Fabrication of Bulk Nanocrystalline Fe-C Alloy by Spark Plasma Sintering of Mechanically Milled Powder

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Scripta Materialia 53 (2005) 863–868

www.actamat-journals.com

Fabrication of bulk nanocrystalline Fe–C alloy by spark


plasma sintering of mechanically milled powder
H.W. Zhang, R. Gopalan, T. Mukai, K. Hono *

National Institute for Materials Science, 1-2-1 Sengen, Tsukuba 305-0047, Japan

Received 17 May 2005; received in revised form 30 May 2005; accepted 31 May 2005
Available online 28 June 2005

Abstract

Fully dense bulk nanocrystalline Fe–0.8wt.%C alloy was synthesized by spark plasma sintering of mechanically milled Fe–C
nanocrystalline powder. The sample sintered at 600 C was composed of 150 nm ferrite grains with nanocrystalline cementite
dispersoids, whose compression yield strength, fracture strength, and plastic strain were 1900 MPa, 3500 MPa, and 40%, respectively.
 2005 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

Keywords: Mechanical milling; Nanocrystal; Spark plasma sintering; Severe plastic deformation; Fe–C

1. Introduction theoretical density. Numerous traditional consolidation


techniques, such as hot pressing [10], hot isostatic press-
Various unconventional properties such as ultrahigh ing [11] and hot-extrusion [12] can yield nearly fully
strength, high hardness, superplasticity and unique mag- dense compacts [13]; however, recrystallization and
netic properties were reported for nanocrystalline mate- grain growth during high temperature consolidation
rials whose grain size is less than 100 nm [1]. During the processes retard the synthesis of nanocrystalline micro-
past two decades, many techniques have been developed structure. Although shock wave compaction [14] can
to synthesize nanocrystalline materials, including inert densify powders without substantial change in micro-
gas condensation and consolidation [2], severe plastic structure and composition, this is not a suitable method
deformation [3], crystallization of amorphous solid [4], for industrial applications. An artifact-free bulk nano-
electrodeposition [5], and surface mechanical attrition crystalline Cu with a good combination of strength
treatment [6–9]. However, producing industrially useful (770 MPa) and ductility (30%) was developed very re-
bulk materials with the microstructure on a nanometer cently by mechanical milling and in situ consolidation
scale is still a challenge for materials engineering. One [15]; however, the uncontrollable sample shape and size
of the widely used promising methods to produce com- may restrict its engineering applications. A process that
mercial bulk nanomaterials with a grain size of less than can be applied to produce bulk nanocrystalline materials
100 nm is ball milling accompanied by a subsequent at moderate temperature with low applied load and fast
consolidation process. Due to the metastable nature of consolidation speed is necessary for potential engineer-
the nanocrystalline materials, the consolidation process ing applications.
must be able to eliminate extensive grain growth and Spark plasma sintering (SPS) can achieve a rapid
retain the nanoscale grain size while obtaining a near heating rate with a pulsed electrical discharge that is
required to densify powders rapidly without substantial
*
Corresponding author. grain growth. The SPS method has been used previously
E-mail address: kazuhiro.hono@nims.go.jp (K. Hono). to consolidate a large variety of ceramic and metal

1359-6462/$ - see front matter  2005 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.scriptamat.2005.05.039
864 H.W. Zhang et al. / Scripta Materialia 53 (2005) 863–868

