Você está na página 1de 4

Available online at www.sciencedirect.

com

Scripta Materialia 64 (2011) 868871


www.elsevier.com/locate/scriptamat

Micromechanical characterization of casting-induced


inhomogeneity in an Al0.8CoCrCuFeNi high-entropy alloy
Zhiyuan Liu,a Sheng Guo,a Xiongjun Liu,a Jianchao Ye,a Yong Yang,a Xun-Li Wang,b
Ling Yang,b Ke Anb and C.T. Liua,
a

Department of Mechanical Engineering, The Hong Kong Polytechnic University, Hung Hom, Kowloon, Hong Kong, PR China
b
Neutron Scattering Science Division, Oak Ridge National Laboratory, Oak Ridge, TN 37831, USA
Received 14 December 2010; revised 12 January 2011; accepted 12 January 2011
Available online 15 January 2011

The microstructural features and micromechanical behavior of individual phases in a cast Al0.8CoCrCuFeNi high-entropy alloy
(HEA) were characterized by high-resolution scanning electron microscopy and micro-compression tests. Use of neutron diraction
enabled the detection of a new phase which was otherwise unobservable by conventional X-ray diraction. The identied phase constitution agreed well with the compositional analysis and the micro-compression results. The delicate microscale characterization of
individual phase provides new insights for the design of novel HEAs with desirable mechanical properties.
2011 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
Keywords: High-entropy alloy; Solidication microstructure; Neutron diraction; Compression test

High-entropy alloys (HEAs) provide a novel alloy concept that signicantly expands the scope of traditional alloy design [13]. HEAs typically consist of at
least ve principal metallic elements in near-equimolar
ratios, and they often form a single solid solution structure, instead of many intermetallic compounds as expected from general physical metallurgy principles.
HEAs show great potential for engineering applications
due to their high hardness, wear resistance, high-temperature softening resistance and oxidation resistance [2,4].
The commonly used alloying elements include face-centered cubic (fcc)-type Cu, Al, Ni, body-centered cubic
(bcc)-type Fe, Cr, Mo, V and hexagonal close packed
(hcp)-type Ti, Co [510]. Over the years considerable efforts have been devoted to the development of new HEA
systems with improved mechanical and functional properties. However, few studies have been carried out to
investigate the mechanical inhomogeneity of these
HEAs even though inhomogeneity is a common issue
in the cast structure for crystalline materials, and the
existing experimental observations have generally revealed the dendritic structure and compositional segregation in various HEA systems [6,11].

Corresponding

author. Tel.: +852 2766 6644; fax: +852 2365


4703; e-mail: mmct8tc@inet.polyu.edu.hk

The AlxCoCrCuFeNi (in atomic proportion) system is


a widely studied HEA system, and the phase constitution
in the as-cast material can be accurately adjusted by controlling the Al addition [6]. When x 6 0.5, only one single
fcc solidsolution phase is observed [6]; bcc solidsolution
phase starts to appear at x = 0.8, and at x > 2.8 a single
bcc phase is obtained although some minor phases (not
detectable via conventional X-ray diraction (XRD)
techniques) do exist [6]. The mechanism behind the addition of fcc-type Al promoting the transformation between
fcc and bcc phases is still unclear, although presumably
this can be rationalized as the alloying of the larger Al
atoms lowers the atomic packing eciency [12]. Previous
reports suggest that the lattice parameters for both fcc
and bcc phases increase with increasing x, but signicant
increase of hardness is seen only at x > 1.0 [6,13]. The formation of bcc phase will enhance the strength of the originally purely fcc solid solution, but it also causes the
embrittlement issue at ambient temperature [14]. Naturally, it is of interest to study the mechanical behavior
of Al0.8CoCrCuFeNi (the composition at which the bcc
phase just starts to appear) from the structureproperty
correlation perspective. In addition, a secondary fcc
phase has been detected in as-cast Al1.0CoCrCuFeNi
[11,15] by conventional XRD, but not yet in Al0.8CoCrCuFeNi. In this work we used neutron diraction to successfully detect the secondary fcc phase. The existence of

1359-6462/$ - see front matter 2011 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.scriptamat.2011.01.020

869

Z. Liu et al. / Scripta Materialia 64 (2011) 868871


(a)
Intensity (cps)

fcc
bcc

30

(b) 30

40

50

60
70
2 (degree)

