+ All Categories
Home > Documents > [NATO Science Series] Properties and Applications of Nanocrystalline Alloys from Amorphous...

[NATO Science Series] Properties and Applications of Nanocrystalline Alloys from Amorphous...

Date post: 05-Dec-2016
Category:
Upload: marcel
View: 213 times
Download: 0 times
Share this document with a friend
12
Al-RARE EARTH-TRANSITION METAL ALLOYS: FRAGILITY OF MELTS AND RESISTANCE TO CRYSTALLIZATION L. BATTEZZATI, M. KUSÝ, M. PALUMBO, and V. RONTO Università di Torino, Dipartimento di Chimica IFM P. Giuria 9, 10125 Torino, Italy Corresponding author: L. Battezzati, e-mail: [email protected] Abstract: The stability of Al-transition metal (TM)-rare earth (RE) alloys is considered with reference to transport and thermodynamic properties of the melt. The mobility of species or groups of atoms is described above the melting point and in the proximity of the glass transition. Viscosity is analysed at first in the frame of the strong/fragile classification of liquids. Al-based glass-formers are shown to be fragile systems. Evidences that for some molecular and metallic glass-formers the Stokes-Einstein equation does not hold below and above T g are reviewed. The alloys are, however, resistant to crystallization on rapid quenching. Some of them display also peculiar devitrification behaviour, often, but not always, with formation of primary compact nanocrystals. Analysing the transformation paths it can be inferred that mobility is belated by composition gradients in the amorphous matrix. The resistance to crystallization may result from the shape and relative position of free energy curves because crystal and liquid phases are such that the driving force for nucleation of intermetallics does not increase steadily but tends to level off on undercooling. 1. INTRODUCTION Al-based amorphous alloys occur in systems containing at least a transition metal (TM) and a rare earth (RE) [1, 2]. Contrary to the general rule for metallic glasses, they can be obtained in composition ranges where no deep eutectic exists. In most cases their composition falls off a shallow eutectic and corresponds to a field of a phase diagram where Al coexists with intermetallic compounds and there is a steep liquidus curve [3, 4]. A primary intermetallic compound starts the equilibrium solidification. Because a driving force for crystallization builds up below the liquidus temperature, it may even be surprising that these melts can undercool to the extent of forming glasses. On the other hand, rapid solidification results either in amorphisation or in the production of a two-phase material containing Al crystals in an amorphous matrix. This indicates that the melt can be undercooled well below the eutectic temperature at which a driving force for Al nucleation appears. At variance, Al crystallization may occur in a trans- 267 B. Idzikowski et al. (eds.), Properties and Applications of Nanocrystalline Alloys from Amorphous Precursors, 267–278. © 2005 Kluwer Academic Publishers. Printed in the Netherlands.
Transcript
Page 1: [NATO Science Series] Properties and Applications of Nanocrystalline Alloys from Amorphous Precursors Volume 184 || Al-Rare Earth-Transition Metal Alloys: Fragility of Melts and Resistance

Al-RARE EARTH-TRANSITION METAL ALLOYS:

FRAGILITY OF MELTS AND RESISTANCE

TO CRYSTALLIZATION

L. BATTEZZATI, M. KUSÝ, M. PALUMBO, and V. RONTO

Università di Torino, Dipartimento di Chimica IFM P. Giuria 9, 10125 Torino, Italy

Corresponding author: L. Battezzati, e-mail: [email protected]

Abstract: The stability of Al-transition metal (TM)-rare earth (RE) alloys is considered with

reference to transport and thermodynamic properties of the melt. The mobility of

species or groups of atoms is described above the melting point and in the proximity

of the glass transition. Viscosity is analysed at first in the frame of the strong/fragile

classification of liquids. Al-based glass-formers are shown to be fragile systems.

