You are on page 1of 4

Available online at www.sciencedirect.

com

ScienceDirect
Scripta Materialia 87 (2014) 5–8
www.elsevier.com/locate/scriptamat

Effect of Er additions on the precipitation strengthening


of Al–Hf alloys
H. Wu, S.P. Wen,⇑ K.Y. Gao, H. Huang, W. Wang and Z.R. Nie⇑
School of Materials Science and Engineering, Beijing University of Technology, Beijing 100124, People’s Republic of China
Received 13 January 2014; revised 24 April 2014; accepted 9 June 2014
Available online 18 June 2014

Precipitation strengthening was investigated in dilute Al–Er, Al–Hf and Al–Er–Hf alloys. Combined addition of 0.045 at.% Er
and 0.18 at.% Hf in Al gives a maximum hardness of 640 MPa, which is significantly higher than that of the Al–0.18Hf and Al–
0.045Er alloys. The remarkable strengthening can be attributed to the addition of Er, which accelerates the precipitation kinetics
and stimulates the decomposition of Al–Hf. The effect of Er and Hf on precipitation strengthening in Al–Er–Hf alloys is discussed.
Ó 2014 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

Keywords: Aluminum alloy; Precipitation; Erbium; Hafnium

The group IV transition elements (Ti, Zr, Hf) are Hf alloys due to the low driving force for precipitation
extremely slow diffusers in Al, and thus Al–transition of Al3Hf and the sluggish diffusivity of Hf.
element systems show particular promise for developing Precipitation hardening and thermal stability can be
thermally stable Al-based alloys [1]. Ordered metastable further enhanced by the joint addition of Sc and Hf in
L12-structured Al3Zr precipitates from supersaturated Al due to the formation of L12-structured Al3(Sc,Hf)
Al–Zr solid solution during post-solidification aging; precipitates [7,8]. Sc and Hf segregate to the dendrite
this precipitate is coherent with the Al matrix and ther- peripheries and interiors, respectively, which results in
mally stable at elevated temperature [2,3]. Nevertheless, homogeneous distribution of precipitates after aging
the maximum solubility of Zr in Al is only 0.083 at.% [8,9]. There is a eutectic reaction in Al with the addition
and therefore the volume of precipitates is limited. Hf of heavy rare-earth (RE) elements Er and Yb, resulting
shows many similarities to Zr although it is a relatively in a stable L12-structured Al3Er and Al3Yb phase in a-
rarely used alloying addition. The maximum solubility Al [10], which is similar to that in Al with addition of Sc.
of Hf in Al is 0.186 at.%, which is larger than that of In the last few years, Er and Yb have become relatively
Zr [4]. Metastable cubic L12-structured Al3Hf precipi- frequently used as alloying elements in Al-alloy to sub-
tates first decompose from the supersaturated Al–Hf stitute for the expensive element Sc [11,12]. Further-
solid solution—and then transform to the equilibrium more, it has been reported that Er and Yb can induce
tetragonal structure after prolonged aging at elevated precipitation in Al–Zr alloys [13–15]. For instance, a sig-
temperatures (>400 °C) [5]. However, in Al–1 wt.% Hf nificant precipitation strengthening is obtained in Al–
alloy, no obvious homogeneous distribution of precipi- Er–Zr alloys, which can be attributed to the acceleration
tates can be observed and the number density of the pre- of the precipitation kinetics of Al–Zr by a synergistic
cipitates is relatively low, even after thousands of hours effect of Zr and Er [13,15]. If the same synergistic effect
of aging at 450 °C [6]. This indicates that Al3Hf disper- can be achieved in Al–Er–Hf alloy, a precipitation hard-
soids are extremely difficult to precipitate in binary Al– ening effect may also be obtained. Thus far, however,
the effect of Er and Hf on precipitation strengthening
in Al–Er–Hf has never been reported. In this paper,
the isochronal and isothermal age behavior will be
⇑ Corresponding authors. Tel.: +86 10 67396439; fax: +86 10 investigated in Al–Er–Hf to clarify the effect of Er on
67391536 (S.P. Wen); e-mail addresses: wensp@bjut.edu.cn; the precipitation of Al–Er–Hf alloys.
zrnie@bjut.edu.cn

http://dx.doi.org/10.1016/j.scriptamat.2014.06.005
1359-6462/Ó 2014 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
6 H. Wu et al. / Scripta Materialia 87 (2014) 5–8