80

90

100

Experimental
Refinement
Difference
FCC1
FCC2
BCC

25
Normalized Intensity

three phases (two fcc phases and one bcc phase) can be
well correlated with the three distinct regions observed
in the microstructure. We also performed careful microcompression tests to characterize these three regions to
obtain their individual mechanical properties. To the best
of our knowledge, this is the rst time a micro-compression test has been used to identify the mechanical properties of the constitutive phase components in HEAs.
The target alloy used in this work has the nominal composition of Al0.8CoCrCuFeNi (in atomic proportion). The
alloys were prepared by arc-melting a mixture of the constituent elements with purity better than 99.9% in a Ti-gettered high-purity argon atmosphere. Repeated melting was
carried out at least ve times to improve the chemical
homogeneity of the alloy. The molten alloy was drop-cast
into a 10 mm diameter copper mold. The phase constitution of the alloy was examined by both XRD using Co radiation (Bruker AXS D8 Discover) and neutron diraction
using the VULCAN diractometer (Spallation Neutron
Source, Oak Ridge National Laboratory, TN, USA [16]).
Because neutrons are highly penetrating, the neutron diffraction measurements are representative of the bulk,
rather than from the surface. As a state-of-the-art engineering diractometer, VULCAN can be exibly congured in
either high-intensity or high-resolution mode [17,18]. The
present measurements were made in the high-resolution
mode, in which a Dd/d  0.2% resolution is maintained
over a wide range of d-spacing. This allowed easy identication of the multiple phases in the sample. The microstructure of the alloy was characterized by scanning electron
microscopy (SEM) and energy dispersive spectrometry
(EDS) using a Leo 1530 FEG microscope operated at
5 kV. For the microstructure observation, the sample surface was sequentially polished down to 0.1 lm grit alumina
suspension nish and then etched with aqua regia solution.
To investigate the micromechanical behavior, micropillars
with a top diameter of 0.8 lm and aspect ratios ranging
from 2:1 to 5:1 were fabricated using a Quanta 200 3D
dual-beam scanning electron microscope/focused ion beam
(FIB) system (FEI Company, Hillsboro, OR, USA).
Following the well-established sequential-milling approach [19,20], the micropillars were prepared with great
care particularly in the interdendritic regions, to ensure
that the as-cut micropillars came exclusively from the
target regions (see below). The micro-compression tests
of the micropillars were carried out using a low-load
Triboindentere Nanoindentation system (Hysitron
Inc., Minneapolis, MN, USA) equipped with a 10 lm
at-end diamond punch at a load-controlled mode. To
ensure the reproducibility of the micro-compression
data, at least three micropillars were fabricated and
compressed from each characteristic region. Vickers
hardness was also measured by applying a load of 1 kg
for 10 s using a Future-Tech microhardness tester.
Figure 1a shows the XRD pattern of the as-cast
Al0.8CoCrCuFeNi. It is seen that the alloy has a simple
solid solution structure in which one fcc and one bcc
phase were identied. Judging from the relative intensity, the main phase of the alloy is fcc. This is in agreement with the previous results reported by Tong et al.
[6]. It should be mentioned here that due to the dendritic
structure (see below), the XRD intensities present texture feature and hence some characteristic peaks are

20
15
10
5
0
0.5

1.0

1.5
D-spacing (Angstrom)

2.0

2.5

Figure 1. (a) XRD pattern of the as-cast Al0.8CoCrCuFeNi alloy; (b)


neutron diraction pattern of the as-cast Al0.8CoCrCuFeNi alloy.

abnormally weak, although the intensity decrease could


also be caused by the highly distorted atomic planes in
the solid solutions [21]. To ascertain that the bcc phase
also exists in the bulk, the alloy was examined using neutron diraction and the pattern is shown in Figure 1b.
Interestingly, apart from one fcc and one bcc phase
identied using conventional XRD, a secondary fcc
phase was also detected. To dierentiate these two fcc
phases, they are termed fcc1 and fcc2 hereinafter. The
fcc2 phase is the minor phase judging from its volume
fraction. Based on the neutron diraction results, the
lattice parameters for the fcc1, fcc2 and bcc phases are
, respectively.
estimated to be 3.603, 3.631 and 2.874 A
It is worth pointing out that the fcc1 and fcc2 phases
were also identied in the as-cast Al0.5CoCrCuFeNi alloy using neutron diraction (results not given here),
and the lattice parameters for the fcc1 and fcc2 (3.596
, respectively) are almost the same as those
and 3.628 A
in Al0.8CoCrCuFeNi. In addition, the lattice parameters
for the fcc1, fcc2 and bcc phases in Al0.8CoCrCuFeNi
measured in this work are very close to those of the corresponding phases in Al1.0CoCrCuFeNi (3.59, 3.62 and
, respectively) reported by Singh et al. using XRD
2.87 A
[11]. It is noted here that our measurement of the lattice
parameters together with those from Singh et al. are different to those reported by Tong et al. [6] for the same
alloy compositions, in which our measured lattice
parameters are less sensitive to the Al composition. This

Figure 2. SEM images of the as-cast Al0.8CoCrCuFeNi alloy: (a) low magnication image; (b) magnied image of the indicated zone in (a)
showing the three distinct regions A, B and C. (c and d) Magnied
images for regions B and C, respectively.

