You are on page 1of 11

Journal of Alloys and Compounds 832 (2020) 154997

Contents lists available at ScienceDirect

Journal of Alloys and Compounds


journal homepage: http://www.elsevier.com/locate/jalcom

Ultrahigh thermal stability and hardness of nano-mixed fcc-Al and


amorphous phases for multicomponent Al-based alloys
Y. Jin a, A. Inoue a, b, c, d, e, *, F.L. Kong b, S.L. Zhu a, **, F. Al-Marzouki d, A.L. Greer f, ***
a
School of Materials Science and Engineering, Tianjin University, Tianjin, 300072, China
b
International Institute of Green Materials, Josai International University, Togane, 283-8555, Japan
c
Institute of Massive Amorphous Alloy Science, China University of Mining Technology, Xuzhou, 221116, China
d
Department of Physics, King Abdulaziz University, Jeddah, 22254, Saudi Arabia
e
MISiS, National University of Science and Technology, Moscow, 119049, Russia
f
Department of Materials Science and Metallurgy, University of Cambridge, Cambridge, CB3 0FS, UK

a r t i c l e i n f o a b s t r a c t

Article history: Fcc-Al (a-Al) þ amorphous phase mixtures with ultrahigh hardness are formed by a heating-induced
Received 28 January 2020 reverse transition from the primary precipitates of a-Al and AlxMy in residual amorphous phase for
Received in revised form melt-spun Al84Y9Ni4Co1.5Fe0.5TM1 (TM ¼ V, Nb, Cr, Mo, Mn, Fe, Co, Ni or Cu) amorphous alloys. The
10 March 2020
resulting particle diameter and volume fraction of the a-Al phase are 5e15 nm and 60e70%, and no
Accepted 29 March 2020
Available online 3 April 2020
internal defects are observed. The solute content in the a-Al phase is 8e11% for the TM ¼ V or Cu alloys.
In contrast, the reverse transition is not found for the TM ¼ Zr, Ag or Au alloys, presumably because these
elements have atomic radii larger than Al and positive heats of mixing with other solute elements
Keywords:
Thermal stability
hindering the atomic rearrangements to decompose the AlxMy compound. The [a-Al þ amorphous]
Hardness phase mixture is maintained up to about 700 K and exhibits a Vickers hardness of 550e580, much higher
Multicomponent than for the corresponding crystalline alloys. The high hardness at elevated temperature is due to the
Al-based amorphous alloy coexistence of perfect crystal a-Al and high solute-content residual amorphous phase. For a cast conical
rod of the TM ¼ Cu alloy, the microstructure is fully amorphous up to a diameter of 0.82 mm, and is an
[a-Al þ amorphous] phase mixture at larger diameters up to 1.2 mm. The formation of the highly stable
[a-Al þ amorphous] phase mixtures by the heating-induced reverse transition, as well as the bulk for-
mation (in cast rods) of similar phase mixtures for the TM ¼ V, Nb, Cr, Mo, Mn, Fe, Co, Ni or Cu alloys is
promising for future structural and coating materials owing to their high hardness and high elevated-
temperature strength.
© 2020 Elsevier B.V. All rights reserved.

1. Introduction corrosion resistance, and high anti-erosion performance [1e3].


With these advantages, they have recently been used as coating
Savings of energy and natural resources motivate the develop- materials in, for example, high-speed patrol boats and hydroplanes
ment of higher-performance alloys that permit miniaturization and [4], applications that benefit from the improved glass-forming
light-weighting. There is a particularly clear need for alloys with ability to over 1 mm in maximum diameter and improved spray-
high specific strength that can be maintained at elevated temper- coating techniques. For high strength and good corrosion resis-
atures. Aluminum-based glassy alloys have attracted a steady in- tance, mixed amorphous þ fcc a-Al phase alloys are also attractive
terest because they exhibit high specific strength, high wear and because these can be formed over wider composition ranges, at
lower cooling rates and under a variety of preparation conditions.
Furthermore, they have higher strength, much higher heat-
resistant strength, and nearly the same corrosion resistance in
* Corresponding author. School of Materials Science and Engineering, Tianjin
comparison to glassy single-phase alloys [5,6]. It is therefore
University, Tianjin, 300072, China.
** Corresponding author. important to search for new heat-resistant alloys that can be
*** Corresponding author. readily prepared with the desirable [amorphous þ a-Al] phase
E-mail addresses: inoue@jiu.ac.jp (A. Inoue), slzhu@tju.edu.cn (S.L. Zhu), alg13@ mixture.
cam.ac.uk (A.L. Greer).

https://doi.org/10.1016/j.jallcom.2020.154997
0925-8388/© 2020 Elsevier B.V. All rights reserved.
2 Y. Jin et al. / Journal of Alloys and Compounds 832 (2020) 154997

