Вы находитесь на странице: 1из 9

Progress in Natural Science: Materials International 29 (2019) 32–40

HOSTED BY Contents lists available at ScienceDirect

Progress in Natural Science: Materials International


journal homepage: www.elsevier.com/locate/pnsmi

Original Research

Effect of sintering temperature on phase evolution of Al86Ni6Y4.5Co2La1.5 T


bulk amorphous composites synthesized via mechanical alloying and spark
plasma sintering
Ashutosh Sahua, Ram Sajeevan Mauryab, Tapas Lahaa,

a
Department of Metallurgical and Materials Engineering, Indian Institute of Technology, Kharagpur 721302, India
b
Department of Metallurgical and Materials Engineering, National Institute of Technology, Rourkela 769008, India

ARTICLE INFO ABSTRACT

Keywords: Al86Ni6Y4.5Co2La1.5 amorphous powders were synthesized by mechanical alloying for 200 h. Subsequent con-
Al based glassy alloy solidation was performed via spark plasma sintering in the temperature range of 250 °C to 500 °C at the pressure
Mechanical alloying of 500 MPa. The role of viscous flow on densification was investigated by studying the viscosity change of the
Spark plasma sintering amorphous phase at different consolidation temperatures. The decrease in viscosity at higher sintering tem-
Intermetallic phase
peratures resulted in better particle bonding and densification of consolidated samples. The formation of only
Viscosity
FCC Al was observed in the consolidated samples at sintering temperatures ≤ 300 °C and the intermetallic
phases formed at temperatures ≥ 400 °C. The mechanical properties of the bulk samples were measured by
Vickers microhardness and nanoindentation tests. The testing results showed that the average values of mi-
crohardness, nanohardness and elastic modulus of the sample consolidated at 500 °C were 3.06 ± 0.14 GPa,
4.85 ± 1.14 GPa and 89.53 ± 9.25 GPa, respectively. The increase in hardness and elastic modulus of the
higher temperature consolidated samples is attributed to the improvement in particle bonding, densification and
distribution of various hard intermetallic phases in the amorphous matrix.

1. Introduction Al-Ni-Y-Co-La alloy compositions by partially replacing Ni and Y with


Co and La and reported improvement in thickness of glassy rods from
In recent years there has been remarkable progress in fabrication of 450 µm (Al86Ni8Y6) to 1 mm (Al-Ni-Y-Co-La) synthesized via copper
amorphous metallic alloys/bulk metallic glasses (BMGs), which is at- mold casting technique route. However, the dimensions of these Al
tributed to more than twice tensile and compressive strength along with based metallic glasses are still not large enough for structural applica-
high fracture and corrosion resistance compared to their crystalline tions.
counterparts [1]. The enhanced mechanical, chemical and physical The problem of dimensional limitation could be overcome by
properties of BMGs are due to the absence of long range periodicity of adopting powder metallurgy route, which has limited phase diagram
atoms and various crystalline defects such as dislocations, grain restrictions and offers amorphization of wide range of compositions.
boundaries and antiphase boundaries [1]. In this context, Al-based Out of several powder metallurgical processing techniques, mechanical
BMGs are reported to have very high specific strength compared to alloying is simple and cost-effective to produce amorphous powders.
other metal based BMGs and could be potentially used in automobile Further, consolidation of amorphous powder via spark plasma sintering
and aerospace industries. However, the requirement of high cooling (SPS) technique could produce large dimension Al-based BMGs as SPS
rates (105–106 k/s) limits the thickness of Al based metallic glasses in offers very fast heating rates (even up to 1000 °C/min) with simulta-
the lower range (1 mm, max.) attributed to off-eutectic nature and low neous application of high uniaxial pressure and DC pulse current [7].
reduced glass transition temperature (Trg=Tg/Tl) of the Al-based glass Generation of plasma waves, spark impact pressure and application of
forming compositions [2,3]. To optimize the composition, Ma et al. high consolidation pressure during SPS along with removal of surface
proposed Al-Ni-Y (Al86Ni9Y5) as good glass forming alloys based on oxide layer from the powder particles, result in better inter-particle
efficient cluster packing (ECP) model [4,5]. Yang et al. [6] developed bonding in very less time [7,8]. Thus, high density samples could be

Peer review under responsibility of Chinese Materials Research Society.



Corresponding author.
E-mail address: laha@metal.iitkgp.ac.in (T. Laha).

https://doi.org/10.1016/j.pnsc.2019.01.009
Received 9 June 2018; Received in revised form 14 January 2019; Accepted 18 January 2019
Available online 14 February 2019
1002-0071/ © 2019 Chinese Materials Research Society. Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/BY-NC-ND/4.0/).
A. Sahu, et al. Progress in Natural Science: Materials International 29 (2019) 32–40

