You are on page 1of 3

Materials Letters 139 (2015) 194–196

Contents lists available at ScienceDirect

Materials Letters
journal homepage: www.elsevier.com/locate/matlet

Amorphous phase decomposition in Al–Ni–RE system alloys


G. Abrosimova n, A. Aronin, A. Budchenko
Institute of Solid State Physics RAS, Chernogolovka, Russia

art ic l e i nf o a b s t r a c t

Article history: The structural changes in Al-based metallic glass at heating and deformation have been studied by X-ray
Received 15 September 2014 diffraction. It has been shown that the amorphous phase of Al–Ni–La and Al–Ni–Y alloys may decompose
Accepted 14 October 2014 at heating and deformation with formation of new amorphous phases with different chemical
Available online 23 October 2014
compositions. Both heating and deformation increase the parameters of diffusion mass transfer which
Keywords: facilitates the atomic redistribution processes.
Amorphous phase & 2014 Published by Elsevier B.V.
Decomposition
Deformation
Short range order

1. Introduction during low-temperature annealing: the initially homogeneous amor-


phous phase decomposed into regions with different component
Amorphous alloys attract particular interest as materials with concentrations which later crystallized into two solid zirconium
unusual properties different from those of crystalline alloys of solutions in iron (zirconium–iron solutions) of different concentra-
similar composition and as systems far from equilibrium. Studies tions. This produced a nanocrystalline structure in contrast to
on structure of amorphous alloys have been carried out over many annealing at higher temperatures when crystallization resulted in
years [1–10], but the renewed interest to these materials is due to significantly larger crystals. Decomposition of the amorphous phase
the fact that they are precursors for fabrication of nanocrystalline with formation and co-existence of several amorphous phases can
materials with diverse composition [11–15]. Research studies occur both before and during the crystallization process [23,24]. The
[2,16–19] have also shown that the structure formed by crystal- scale of the chemical composition change in the amorphous matrix
lization can undergo fundamental changes depending on the can be appreciably different. In the Fe–Zr system it is 10–30 nm; but,
processes or structural changes proceeding in the amorphous for instance, in Ni–Mo–B system alloys decomposition occurs on a
phase prior to crystallization. scale of more than 50 nm [19]; in Cu–Zr–Ti system alloy it is 2–5 nm
The structure of amorphous metallic alloys produced by rapid [25]; in the latter case subsequent crystallization gives rise to very
quenching is usually homogeneous. It is known, however, that fine nanocrystalline structure. Chemical decomposition of the amor-
sometimes immediately after production or as a result of some phous phase and resulting formation of regions with different short-
influence, the amorphous structure changes or becomes inhomo- range order types involve changes in the morphology and size of
geneous. For instance, formation of inhomogeneous amorphous crystallization precipitations and, in some cases, phase composition
phases (co-existence of two amorphous phases) was observed in changes [26]. Changes occurring within the amorphous state could
Pd74Au8Si18 [20], Pd40.5Ni40.5P19 [21], Be40Ti24Zr36 [22] and some significantly influence the parameters of the crystal structure form-
other amorphous alloys. Such phases have generally no distinct ing upon subsequent heating [16]. The data available indicate that
interface. As a rule, the presence of an inhomogeneous amorphous decomposition of the amorphous phase prior to the onset of
structure is established by the data of electron microscopy analysis crystallization brings about further formation of nanostructure even
or by deformation of the main diffusion maximum on X-ray in amorphous alloys that crystallize during conventional annealing
diffraction patterns. The process of structural changes of the with formation of fairly large crystals.
amorphous phase remains practically unstudied. Among the few Besides heating, deformation is another way of introducing
works one could mention [18] that reported observation of energy into the system. The most recent trends are application of
continuous decomposition of the amorphous phase in Fe–Zr alloy severe plastic deformation (SPD) and formation of novel struc-
tures. The use of the SPD method allowed formation of a
nanocrystalline structure in systems where it cannot be formed
n
Corresponding author. Tel.: þ 7 496 5224685. by heat treatment [27–29]. Therefore, of interest is research on the
E-mail address: gea@issp.ac.ru (G. Abrosimova). influence of various external actions on structural changes of the