In 2017, Han et al. found a reverse transition in which the vol- 3. Results
ume fraction of amorphous phase (am) increases on going from a
three-phase mixture [a-Al þ AlxMy (M ¼ Y, Ni, Co, Fe, Pd) þ residual Fig. 1(a) and (b) show X-ray diffraction patterns and DSC curves,
(am’)] to a two-phase mixture [a-Al þ residual (am00 )]. This occurs respectively, of as-spun Al84Y9Ni4Co1.5Fe0.5TM1 (TM ¼ Zr, V, Nb, Cr,
during heating of Al-based amorphous alloys around the compo- Mo, Mn, Cu, Ag or Au) alloy ribbons. All of the X-ray patterns show
sition Al84Y9Ni4Co1.5Fe0.5Pd1 (at%), and the desirable two-phase broad haloes without any appreciable crystalline peaks. However,
mixture [a-Al þ am00 ] is retained at high temperatures up to some of the halos (such as TM ¼ Ag and Nb) are not so symmetrical,
650 K [7]. This phase mixture also retains high strength up to these which are considered due to the presence of a small number of
temperatures, which is significantly different from the poor heat clusters or nanocrystalline phases. The position of the maximum of
resistance associated with many Al-based amorphous alloys of the main diffraction halo is not fixed, and appears to shift to lower
lower crystallization temperature. The [a-Al þ am00 ] alloy type is angle for the alloys containing TM elements with larger atomic
thus attractive as a new type of coating for high-temperature ap- radius, as seen in particular for the Zr-containing alloy; this in-
plications. So far, however, all the reports of the reverse transition dicates that the TM elements are incorporated into the amorphous
([a-Al þ AlxMy þ am’] / [a-Al þ am00 ]) are for compositions con- phase in each case. The DSC curves show an endothermic peak due
taining some noble metal. In the present work, we seek new Al- to the glass transition, followed by a narrow supercooled-liquid
based alloys showing the reverse transition, while having only region and then two or three exothermic peaks due to the phase
inexpensive components. We find that a similar reverse transition transitions for all the Al-based glassy alloys except the Mn-
does occur for Al84Y9Ni4Co1.5Fe0.5TM1, (TM ¼ V, Nb, Cr, Mo, Mn, Fe, containing alloy, as shown in Fig. 1 (b). The second exothermic
Co, Ni or Cu) multicomponent amorphous alloys which can be peak marked with arrows has always much lower area. The DSC
regarded as less expensive and therefore more attractive for data indicate that the crystallization is completed through two or
application. three stages.
It is also important to derive a general rule for the appearance of Fig. 2(aef) show X-ray diffraction patterns of the V-, Nb-, Cr-,
such an unusual reverse transition for multicomponent Al-based Mo-, Mn- and Cu-containing glassy alloys, annealed for 900 s at
amorphous alloys by examining the heating-induced phase temperatures just above the first, second and third exothermic
decomposition. This work examines the possibility of forming the peaks. As identified in the corresponding diffraction patterns, the
[a-Al þ am00 ] phase mixture during crystallization for a very wide first exothermic peak is due to the change from an amorphous
range of melt-spun amorphous alloys: Al84Y9Ni4Co1.5Fe0.5TM1 phase to fcc-Al (a-Al) þ AlxMy (M ¼ Y, Ni, Co, Fe and V, Nb, Cr, Mo,
(TM ¼ Zr, V, Nb, Cr, Mo, Mn, Fe, Co, Ni, Cu, Ag or Au). The micro- Mn or Cu) þ residual amorphous (am’) phases, while the second
structural features, thermal stability and mechanical properties of peak originates from the transition of this three-phase mixture to
the [a-Al þ am00 ] samples are characterized, and the dominant an [a-Al þ am00 ] mixture for all the alloys. It is notable that the
factors underlying the reverse transition are investigated. multicomponent AlxMy compound disappears, and the volume
fractions of a-Al and amorphous phase increase during the second-
stage exothermic reaction. The third exothermic peak corresponds
to the complete crystallization of the sample, with the transition
from [a-Al þ am00 ] to [a-Al þ Al3Y þ Al9(Fe, Co, Ni, TM)2 (TM ¼ V, Nb,
2. Experimental procedures Cr, Mo, Mn or Cu)] [10,11]. A similar heating-induced reverse
transition from [a-Al þ AlxMy (M ¼ Y, Fe, Co, Ni, TM) þ am’] to [a-
Multicomponent Al-based alloys with compositions of Al84Y9- Al þ am00 ] is also recognized for Fe, Co and Ni as the added TM
Ni4Co1.5Fe0.5TM1 (TM ¼ Zr, V, Nb, Cr, Mo, Mn, Fe, Co, Ni, Cu, Ag or element, as shown in Fig. 3(aec). The X-ray diffraction patterns
Au) were chosen because the formation of [a-Al þ amorphous] show clearly the sequence [am] / [a-Al þ AlxMy þ am’] / [a-
phase mixtures was expected to give higher mechanical strength Al þ am00 ] / [a-Al þ Al3Y þ Al9(Fe, Co, Ni, TM)2] with increasing
and higher thermal stability [5,6]. Transition-metal (TM) elements temperature. Among all the alloys, the metastable [a-Al þ am00 ]
were selected by considering the atomic size mismatches to Al and phase mixture is maintained over the wide range of 620e650 K
the heat of mixing with the constituent base elements (Al, Y, Ni, Co and, most notably up to the high temperatures of 600e660 K for
and Fe) [8,9]. The alloy compositions are given as nominal atomic the Nb-, Cr- and Mo-containing alloys. These features indicate the
percentages. Mixed alloy ingots were prepared by arc-melting the possibility that these Al-based alloys would retain extremely high
mixtures of pure metals with purities above 99.9 wt% under an heat-resistant strength even at temperatures above 700 K.
argon atmosphere. Alloy ribbons with a thickness of about 0.03 mm To confirm the reverse transition from [a-Al þ AlxMy þ am’] to
and width of 1.5e1.7 mm were prepared by single-roller melt- [a-Al þ am00 ], bright-field TEM images and selected-area electron
spinning. Conical alloy rods were also prepared by injection copper- diffraction patterns of the V-, Nb-, Cr- and Cu-containing glassy
mold casting. Amorphicity was evaluated by X-ray diffraction, op- alloys annealed for 900 s at 640, 630, 640 and 640 K, respectively,
tical microscopy (OM) and transmission electron microscopy corresponding to the temperature just above the second
(TEM). The thermal stability associated with glass transition, exothermic peak, are shown in Fig. 4(aeh). A fine dispersion of
supercooled liquid region and crystallization was examined by precipitates in the amorphous matrix can be identified as fcc a-Al
differential scanning calorimetry (DSC) at a heating rate of phase. The average diameter of the a-Al particles is about 7 nm for
0.67 K s1. The microstructures developed upon crystallization TM ¼ V, 5 nm for TM ¼ Nb, 7 nm for TM ¼ Cr and about 15 nm for
were examined by X-ray diffraction, TEM combined with energy- TM ¼ Cu, indicating a tendency for the diameter to decrease with
dispersion X-ray (EDX) analysis, high-resolution TEM and nano- increasing melting temperature and/or atomic radius of the TM
beam diffraction analysis, and crystallization kinetics was evalu- element.
ated by DSC at heating rates of 0.167e0.667 K s1. Hardness was We further tried to confirm the coexistence of only [a-Al þ am00 ]
measured using a Vickers hardness indenter with a load of 0.49 N. phases for the Cr- and Cu-containing alloys annealed after the
Bending ductility was evaluated by a simple bend test. Shear second exotherm by high-resolution TEM (HRTEM) and nanobeam
markings around the Vickers indents and on the outer surface of diffraction methods. Fig. 5(aeh) show HRTEM images, nanobeam
bent ribbons were examined by OM and scanning electron micro- diffraction patterns and EDX profiles with analytical sites and
scopy (SEM). compositions. The fringe contrast, as well as the atomic
Y. Jin et al. / Journal of Alloys and Compounds 832 (2020) 154997 3