consolidated from amorphous powders in few minutes of sintering via restrict cold welding of powders during milling. Additionally, the ball
SPS without causing much devitrification. High heating rates help in milling was given 15 min break after every 30 min of rotation to avoid
removing particle surface oxide layers, retaining amorphous phase and any possibility of heat accumulation in the vial. The entire powder
decreasing activation energy of viscous flow in sintering [9,10]. Li et al. handling process was carried out inside a glovebox (Labstar, M-Braun
have reported higher retention of amorphous phase in Inertgas - systeme GmbH, Germany) under high purity Ar atmosphere
Al86Ni6Y4.5Co2La1.5 bulk alloys by increasing the heating rate from (< 20 ppm of O2 level) to avoid air contamination.
5 °C/min to 40 °C/min in SPS consolidation of gas atomized amorphous Mechanically alloyed amorphous powders were consolidated via
powders [9]. Liu et al. have reported decrease in activation energy for SPS (SPS 625, Fuji Electronic Industrial Co. Ltd., Japan) using tungsten
viscous flow from 83.1 KJ/mol at 20 K/min to 57.5 KJ/mol at 140 K/ carbide die (10 mm internal diameter) and punch (10 mm diameter)
min heating rate in SPS consolidation of Ti40.6Zr9.4Cu37.5Ni9.4Al3.1 set. A hydraulic press was used for pressing the powder in the die before
amorphous powder [10]. placing inside the SPS chamber. High uniaxial pressure (500 MPa) and
SPS parameters (viz. pressure, temperature and time) play decisive heating rate (100 °C/min) were applied to simultaneously achieve
role in densification mechanism, microstructural phase evolution and better densification and amorphous phase retention. The consolidation
mechanical properties of consolidated amorphous alloys [8,11–14]. temperatures were varied from 250 °C to 500 °C based on the transition
Higher consolidation pressure assists in deforming the particles, im- temperatures obtained from thermal analysis of the amorphous pow-
proving viscous flow of amorphous phase, preserving higher amount of ders. Holding time was kept as 15 min. Vacuum pressure of 26 Pa was
amorphous phase and removing oxide layers from particle surface created inside SPS chamber during consolidation to avoid oxidation of
during sintering [8,11,12]. Li et al. reported improvement in sinter powders during sintering. SPS instrument generated punch displace-
density of Al86Ni6Y4.5Co2La1.5 bulk alloys from 2.85 gm/cm3 to ment data with respect to temperature and time was used to analyse the
3.32 gm/cm3 by increasing consolidation pressure from 200 MPa to sintering behaviour of the mechanically alloyed powders.
600 MPa, attributed to removal of brittle surface oxide layers and better
viscous flow during sintering [8]. Similarly, Paul et al. have also re- 2.2. Characterization
ported improvement in densification from 80% to 98% of theoretical
density of SPS consolidated Fe48Cr15Mo14Y2C15B6 bulk alloys by in- X-ray diffraction (Bruker D8 Advance diffractometer, Germany) was
creasing consolidation pressure from 20 MPa to 70 MPa, attributed to performed using CuKα (λ = 1.54 Å) radiation to study the phase evo-
better viscous flow and deformation of powder particles [11]. Present lution in the various interval milled powders and various temperature
authors have already reported for Al86Ni8Y6 amorphous alloy that high consolidated bulk samples. Scanning electron microscope (Carl Zeiss,
SPS consolidation pressure could be very much effective in improving Merlin FE-SEM) was employed to study the morphology of various in-
relative density along with retaining higher volume fraction of amor- terval milled powders and consolidation behaviour of the consolidated
phous phase [12]. On the other hand, higher sintering temperature bulk alloys. Image analysis was conducted on 10–12 SEM images with
leads to increase in viscous flow and better atomic diffusion during the software (Leica MW Image Analysis System with Leica Application
sintering leading to higher density bulk alloys at the expense of partial Suite - V3.8, Leica Microsystem GmbH, Germany) to estimate the
devitrification of amorphous phase [13,14]. Sasaki et al. observed an average particle size of different interval milled powders. High re-
increase in relative density from 92.2% to 99.9% in Al85Ni10La5 solution transmission electron microscopy (TEM, JEM-2100 LaB6,
amorphous alloys by increasing consolidation temperature from 473 K 200 kV, JEOL USA, Inc.) was carried out to validate the crystalline and
to 753 K, even after decreasing the consolidation pressure and sintering amorphous nature of the milled powders and the bulk samples.
time [13]. Similarly, Deng et al. found an increase in density from Ultrasonically dispersed milled powders were placed on a Cu grid to
3.02 gm/cm3 to 3.34 gm/cm3 by increasing consolidation temperature perform TEM of the powders. The TEM samples of the sintered products
from 250 °C to 400 °C during consolidation of Al86Ni7Y4.5La1.5Co1 par- were prepared using conventional method of metallographic polishing,
tially amorphous powders [14]. dimpling and ion milling. Differential scanning calorimetry (DSC 404
In the present research work Al86Ni6Y4.5Co2La1.5 (at%) alloy com- F1 Pegasus, NETZSCH GmbH, Germany) was performed at a heating
position was chosen, attributed to the better glass forming ability (GFA) rate of 10 °C/min to find out the various phase transitions in the
via rapid solidification route of this composition compared to the other amorphous powder.
Al-based glass forming compositions [6,15]. An attempt has been made Mechanical properties of the sintered samples were evaluated by
to fully amorphize Al86Ni6Y4.5Co2La1.5 (at%) powders via mechanical performing (i) Vickers microhardness (UHL VMHT-001, Walter Uhl,
alloying. Consolidation of the amorphous powders was carried out by Germany) test at various loads of 100 gf, 300 gf, 500 gf and 1000 gf and
SPS at varying temperatures at a constant pressure of 500 MPa to obtain (ii) nanoindentation (Triboindenter TI 950, Hysitron Inc., USA) test at
high density bulk amorphous alloys. The effect of consolidation tem- 5000 µN load using a standard Berkovich tip (TI-0039, Hysitron Inc.,
perature on viscous flow of mechanically alloyed Al86Ni6Y4.5Co2La1.5 USA) with total included angle of 142.3° and tip radius of 100 nm. The
amorphous powders during sintering was studied in this work. loading, holding and unloading times were 10 s each.