870

Z. Liu et al. / Scripta Materialia 64 (2011) 868871

Table 1. Chemical composition and yield strength for the three distinctive regions in the as-cast Al0.8CoCrCuFeNi alloy.
Element (at.%)
Nominal
Region A
Region B
Region C

Structure

Al

Co

Cr

Cu

Fe

Ni

Yield strength (MPa)

fcc1
fcc1 + bcc
fcc1 + fcc2

13.79
10.39
22.82
14.20

17.24
21.26
13.97
6.62

17.24
22.21
13.54
5.16

17.24
8.56
15.76
53.83

17.24
22.12
12.52
6.11

17.24
15.47
21.39
14.08

764 47
958 80
825 60

is most likely due to the fact that Tong et al. only identied one fcc phase in their XRD analyses, and the hidden fcc peaks would cause the inaccurate peak tting
from which the lattice parameters were calculated. The
lattice parameters reect the lattice distortion information and our results can better account for the hardness
variation as a function of Al addition: from x = 0.5 to
0.8, the alloying of Al causes the partial transformation
from fcc to bcc phase while the fcc lattice remains nearly
invariant; at x = 0.8, the main phase is still fcc and the
amount of bcc phase is insucient to cause signicant
strengthening (the Vickers hardness obtained using
1 kg load for Al0.5CoCrCuFeNi and Al0.8CoCrCuFeNi
are 258 and 280, respectively); however, when x reaches
1.0, the main phase becomes bcc [11], and the large
amount of harder bcc phases causes the hardness to increase to 531.
Figure 2 shows the microstructures of the as-cast alloy.
Dendritic and interdendritic structures typical of cast
HEAs [6] are clearly observed. According to Tong et al.
[6], there are only two distinctive regions in the microstructure and the dendritic regions have a fcc structure,
while the interdendritic regions have a mixed fcc + bcc
structure. However, our high-resolution SEM observation, Figure 2, apparently shows the existence of three
distinctive regions: a dendritic region (A) and two interdendritic regions (B and C). The dendritic region A has
a relatively homogeneous contrast, which is consistent
with that reported by Tong et al. [6], and corresponds to
the single fcc structure. In contrast, regions B and C appear to contain more than one phase. Based on the microstructural features observed in Figure 2, the forming
process of these three distinctive regions can be envisaged
as follows: when the liquid cools down, the primary phase
forms and this is region A. The remaining liquid then separates into two compositionally dierent liquids, Cu-rich
liquid and Cu-depleted liquid, and these two liquid phases
then solidify separately. The Cu-depleted liquid decomposes into the mixed bcc and fcc phases (region B) probably following a eutectic reaction as suggested by Tong
et al. [6], while the Cu-rich liquid solidies and then further decomposes into two fcc phases via the spinodal
decomposition as indicated by the modulated structure
seen in Figure 2d. The Cu-depleted liquid solidies at

higher temperature compared to the Cu-rich liquid, and


this can account for the coarser microstructure in region
B than in region C. Although apparently more detailed
work needs to be carried out to conrm the hypothesis
of the above solidication process, it is a reasonable one
which can explain the as-obtained microstructure well.
Since it is generally perceived that the bcc phase is a harder phase compared with the fcc phase, region B would
have a higher hardness (or strength) than that of region
A. This is in fact supported by the micro-compression
tests discussed below. From the neutron diraction and
composition analyses we already know that the fcc1 is
the main phase, and we can further infer that region A
has the fcc1 structure, region B has the fcc1 + bcc structure and region C has the fcc1 + fcc2 structure.
The average chemical compositions for the three
regions from multiple-point EDS analysis are listed in
Table 1. It can be seen that these three regions have
clearly dierent chemical compositions: region A is enriched in Co, Cr and Fe but decient in Al and Ni and
Cu; region B is enriched in Al and Ni but decient in
the other elements; region C is signicantly enriched in
Cu but decient in Co, Cr and Fe. Ni tends to accompany Al in the interdendritic regions due to the large negative mixing enthalpy (DHmix) between these elements
[22]. This possibly results in the formation of NiAl-type
(B2) phase in region B [6,11]. Cu has a tendency to segregate from other alloying elements because it has a positive DHmix with other elements (apart from Al) [22].
In addition to identifying the phase composition for
the three distinctive regions, we have gone further and
characterized their individual mechanical properties
using the well-established micro-compression technique
[23,24]. Micropillars were machined from the three regions using the FIB technique and then compressed
using a at-end diamond punch under the load-controlled condition. Typical micropillars before and after
the micro-compression tests are shown in Figure 3, and
representative nominal stressstrain curves are plotted
in Figure 4. Here the stress is taken as the applied load
divided by the initial top area of the pillar, and the strain
is taken as the indenter displacement divided by the
initial height of the pillar. As shown in Figure 4, all specimens exhibit a linear response upon loading before the
1400
Region A
Region B
Region C