http://dx.doi.org/10.1016/j.matlet.2014.10.076
0167-577X/& 2014 Published by Elsevier B.V.
G. Abrosimova et al. / Materials Letters 139 (2015) 194–196 195

amorphous phase. The objective of the present work was this kind
of research in the course of low-temperature annealing and
deformation of amorphous Al–Ni–RE (RE ¼La or Y) alloys.
Alloy ingots of Al87Ni8La5 and Al87Ni10Y3 composition were
prepared by arc melting as preliminarily master alloys. Amorphous
ribbons about 30 мm thick and  20 mm wide were made by the
single roller melt spinning technique in a helium atmosphere. The
samples were heated and annealed at different temperatures.
Deformation was accomplished by the multiple rolling technique
using a VEB Schwermaschinenbau four-roll laboratory mill. The
deformation value (compression) varied from 10% to 30%.
The structure of the as-prepared and annealed samples was
studied by X-ray diffraction using a SIEMENS D-500 diffractometer
with Co Kα-radiation. Special computer programs were used for
smoothing and background correction.
The as-prepared quenched alloys were amorphous. The X-ray
diffraction patterns exhibited only wide halos typical of an
amorphous phase; no reflections from crystalline phases were
observed. According to DSC data, the onset temperature of the first
crystallization stage is 228 1С; aluminum nanocrystals thereby
start forming in the amorphous phase. Isothermal annealing
within the amorphous state was carried out at 150 1С, which is
appreciably less than the crystallization onset temperature.
The structure of the amorphous phase changes during anneal-
ing, which is revealed in the change of the diffusion maximum in
the X-ray diffraction patterns. The first diffusion maximum in the
X-ray diffraction pattern of the original amorphous ribbon is
symmetrical and after annealing it shows a shoulder on the large
angle side, the degree of the maximum distortion increasing with
annealing time. Fig. 1 presents the X-ray diffraction pattern of as-
prepared sample (a) and of the sample annealed for 25 h (first
diffusion peak region) (b). It is seen that subsequent to heat
treatment the diffusion maximum is a superposition of two
maxima. The positions of the scattering maxima are known to
determine the radius of the first coordination sphere R1,
R1 ¼ 7:73=ðS1 Þmax ¼ 14:06=ðS2 Þmax ¼ 20:46=ðS3 Þmax ……;

where (S1)max (S2)max (S3)max ¼4πsin θ/λ is the wave vector corre-
sponding to the first (second, third,…) maximum of the intensity
curve, θ the scattering angle, and λ the radiation wavelength [30].
Thus, the two diffusion maxima point to the presence of amor- Fig. 1. X-ray diffraction pattern of as-prepared amorphous Al87Ni8La5 alloy (a) and
phous matrix regions with different radii of the first coordination sample annealed at 150 1С for 25 h (b); 1 denotes the experimental spectrum, 2 the
sphere. The difference in the angular positions of the diffusion summary curve, 3 and 4 lines are the diffusion halos from the first and second
amorphous phases, respectively.
maxima is indicative of the formation of amorphous regions with
different chemical compositions of the two amorphous phases.
The maximum located at the smaller angles corresponds to the
amorphous phase with a large radius of the first coordination
sphere (or the largest interatomic distance in the amorphous
phase). Since in the system in question the largest atoms are
those of lanthanum (radii RNi ¼0.124 nm, RAl ¼ 0.143 nm, RLa ¼
0.188 nm), the amorphous phase is lanthanum-enriched. Hence,
the right halo corresponds to the amorphous phase with a smaller
radius of the first coordination sphere that is nickel-enriched or/
and lanthanum-depleted. The radius of the first coordination
sphere is determined by the position of the first diffusion peak
in the X-ray diffraction pattern using the Ehrenfest equation [30]
2R1 sin θ ¼ 1:23λ;