Fig. 1. (a) X-ray diffraction patterns and (b) DSC curves of melt-spun Al84Y9Ni4Co1.5Fe0.5TM1 (TM ¼ Zr, V, Nb, Cr, Mo, Mn, Cu, Ag or Au) alloy ribbons. The arrows in (a) and (b)
present the summit position of the maximum X-ray diffraction peak and the peak temperature of the second exothermic peak, respectively.

Fig. 2. X-ray diffraction patterns of (a) Al84Y9Ni4Co1.5Fe0.5V1, (b) Al84Y9Ni4Co1.5Fe0.5Nb1, (c) A84Y9Ni4Co1.5Fe0.5Cr1, (d) Al84Y9Ni4Co1.5Fe0.5Mo1, (e) Al84Y9Ni4Co1.5Fe0.5Mn1 and (f)
Al84Y9Ni4Co1.5Fe0.5Cu1 alloys annealed for 900 s at different temperatures.

configurations in the a-Al phase, does not indicate defects, sug- am00 phase (d and h) regions of the Cr- and Cu-containing alloys
gesting that this nanoscale phase formed through the reverse obtained by annealing for 900 s at 640 K. The compositions are
transition has a perfect crystal structure. The nanobeam diffraction evaluated to be approximately Al89.4Y3.2Ni3.8Co2.1Fe0.6Cr0.9 and
patterns confirm that the precipitates are a-Al. Based on the data in Al91.8Y0.9Ni5.2Co0.9Fe1Cu0.2 (at%) for the a-Al phase, and Al77.6Y11.5-
Figs. 2, 4 and 5, the lattice parameter of the a-Al phase is evaluated Ni7.0Fe0.4Co2.2Cr1.3 and Al78.8Y11.4Ni5.5Co2.6Fe0.7Cu1.0 for the am00
to be 0.4053 nm for the Cr- and Cu-containing alloys, and no phase. The am00 phase has much higher Y, Ni and Co contents than
appreciable change in the lattice parameter with TM elements is the a-Al phase, suggesting that the am00 phase was formed mainly
recognized. This value is slightly larger than that (0.4049 nm) for from the AlxMy compound phase regions. The other solute elements
pure Al, indicating that the a-Al contains a small amount of Y with a except Y are also enriched into the am00 phase as shown in Fig. 5 (d
much larger atomic radius than Al, in addition to the other solute and h). Nevertheless, the a-Al phase has an overall solute content of
elements with smaller atomic radii than Al. The perfect crystal a-Al about 10.6 at% (Fig. 5 c) for TM ¼ Cr, and 8.2 at% (Fig. 5 g) for
phase is surrounded by the residual amorphous phase and no AlxMy TM ¼ Cu, values which are much higher than the equilibrium solid-
intermetallic compound is recognized, even at the interface be- solubility limit (near zero at%) at room temperature for Al-(Fe, Co,
tween the a-Al and am00 phases. Although the a-Al particle diam- Ni, TM) binary alloys [12]. A similar partitioning of Y was also
eter for the TM ¼ Cu alloy is about twice that for the TM ¼ Cr alloy, recognized for the Nb-containing alloy, being independent of pe-
the a-Al particles are still defect-free and the AlxMy compound is riodic group number and atomic size of the TM elements.
still absent. Fig. 5 shows the EDX profiles of the a-Al (c and g) and To investigate the ease of the reverse amorphization in
4 Y. Jin et al. / Journal of Alloys and Compounds 832 (2020) 154997