2. Experimental procedure 3. Results and discussion

2.1. Synthesis of amorphous powders and bulk samples 3.1. Microstructure characterization of the milled powders

Elemental Al (99.5%, −44 µm), Ni (99.996%, −125 µm), Y (99.9%, SEM images and the corresponding powder size distribution graphs
−420 µm), Co (99.5%, −44 µm) and La (99.9%, −44 µm) powders shown in Fig. 1 represent the morphology and size distribution of 1 h,
were procured from Alfa Aesar and Strem Chemicals (MA, USA) and 50 h, 100 h, 150 h and 200 h mechanically alloyed Al86Ni6Y4.5Co2La1.5
used to prepare a powder blend of Al86Ni6Y4.5Co2La1.5 (at%). The (at%). The average particle sizes were found to be 18.28 ± 2.34 µm,
powder blend was subjected to mechanical alloying in a high energy 3.09 ± 1.2 µm, 2.35 ± 0.84 µm, 1.69 ± 0.64 µm and
planetary ball mill (Retsch GmbH, Germany, Model: PM200) using 0.94 ± 0.34 µm, respectively, as estimated by quantitative image
hardened steel vials (125 ml capacity) and balls (10 mm diameter). Ball analysis performed on 10 field of views for each samples. Comparing
to powder weight ratio (BPR) and disc rotation speed were kept at 15:1 the particle size of 1 h milled powders with that of the 50 h milled
and 300 RPM, respectively. Toluene (80 ml) was used for wet grinding powders, drastic (~85%) reduction in particle size was observed in 50 h
to prevent agglomeration, air contamination and heat generation. ones, whereas there was only ~24% reduction in particle size with
Stearic acid (1 wt%) was used as process controlling agent (PCA) to further increase in milling time from 50 h to 100 h. This is also observed

33
A. Sahu, et al. Progress in Natural Science: Materials International 29 (2019) 32–40

Fig. 1. SEM images of (a) 1 h, (c) 50 h, (e) 100 h, (g) 150 h and (i) 200 h milled Al86Ni6Y4.5Co2La1.5 powder, and (b), (d), (f), (h) and (j) their corresponding
histograms showing particle size distribution in the milled powders.

from the corresponding particle size distribution plots that maximum in dislocation density, work hardening and grain refinement which
number of particles lie in the range of 15–20 µm in the 1 h milled slowed down the process of reduction in particle size. Thus, there was
powders, which shifted to the range of 2–3 µm in case of 50 h milled lesser degree of decrement in particle size in the milling time range of
powders. Initially powder particles were of larger size and ductile in 50–200 h. The size distribution plots corresponding to 100 h, 150 h and
nature and could absorb higher amount of impact energy during milling 200 h milled powders show the shift of particle size range (where the
leading to flattening of these particles. Flat particle surfaces possess maximum number of particles lies) from 3 to 4 µm to 1–1.5 µm and
higher tendency to get layered, agglomerated and cold welded when 0.5–1 µm, respectively. At a certain stage of milling (200 h in the pre-
come in contact during milling, attributed to higher surface area sent study), various defects viz. dislocations, voids, anti-phase
leading to faster diffusion [16]. In this process of mechanical de- boundary etc. along with the lattice strain reached to a critical value
formation, new defects viz. dislocations, grain boundaries and anti- and destabilize the atomic periodicity resulting in amorphous structure
phase boundaries are generated and gradually make the particles brittle [16]. The process of powder amorphization is discussed in detail with
through work hardening [16]. Accordingly, repeated flattening, cold the help of XRD results in the following section.
welding, fracturing and fragmentation during milling led to saturation Fig. 2 shows the XRD graphs of 1 h, 30 h, 50 h, 100 h and 200 h

Fig. 2. (a) XRD patterns showing various phase evolution and process of amorphization in 1 h, 30 h, 50 h, 100 h and 200 h of milled Al86Ni6Y4.5Co2La1.5 powders and
(b) Exploded view of the amorphous hump observed in the XRD pattern of 200 h milled powders.