Evidences that for some molecular and metallic glass-formers the Stokes-Einstein

equation does not hold below and above Tg are reviewed. The alloys are, however,

resistant to crystallization on rapid quenching. Some of them display also peculiar

devitrification behaviour, often, but not always, with formation of primary compact

nanocrystals. Analysing the transformation paths it can be inferred that mobility is

belated by composition gradients in the amorphous matrix. The resistance to

crystallization may result from the shape and relative position of free energy curves

because crystal and liquid phases are such that the driving force for nucleation of

intermetallics does not increase steadily but tends to level off on undercooling.

1. INTRODUCTION

Al-based amorphous alloys occur in systems containing at least a transition metal

(TM) and a rare earth (RE) [1, 2]. Contrary to the general rule for metallic glasses, they

can be obtained in composition ranges where no deep eutectic exists. In most cases their

composition falls off a shallow eutectic and corresponds to a field of a phase diagram

where Al coexists with intermetallic compounds and there is a steep liquidus curve [3,

4]. A primary intermetallic compound starts the equilibrium solidification. Because a

driving force for crystallization builds up below the liquidus temperature, it may even

be surprising that these melts can undercool to the extent of forming glasses. On

the other hand, rapid solidification results either in amorphisation or in the production

of a two-phase material containing Al crystals in an amorphous matrix. This indicates

that the melt can be undercooled well below the eutectic temperature at which a driving

force for Al nucleation appears. At variance, Al crystallization may occur in a trans-

267

B. Idzikowski et al. (eds.), Properties and Applications of Nanocrystalline Alloys from Amorphous Precursors, 267–278.© 2005 Kluwer Academic Publishers. Printed in the Netherlands.

Page 2: [NATO Science Series] Properties and Applications of Nanocrystalline Alloys from Amorphous Precursors Volume 184 || Al-Rare Earth-Transition Metal Alloys: Fragility of Melts and Resistance

268 L. Battezzati et al.

formation of the glass. DSC analyses reveal that Al precipitation from glassy matrices

takes place at temperatures just above 150°C [1-4].

As for intermetallics, a contrasting behaviour is found for two groups of alloys. In

the binary Al-RE and the RE-rich ternary systems, metastable compounds with different

structures are formed on quenching from the melt and on heating the glass [5].

The possibility of nucleating various phases can represent a “confusion principle” for

the alloy helping in glass formation. The production of Al nanocrystals in spite of

a substantial driving force for nucleation of other phases, can be the outcome of homo-

geneous fluctuations in the undercooled melt or glass and be driven by the higher

mobility of Al with respect to transition and rare earth elements. For the other group of

alloys, those rich in transition metal or ternaries containing Fe, a “confusion principle”

may still operate since several intermetallic phases exist. However, the same phases are

found both during quenching, in competition with the glass, and during devitrification.

The driving force for their formation should predominate in any case as suggested by

the higher liquidus point with respect to the other group of alloys. They can then trigger

crystallization of Al as observed in the foundry practice of inoculation.

The formation of glassy or nanocrystalline phases is, therefore, the outcome of

a subtle interplay of thermodynamic [6], i.e. driving forces for transformation, and

kinetic, i.e. atomic mobility, factors. Progress in understanding these issues is reviewed

by considering transport and thermodynamic properties of the melt.

2. VISCOSITY OF UNDERCOOLED LIQUIDS.

STRONG AND FRAGILE MELTS

Non-glass forming liquids usually have = 10 3 Pa s at the equilibrium melting

point, Tm, whereas, for metallic glass-formers, is of the order of 10 1 Pa s at Tm.

During the undercooling of glass-forming liquids, viscosity shows an extraordinary

increase up to 1012 Pa s in correspondence of the glass transition, Tg. As a conse-

quence, their mobility is reduced, nucleation is retarded and the system can be vitrified

[6].