Fig. 3. X-ray diffraction patterns of (a) Al84Y9Ni4Co1.5Fe1.5, (b) Al84Y9Ni4Co2.5Fe0.5 and (c) Al84Y9Ni5Co1.5Fe0.5 alloys annealed for 900 s at different temperatures.

comparison with the first-stage crystallization, the activation en- and the third stages than between the first and the second stages.
ergies for the three stages of decomposition were evaluated by the We also examined the possibility of forming a similar [a-
Kissinger analysis [13]. As examples, Fig. 6(a) and (b) show the Al þ am00 ] phase mixture from [a-Al þ AlxMy þ am’] in the other
Kissinger plots of the first, second and third exothermic peaks for alloy series of Al84Y9Ni4Co1.5Fe0.5TM1 (TM ¼ Zr, Ag or Au). The se-
Al84Y9Ni4Co1.5Fe0.5TM1 (TM ¼ Nb or Cu) amorphous alloys, together lection of these TM elements is because they have larger atomic
with the corresponding DSC curves measured at different heating radii than those of Al and the other constituent elements [8] and
rates in Fig. 6 (c). Three exothermic peaks are seen in the DSC include atomic pairs with positive heats of mixing with the other
curves. From the clear straight lines in the plots, the activation elements as exemplified for ZreY, AgeCo, AgeFe, Au-Co and AueFe
energies for the first, second and third stages are estimated to be [9]. As shown by the X-ray diffraction patterns in Fig. 7(aec), no
286, 247 and 196 kJ mol1, respectively, for TM ¼ Nb, and 252, 227 reverse transition from [a-Al þ AlxMy þ am’] to [a-Al þ am00 ] is
and 174 kJ mol1, respectively, for TM ¼ Cu. The activation energies recognized for the TM ¼ Zr, Ag or Au alloys.
are slightly larger for the Nb-containing alloy and the difference Fig. 8 shows the changes in the structure, Vickers hardness (Hv)
seems to reflect much higher melting temperature and larger and bending plasticity with annealing temperature for Al84Y9Ni4-
atomic size for Nb than for Cu [14]. Although the activation energies Co1.5Fe0.5TM1 (TM ¼ V, Nb, Cr, Mo, Mn or Cu) alloy ribbons, together
decrease in the order of first / second / third stage, the stability with previous data on the Al84Y9Ni4Co1.5Fe0.5Pd1 alloy [7]. The Hv is
of the metastable [a-Al þ am00 ] phase mixture is estimated to be 360e430 in the as-spun glassy state, increases rapidly in the first-
relatively high because of the high decomposition temperature as stage crystallization to [a-Al þ AlxMy þ am’] and further to
well as the much larger temperature interval between the second maximum values of 550e580 at the temperatures around the
Y. Jin et al. / Journal of Alloys and Compounds 832 (2020) 154997 5

Fig. 4. TEM bright-field images and selected-area electron diffraction patterns of (a, b) Al84Y9Ni4Co1.5Fe0.5V1, (c, d) Al84Y9Ni4Co1.5Fe0.5Nb1, (e, f) Al84Y9Ni4Co1.5Fe0.5Cr1 and (g, h)
Al84Y9Ni4Co1.5Fe0.5Cu1 alloys annealed for 900 s at 640 K, 630 K, 640 K and 640 K, respectively. The arrows present a-Al precipitates.

Fig. 5. HRTEM images, nanobeam diffraction patterns and EDX profiles with analytical sites and compositions of (aed) Al84Y9Ni4Co1.5Fe0.5Cr1 and (eeh) Al84Y9Ni4Co1.5Fe0.5Cu1
alloys annealed for 900 s at 640 K. The analyses were made in a-Al phase region for Fig. 5 (c) and (g) and the residual amorphous phase for Fig. 5 (d) and (h).

second-stage reverse amorphization reaction to [a-Al þ am00 ], and different among the TM elements. The absence of a significant
then decreases significantly to 450e500 in the third decomposition difference in the maximum hardness values with TM elements
stage to [a-Al þ Al3Y þ Al9(Fe,Co,Ni,TM)2]. The Cr- and Mo- indicates that the metastable two-phase mixture [a-Al þ am00 ]
containing alloys also exhibit the highest hardness values in the plays a dominant role in the achievement of the high hardness.
annealing temperature range just before the phase decomposition Fig. 9 shows SEM and optical micrographs revealing the shear
from [a-Al þ am00 ] to the three crystalline phases. It is notable that markings on the outer surface of bent ribbons and around the
the highest Vickers hardness is obtained for the metastable [a- Vickers hardness indents, for Al84Y9Ni4Co1.5Fe0.5TM1 (TM ¼ Cr or
Al þ am00 ] phase mixture, and that the fully crystalline phase Cu) ribbons in as-spun and annealed states. All the as-spun ribbons
mixtures including compounds exhibit much lower hardness. This have good plasticity: they can be bent through 180 without frac-
feature is recognized for all the present Al-based alloys except the ture, as shown in Fig. 9 (a, c). Although the good plasticity can be
TM ¼ Zr, Ag or Au alloys, though the hardness values are slightly maintained nearly up to the glass-transition temperature, heating
6 Y. Jin et al. / Journal of Alloys and Compounds 832 (2020) 154997