powder materials to high densities [16]. The SPS method microhardness test was conducted by means of a Mitu-
was used very recently to synthesize nanocrystalline toyo Mvk-H2 model microhardness testing machine
Fe–Al with a near theoretical density with controlled with a load of 50 gf (0.49 N) at a time of 10 s.
grain growth [17]. The unique boundary cleaning effect
of the SPS process makes the powder handling more
convenient [18]. 3. Results and discussion
In the present investigation, we selected Fe–C powder
as the precursor to fabricate dense bulk nanocrystalline Fig. 1(a) shows a TEM bright field image and its cor-
materials by means of the SPS method, because a recent responding selected area electron diffraction (SAED)
study showed that the Fe–C nanocrystalline microstruc- pattern obtained from the Fe–C powder that were
ture that was produced by mechanical milling exhibits mechanically milled for 100 h. The microstructure of
strong resistance to grain growth by heating [19]. In the as-milled powder is composed of randomly oriented
addition, the Fe–C system has been used in a wide range nanograins of body-centered cubic Fe. The histogram of
of industrial applications and is not susceptible to media the grain size distribution, Fig. 1(b), is characterized by
pollution during mechanical milling with stainless steel a normal logarithmic distribution with a narrow size
balls. distribution of 1–5 nm. The mean grain size is approxi-
mately 2 nm. This microstructural feature is very
similar to that of the mechanically milled pearlite
2. Experimental reported by Ohsaki et al. [19]. Their high resolution
TEM observation revealed that little dislocation stays
Fe–0.8wt%C alloy powder was prepared by ball mill- within the grains of the nanocrystalline ferrite, and their
ing iron (99.99% purity) and graphite powder (99.999%
purity) in a Fritsch P-6 planetary ball mill. Stainless steel
media (10 mm in diameter) were used with a ball to a
charge ratio of 10:1. Powder handling was conducted
in an argon glove box to avoid contamination. Milling
was conducted at 250 rpm for 100 h, with an intermedi-
ate halt for 2 h after every 20 h milling to reduce the
temperature rise. The as-milled Fe–C powder was
packed in a graphite die with an inner diameter of
10 mm and was densified by a SPS apparatus, Sumi-
tomo Coal Mining Company Model 1050, in a vacuum
of <10 3 Pa at a load of 5.5 kN (70 MPa) for 10 min.
Temperatures of 400, 500 and 600 C were selected to
compare the consolidation behavior, mechanical prop-
erties, and constituent phases. A thermocouple placed
in the middle part of the graphite die was used to control
the temperature during the SPS treatment. The end-
products of approximately B10 mm · 5 mm were pol-
ished with SiC paper to remove surface contamination
from the graphite die and foil before conducting struc-
tural and property characterization. The density was
determined by the Archimedes method with water as a
liquid medium. The constituent phases in the as-milled
powders and SPS samples at various temperatures were
examined by X-ray diffraction using CuKa radiation.
The grains size and the microstrain in the Fe–C alloy
under different treatments were estimated by the Scherrer
function based on the experimental full-width at a
half-maximum on each diffraction peak [20]. Micro-
structural examinations were conducted by transmission
electron microscopy (TEM) using Philips CM200. Com-
pression tests were performed at a quasi-static strain
rate of 1 · 10 4 in an Instron machine using rod samples
of B2.5 mm · 5.0 mm. Because of this sample size Fig. 1. TEM observations of mechanical milled Fe–C alloy and the
limitation, only compression tests were performed. The corresponding grain size distributions.
H.W. Zhang et al. / Scripta Materialia 53 (2005) 863–868 865