Table 1. Compositions of the experimental alloys (at.%). in the temperature range between room temperature and
Samples Nominal Verified 550 °C.
composition composition The precipitation hardening behavior of Al–Er–Hf is
Er Hf Er Hf distinct from that of Al–Er and Al–Hf. With the addi-
tions of Hf in Al–Er, the rate of strengthening between
Al–0.045Er 0.045 — 0.048 —
100 and 250 °C decreases, and the peak hardness is also
Al–0.18Hf — 0.18 — 0.175
Al–0.045Er–0.1Hf 0.045 0.1 0.04 0.11
delayed. An analogous delay of the peak hardness was
Al–0.045Er–0.18Hf 0.045 0.18 0.044 0.185 reported in Al–Er–Zr [13,17] and Al–Yb–Zr [14] with
increasing amount of Zr. For the Al–0.045Er–0.18Hf
alloy, there is an evident increase in hardness between
Four experimental alloys were cast from the commer- 250 and 475 °C, and the peak hardness of 609 ±
cially pure Al (Fe: 0.04 wt.%; Si:0.03 wt.%), Al–6 wt.% 7 MPa at 475 °C is significantly higher than the peak
Er and Al–4 wt.% Hf master alloys. The starting materi- hardness of other three alloys. When the temperature
als were melted in a crucible furnace by heating to is above 500 °C, the microhardness continuously
790 °C. After thorough stirring, the melt was poured into decreases as a result of the coarsening and dissolution
an iron mold to produce a 30  90  150 mm3 plate and of the precipitates.
quickly cooled to ambient temperature. The composi- Figure 2 demonstrates the microhardness variation as
tions of the alloys are given in Table 1 (all compositions a function of aging time for all the experimental alloys
are in at.% unless otherwise noted), which were verified during isothermal aging at 350 °C. The Al–0.18Hf alloy
by inductively coupled plasma-atomic emission spectros- does not exhibit precipitation hardening after aging for
copy. The castings were homogenized at 640 °C for 24 h 500 h. The isochronal and isothermal aging hardening
and then water quenched to ambient temperature. Two behaviors indicate that the binary Al–Hf alloy has no
separate aging studies were conducted: (i) isochronal significant precipitation strengthening. In the Al–Hf sys-
aging in 3 h steps of 25°Cfor temperatures ranging from tem, the maximum solid solubility of Hf in Al is 0.186
150 to 600 °C; and (ii) isothermal aging at 350 °C for dif- at.% at the peritectic temperature of 662.2 °C [4]. This
ferent times ranging from 0.1 to 500 h. solid solubility, however, only diminishes slightly with
Vickers microhardness measurements were performed decreasing temperature (the solid solubility of Hf is
on polished samples using a load of 200 g and a dwell 0.1 at.% at 350 °C [4]). Therefore, the driving force
time of 10 s. At least 10 indentations were made per sam- of the precipitation is low and the precipitation process
ple across several grains. The microstructural evolution is very slow due to low diffusivity, which limits the
was investigated by transmission electron microscopy potential for precipitation strengthening. Thus, the Al–
(TEM) using a JEOL 2100 microscope TEM foils were 0.18Hf alloy exhibits negligible age hardening. Indeed,
prepared by twin-jet polishing with an electrolyte solu- most of the reported precipitation hardening in Al–Hf
tion of 30% nitric acid and 70% methanol at the voltage alloys were produced by non-equilibrium cast supersat-
of 15 V DC and the temperature below 25 °C. urated alloys [5,18]. The addition of trace alloying ele-
Figure 1 displays the precipitation behavior of ments, e.g. Si or Fe, to Al–Hf alloy in order to
Al–0.045Er, Al–0.18Hf, Al–0.045Er–0.1Hf and increase the driving force and speed up the precipitation
Al–0.045Er–0.18Hf during isochronal aging as deter- process of Al3Hf has been reported [16,19]. The results
mined by microhardness. The hardness of Al–0.045Er show that the distribution of spherical Al3Hf dispersoids
increases between 150 and 250 °C and achieves a peak in Al–1.1 wt.% Hf–0.15 wt.% Si alloy was still heteroge-
of 422 ± 15 MPa at 250 °C, then decreases rapidly at neous and there was no evident precipitation hardening.
higher temperature. However, the homogenized Al– As shown in Figure 2, there is a maximum hardness
0.18Hf has no strengthening effect, and the microhard- of 392 ± 11 MPa for Al–0.045Er alloy after aging for
ness maintains nearly constant over the whole tempera- 10 min; then the hardness decreases rapidly with over-
ture range. Hallem [16] also reported that the hardness aging, which is behavior similar to that of Al–0.03 Er
of Al–0.95 wt.% Hf alloy was fairly constant after aging aged at 400 °C [17]. The hardening rate of Al–
0.045Er–0.1Hf is slower and the peak hardness of