34
A. Sahu, et al. Progress in Natural Science: Materials International 29 (2019) 32–40

milled powders. XRD peaks corresponding to all the elements could be


clearly observed in the pattern of 1 h milled powders. XRD pattern of
30 h milled powders revealed a slight peak broadening and intensity
reduction along with evolution of some new peaks at 2θ angle of 29.66°
and 55.66° compared to 1 h milled powders. These new peaks indicated
formation of intermetallic phases such as Ni2Y3 and La7Ni3. This could
be attributed to the repeated flattening and layering of particles due to
cold welding which led to diffusion of atoms across the layers leading to
formation of complex intermetallic phases [17]. However, the XRD
peaks related to these intermetallic phases disappeared and peak in-
tensity reduced when the milling time was increased to or beyond 50 h.
Formation and disappearance of such complex metastable phases
during mechanical alloying is a common phenomenon and has also
been reported by several other researchers including the present au-
thors [16–18]. Progressive deformation continuously decreased the
crystallite size, increased lattice strain and strain energy, and thus im-
parted nanocrystallinity in the powders leading to more peak broad-
ening which is clear from the XRD pattern of 100 h milled powders.
Continuous deformation of powder particles increased strain energy of Fig. 4. DSC thermogram of the 200 h mechanically alloyed amorphous powders
the system by introducing various defects and thus finally destabilized showing various phase transitions at different temperatures upon heating.
the atomic periodicity leading to formation of amorphous phase after
200 h of milling which was confirmed by the presence of a broad hump
powders and some other Al-based amorphous systems [19–21].
in the corresponding XRD pattern. This was further confirmed by the
TEM analysis performed on 200 h milled powders, the results of which
are shown in Fig. 3. Featureless TEM image without any crystalline 3.2. Consolidation behaviour of mechanically alloyed amorphous powders
phases (Fig. 3a) and the corresponding diffused ring in the selected area
diffraction (SAD) pattern (Fig. 3b) along with absence of long range The punch displacement data with respect to the sintering tem-
atomic periodicity in the high resolution TEM image (Fig. 3c) collec- perature and time obtained during the spark plasma sintering was
tively corroborate the completion of amorphization of analysed to understand the sintering behaviour of the mechanically
Al86Ni6Y4.5Co2La1.5 alloy powders milled for 200 h. alloyed powders. The instantaneous relative density at a certain tem-
Fig. 4 shows DSC thermogram of the mechanically alloyed perature was calculated using the following equation.
Al86Ni6Y4.5Co2La1.5 amorphous powders. The endothermic event pre- h0
ceded by an exothermic event indicates the glass transition temperature = 0
ht (1)
(Tg ) at 210 °C in the DSC thermogram, which was followed by onset of
the first crystallization event at 258 °C (Tx1) as shown in the figure. where, 0 and h 0 are the green density and height of the compacted
Thus, a wider supercooled liquid region (Tx1 Tg = 48 °C) is obtained powder respectively, after applying 500 MPa pressure at room tem-
compared to various other Al-based amorphous systems synthesized by perature. The instantaneous height of the compact during sintering is
both rapid solidification and mechanical alloying methods, exhibiting given by ht .
the supercooled liquid region varying in the range of 10–39 °C [19–22]. The change in SPS punch displacement and the corresponding in-
Three exothermic peaks in the present DSC curve reveal three stages of stantaneous relative density at four different consolidation tempera-
devitrification. First crystallization event (Tx1) indicated partial forma- tures (viz. 250 °C, 300 °C, 400 °C and 500 °C) during sintering of the
tion of nanocrystalline FCC Al. Second crystallization event starts at mechanically alloyed amorphous powders has been shown in Fig. 5a.
385 °C (Tx2 ) and corresponds to the growth of the nanocrystalline FCC The inclined segment of the curves represents the punch displacement
Al formed in earlier stage. The third stage of crystallization which starts and the corresponding instantaneous relative density during tempera-
at 426 °C (Tx3 ), corresponds to the transformation of residual amorphous ture ramp and the vertical section at the end for each sintering tem-
phase, growth of nanocrystalline FCC Al phase along with the formation perature represents the same during the holding time of 15 min. The
of some intermetallic phases. Similar, devitrification behaviour has punch displacement and relative density increased almost linearly
been reported in gas atomized Al86Ni6Y4.5Co2La1.5 amorphous alloy during temperature ramp from room temperature to the sintering
temperatures.
Different phases evolved in the amorphous powder during tem-
perature ramp and holding time, can affect the viscous flow of amor-
phous phase close to or above Tg and sintering mechanism. Thus, to
study the sintering mechanism at various sintering temperatures it is
essential to understand the phase evolution during temperature rise and
at the initial stage of holding. Hence, XRD analysis was carried out on
the bulk sample consolidated at 500 °C for 3 min, whereas the holding
time during normal sintering was 15 min. The XRD pattern (Fig. 5b)
indicates that the amorphous phase was mostly retained after 3 min of
sintering except formation of only trace amount of nanocrystalline FCC
Al. However, formation of various intermetallic phases has been ob-
served in 15 min sintered samples (discussed earlier in Section 3.3).
Thus, it could be envisaged that viscous flow of the amorphous powder
has played crucial role in densification without noticeable devitrifica-
Fig. 3. (a) TEM micrograph, (b) corresponding SAED pattern exhibiting dif- tion during the period of temperature rise and at the initial stage of
fused ring and (c) absence of long range periodicity in HRTEM image revealing holding at all sintering temperatures.
the amorphous nature of 200 h milled Al86Ni6Y4.5Co2La1.5 powders. Viscous flow of atoms at various sintering temperatures can be

35
A. Sahu, et al. Progress in Natural Science: Materials International 29 (2019) 32–40

Fig. 5. (a) Variation in punch displacement and the corresponding instantaneous relative density with temperature during spark plasma sintering, (b) XRD pattern of
an SPS consolidated sample sintered with 3 min holding time and (c) variation of linear shrinkage with time at different sintering temperatures at the initial stage
(first 100 s) of holding.

explained with the help of Frenkel's equation [10,23], which gives the =
3
(5)
4Dm
relation between linear shrinkage and viscosity as shown in the fol-
lowing equation. Eq. (5) clearly shows the inverse relation between the viscosity and
h slopes of the linear shrinkage curves. Viscosity at 250 °C, 300 °C, 400 °C
3
= t and 500 °C were calculated to be 10.12 × 1010 Pa S, 7.6 × 109 Pa S,
h0 4D (2)
4.34 × 109 Pa S and 3.4 × 109 Pa S, respectively. The decrease in
where, Δh/h0 is the linear shrinkage, γ is the surface energy, D is the viscosity with increase in sintering temperature led to increase in vis-
average particle size, η is the viscosity which is constant at a certain cous flow. Thus, the linear shrinkage increased at higher temperatures
temperature and t is the time. The value of γ was considered as 1.5 J/m2 (Fig. 5c), which assisted in decreasing the porosity, and thus in im-
adopted in this study, as the surface energy of most of the metals vary in proving the densification.
the range of 1–2 J/m2 [24]. Assuming γ, D and η are constants at a fixed Liu et al. [10] have reported two stage densification mechanism
sintering temperature, the equation reduces to the form during SPS of Ti-based amorphous powders. The first stage was at-
tributed to the viscous flow in the supercooled liquid region, whereas
y = kt (3) the second stage was dominated by volume diffusion in various crys-
3 talline phases which evolved above the supercooled liquid region.
where, y = h and k = 4D .
h
0 However, in the present work, the amorphous phase was mostly re-
Therefore, at two instantaneous sintering times t1 and t2, k can be tained during the temperature ramp and at the initial stage of holding.
written as, Thus, only a single stage sintering mechanism caused by viscous flow in
y2 y1 the amorphous phase predominated. Here it should be mentioned that
k= various intermetallic phases were formed during 15 min of hold at
t2 t1 (4)
sintering temperatures ≥ 400 °C (discussed in Section 3.3), which
The values of 2 1 can be estimated from the slopes, ‘m’ of the
y y
should have affected the densification process. However, the viscous
t2 t1
linear shrinkage vs time curves at different temperatures at the initial sintering model was applied here to understand the contribution of
stage of sintering (first 100 s) as shown in Fig. 5c. Thus, viscosity can be viscous flow in the densification process. Retention of amorphous phase
expressed as: during temperature ramp and at initial holding period at high