Fig. 6. DSC curves measured at different heating rates and the corresponding Kissinger plots of the first-, second- and third-exothermic peaks for (a, b) Al84Y9Ni4Co1.5Fe0.5Nb1 and
(c, d) Al84Y9Ni4Co1.5Fe0.5Cu1 glassy alloys.

to higher temperature leads to embrittlement of the ribbons, which confirmed that coexistent [a-Al þ am] phases in the diameter range
then fracture during bending. In contrast, whatever the annealing 1.0e1.2 mm, and then coexistent [a-Al þ Al3Y þ Al9(Fe, Co, Ni,
treatment, no cracking is detected around Vickers hardness in- Cu)2 þ unknown phases] at larger diameters up to about 3 mm. We
dents, and the observed shear traces indicate plasticity under this further confirmed the formation of [a-Al þ am] phases without any
loading condition. other phases in the region with a diameter of about 1 mm because
there are no previous data on the formation of nanoscale mixed [a-
Al þ am] phases in cast Al-based alloy rods. Fig. 11(aed) show
4. Discussion bright-field TEM image, selected-area electron diffraction pattern,
high-resolution TEM images and nanobeam diffraction patterns
The Al84Y9Ni4Co1.5Fe0.5TM1 alloys with (TM ¼ V, Nb, Cr, Mo, Mn, taken from [a-Al] and [am] phases, respectively. The cast structure
Fe, Co, Ni, Cu or Pd [7]) show the reverse transition, and those with consists of [a-Al þ am] phases and the particle size and inter-
(TM ¼ Zr, Ag or Au) do not. The reverse transition appears to be particle spacing of [a-Al] phase are about 30 nm and 15e20 nm,
limited to the TM elements which have smaller atomic radii than Al respectively, which are considerably larger than those (about
as well as negative or nearly zero heat of mixing with the other 7e15 nm and 10e15 nm, respectively) for [a-Al] phase obtained by
constituent elements [7]. For all the studied alloys, however, AlxMy the annealing-induced reverse transition of amorphous phase. No
is a primary precipitation phase that is metastable. appreciable other phases are observed in the region, indicating
The absence of the reverse transition for (TM ¼ Zr, Ag or Au) is clearly that the coexistent [a-Al þ am] phases are formed in the
presumably due to the difficulty of phase decomposition of the 1 mm diameter region of the conical cast alloy rod, in agreement
AlxMy compound in these alloys. This may be because these ele- with the X-ray diffraction data. The EDX data in Fig. 11(e and f)
ments have atomic radii larger than those for any of the other indicate that the total solute content is about 7.3 at% for the [a-Al]
constituent elements including Al [7], as well as large positive heats phase and 20.4 at% for the [am] phase. In comparison with the EDX
of mixing with other solute elements [9], as exemplified by ZreY, data of the annealing-induced [a-Al þ am] phases shown in Fig. 5,
Ag-(Ni, Co or Fe) and Au-(Ni, Co or Fe) pairs [9]. The larger atomic the solute content of [a-Al] phase in the cast alloy is nearly the same
radii can suppress the mobility of the constituent elements thus as that for [a-Al] phase in the annealed alloy, while the solute
impeding the reverse transition, while the positive heat of mixing content in [am] phase is lower for the cast alloy, indicating that the
can also suppress the formation of a compositionally uniform AlxMy as-cast mixed phase structure is in a higher level of metastable
compound, impeding its decomposition to a-Al þ amorphous state. It is notable that metastable microstructures of [am] or [a-
phases. The ultimate decomposition of the AlxMy phase in these Al þ am] phases are formed even for the alloy rod with a diameter
three alloys occurs when the atomic mobility is sufficient at higher of about 1.2 mm, indicating that the microstructure of choice,
temperatures just below the third exothermic peak, and it proceeds namely coexistent [a-Al þ am] phases with high strength, can be
directly to the equilibrium phases (Fig. 7). obtained even in cast rod form. As shown in Fig. 10 (d), the Vickers
We examined the possibility of forming a bulk rod, fully amor- hardness is 420e430 for the amorphous single phase, increases to
phous or with [am þ a-Al] phases, for the Al84Y9Ni4Co1.5Fe0.5Cu1 about 510 for the [a-Al þ am] phases, and then decreases signifi-
alloy which can be regarded as attractive for future application. cantly to about 265 for the completely crystalline microstructure.
Fig. 10(aec) show photographs of the rod prepared by casting into a The Hv values of the [am] as well as the [a-Al þ am] phases are
mold with a conical inner cavity, and optical micrographs of regions nearly the same as those for the as-spun amorphous ribbons and
of the longitudinal cross-section of the rod. The cast microstructure the annealed ribbons with metastable [a-Al þ am00 ] microstructure.
is fully glassy for rod diameters less than about 0.82 mm and shows It is promising that the high hardness and high thermal stability
a [a-Al þ am] mixture for diameters up to about 1.2 mm. From the characteristic of the [a-Al þ am] phase mixture can be exploited
X-ray-diffraction patterns of the longitudinal cross-section, we also
Y. Jin et al. / Journal of Alloys and Compounds 832 (2020) 154997 7