where R1 is the radius of the first coordination sphere, 2θ the angle


of reflection, and λ the radiation wavelength. Following annealing
for 25 h the radii of the first coordination sphere of the two
amorphous phases were equal to 2.98 Å and 2.54 Å.
Fig. 2 shows the difference in the angular position of the
diffusion maxima corresponding to the two amorphous phases
as a function of annealing time of the amorphous sample. It is seen Fig. 2. Difference in position of diffusion peaks as a function of annealing time of
that at the initial stage of annealing the decomposition of the amorphous Al87Ni8La5 alloy at 1501С.
196 G. Abrosimova et al. / Materials Letters 139 (2015) 194–196

According to our estimates [32], the share of the material in


shear bands is only several percent and their contribution to
scattering is insignificant. Therefore, though, in principle, scatter-
ing from shear bands could bring about broadening or distortion of
the diffusion maximum, yet such effects are not observed owing to
the small share of the regions. Hence, as in the case of heat
treatment, the discovered change in the shape of the diffusion
maximum revealed by the X-ray diffraction patterns is conditioned
by formation of regions with a type of short-range order and
composition different from those of the matrix. Naturally this does
not exclude a different structure of the material in shear bands
but, owing to the negligible contribution to scattering, its share
can hardly be estimated.
Thus, two amorphous halos testify that the sample contains
regions of different compositions. Analysis of the atom size and
composition of the metallic glass enables the conclusion that the
two diffusion maxima correspond to yttrium (and/or nickel)-
enriched and -depleted regions. As a result, deformation of amor-
phous Al87Ni10Y3 alloy involves a distinct decomposition of the
Fig. 3. X-ray diffraction pattern of amorphous Al87Ni10Y3 alloy samples subsequent amorphous phase into regions of different chemical compositions.
to deformation; 1 denotes the experimental spectrum, 2 the summary curve,
From the results presented it follows that heating as well as
3 and 4 lines are the diffusion halos from the first and second amorphous phases,
respectively. deformation of amorphous alloys can lead to decomposition of the
amorphous phase. Both processes increase the parameters of
diffusion mass transfer which facilitates the atomic redistribution
amorphous phase proceeds more intensively (the difference in the processes. Hence, under certain conditions temperature and defor-
angular position of the peaks increases significantly), but after a mation can have similar effects on amorphous matrix structure.
lapse of time (about 12 h) the decomposition process is finished
and there are no further changes in the positions of the maxima.
Thus, isothermal annealing of amorphous Al87Ni8La5, alloy References
leads to decomposition of the amorphous phase: formation of
regions of different chemical compositions and, probably, with [1] Greer AL. Acta Metall 1992;30:171.
different types of short-range order. As mentioned above the [2] Abrosimova GE, Aronin AS, Pankratov SP, Serebryakov AV. Scr Metall
result of heat-induced decomposition of the amorphous phase 1980;14:967.
[3] Khan Y, Sostarich M. Z Metallk 1981;72:266.
was observed earlier, yet, here we report the first observation of [4] Mak A, Samwer K, Johnson WL. Phys Lett 1983;98A:353.
the decomposition process (time-related change of the coordina- [5] Mehra M, Schulz R, W.Johnson WL. J Non-Cryst Solids 1984;61-62:859.
tion sphere radii). [6] Yavari AR. Acta Metall 1988;36:1863.
[7] Abrosimova GE, Aronin AS, Asadchikov VE, Pankratov SP, Serebryakov AV. Fiz
Heat-induced decomposition of the amorphous phase was also