36
A. Sahu, et al. Progress in Natural Science: Materials International 29 (2019) 32–40

Table 1
Absolute density, relative density, nanohardness and elastic modulus of various temperature SPS consolidated bulk samples.
Sintering Density Relative density Relative density Hmax (GPa) Hmin Havg (GPa) Emax (GPa) Emin Eavg (GPa)
temperature (°C) (gm/cc) calculated using Eq. (1) measured using (GPa) (GPa)
(%) Archimedes principle (%)

250 2.69 82 84 3.26 0.91 2.06 ± 0.68 52.65 34.22 49.16 ± 10.78
300 2.79 85 87 3.49 1.6 2.38 ± 0.46 62.42 42.53 51.68 ± 5.17
400 2.92 89 92 5.3 1.85 3.08 ± 0.77 86.2 47.33 63.6 ± 9.38
500 3.08 94 95 7.66 2.7 4.85 ± 1.08 115.4 77.34 86.16 ± 9.25

consolidation temperatures can be attributed to the effect of high


consolidation pressure and heating rate, which is discussed in the fol-
lowing section.
High consolidation pressure generates high stress concentration at
the particle asperities and causes fracturing and fragmentation of brittle
amorphous particles, thus provides better surface contacts [8]. Com-
bined effect of high consolidation pressure and temperature ensured
better mass flow of atoms leading to improved inter-particle bonding
and higher relative density [25]. Table 1 lists the densities and relative
densities of different temperature sintered samples. Relative densities
calculated by Eq. (1) were 82% and 94% for 250 °C and 500 °C con-
solidated alloys, respectively. The relative densities were also verified
by Archimedes’ principle and the values obtained are comparable to
that obtained from Eq. (1).
Back scattered electron (BSE) images revealing the particle bonding
in the microstructures of the bulk alloys are shown in Fig. 6. The vo-
lume fraction and size of porosities are very high in 250 °C consolidated
alloy as can be observed in Fig. 6a. The amounts of the same have
decreased and the particle bonding improved in the higher temperature Fig. 7. XRD patterns of the samples consolidated from mechanically alloyed
consolidated alloys (Fig. 6b, c and d). This is in support with the re- amorphous powders at 250 °C, 300 °C, 400 °C and 500 °C showing partial de-
lative density results obtained by Eq. (1) and Archimedes’ principle, as vitrification during spark plasma sintering.
discussed earlier.
consolidated at ≤ 300 °C show only minor amounts of nanocrystalline
FCC Al in an amorphous matrix, confirmed by the XRD peak at ~45°
3.3. Microstructural analysis of consolidated samples
overlaying the broad amorphous hump in 2θ range of 30–50°. The XRD
peak intensity related to nanocrystalline FCC Al increased with increase
Fig. 7 shows the XRD patterns of the bulk samples consolidated at
in sintering temperature indicating formation of more amount of FCC
250 °C, 300 °C, 400 °C and 500 °C. The XRD patterns of the samples

Fig. 6. Back scattered electron images showing inter-particle bonding and various precipitated phases (marked by yellow arrows) in samples consolidated from
mechanically alloyed amorphous powders at (a) 250 °C, (b) 300 °C, (c) 400 °C and (d) 500 °C, respectively keeping the sintering time fixed as 15 min.