1200

Stress (MPa)

1000
800
600
400
200

0.05
0

Figure 3. SEM images of the micropillars before (a, c and e) and after
(b, d and f) the micro-compression tests. (a and b), (c and d) and (e and
f) correspond to regions A, B and C, respectively.

Strain

Figure 4. Stressstrain curves for the micropillars fabricated in the


three dierent regions as shown in Figure 2.

Z. Liu et al. / Scripta Materialia 64 (2011) 868871

yielding point, i.e. the rst observable strain burst appears [25]. The extracted yield strengths for the three regions are listed in Table 1. The micropillars cut from the
interdendritic region B have the highest yield strength of
958 MPa, which is about 20% higher than that of the
dendritic region A (764 MPa). The much higher
strength of region B most probably derives from the precipitation of the bcc phase in this region. Region C has a
slightly higher yield strength (825 MPa) than that of region A, in agreement with our previous assumption that
this region has a mixed two-fcc-phase structure resulting
from the spinodal decomposition. It is noted here that
the single-phase region A has a high yield strength of
764 MPa, which is much higher than the strength (i.e.
ultimate tensile strength) of any constituent metal element (Al, 47; Co, 255; Cr, 483; Cu, 220; Fe, 289; Ni,
407 MPa [26]). In addition, the yield strength presents
an apparent size eect in that the yield strengths measured from the micropillars are much higher than that
from the macrocompression test (r0.2  470 MPa), and
the larger the size of the micropillars, the closer the yield
strength measured from the micropillars and the macrosamples. The mechanism behind the size eect in HEAs
should be dierent to the dislocation starvation responsible for the size eect in single crystals [19,27]. More detailed results and discussions on the size eect in HEAs
will be reported elsewhere.
After yielding, the pillars deform plastically via the
emission of sporadic strain burst under the load-controlled condition, separated by elastic-like deformation
segments. Similar phenomena have also been observed
from the micro-compression of a variety of single-crystal
metals [28,29]. From the nominal stressstrain curves, it
is noticed that smaller strain bursts tend to occur for regions B and C compared to those in region A. This could
be attributed to the blocking and storage of dislocations
by the phase boundaries in the two-phase regions, B and
C, and the interaction between the high density of dislocations and the phase boundaries can account for the
much less jumpy deformation behavior [30].
In summary, microstructural features and micromechanical behavior for each individual phase in a cast
Al0.8CoCrCuFeNi alloy have been determined using
high-resolution SEM, together with delicate FIB-assisted
micro-compression tests. Both compositional and micromechanical analyses revealed the existence of three distinct regions, in agreement with the phases identied by
neutron diraction. The bcc-phase-containing region
clearly has a higher yield strength than the other two regions. The understanding of the mechanical behavior of
individual phase provides important input for the design
of new HEAs with desirable mechanical properties.
This research was supported by the internal
funding from HKPU. The Spallation Neutron Source
is operated with support from the Division of Scientic
User Facilities, Oce of Basic Energy Sciences, US
Department of Energy, under contract DE-AC0500OR22725 with UT-Battelle, LLC.
[1] J.W. Yeh, Ann. Chim.-Sci. Mat. 31 (2006) 633.
[2] W.H. Wu, C.C. Yang, J.W. Yeh, Ann. Chim.-Sci. Mat. 31
(2006) 737.