three-dimensional atom probe (3DAP) analysis showed


that carbon atoms were strongly segregated at the grain
boundaries, and only approximately 1.0 at.%C was dis-
solved within the ferrite grains. Since the nanocrystalline
ferrite in Fig. 1 has very similar features to that in the
previous work, a similar carbon distribution is expected
in the present sample.
The as-milled Fe–C powder was consolidated subse-
quently into bulk samples by SPS at 400, 500 and
600 C, respectively, for 10 min. The average grain size
in the various stages of the samples estimated by XRD
is summarized in Table 1 together with the values of
microhardness and relative density. A relative density
of 98.5% and an average microhardness of 6.2 GPa were
achieved in the sample sintered at 600 C (see Table 1). Fig. 3. XRD profiles of the Fe–C alloys before and after SPS
treatment at 400, 500 and 600 C.
The relative density and the average microhardness de-
creased with decreasing temperature. Sintering at
400 C and 500 C led to relative densities of 78% and observed in the SPS samples together with the diffrac-
89%, respectively, and the average microhardness of tion peaks that are attributed to the presence of cement-
2.5 GPa and 5.97 GPa, respectively. Compression true ite, indicating that grain growth, release of the
stress–strain plots of the samples sintered at 500 C and microstrain, and precipitation of cementite occurred
600 C are shown in Fig. 2. The sample sintered at during the SPS process. Table 1 summarizes the grain
500 C shows a yield strength of around 2000 MPa and sizes and the microstrains of the samples estimated from
a failure strength of 2200 MPa after work hardening with the XRD results.
a plastic strain of 0.05. The sample sintered at 600 C Fig. 4 shows the TEM observations and the corre-
shows a yield strength of 1900 MPa and a failure strength sponding grain size distributions of the samples sintered
of 3500 MPa with a large plastic strain of about 0.40. at (a) 400 C, (b) 500 C, and (c) 600 C. The micro-
XRD profiles of the samples are shown in Fig. 3. structure of the sample sintered at 400 C (Fig. 4(a)) is
Compared to the broad diffraction peaks from the characterized by an uniformly distributed nanometer-
as-milled Fe–C powder, sharper diffraction peaks are scale ferrite and cementite, whose random orientations
are confirmed by the continuous diffraction rings in
the SAED pattern. The dark field image excited by the
Table 1 cementite spot indicated by the arrowhead shows that
Grain size and lattice strain estimated by X-ray diffraction and Vickers
hardness and relative density of the samples
nanoscale cementite particles precipitate along the grain
boundaries of the nanoferrite. The grain size distribu-
Grain Lattice Hardness Relative
size [nm] strain [%] [GPa] density [%]
tion indicates a narrow range of 5–100 nm with a mean
value of 30 nm, which is very close to the XRD measure-
As-milled 5 ± 0.2 0.766 ± 0.08 – –
400 C SPS 34 ± 5.3 0.278 ± 0.06 2.5 ± 1.2 78
ment result in Table 1 (34 nm). Compared to the sample
500 C SPS 57 ± 8.2 0.237 ± 0.058 5.97 ± 1.0 89 sintered at 400 C, larger grains are visible in the sample
600 C SPS 151 ± 15 0.124 ± 0.008 6.2 ± 0.3 99 sintered at 500 C (Fig. 4(b)), in which the constituent
phases were confirmed to be the same as those in the
sample sintered at 400 C. The corresponding SAED
pattern illustrates that these grains also show random
orientations. The dark field image excited by the cement-
ite spot (indicated by the arrowhead) indicates that the
amount of cementite precipitates increased in this sam-
ple compared to the one sintered at 400 C. The histo-
gram shows a rather broad grain size distribution of
10–250 nm with an average value of 50 nm. Large
ferrite and cementite grains are observed in the sample
sintered at 600 C (Fig. 4(c)). The grain size distribution
has a wide range of 50–450 nm, with an average size of
about 150 nm.
This work has demonstrated that fully dense bulk
Fig. 2. Compressive true stress–strain curves of 500 C and 600 C Fe–C alloys with a grain size of 150 nm containing a
SPS treated samples. high density of cementite dispersoids can be produced
866 H.W. Zhang et al. / Scripta Materialia 53 (2005) 863–868

Fig. 4. TEM observations and the corresponding grain size distributions of (a) 400 C, (b) 500 C and (c) 600 C SPS treated samples.