37
A. Sahu, et al. Progress in Natural Science: Materials International 29 (2019) 32–40

where, vamor is the volume fraction of the retained amorphous phase in


the sintered samples. Aamor and Acryst are the areas under the curves of
amorphous and crystalline phase peaks in the XRD patterns, respec-
tively. The estimated retained amorphous phase in 250 °C, 300 °C,
400 °C, 500 °C – sintered samples were 98%, 95%, 91%, 80%, ap-
proximately respectively.
Suppression of devitrification during temperature ramp up to 500 °C
and 3 min sintering time (Fig. 5b) as discussed in the previous section
(Section 3.2) can be attributed to the lower sintering time (3 min) in
that particular case and application of very high pressure (500 MPa)
[8,9]. The present authors have reported earlier about the retention of
high volume fraction of amorphous phase during SPS consolidation of
Al86Ni8Y6 amorphous powder at higher consolidation pressure attrib-
uted to suppression of long range diffusion [12]. Jiang et al. [26] and
Jiang et al. [27] have also reported that high pressure restricted long
range diffusion, increased crystallization temperature and thus higher
extent of thermal activation was required for crystallization in
Zr41.2Ti13.8Cu12.5Ni10Be22.5 and Fe72P11C6Al5B4Ga2 metallic glasses.
Besides, high heating rate reduces the ramping time to reach the pre-
decided sintering temperature and assist in suppressing devitrification.
Li et al. have reported complete retention of amorphous phase at higher
heating rate (40 °C/min) compared to formation of FCC Al at lower
heating rate (5 °C/min) during SPS consolidation of gas atomized
Al86Ni6Y4.5Co2La1.5 amorphous powder [9]. Thus, the combined effect
of high consolidation pressure and steep heating rate (100 °C/min) re-
tained the amorphous phase predominantly and only a trace amount of
FCC Al formed with 3 min hold (Fig. 5b), whereas complete devi-
trification of the amorphous phase occurred during temperature up to
500 °C rise at comparatively a very slow heating rate of 10 °C/min as
seen earlier in the DSC results (Fig. 4).
Fig. 8. TEM analysis of sample consolidated at 500 °C revealing (a) distribution
The intermetallic phases were formed at consolidation temperatures
of various crystalline phases in the amorphous matrix, (b) crystalline diffraction
≥ 400 °C during longer duration (15 min) holding as seen earlier in the
spots overlaying the amorphous diffused rings in the SAED pattern and (c)
lattice fringes of Al13Co4 corresponding to (212) plane in the amorphous ma- corresponding XRD pattern (Fig. 7). Devitrification of amorphous phase
trix. during SPS is affected by three primary factors viz. (i) consolidation
temperature, (ii) consolidation pressure and (iii) holding time. Crys-
talline phase formation during sintering at different consolidation
Al. The sample consolidated at 400 °C exhibited peaks related to in-
temperatures can be explained by Stokes-Einstein equation [28,29]. At
termetallic phases in 2θ range of 30–50°, 65° and 75° along with the
low temperatures (ideally below Tg ) viscosity of the amorphous phase is
nanocrystalline FCC Al and amorphous phase. When the sintering
very high and diffusion of the constituent atoms are restricted. How-
temperature was increased to 500 °C, XRD peak intensity of the inter-
ever, the smaller constituent elements show hopping like diffusion be-
metallic phases such as Al4Ni3, Al3Ni5, La2Ni3, Y3Co2, NiY3 and Al13Co4
haviour and get decoupled. Yu et al. [28] have reported hopping like
increased exhibiting more amount of intermetallic phase formation
diffusion behaviour of P and Be atoms at low temperatures in Pd- and
compared to 400 °C - sintered sample. Formation of these crystalline
Zr-based metallic glasses with the help of NMR technique. In the pre-
phases in the sintered alloys is also observed in the corresponding back
sent study, Al atoms were decoupled and accumulated to form nano-
scattered electron (BSE) images as white spots (marked by yellow ar-
crystalline FCC Al at sintering temperatures ≤ 300 °C. At higher tem-
rows) in Fig. 6. These white spots are absent in the BSE images of the
peratures (ideally above Tg ), viscosity of the amorphous phase becomes
samples sintered at ≤ 300 °C (Fig. 6a, b). Only trace amounts of in-
low and atomic mobility increases. The diffusion kinetics becomes
termetallic phases are observed in the 400 °C - sintered sample (Fig. 6c).
faster where twin-like coupling mechanism dominates, and the cou-
Amount of the intermetallic phase formation increased at 500 °C and
pling of different radii elements results in formation of complex crys-
the uniform distribution of these phases can be observed in the BSE
talline phases [29]. Thus, various intermetallic phases formed at sin-
image provided in Fig. 6d. Further TEM analysis revealed the dis-
tering temperatures ≥ 400 °C. In addition, the localized high
tribution of different crystalline phases in amorphous matrix of 500 °C -
temperature zones created during SPS also influence the formation of
sintered sample (Fig. 8). It is observed from Fig. 8a that very fine
intermetallic phases. The intense joule heating due to high current
(5–50 nm) nanocrystalline phases are distributed in the amorphous
density at particle surface asperities lids to formation of localized high
matrix. Corresponding SAED pattern (Fig. 8b) possesses bright diffrac-
temperature zones depending on particle morphology and area of
tion spots along with the diffused rings confirming the distribution of
contact [7,30] and various complex intermetallic phases may form at
nanocrystalline phases in amorphous matrix. The HRTEM image
these high temperature zones. The number of such high temperature
(Fig. 8c) of the selected area of Fig. 8a shows both the periodic and
zones increases at higher sintering temperatures due to increase in DC
random arrangement of atoms, related to nanocrystalline Al13Co4 pre-
pulse current. Thus, the amount of intermetallic phases also increased
cipitate (corresponding to the d212 plane with a spacing of 3.4 Å) and
at higher consolidation temperatures. Deng et al. have reported for-
amorphous phase, respectively.
mation of nanocrystalline phases such as FCC Al, Al4NiY and La7Ni3
The retained amorphous phase fraction in the sintered samples can
during SPS of Al86Ni7Y4.5La1.5Co1 partially amorphous powders, at-
be found by the following formula.
tributed to rise in local temperature [14]. In the previous studies of
present authors, formation of various intermetallic phases during SPS
Aamor consolidation of Al86Ni8Y6 and AL86Ni6Y6Co2 amorphous powders at
vamor =
Aamor + Acryst (6) higher sintering temperatures was also found [18,31]. In addition to the

