrains

were mostly dislocation free.[25] This can be attributed to a

more planar dislocation slip mechanism due to the lower

SFE energy for Ni (�130mJ �m�2[36]) compared to pure Al

(166mJ �m�2[37]) and also to the higher melting point of the

former material. In Al the easy cross-slip and climb of

dislocations yield dynamic recovery and/or recrystallization

during DPD-processing.

The stress–strain curves indicate a significant increase in

the flow stress compared to the non-impacted sample. This

behavior can be attributed to the combined effect of the high

1032 http://www.aem-journal.com � 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

dislocation density and the reduced grain

size as have been discussed by Hughes

and Hansen[28,29] and by Krasilnikov.[40]

The present TEM observations revealed

that the average grain size is almost identical

in the whole disk. However, the average

grain size obtained by EBSD was different,

therefore, it is hypothesized that the average

misorientation angle across the grain or

subgrain boundaries changes along the

radius of the impacted disk, which may

contributes to the observed difference in the

mechanical behavior of the samples obtained

from the locations shown in Figure 1 (e.g.,

H1 and H2). In addition, the intensity of the

h220i fiber texture is stronger in region 1

than in region 2. In a face centered cubic

materials such as Ni, the number of activated

dislocation slip systems strongly depends

on the crystallographic orientation relative to the loading

direction.[41] For samples V the loading direction is parallel

to the h220i fiber texture component that yields an activation

of a relatively large number of slip systems resulting in

lower flow stress than for specimens H. This effect depends on

the location in the impacted disk due to the change of the

strength of texture. Similar effect of strain-path change

was recently reported in the case of commercial purity

Al.[26] Additionally, the change in loading direction in H-type

samples may lead to destabilization of the dislocation

structure formed during DPD-processing and also to the

variation of the crystallographic texture. These issues may

contribute to a macroscopic softening/hardening transient

effects.

4. Conclusions

Bulk coarse-grained nickel with purity higher than

98.4wt% was processed by a DPD technique. The impact

deformation caused significant changes in the microstructure

which were studied by EBSD, TEM, and XLPA. The following

observations were made:

(i) T

he microstructure after DPD consists of lamellar grains

with subgrain structure.

(ii) A significant reduction of grain size from about 9mm to

230 nm was observed by TEM.

(iii) X LPA investigation revealed a relatively large dislocation

density (13� 2� 1014m�2) and a crystallite size of

110� 10 nm which is lower than the grain sizes deter-

mined by EBSD or TEM.

(iv) T he compressive yield strength increased by a factor of

5–6 due to DPD. The maximum flow stress varied

between 645 and 780MPa and its value depended on

the location in the impacted disk and the loading direc-

tion. This behavior can be attributed to the heterogeneous

ADVANCED ENGINEERING MATERIALS 2012, 14, No. 11

COM

L. Farbaniec et al./High Purity Ultrafine-Grained Nickel

ADVA

microstructure and the varying strength of the h220i fibertexture along the diameter of the impacted disk.

MUNIC

AT

Received: January 30, 2012

Final Version: June 19, 2012

Published online: July 16, 2012

ION

[1] E. O. Hall, Proc. Phys. Soc. Lond. B 1951, 64, 747.

[2] N. J. Petch, J. Iron Steel Inst. 1953, 174, 25.

[3] H. Gleiter, Prog. Mater. Sci. 1989, 33, 223.

[4] H. Gleiter, Nanostruct. Mater. 1992, 1, 1.

[5] H. Gleiter, Nanostruct. Mater. 1995, 6, 3.

[6] R. Z. Valiev, R. K. Islamgaliev, I. V. J. Alexandrov, Prog.

Mater. Sci. 2000, 45, 103.

[7] A. Azushima, R. Kopp, A. Korhonen, D. Y. Yang,

F. Micari, G. D. Lahoti, P. Groche, J. Yanagimoto,

N. Tsuji, A. Rosochowski, A. Yanagida, CIRP Ann.

Manuf. Technol. 2008, 57, 716.

[8] Y. Ito, Z. Horita, Mater. Sci. Eng. A 2009, 503, 32.

[9] A. P. Zhilyaev, A. A. Gimazov, E. P. Soshnikova,

A. Revesz, T. G. Langdon, Mater. Sci. Eng. A 2008,

489, 207.

[10] T. Hebesberger, H. P. Stuwe, A. Vorhauer, F. Wetscher,

R. Pippan, Acta Mater. 2005, 53, 393.

[11] Y. Todaka, J. Sasaki, T. Moto, M. Umemoto, Scr. Mater.

2008, 59, 615.