from mechanically milled nanocrystalline ferrite powder densification to a near theoretical density can be
by SPS at a modest load of 70 MPa in a very short time. achieved in the SPS process for a sort sintering time with
The sample sintered at 600 C exhibited a good combi- only 10% of the load that is required for hot pressing.
nation of strength and ductility (yield strength of around These advantageous intrinsic properties of the SP
2000 MPa, plastic strain of 40%) and uniform strain sintered materials result in good ductility with a plastic
hardening leading to a failure strength over 3500 MPa. strain of around 40%. However, localized shear defor-
Nanocrystalline metals and alloys exhibit an en- mation inside shear bands usually causes limited ductil-
hanced yield strength following the Hall–Petch effect. ity as well as the diminishing of strain hardening in
Table 2 summarizes the mechanical properties reported nanocrystalline metallic materials [31]. Most of the
in ultrafine grained iron and nanocrystalline irons pro- nanocrystalline (<100 nm) and ultrafine-grained (<300
duced by various methods. The yield strength of the nm) Fe and Fe alloys were reported to show little or
severely deformed ultrafine grained Fe is less than no strain hardening, or even strain softening, in which
800 MPa with plastic strains of 20–30% [21,22]. How- the shear banding was attributed to the dominant defor-
ever, powder-consolidated Fe [23–25] and Fe–C nano- mation mechanism [21–30]. The present uniform strain
crystalline alloys [26–28] show much higher strengths hardening effect as well as the plastic strain over 40%
ranging from 1600 MPa to 2500 MPa, but their plastic in the compressive stress–strain curve of the sample sin-
strains are less than 10%. One of the main reasons for tered at 600 C may imply that the deformation mecha-
the low ductility is the poor sample quality, for example, nism would not be dominated by the inhomogeneous
resident porosities, impurities and poor metallurgical deformation inside shear bands. The absence of local-
bonding between particles. In the SPS process, good ized shear deformation in the present nanocrystalline
metallurgical bonding between particles can be achieved Fe–C alloy might be partially attributed to the cementite
by the pulsed field activation process, which is known to precipitates inside the grain and along the grain bound-
have an efficient oxide removal effect [18]. In addition, aries. These would act as the obstacles of dislocation
H.W. Zhang et al. / Scripta Materialia 53 (2005) 863–868 867

Table 2
Comparisons of the mechanical properties of bulk ultrafine grained and nanocrystalline Fe and Fe–C alloys prepared by different processes
Materials Processing Relative Grain Yield Fracture Plastic Ref.
density size (nm) strength strength strain (%)
(%) (MPa) (MPa)
Fe ECAP 100 150–300 700 700 20 [21]
Fe ECAP 100 200–400 800 800 30 [22]
Fe CC (1.4 GPa, 10 h) + HP (850 MPa, 3 h, 590 C) >99 268 1600 1480 12 [23]
Fe CC (1.4 GPa, 10 h) + HP (850 MPa, 3 h, 500 C) >99 138 2300 2000 4 [23]
Fe CC (1 GPa, 4 h) + HP (750 MPa, 30 min, 410 C) >99 80 2500 2000 6 [24]
Fe–2C HP (200 MPa, 3 h, 580 C) 80 – 1690 1690 1 [26]
Fe–1.87C HP (50 MPa, 25 min, 800 C) 92 75 1215 1544 1 [27]
Fe–10Cu HP (170 GPa, 30 min, 600 C) >99 250 1700 1600 3 [28]
*
Fe–0.8C SPS (70 MPa, 10 min, 500 C) 89 50 2000 2200 8
*
Fe–0.8C SPS (70 MPa, 10 min, 600 C) 99 150 1900 3500 40
ECAP: equal channel angular pressing, CC: cold compaction, HP: hot pressing, SPS: spark plasma sintering.
*
Present investigation.