38
A. Sahu, et al. Progress in Natural Science: Materials International 29 (2019) 32–40

Fig. 9. Variation in Vickers microhardness values of SPS consolidated samples


sintered at varying temperatures and tested at 100 gf, 300 gf, 500 gf and 1000 gf
Fig. 10. Variation in nanohardness and elastic modulus values obtained by
loads.
carrying out nanoindentation at peak load of 5000 µN on samples, sintered at
different temperatures, resulting from the nanocrystalline precipitate dis-
above factors, the presence of high defect density in the mechanically tributed amorphous matrix.
alloyed powder reduces the activation energy for mass transport during
sintering and promotes devitrification. (H = 0.151 G), where ‘H’ and ‘G’ are the Vickers hardness and shear
High consolidation pressure also affects intermetallic phase forma- modulus, respectively [37]. Thus, the observed variation of nanohard-
tion during SPS, where long distance atomic movement get suppressed ness and elastic modulus from maximum to minimum values can be
under high pressure [26,27]. However, due to short range atomic dif- attributed to the presence of hard intermetallic nano-precipitates and
fusion atoms get paired up according to Miracle's topological criteria comparatively soft nanocrystalline FCC Al in the harder amorphous
[32]. Thus, the combined effect of localized high temperature zones in matrix. The calculated H/E ratio from nanoindentation data points lies
high temperature sintering (≥ 400 °C) and the short-range ordering in the range of 0.026–0.066. This signifies the presence of amorphous
under high consolidation pressure during longer holding time (15 min) matrix along with crystalline phases as this ratio for most of the bulk
resulted in the formation of intermetallic phases. metallic glasses lies near ~0.05 and increases with increasing crystal-
line phase fraction [38].
3.4. Mechanical property evaluation of the consolidated alloys The observed decrease in microhardness with increasing load and
the difference in hardness values between microhardness and nano-
Fig. 9 shows the variation in Vickers microhardness of the SPS hardness can be ascribed to indentation size effect (ISE) which is a well-
consolidated samples tested at 100 gf, 300 gf, 500 gf and 1000 gf loads. known phenomenon in crystalline materials [39]. ISE is also reported to
The samples sintered at 250 °C, 300 °C, 400 °C and 500 °C exhibited be observed in metallic amorphous materials [40]. Jang et al. carried
microhardness values of 2.03 ± 0.12 GPa, 2.12 ± 0.16 GPa, out a series of nanoindentation experiments with different indenters
2.35 ± 0.17 GPa and 3.06 ± 0.14 GPa, respectively, at 100 gf load. and reported that glassy alloys undergo strain softening rather than
The increase in hardness for higher temperature consolidated samples is strain hardening during deformation attributed to shear localized zone
mainly attributed to (i) the better inter-particle bonding, (ii) higher [40,41]. The rearrangement of atoms accommodates local shear strain
relative density, (iii) retained amorphous phase and (iv) uniformly which leads to formation of shear transformation zones (STZs) and
distributed hard intermetallic phases in the amorphous matrix. Sasaki strain induced softening during deformation of glassy alloys. At lower
et al. and Deng et al. also reported improvement in hardness of higher loads, comparatively lesser number of STZs are formed, which reduces
temperature sintered bulk alloys consolidated from Al-based amor- the amount of free volume involvement and hardness increases and that
phous/partially amorphous gas atomized powders, attributed to higher is the reason ISE was observed during microhardness testing of the
relative density and nanocrystalline phase formation [13,14]. spark plasma sintered samples in the present study.
Fig. 10 represents the variation of nanoindentation hardness and
elastic modulus of the samples consolidated at various sintering tem- 4. Conclusions
peratures. The maximum, minimum and average values of nanohard-
ness and elastic modulus of the 250 °C, 300 °C, 400 °C and 500 °C con- Summarily, this research work gives a detailed description of the
solidated alloys are listed in Table 1. The observed improvement in synthesis of mechanically alloyed Al86Ni6Y4.5Co2La1.5 amorphous
nanohardness of bulk alloys with increasing consolidation temperature powders and the governing mechanisms in consolidating the powders
can be ascribed to the same reasons as described earlier while ex- via spark plasma sintering. Mechanical alloying for 200 h results in
plaining the increase in microhardness values. The variation of nano- complete amorphization of Al86Ni6Y4.5Co2La1.5 alloy system with an
hardness and elastic modulus with indentation depth (Fig. 10) can be average particle size of 0.94 ± 0.34 µm. The amorphous powders ex-
explained as follows. hibit glass transition and the first crystallization temperatures at 210 °C
Hardness measured by nanoindentation test of bulk nanocrystalline and 258 °C, respectively with a supercooled liquid region of 48 °C. The
Al samples varies in the range of 2–3.5 GPa [33–35]. Guo et al. have DSC curve reveals three-stage crystallization behaviour of the amor-
[36] reported nanohardness of Al86Ni6Y4.5Co2La1.5 melt spun amor- phous powder during heating, corresponding to the formation of FCC
phous ribbon and copper mold cast BMG are ~4.6 GPa and ~5 GPa Al, grain growth, devitrification of residual amorphous phase and the
respectively. The hardness of various Al-Ni, Al-Y, Al-Co and Al-La based formation of various intermetallic phases. The amorphous phase is
intermetallic compounds lie in the range of 9–13 GPa, 5–10 GPa, mostly retained during temperature ramp and at the initial stage of
10–17 GPa and 3–8 GPa, respectively, calculated by intrinsic correla- sintering. The viscosities calculated at 250 °C, 300 °C, 400 °C and 500 °C
tion between hardness and shear modulus of polycrystalline materials are 10.12 × 1010 Pa S, 7.6 × 109 Pa S, 4.34 × 109 Pa S and 3.4 × 109