Figure 1. Evolution of Vickers microhardness during isochronal aging


for Al–0.045Er, Al–0.18Hf, Al–0.045Er–0.1Hf and Al–0.045Er–0.18Hf Figure 2. Vickers microhardness vs. aging time at 350 °C for Al–
alloys. 0.045Er, Al–0.18Hf, Al–0.045Er–0.1Hf and Al–0.045Er–0.18Hf alloys.
H. Wu et al. / Scripta Materialia 87 (2014) 5–8 7

360 ± 4 MPa is lower than that of the Al–0.045Er alloy. radius of 3.4 ± 0.3 nm. The microhardness increases by
With prolonged aging, the hardness decreases more 340 MPa from 100 to 475 °C, owing to the precipitation
slowly than that of Al–0.045Er; however, Al–0.045Er– of the dense, homogeneously distributed spherical
0.18Hf alloy achieves a peak hardness of 640 ± 8 MPa particles.
after aging for 100 h, and exhibits only a slight over- Figure 3e shows a TEM micrograph of Al–0.045Er–
aging after aging for 500 h. When aged isothermally at 0.18Hf aged to 550 °C; the average radius of the precip-
the same temperature of 350 °C, the peak hardness of itates, 15.5 ± 0.8 nm, is larger and the number density of
Al–0.04Er–0.08Zr is only 560 MPa after aging for the precipitates is less than those aged at 475 °C, indicat-
70 h [13], which is much lower than the peak hardness ing the coarsening of the precipitates, which is consistent
of the present Al–0.045Er–0.18Hf alloy. For the iso- with the corresponding sharp decrease in microhard-
chronal aging, the peak hardness of Al–0.045Er– ness. The inset micrograph of the precipitates in Figure
0.18Hf is 609 MPa, which is also larger than the 3e exhibits a dark core surrounded by a brighter shell.
isochronal aging peak hardness of 505 MPa for the The composition of the precipitate detected by energy-
Al–0.04Er–0.08Zr alloy [13]. dispersive spectroscopy (EDS) consists of Al, Er and
In order to further understand the precipitation Hf, as shown in Figure 3f. Characteristic L12 superlat-
strengthening behavior, the corresponding microstruc- tice reflections of the precipitates are visible in the inset
tures of the two ternary alloys were characterized using selected-area diffraction (SAD) pattern in Figure 3d and
TEM. Very few precipitates are detected in e. In published studies of Al–Er–Zr [13], Al–Sc–Er–Zr
Al–0.045Er–0.1Hf after isochronal ageing to 350 °C, as [20] and Al–Sc–Hf [7] alloys, Er and Sc are found to seg-
shown in Figure 3a, and the corresponding hardness is regate to the core of the precipitates, while Zr and Hf
only 350 ± 5 MPa. Figure 3 shows micrographs pf three segregate to the periphery of the precipitates due to
various aging stages of Al–0.045Er–0.18Hf: under-aged the difference in diffusion coefficient between Er, Sc
(Fig. 3a), peak-aged (Fig. 3c,d) and over-aged (Fig. 3e), and Zr, Hf. Therefore, the precipitates in the present
respectively. The fine precipitates are distributed homo- Al–Er–Hf should be L12 structured Al3(Hf,Er), with
geneously in Figure 3b. The number density of the an Er-rich core and an enveloping Hf-rich shell, but this
precipitates is much higher than that of Al–0.045Er– hypothesis needs to be verified in further studies.
0.1Hf alloy, and these precipitates are coherent with Figure 4a and b are bright-field and dark-field images
the Al matrix, as determined from the Ashby–Brown of Al–0.045Er–0.18Hf alloy after isothermal aging at
strain-field contrast. Figure 3c and d are, respectively, 350 °C for 100 h. The average diameter of the precipi-
dark-field and bright-field TEM images of Al–0.045Er– tates is 2.3 ± 0.2 nm, and the corresponding microhard-
0.18Hf after aging to 475 °C, showing a homogeneous ness reaches the peak. Figure 4c and d are dark-field and
distribution of spherical precipitates with an average bright-field micrographs of precipitates after aging for
321 h; these dense dispersed particles have an average
diameter of 3.5 ± 0.4 nm. The average diameter of the
precipitates increases by only 1.2 nm from 100 to
321 h, indicating that the Al3(Hf,Er) precipitates possess
an outstanding resistance to growth and coarsening,
owing to the sluggish diffusibility of the Hf. In addition,
Hf reduces the lattice mismatch between the precipitates
and Al matrix [7,21]. Since the small lattice mismatch
decreases the interfacial free energy of precipitates/
matrix which is the driving force of the coarsening
[22], the addition of Hf retards the growth and coarsen-
ing of Al3(Hf,Er) precipitates.
Combined additions of 0.045Er and 0.18Hf in Al
results in a pronounced strengthening effect compared
with binary Al–0.045Er and Al–0.18Hf alloys. This
results from the formation of coherent L12-ordered
Al3(Hf,Er) precipitates with a structure consisting of an
Er-enriched core surrounded by an Hf-enriched shell.
Similar core–shell structures have been reported exten-
sively in Al–Er-Zr [13], Al–Sc–RE [11,23,24] and Al–
Sc–TM (transition metal) alloys [7,25,26]. By comparing
the precipitation behavior of Al–Er and Al–Er–Hf alloys
in Figures 1 and 2, we can find that the initial precipita-
tion strengthening decreases with increasing Hf concen-
tration, indicating that Hf will retard the precipitation
of Al3Er, and then slow down the initial precipitation
Figure 3. TEM micrographs during isochronal aging: (a) bright-field strengthening, as evidenced by the TEM micrographs
image of Al–0.045Er–0.1Hf aged to 350 °C; (b, d, e) bright-field image in Figure 3a and b. Due to the large supersaturation of
of Al–0.045Er–0.18Hf aged to 400, 475 and 550 °C, respectively; (c) Hf in Al–0.045Er–0.18Hf, the strengthening effect is sig-
corresponding dark-field image of Al–0.045Er–0.18Hf aged to 475 °C; nificant. Thus, the isochronal aging peak hardness of Al–
(f) EDS pattern of precipitates of Al–0.045Er–0.18Hf aged to 550 °C. 0.045Er–0.18Hf is 100 and 250 MPa larger than the iso-
8 H. Wu et al. / Scripta Materialia 87 (2014) 5–8