gliding and grain rotations, which would inactivate the try of Science and Technology, Korea through the
shear band formation and transmission since lattice Korea Institute of Science and Technology and by the
reorientation and grain rotation usually facilitate the grant-in-aid of scientific research, Priority Area ‘‘Science
formation of shear bands and their transmission [32]. of Metallic Glass’’ of MEXT, Japan.
The appearance of the large strain hardening effect in
this sample indicates that dislocation motions are active
in this two phase nanocrystalline material, which is unu- References
sual for the nanocrystalline Fe with a grain size smaller
than 100 nm. Improved work hardening effect due to the [1] Lu K. Mater Sci Eng R 1996;16:161.
presence of cementite particle was also recently reported [2] Gleiter H. Prog Mater Sci 1989;33:223.
in the ultrafine grained steel with finely dispersed [3] Valiev RZ, Korznikor AV, Mulyukov RR. Mater Sci Eng A
1993;168:141.
cementite particles [33]. The enhanced ductility due to [4] Yoshizawa Y, Oguma S, Yamauchi K. J Appl Phys 1988;64:6044.
the second phase particles interacting with shear bands [5] Erb U, El-Sherik AM, Palumbo G, Aust KT. Nanostruct Mater
was also reported in bulk metallic glasses [34] as well 1993;2:383.
as in nanocrystalline metals [35]. Detailed TEM work [6] Lu K, Lu J. J Mater Sci Technol 1999;15:193.
is in progress to understand the unique deformation [7] Tao NR, Sui ML, Lu J, Lu K. Nanostruct Mater 1999;11:433.
[8] Zhang HW, Hei ZK, Liu G, Lu J, Lu K. Acta Mater
mechanism that was observed in this two phase bulk 2003;51:1871.
nanocrystalline alloy. [9] Wu X, Tao N, Hong Y, Xu B, Lu J, Lu K. Acta Mater
2002;50:2075.
[10] Dogan C, Rawers J, Govier D, Korth G. Nanostruct Mater
4. Summary 1994;4:631.
[11] Lillo TM, Korth G. Nanostruct Mater 1998;10:95.
[12] Livine Z, Munitz A, Rawers J, Fields R. Nanostruct Mater
Bulk nanocrystalline Fe–C alloys with a grain size of 1998;10:503.
150 nm containing a few nanometer cementite disper- [13] Rawers J. Nanostruct Mater 1999;11:512.
soids were fabricated from the mechanically milled Fe [14] Gourdin WH. J Appl Phys 1984;55:172.
and graphite powder by SPS. Yield strength of [15] Youssef KM, Scattergood RT, Murty KL, Koch CC. Appl Phys
Lett 2004;85:929.
2000 MPa, fracture strength of 3500 MPa, and a plastic [16] Groza JR, Zavaliangos A. Mater Sci Eng A 2000;287:171.
strain over 40% were obtained. The short processing [17] Paris S, Gaffet E, Bernard F, Munir ZA. Scripta Mater
period of several minutes and a low applied load of tens 2004;50:691.
of MPa that are required in the SPS process may enable [18] Risbud SH, Groza JR, Kim MJ. Philos Mag 1994;69B:525.
possible extension of the technique to industrial [19] Ohsaki S, Hono K, Hidaka H, Takaki S. Scripta Mater
2005;52:271.
production. [20] Klug HP, Alexander LE. X-ray diffraction procedures for
polycrystalline and amorphous materials. New York, NY: Wiley;
1974. p. 661.
Acknowledgements [21] Wei Q, Kecskes L, Jiao T, Hartwig KT, Ramesh KT, Ma E. Acta
Mater 2004;52:1859.
[22] Han BQ, Lavernia EJ, Mohamed FA. Metal Mater Trans
This work was supported in part by the Center for 2003;34A:71.
Nanostructured Materials Technology (CNMT) under [23] Jia D, Ramesh KT, Ma E. Acta Mater 2003;51:3495.
the 21st Century Frontier R&D Programs of the Minis- [24] Wei Q, Jia D, Ramesh KT, Ma E. Appl Phys Lett 2002;81:1240.
868 H.W. Zhang et al. / Scripta Materialia 53 (2005) 863–868

[25] Jia D, Ramesh KT, Ma E. Scripta Mater 2000;42:73. [31] Ma E. Scripta Mater 2003;49:663.
[26] Munitz A, Fields RJ. Powder Metall 2001;44:139. [32] Harren SV, Deve HE, Asaro RJ. Acta Metall 1988;36:2435.
[27] Rawers J, Krabbe R, Duttlinger N. Mater Sci Eng 1997;230A:139. [33] Song R, Ponge D, Raabe D. Scripta Mater 2005;52:1075.
[28] Carsley JE, Milligan WW, Hackney SA, Aifantis EC. Metall [34] Eckert J, Koch CC. Nanostructured materials; processing, prop-
Mater Trans 1995;26A:2479. erties, and applications. William Andrews: Norwich, NY; 2002.
[29] Jain M, Christman T. Acta Metall Mater 1992;42:1901. p. 423.
[30] Carsley JE, Fisher A, Milligan WW, Aifantis EC. Metall Mater [35] Zimmerman AF, Palumbo G, Aust KT, Erb U. Mater Sci Eng
Trans 1998;29A:2261. 2002;328A:137.

You might also like