871

[3] J.W. Yeh, S.K. Chen, S.J. Lin, J.Y. Gan, T.S. Chin, T.T.
Shun, C.H. Tsau, S.Y. Chang, Adv. Eng. Mater. 6 (2004)
299.
[4] X.F. Wang, Y. Zhang, Y. Qiao, G.L. Chen, Intermetallics
15 (2007) 357.
[5] H.Y. Chen, C.W. Tsai, C.C. Tung, J.W. Yeh, T.T.
Shun, C.C. Yang, S.K. Chen, Ann. Chim.-Sci. Mat. 31
(2006) 685.
[6] C.J. Tong, Y.L. Chen, S.K. Chen, J.W. Yeh, T.T. Shun,
C.H. Tsau, S.J. Lin, S.Y. Chang, Metall. Mater. Trans.
36A (2005) 881.
[7] J.W. Yeh, S.K. Chen, J.Y. Gan, S.J. Lin, T.S. Chin, T.T.
Shun, C.H. Tsau, S.Y. Chang, Metall. Mater. Trans. 35A
(2004) 2533.
[8] C.Y. Hsu, W.R. Wang, W.Y. Tang, S.K. Chen, J.W.
Yeh, Adv. Eng. Mater. 12 (2010) 44.
[9] G.Y. Ke, S.K. Chen, T. Hsu, J.W. Yeh, Ann. Chim.-Sci.
Mat. 31 (2006) 669.
[10] Y. Zhang, G.L. Chen, L. Gan, J. ASTM Int. 7 (2010)
JAI102527.
[11] S. Singh, N. Wanderka, B.S. Murty, U. Glatzel, J.
Banhart, Acta Mater. 59 (2011) 182.
[12] F.J. Wang, Y. Zhang, G.L. Chen, J. Alloy Compd. 478
(2009) 321.
[13] C.J. Tong, M.R. Chen, S.K. Chen, J.W. Yeh, T.T. Shun,
S.J. Lin, S.Y. Chang, Metall. Mater. Trans. 36A (2005)
1263.
[14] C.W. Tsai, M.H. Tsai, J.W. Yeh, C.C. Yang, J. Alloy
Compd. 490 (2010) 160.
[15] U.S. Hsu, U.D. Hung, J.W. Yeh, S.K. Chen, Y.S. Huang,
C.C. Yang, Mater. Sci. Eng. A 460 (2007) 403.
[16] T.E. Mason, D. Abernathy, I. Anderson, J. Ankner, T.
Egami, G. Ehlers, A. Ekkebus, G. Granroth, M. Hagen,
K. Herwig, J. Hodges, C. Homann, C. Horak, L.
Horton, F. Klose, J. Larese, A. Mesecar, D. Myles, J.
Neuefeind, M. Ohl, C. Tulk, X.L. Wang, J. Zhao, Physica
B 385386 (2006) 955.
[17] X.L. Wang, T.M. Holden, G.Q. Rennich, A.D. Stoica,
P.K. Liaw, H. Choo, C.R. Hubbard, Physica B 385
(2006) 673.
[18] X.L. Wang, T.M. Holden, A.D. Stoica, K. An, H.D.
Skorpenske, A.B. Jones, G.Q. Rennich, E.B. Iverson,
Mater. Sci. Forum 652 (2010) 105.
[19] J.R. Greer, W.C. Oliver, W.D. Nix, Acta Mater. 53
(2005) 1821.
[20] Y. Yang, J.C. Ye, J. Lu, F.X. Liu, P.K. Liaw, Acta
Mater. 57 (2009) 1613.
[21] J.W. Yeh, S.Y. Chang, Y.D. Hong, S.K. Chen, S.J. Lin,
Mater. Chem. Phys. 103 (2007) 41.
[22] A. Takeuchi, A. Inoue, Mater. Trans. 46 (2005)
2817.
[23] M.D. Uchic, D.A. Dimiduk, Mater. Sci. Eng. A 400
(2005) 268.
[24] J.C. Ye, J. Lu, Y. Yang, P.K. Liaw, Intermetallics 18
(2010) 385.
[25] H. Bei, S. Shim, E.P. George, M.K. Miller, E.G. Herbert,
G.M. Pharr, Scripta Mater. 57 (2007) 397.
[26] D. Henkel, A.W. Pense, Structure and Properties of
Engineering Materials, fth ed., McGraw-Hill, New York,
2001.
[27] W.D. Nix, J.R. Greer, G. Feng, E.T. Lilleodden, Thin
Solid Films 515 (2007) 3152.
[28] M.D. Uchic, P.A. Shade, D.M. Dimiduk, Annu. Rev.
Mater. Res. 39 (2009) 361.
[29] E.M. Nadgorny, D.M. Dimiduk, M.D. Uchic, J. Mater.
Res. 23 (2008) 2829.
[30] S.I. Rao, D.M. Dimiduk, T.A. Parthasarathy, M.D.
Uchic, M. Tang, C. Woodward, Acta Mater. 56 (2008)
3245.

Você também pode gostar