sition of Al–Er and slows down the growth and


coarsening of the precipitates.

The work is supported in part by the National


Key Basic Research & Development Plan Project (No.
2012CB619503), National Natural Science Foundation
Project (Nos. 51201003 and 51101001), the National
High Technology Research and Development Program
(No. 2013AA031301), the International Science & Tech-
nology Cooperation Program of China (2013DFB
c50170), and the Beijing Municipal Science Foundation
of Education Commission (KM201310005002).

[1] K.E. Knipling, D.C. Dunand, D.N. Seidman, Z. Metal-


lkd. 97 (2006) 246–265.
[2] K.E. Knipling, D.C. Dunand, D.N. Seidman, Acta Mater.
Figure 4. Bright-field and dark-field TEM micrographs of Al–0.045Er– 56 (2008) 1182–1195.
0.18Hf alloy isothermally aged at 350 °C for: (a,b) 100 h; (c,d) 321 h. [3] K.E. Knipling, D.C. Dunand, D.N. Seidman, Metall.
Mater. Trans. A 38 (2007) 2552–2563.
[4] T. Wang, Z.P. Jin, J.C. Zhao, J. Phase Equilib. 23 (2002)
416–423.
chronal aging peak hardness of Al–0.04Er–0.08Zr [13] [5] N. Ryum, J. Mater. Sci. 10 (1975) 2075–2081.
and Al–0.06Er–0.06Zr [17], respectively, even if Hf is [6] S. Hori, N. Furushiro, W. Fujitani, J. Japan Inst. Light
seven times more expensive than Zr [27]. Again, although Metals 31 (1981) 649–654.
the precipitation process of Al–Er–Hf alloy is relatively [7] H. Hallem, W. Lefebvre, B. Forbord, F. Danoix, K.
slow, it shows a strength comparable to that of Al– Marthinsen, Mater. Sci. Eng. A 421 (2006) 154–160.
0.06Sc–0.02Er [11] and Al–0.1Sc–0.17Hf [16]. Addition- [8] H. Hallem, B. Forbord, K. Marthinsen, Mater. Sci.
ally, Er and Hf are much cheaper than Sc, thus substitut- Forum 28 (2004) 825–831.
ing Er and Hf for the costly Sc contributes to reducing [9] V.V. Zakharov, T.D. Rostova, Met. Sci. Heat Treat. 49
the cost of these alloys. (2007) 435–442.
[10] O.I. Zalutskaya, V.R. Ryabov, I.I. Zalutskii, Dopov.
Base on the above experimental results, we can deduce
Akad. Nauk. Ukr. A 31 (1969) 255–259.
that the enhanced strengthening effect of Al–Er–Hf [11] R.A. Karnesky, D.C. Dunand, D.N. Seidman, Acta
alloys must be attributed to the synergistic effect of Er Mater. 57 (2009) 4022–4031.
and Hf. On one hand, Hf can retard the decomposition [12] M.E. Van Dalen, D.C. Dunand, D.N. Seidman, Acta
of Al–Er due to the interaction between Er and Hf and Mater. 59 (2011) 5224–5237.
the enveloping Hf-enriched shell as a barrier can retard [13] S.P. Wen, K.Y. Gao, Y. Li, H. Huang, Z.R. Nie, Scr.
the growth of the precipitates. Moreover, the Hf- Mater. 65 (2011) 592–595.