Fig. 7. X-ray diffraction patterns of Al84Y9Ni4Co1.5Fe0.5TM1 (TM ¼ Zr (a), Ag (b) or Au (c)) alloys annealed for 900 s at different temperatures including the first-, second- and third-
exothermic peak temperatures.

even for the bulk Al-based structural and coating materials with reported for the a-Al þ residual amorphous phases in Al88Y2Ni9M1
high elevated-temperature strength. (M ¼ Mn or Fe) [17], Al88Y12-xNix [18] and Al88Ni9Ce2Fe1 [19]
As shown in Fig. 8, the metastable [a-Al þ am00 ] alloys exhibit amorphous alloys obtained by cooling rate control. In earlier works,
very high Hv values of 550e580 in the high annealing temperature the high strength of a-Al was attributed to the formation of su-
range of 620e680 K. The solute content is analyzed to be about persaturated solid solution nanocrystals containing a large amount
10 at% for a-Al and about 22 at% for am00 and the volume fractions of of solute elements [18] and having perfect crystal without appre-
a-Al and am00 phases are measured to be approximately 55% and ciable internal defects [19]. The present a-Al phase formed by the
45%, respectively, from the TEM images. It is important to investi- reverse transition appears very similar to that formed directly from
gate why these simple mixed-phase alloys exhibit high Hv values the supercooled liquid phase [17e19], though the annealing tem-
exceeding those for phase mixtures that include intermetallic perature is considerably higher for the present alloys. The forma-
compounds. The Hv has been reported to be about 250 for tion of the perfect crystal a-Al supersaturated solution phase as
Al88Y2Ni10-xMx (M ¼ Fe or Co) amorphous alloys [15] with similar well as the achievement of the ultrahigh strength appears to be
solute content as for the am00 phase in the present work. Applying independent of the annealing process and alloy composition. On
the rule of mixtures, the Hv of the a-Al phase is estimated to be the other hand, the Hv of the as-cast amorphous þ a-Al phase alloys
about 820 which is about 10 times the highest value reported for is as low as about 350, as shown in Fig. 10 (d). The Al phase has
conventional age-hardening Al 7075 alloys [16]. Furthermore, this much larger particle size of about 30 nm and appears to include a
Hv value is almost comparable to the Hv value (about 800) for a-Al high density of internal defects. The much lower Hv values for the
phase with a size of about 3e10 nm which has been previously as-cast alloys are presumed to be due to the much lower strength of
8 Y. Jin et al. / Journal of Alloys and Compounds 832 (2020) 154997

Fig. 8. Changes in the structure, Vickers hardness (Hv) and bending ductility with annealing temperature (Ta) for Al84Y9Ni4Co1.5Fe0.5TM1 (TM ¼ V, Nb, Cr, Mo, Mn or Cu) alloys. The
previous data of Al84Y9Ni4Co1.5Fe0.5Pd1 alloy are also shown for comparison [6].

Fig. 9. Scanning electron micrographs of the outer bent surface and optical micrographs of the slip markings around Vickers hardness indentation trace for (a, b) Al84Y9Ni4-
Co1.5Fe0.5Cr1, and (c, d) Al84Y9Ni4Co1.5Fe0.5Cu1 alloys.

Al phase containing internal defects as compared with the perfect caused by the absence of internal defects for a-Al phase as well as
crystal Al phase as well as the much lower Hv values of the lower the high thermal stability of the am00 phase, itself resulting from the
solute-content amorphous phase without distinct solute element high solute contents.
partitioning. Thus, the present high Hv values are presumed to Based on the present experimental data, schematic illustrations
originate from the combination of two factors; (1) the formation of of continuous-cooling-transformation (CCT) and continuous-
nanoscale a-Al phase without appreciable internal defects, and (2) heating-transformation (CHT) curves and the interrelation of the
the formation of a residual amorphous matrix with a solute content Gibbs free energies of the constituent phases, composition and
much higher than the nominal overall content in the alloy. In temperature for Al84Y9Ni4Co1.5Fe0.5Cu1 glassy alloy are shown in
addition, the maintenance of the metastable nanoscale [a- Fig. 12(a) and (b). On cooling, and considering the effects of larger
Al þ am00 ] phase mixture in the high and wide annealing temper- rod diameter, implying lower cooling rate, the fastest solidification
ature range of 620e680 K is also due to the low atomic diffusivity is to a single amorphous phase, then above a critical diameter of
Y. Jin et al. / Journal of Alloys and Compounds 832 (2020) 154997 9

Fig. 10. (aec) Photos and optical micrographs of a wedge-shape Al84Y9Ni4Co1.5Fe0.5Cu1 alloy rod prepared by injection-type copper mold casting, and (d) Vickers hardness taken
from the longitudinal cross section of the cast alloy rod, and (e) the relation between the as-cast structure and the wedge-shaped cone rod prepared by injection copper mold
casting.