Metalov Metallovedenie 1986;62:496.
observed in the amorphous Al–Ni–Y system alloys. The alloy [8] Yavari AR, Osamura K, Pkuda H, Amemia Y. Phys Rev 1988;B 37:7759.
samples were subjected to plastic deformation by way of multiple [9] Abrosimova GE, Aronin AS, Ignat'eva EYu, Molokanov VV. J Magn Magn Mater
rolling. Fig. 3 presents the X-ray diffraction pattern of the amor- 1999;203:169.
[10] Cremaschi V, Arcondo B, Sirkin H, Vazquez M, Asenjo A, Garcia JM, et al. J
phous Al87Ni10Y3, alloy sample after 35% deformation. It is seen Mater Res 2000;15:1936.
that the first diffusion maximum is asymmetric and, as in the case [11] He Y, Poon JF, Shiflet GY. Science 1988;241:1640.
of heat treatment, it is a superposition of at least two maxima. [12] Inoue A, Ochiai T, Horio Y, Masumoto T. Mater Sci Eng 1994;A179–A180:649.
[13] Kim YH, Inoue A, Masumoto T. Mater Trans 1991;JIM 32:331.
As is known, plastic deformation of amorphous alloys is [14] Louzguine DV, Inoue A. J Non-Cryst Solids 2002;311:281.
achieved by formation and propagation of shear bands. According [15] Abrosimova GE, Aronin AS, Bezrukov AV, Pankratov SP, Serebryakov AV.
to the literature data [31], shear bands are oriented at an angle of Metallofisika 1982;4:69–73 (in Russian).
[16] Abrosimova GE, Aronin AS. Phys Solid State 1998;40:1603.
557 51 to the deformation direction. Another characteristic fea- [17] Aronin AS, Ivanov SA, Yakshin AE, et al. Phys Solid State 1991;33:2527.
ture of plastic deformation of amorphous alloys is increasing [18] Aronin AS, Abrosimova GE, Kir'janov YuV. Phys Solid State 2001;43:2003.
concentration of free band volume, i.e., increasing mean intera- [19] Abrosimova GE, Aronin AS, Ignat'eva EYu. Phys Solid State 2006;48:563.
[20] Chou C-P, Turnbull DJ. Non-Сryst Solids 1975;17:169.
tomic distance. Since shear bands are distinguished by disordered
[21] Chen HS. Mater Sci Eng 1976;23:151.
structure of the amorphous phase and low density regions in [22] Tanner L, Ray R. Scr Metall 1980;14:657.
which the mean interatomic distance is somewhat larger than in [23] Abrosimova GE, Aronin AS. Phys Solid State 2009;51:1765.
[24] Pospelova MM, Abrosimova GE. Des Mater Technol 2011;4:26 (in Russian).
the basic amorphous matrix, the structure of the deformed
[25] Aronin AS, Abrosimova GE, Gurov AF, Kir'janov YuV, Molokanov VV. Mater Sci
samples can be regarded as “double-phase”, the first “phase” Eng 2001;А304–306:375.
being the original amorphous alloy and the other “phase” the [26] Abrosimova GE, Aronin AS, Ignat'eva EYu, Molokanov VV. J Magn Magn Mater
material in the shear bands. Then the diffusion peaks in the X-ray 1999;203:169.
[27] Valiev RZ. Paradoxes of severe plastic deformation. In: Zehetbauer MJ, editor.
diffraction patterns make up a superposition of two diffusion Proceedings of the conference on nanomaterials by severe plastic deforma-
maxima from each “phase”. It should be remembered, however, tion–NANOSPD2; 2002. p. 109.
that the share of the material in the shear bands is not large and, [28] Aronin A, Abrosimova G, Matveev D, Rybchenko O. Rev Adv Mater Sci
2010;25:52.
hence, its contribution to scattering will be negligible. The follow- [29] Aronin AS, Abrosimova GE. Mater Lett 2012;83:183.
ing variants of structural changes are feasible: [30] Skryshevskii AF. Structure analysis of liquids and amorphous solids. Moscow:
High School; 1980.
[31] Herman Herbert, editor. Ultrarapid quenching of liquid alloys. New York:
– deformation in shear bands (plastic flow localization regions),
Academic Press; 1981. p. 375.
– deformation of the whole sample (in shear bands as well as [32] Abrosimova GE, Aronin AS, Afonikova NS, Kobelev NP. Phys Solid State
shear-free zones). 2010;52:1892.

You might also like