enriched shell can reduce the lattice parameter mismatch [14] Y. Zhang, W. Zhou, H. Gao, Y. Han, K. Wang, Scr.
and concomitantly the precipitate–matrix interfacial Mater. 69 (2013) 477–480.
energy, which can slow down the coarsening process of [15] H. Li, Z. Gao, H. Yin, H. Jiang, X. Su, J. Bin, Scr.
Mater. 68 (2013) 59–62.
the precipitates. On the other hand, the pre-existing
[16] H. Hallem, Ph.D thesis, Norwegian University of Science
Al3Er precipitates which form below 400 °C offer more and Technology, 2005.
heterogeneous nucleation sites for the sluggish Hf atoms, [17] H. Li, J. Bin, J. Liu, Z. Gao, X. Lu, Scr. Mater. 67 (2012)
thereby increase the driving force and speeding up the 73–76.
precipitation of Al3Hf; thus Er accelerates the precipita- [18] S.K. Pandey, C. Suryanarayana, Mater. Sci. Eng. A 111
tion and stimulates the decomposition of Al–Hf. There- (1989) 181–187.
fore, the enhanced precipitation-strengthening effect and [19] H. Hallem, B. Forbord, K. Marthinsen, Mater. Sci. Eng.
excellent thermal stability of Al–0.045Er–0.18Hf can be A 387–389 (2004) 940–943.
explained by the synergistic effect of Er and Hf. [20] C. Booth-Morrison, D.C. Dunand, D.N. Seidman, Acta
In summary, the precipitation strengthening of Mater. 59 (2011) 7029–7042.
[21] Y. Harada, D.C. Dunand, Mater. Sci. Eng. A 329–331
Al–Er, Al–Hf and Al–Er–Hf alloys was investigated
(2002) 686–695.
Al–0.045Er achieves a peak hardness of 422 MPa. Al– [22] C.J. Kuehmann, P.W. Voorhees, Metall. Mater. Trans. A
0.18Hf exhibits no age hardening due to the low driving 27 (1996) 937–943.
force and slow precipitation process of Al3Hf. [23] R.A. Karnesky, M.E. van Dalen, D.C. Dunand, D.N.
Al–0.045Er–0.18Hf alloy achieves a maximum hardness Seidman, Scr. Mater. 55 (2006) 437–440.
of 640 MPa, which is significantly higher than that of [24] M.E. Krug, A. Werber, D.C. Dunand, D.N. Seidman,
Al–0.045Er and Al–0.18Hf alloys due to the dense Acta Mater. 58 (2010) 134–145.
homogeneously distributed L12-structured Al3(Hf,Er) [25] M.E. van Dalen, D.C. Dunand, D.N. Seidman, Acta
nanoprecipitates. The synergistic effect of Er and Hf Mater. 53 (2005) 4225–4235.
can explain the enhanced precipitation hardness and [26] K.E. Knipling, R.A. Karnesky, C.P. Lee, D.C. Dunand,
D.N. Seidman, Acta Mater. 58 (2010) 5184–5195.
excellent thermal stability of Al–0.045Er–0.18Hf. Er
[27] P.J. Loferski, US Geological Survey 2011, Minerals
accelerates the precipitation process and stimulates the Yearbook: Zirconium and Hafnium, 2013.
decomposition of Al–Hf, while Hf retards the decompo-

You might also like