Fig. 11. (a) Bright-field TEM image, (b) selected-area electron diffraction pattern, HRTEM images and nanobeam electron diffraction patterns taken from (c) fcc-Al and (d)
amorphous phase regions, and EDX profiles, analytical sites and compositions of (e) fcc-Al and (f) amorphous phase regions for a wedge-shape Al84Y9Ni4Co1.5Fe0.5Cu1 alloy rod
prepared by injection-type copper mold casting.

about 0.82 mm the solidification is to [a-Al þ am] and above a concluded that the reverse transition occurs only in the decom-
diameter of about 1.2 mm to [a-Al þ Al3Y þ Al9(Ni,Co,Fe)2 þ an position stages from the amorphous phase. The reverse transition
unknown phase]. Thus, no structural change corresponding to the has been interpreted on the basis of the changes in the relative free
reverse amorphization is seen on continuous cooling. It is energies with alloy composition for the various phases relevant for
10 Y. Jin et al. / Journal of Alloys and Compounds 832 (2020) 154997

Fig. 12. Schematic illustrations of (a) continuous-cooling-transformation (CCT) and continuous-heating-transformation (CHT) curves and (b) the relation among the free energy,
composition and temperature for Al84Y9Ni4Co1.5Fe0.5Cu1 glassy alloy.

the first and second exotherms on heating an 84Al amorphous alloy overall solute content is much higher than the maximum solubility
[7]. On a similar thermodynamic basis, we try to explain the limit for Al-based binary alloys. For the sequence of decomposition
heating-induced reverse transition for the Cu-containing alloy reactions, the activation energy is measured to be
where the alloy contents in the precipitate phases in the first and 252e286 kJ mol1 for the first stage, 227e247 kJ mol1 for the
the second exotherms have been measured. As depicted in Fig. 12 second stage, and 174e196 kJ mol1 for the third stage. There is a
(b), where the alloy is represented as a binary system, with all tendency for the activation energy to be higher for a TM solute with
solutes considered together, the free energy of all the phases de- a higher melting temperature and larger atomic radius. The Vickers
creases with increasing temperature, but that of the supercooled hardness Hv is 360e430 in the as-spun state, increases significantly
liquid decreases faster because of its higher entropy. The faster for the [a-Al þ AlxMy þ am’] phase mixture, and further to
lowering of the free energy of the supercooled liquid permits maximum values of 550e580 for the [a-Al þ am00 ] phase mixture,
reverse amorphization to occur in the second exotherm. In the and then decreases rapidly upon decomposition to the fully crys-
figure, the free energies of the crystalline phases are shown fixed, talline phase mixture. The [a-Al þ am00 ] phase mixture exhibits the
and relative to them the free energy of the supercooled liquid is highest Hv and is significant in developing lightweight materials
shown for the temperatures of the peaks of the first and second with high elevated-temperature strength. That the reverse
exotherms. The free energy is lowered (arrow 1 in the figure) by amorphization is seen only for some TM solutes is presumably due
eutectic crystallization to a-Al and AlxMy in the first exotherm. In to the low mobility of TM solutes with atomic radii larger than Al
the second exotherm, the free energy is further lowered (arrow 2 in and positive heats of mixing. In a conical cast rod with TM ¼ Cu, an
the figure) by the peritectic decomposition of AlxMy to a-Al and amorphous single phase is formed for diameter up to about
supercooled liquid. With further increasing temperature, sufficient 0.82 mm, while the diameter for [a-Al þ am] phases extends to
atomic mobility would permit the formation of the equilibrium about 1.2 mm. The cast [a-Al þ am] alloys exhibit a high hardness of
phases (marked 3 in the figure): Al9(Ni, Co, Fe)2, Al3Y and an un- about 510, which decreases rapidly to 265 in the fully crystalline
known phase. phase mixture. The bulk-forming ability and high thermal stability
of the metastable [a-Al þ am] phase mixture are due to the low
atomic diffusivities in the co-existing perfect crystal a-Al phase and
5. Summary high-solute-content amorphous phase.

The formation, thermal stability and phase decomposition of Declaration of competing interest
melt-spun amorphous Al84Y9Ni4Co1.5Fe0.5TM1 (TM ¼ Zr, V, Nb, Cr,
Mo, Mn, Fe, Co, Ni, Cu, Ag or Au) alloys were examined with the aim We wish to confirm that there are no known conflicts of interest
of developing a-Al þ amorphous phase mixtures with high thermal associated with this publication and there has been no significant
stability and high strength through the heating-induced reverse financial support for this work that could have influenced its
transition from [a-Al þ AlxMy þ am’] to [a-Al þ am00 ] phases in their outcome.
crystallization reactions. For the alloys with TM ¼ V, Nb, Cr, Mo, Mn,
Fe, Co, Ni or Cu, the decomposition sequence is [am] / [a-
CRediT authorship contribution statement
Al þ AlxMy þ am’] / [a-Al þ am00 ] / [a-
Al þ A3Y þ Al9TM2 þ unknown phases] and includes the reverse
Y. Jin: Methodology, Investigation, Writing - original draft,
transition. For the alloys with TM ¼ Zr, Ag or Au, the decomposition
Writing - review & editing. A. Inoue: Conceptualization, Investi-
sequence is [am] / [a-Al þ AlxMy þ am’] / [a-
gation, Writing - original draft, Writing - review & editing, Super-
Al þ Al3Y þ Al9TM2 þ unknown phases], showing no heating-
vision, Funding acquisition. F.L. Kong: Investigation, Writing -
induced reverse amorphization. For these alloys, the solutes (Zr,
review & editing. S.L. Zhu: Data curation, Supervision. F. Al-Mar-
Ag or Au) have atomic radii larger than Al and positive heats of
zouki: Formal analysis. A.L. Greer: Conceptualization, Methodol-
mixing with other constituent elements. The a-Al phase formed by
ogy, Writing - review & editing.
the reverse transition has a nearly spherical morphology with a
diameter of 5e15 nm and no appreciable internal defects. The so-
lute contents in the a-Al and am00 phases in the TM ¼ V, Cr and Cu Acknowledgments
alloys are 8.2e11.0 at% and 21.2e22.4 at%, respectively, and the Y
and Co elements are partitioned into the am00 phase. The a-Al phase The authors are grateful for support from the Recruitment
can be regarded as a metastable supersaturated solid solution, as its Program of Global Experts “1000 Talents Plan” (WQ20121200052),
Y. Jin et al. / Journal of Alloys and Compounds 832 (2020) 154997 11