[12] Y. Iwahashi, Z. Horita, M. Nemoto, T. G. Langdon, Acta

Mater. 1997, 45, 4733.

[13] P. L. Sun, P. W. Kao, C. P. Chang, Mater. Sci. Eng. A 2000,

283, 82.

[14] K. Neishi, Z. Horita, T. G. Langdon, Mater. Sci. Eng. A

2002, 325, 54.

[15] A. P. Zhilyaev, B.-K. Kim, J. A. Szpunar, M. D. Baro,

T. G. Langdon, Mater. Sci. Eng. A 2005, 391, 377.

[16] O. V. Mishin, V. Y. Gertsman, R. Z. Valiev, G. Gottstein,

Scr. Mater. 1996, 35, 873.

[17] S. V. Dobatkin, J. A. Szpunar, A. P. Zhilyaev, J. Y. Cho,

A. A. Kuznetsov, Mater. Sci. Eng. A 2007, 462, 132.

[18] D. H. Shin, W. J. Kim, W. Y. Choo, Scr. Mater. 1999, 41,

259.

[19] S. V. Dobatkin, O. V. Rybal’chenko, G. I. Raab, Mater. Sci.

Eng. A 2007, 463, 41.

NCED ENGINEERING MATERIALS 2012, 14, No. 11 � 2012 WILEY-VCH Verl

[20] W. S. Zhao, N. R. Tao, J. Y. Guo, Q. H. Lu, K. Lu, Scr.

Mater. 2005, 53, 745.

[21] Y. S. Li, N. R. Tao, K. Lu, Acta Mater. 2008, 56, 230.

[22] Y. Z. Tian, S. D. Wu, Z. F. Zhang, R. B. Figueiredo,

N. Gao, T. G. Langdon, Mater. Sci. Eng. A 2011, 528,

4331.

[23] F. Huang, N. R. Tao, K. Lu, J. Mater. Sci. Technol. 2011,

27, 1.

[24] A. Abdul-Latif, G. F. Dirras, S. Ramtani, A. Hocini, Int. J.

Mech. Sci. 2009, 51, 797.

[25] G. Dirras, T. Chauveau, A. Abdul-Latif, S. Ramtani,

Q.-H. Bui, Phys. Status Solidi A 2010, 207, 2233.

[26] G. Dirras, T. Chauveau, A. Abdul-Latif, J. Gubicza,

S. Ramtani, Q. Bui, Z. Hegedus, B. Bacroix, Metall. Mater.

Trans. A 2012, 43, 1312.

[27] B. Bay, N. Hansen, D. Kuhlmann-Wilsdorf, Mater. Sci.

Eng. A 1989, 113, 385.

[28] D. A. Hughes, N. Hansen, Acta Mater. 1997, 45, 3871.

[29] D. A. Hughes, N. Hansen, Acta Mater. 2000, 48, 2985.

[30] D. A. Hughes, Scr. Mater. 2002, 47, 697.

[31] G. Dirras, H. Couque, J. Gubicza, A. Ouarem,

T. Chauveau, P. Jenei, Mater. Sci. Eng. A 2010, 527,

4128.

[32] C. Zener, J. H. Hollomon, J. Appl. Phys. 1944, 15, 22.

[33] Y. S. Li, Y. Zhang, N. R. Tao, K. Lu, Acta Mater. 2009, 57,

761.

[34] G. Ribarik, J. Gubicza, T. Ungar, Mater. Sci. Eng. A 2004,

387–389, 343.

[35] L. Balogh, G. Ribarik, T. Ungar, J. Appl. Phys. 2006, 100,

023512.

[36] L. E. Murr, Scr. Metall. 1972, 6, 203.

[37] L. E. Murr, Interfacial Phenomena in Metals and Alloys,

Addison-Wesley, Reading, MA 1975, 378.

[38] A. P. Zhilyaev, J. Gubicza, G. Nurislamova, A. Revesz,

S. Surinach, M. D. Baro, T. Ungar, Phys. Status Solidi A

2003, 198, 263.

[39] G. Saada, G. Dirras, in Dislocations in Solids, Vol. 15 (Eds:

J. P. Hirth, L. Kubin), Elsevier B.V, North-Holland 2009,

Ch. 90.

[40] N. Krasilnikov, W. Lojkowski, Z. Pakiela, R. Valiev,

Mater. Sci. Eng. A 2005, 397, 330.

[41] B. Clausen, T. Lorentzen, T. Leffers, Acta Mater. 1998, 46,

3087.

ag GmbH & Co. KGaA, Weinheim http://www.aem-journal.com 1033