39
A. Sahu, et al. Progress in Natural Science: Materials International 29 (2019) 32–40

Pa S, respectively. The formation of only FCC Al has been observed at [12] R.S. Maurya, A. Sahu, T. Laha, Mater. Sci. Eng. A 649 (2016) 48–56.
sintering temperatures ≤ 300 °C. Intermetallic phases form at tem- [13] T.T. Sasaki, K. Hono, J. Vierke, M. Wollgarten, J. Banhart, Mater. Sci. Eng. A 490
(2008) 343–350.
peratures ≥ 400 °C. Relative density of bulk sample consolidated at [14] S.S. Deng, D.J. Wang, Q. Luo, Y.J. Huang, J. Shen, Adv. Powder Technol. 26 (2015)
500 °C is ~94%. Higher relative density, better particle bonding and 1696–1701.
formation of hard intermetallic phases in the amorphous matrix of [15] N.C. Wu, L. Zuo, J.Q. Wang, E. Ma, Acta Mater. 108 (2016) 143–151.
[16] C. Suryanarayana, Prog. Mater. Sci. 46 (2001) 1–184.
500 °C consolidated sample result in average Vickers microhardness [17] B.S. Murty, S. Ranganathan, Int. Mater. Rev. 43 (1998) 101–141.
and nanohardness of 3.06 ± 0.14 GPa and 4.45 ± 1.08 GPa, respec- [18] R.S. Maurya, A. Sahu, T. Laha, Mater. Des. 93 (2016) 96–103.
tively with elastic modulus of 89.53 ± 9.25 GPa. The increase in [19] B.J. Yang, J.H. Yao, Y.S. Chao, J.Q. Wang, E. Ma, Philos. Mag. 90 (2010)
3215–3231.
hardness value with the decrease in indentation load is ascribed to the [20] H.W. Yang, W.P. Tong, X. Zhao, L. Zuo, J.Q. Wang, J. Alloy. Compd. 473 (2009)
indentation size effect. 347–350.
[21] X.P. Li, M. Yan, B.J. Yang, J.Q. Wang, G.B. Schaffer, M. Qian, Mater. Sci. Eng. A 530
(2011) 432–439.
Acknowledgement
[22] R.S. Maurya, T. Laha, J. Mater. Sci. Technol. 31 (2015) 1118–1124.
[23] V.V. Dabhade, T.R. Rama Mohan, P. Ramakrishnan, Mater. Res. Bull. 42 (2007)
The author, T. Laha acknowledges the financial support obtained 1262–1268.
from the Science and Engineering Research Board, Department of [24] L. Vitos, A.V. Ruban, H.L. Skriver, J. Kollar, Surf. Sci. 411 (1998) 186–202.
[25] B. Zheng, D. Ashford, Y. Zhou, S.N. Mathaudhu, J.P. Delplanque, E.J. Lavernia, Acta
Science & Technology, Government of India (SB/S3/ME/0044/2013) Mater. 61 (2013) 4414–4428.
and Sponsored Research and Industrial Consultancy, Indian Institute of [26] J.Z. Jiang, T.J. Zhou, H. Rasmussen, U. Kuhn, J. Eckert, C. Lathe, Appl. Phys. Lett.
Technology Kharagpur, India (GAF). 77 (2000) 3553–3555.
[27] J.Z. Jiang, J.S. Olsen, L. Gerward, S. Abdali, J. Eckert, N.S. Boer, L. Schultz,
J. Truckenbrodt, P.X. Shi, J. Appl. Phys. 87 (2000) 2664–2666.
References [28] H.B. Yu, K. Samwer, Y. Wu, W.H. Wang, Phys. Rev. Lett. 109 (2012) 1–5.
[29] V. Zollmer, K. Ratzke, F. Faupel, Phys. Rev. Lett. 90 (2003) 1–4.
[30] S. Nowak, L. Perriere, L. Dembinski, S.T. Nenez, Y. Champion, J. Alloy. Compd. 509
[1] W.H. Wang, C. Dong, C.H. Shek, Mater. Sci. Eng. R 44 (2004) 45–89.
(2011) 1011–1019.
[2] F.Q. Guo, S.J. Poon, G.J. Shiflet, Mater. Sci. Forum 331–337 (2000) 31–42.
[31] R.S. Maurya, A. Sahu, T. Laha, J. Non-Cryst. Solids 453 (2016) 1–7.
[3] H. Yang, J.Q. Wang, Y. Li, Philos. Mag. 87 (2007) 4211–4228.
[32] D.B. Miracle, O.N. Senkov, J. Non-Cryst. Solids 319 (2003) 174–191.
[4] C.S. Ma, J. Zhang, W.L. Hou, X.C. Chang, J.Q. Wang, Philos. Mag. Lett. 88 (2008)
[33] Z.F. Liu, Z.H. Zhang, J.F. Lu, A.V. Korznikov, E. Korznikova, F.C. Wang, Mater. Des.
599–605.
64 (2014) 625–630.
[5] D.B. Miracle, Acta Mater. 54 (2006) 4317–4336.
[34] S. Varam, K.V. Rajulapati, K.B.S. Rao, J. Alloy. Compd. 585 (2014) 795–799.
[6] B.J. Yang, J.H. Yao, J. Zhang, H.W. Yang, J.Q. Wang, E. Ma, Scr. Mater. 61 (2009)
[35] X.K. Sun, H.T. Cong, M. Sun, M.C. Yang, Metall. Mater. Trans. A 31A (2000)
423–426.
1017–1024.
[7] O. Guillon, J. Gonzalez-Julian, B. Dargatz, T. Kessel, G. Schierning, J. Rathel,
[36] H. Guo, C. Jiang, B. Yang, J. Wang, J. Mater. Sci. Technol. 33 (2017) 1272–1277.
M. Herrmann, Adv. Eng. Mater. 16 (2014) 830–849.
[37] X.Q. Chen, H. Niu, D. Li, Y. Li, Intermetallics 19 (2011) 1275–1281.
[8] X.P. Li, M. Yan, G. Ji, M. Qian, J. Nanomater. 2013 (2013) 1–6.
[38] S. Vincent, B.S. Murty, M.J. Kramer, J. Bhatt, Mater. Des. 65 (2015) 98–103.
[9] X.P. Li, M. Yan, H. Imai, K. Kondoh, G.B. Schaffer, M. Qian, J. Non-Cryst. Solids 375
[39] W.D. Nix, H. Gao, J. Mech. Phys. Solids 46 (1998) 411–425.
(2013) 95–98.
[40] J. Jang, B.G. Yoo, Y.J. Kim, J.H. Oh, I.C. Choi, H. Bei, Scr. Mater. 64 (2011)
[10] L.H. Liu, C. Yang, Y.G. Yao, F. Wang, W.W. Zhang, Y. Long, Y.Y. Li, Intermetallics
753–756.
66 (2015) 1–7.
[41] S. Nachum, A.L. Greer, J. Alloy. Compd. 615 (2014) S98–S101.
[11] T. Paul, N. Chawake, R.S. Kottada, S.P. Harimkar, J. Alloy. Compd. 738 (2018)
10–15.

40

Вам также может понравиться