the National Natural Science Foundation of China (51771131), the A.L. Greer, Novel heating-induced reversion during crystallization of Al-based
glassy alloys, Sci. Rep. 7 (2017) 46113.
Deanship of Scientific Research (DSR), King Abdulaziz University,
[8] J.C. Slater, Atomic radii in crystals, J. Chem. Phys. 41 (1964) 3199e3204.
Jeddah, Saudi Arabia (1-1-435/HiCi) and the Ministry of Education [9] A. Takeuchi, A. Inoue, Classification of bulk metallic glasses by atomic size
and Science of the Russian Federation in the framework of the difference, heat of mixing and period of constituent elements and its appli-
program aimed to increase the competitiveness of the National cation to characterization of the main alloying element, Mater. Trans. 46
(2005) 2817e2829.
University of Science and Technology, “MISiS” (No. K2-2019-002). [10] D. Bailey, The structure of two polymorphic forms of YAl3, Acta Crystallogr. 23
ALG acknowledges support from the European Research Council (1967) 729e733.
under the European Union’s Horizon 2020 research and innovation [11] A. Douglas, The structure of Co2Al9, Acta Crystallogr. 3 (1950) 19e24.
[12] H. Okamoto, Desk Handbook: Phase Diagrams for Binary Alloys, 2 ed., ASM
program (grant ERC-2015-AdG-695487: Extend Glass). International, Materials Park, Ohio, 2010.
[13] H.E. Kissinger, Reaction kinetics in differential thermal analysis, Anal. Chem.
References 29 (1957) 1702e1706.
[14] W.D. Callister Jr., D.G. Rethwisch, Fundamentals of Materials Science and
Engineering: an Integrated Approach, 4 ed., John Wiley & Sons, Hoboken, NJ,
[1] A. Inoue, Amorphous, nanoquasicrystalline and nanocrystalline alloys in Al-
2012.
based systems, Prog. Mater. Sci. 43 (1998) 365e520.
[15] A.P. Tsai, A. Inoue, T. Masumoto, Ductile Al-Ni-Zr amorphous alloys with high
[2] A. Inoue, S. Sobu, D.V. Louzguine, H. Kimura, K. Sasamori, Ultrahigh strength
mechanical strength, J. Mater. Sci. Lett. 7 (1988) 805e807.
Al-based amorphous alloys containing Sc, J. Mater. Res. 19 (2004) 1539e1543.
[16] S.W. Kim, D.Y. Kim, W.G. Kim, K.D. Woo, The study on characteristics of heat
[3] A. Inoue, A. Takeuchi, Recent development and application products of bulk
treatment of the direct squeeze cast 7075 wrought Al alloy, Mater. Sci. Eng. A-
glassy alloys, Acta Mater. 59 (2011) 2243e2267.
Struct. 304e306 (2001) 721e726.
[4] C. Fan, X.X. Yue, A. Inoue, C.T. Liu, X.P. Shen, P.K. Liaw, Recent topics on the
[17] Y.H. Kim, A. Inoue, T. Masumoto, Ultrahigh tensile strengths of Al88Y2Ni9M1
structure and crystallization of Al-based glassy alloys, Mater. Res.-Ibero-Am. J.
(M¼Mn or Fe) amorphous alloys containing finely dispersed fcc-Al particles,
22 (2019), e20180619.
Mater. Trans., JIM 31 (1990) 747e749.
[5] Y. Kim, A. Inoue, T. Masumoto, Ultrahigh tensile strengths of Al88Y2Ni9M1 (M¼
[18] Y.H. Kim, A. Inoue, T. Masumoto, Increase in mechanical strength of Al-Y-Ni
Mn or Fe) amorphous alloys containing finely dispersed fcc-Al particles,
amorphous alloys by dispersion of nanoscale fcc-Al particles, Mater. Trans.,
Mater. Trans., JIM 31 (1990) 747e749.
JIM 32 (1991) 331e338.
[6] F.F. Han, A. Inoue, Y. Han, F.L. Kong, S.L. Zhu, E. Shalaan, F. Al-Marzouki, High
[19] A. Inoue, Y. Horio, Y.H. Kim, T. Masumoto, Elevated-temperature strength of
formability of glass plus fcc-Al phases in rapidly solidified Al-based multi-
an Al88Ni9Ce2Fe1 amorphous alloy containing nanoscale fcc-Al particles,
component alloy, J. Mater. Sci. 52 (2017) 1246e1254.
Mater. Trans., JIM 33 (1992) 669e674.
[7] F.F. Han, A. Inoue, Y. Han, F.L. Kong, S.L. Zhu, E. Shalaan, F. Al-Marzouki,

You might also like