A general pattern of the viscosity of glasses and glass-forming liquids over a broad

temperature range involves three regimes. Below Tg the viscosity shows an exponential

behaviour

;exp 101

RT

E(1)

the values of Tg and (Tg) depend on the cooling rate: when it is decreased, (Tg)

increases and Tg decreases. Analogously, an exponential behaviour is found above Tm,

with different pre-exponential and activation energy parameters. Experimental data for

the viscosity of metallic melts cannot be obtained in the whole temperature range

between Tm and Tg due to the tendency of the liquid to crystallize, so the data obtained at

low and high temperature must be fitted using an appropriate function. Here, is de-

scribed by the Vogel-Fulcher-Tamman (VFT) relationship

Page 3: [NATO Science Series] Properties and Applications of Nanocrystalline Alloys from Amorphous Precursors Volume 184 || Al-Rare Earth-Transition Metal Alloys: Fragility of Melts and Resistance

Al-Rare Earth-Transition Metal Alloys 269

)(exp

,0TTBA (2)

where A and B are constants and T0, is the temperature where the viscosity would diverge because the liquid freezes to an amorphous solid.

In order to compare data relative to different systems, the viscosity is plotted as logvs. Tg/T: the so-called Angell plot (Fig. 1) [7]. The few cases found to date are those of SiO2 showing nearly Arrhenian behaviour (Eq. 1), and of o-terphenyl presenting VFT behaviour (Eq. 2). In the Angell classification of glass-forming liquids these represent limiting “strong” and “fragile” behaviour, respectively. From a structural point of view, strong liquids contain tetrahedral three-dimensional networks (e.g. SiO2, GeO2 and BeF2), that confer resistance to thermal disruption of the structure; the fragile systems on the contrary usually do not present strong directional bonding.

Figure 1. The viscosity of various glass-formers reported as a function of normal-ised temperature (Tg/T). The network glass-former SiO2 display “strong” behaviour (upper full line) whereas the molecular o-terphenyl glass-former (lower full line) is “fragile”. Alloys fall in between the two extremes being relatively fragile: Zr46.75Ti8.25Cu7.5Ni10Be22.5 (dashed line); Mg65Cu25Y10 (dotted line); Pd40Ni40P20(dot-dashed line), Al87Ni7Ce6 (full bold line)

The kinetic properties have a thermodynamic counterpart in the rate of entropy loss on undercooling which is faster when the fragility of the melt increases. Using the Kauzmann paradox which implies that the entropy of the liquid and crystal phases (having specific heat difference Cp) [8] becomes equal at temperature, T0, below the experimental Tg,

(3) 0ln0

TdCSSmT

T

pm

fragile systems are found to approach T0, the Kauzmann temperature, more rapidly than the strong ones. So, the fragile/strong behaviour of glass-formers can be quantified by the ratio Tg/T0: the higher the fragility the lower is Tg/T0. This assessment of the thermo-dynamic fragility gives similar results as the ratio Tg/T0, , with T0, the parameter enter-ing the VFT equation, representing the kinetic fragility. The resistance of the strong glass to structural changes is also shown by the small changes in heat capacity taking place at Tg (Cp(liq)/Cp(glass) 1.1), while fragile ones present higher jumps (60-80% of the glass specific heat) [9].

Page 4: [NATO Science Series] Properties and Applications of Nanocrystalline Alloys from Amorphous Precursors Volume 184 || Al-Rare Earth-Transition Metal Alloys: Fragility of Melts and Resistance

270 L. Battezzati et al.

The fragile/strong behaviour of the supercooled liquids can be also classified by means of dimensionless parameters: m and D. The former is defined as the slope of the viscosity curve reported in an Angell plot in the vicinity of Tg:

.)/()(log

gTTg TTddm (4)

Using the VFT equation, the slope of log( ) versus 1/T is given by E(Tg) = RBTg

2/(Tg T0, )2 so, the fragility at the glass transition can be quantitatively expressed as m = BTg/2.3(Tg T0, )2. D is the fragility parameter entering the modified VFT equation,

)(exp

,0

,0

TTDT

A , (4a)

where the product DT0, replaces the parameter B in the conventional expression. Strong glasses present low values of m and high values of D (e. g. m = 20 and D = 100 for SiO2), whereas the reverse occurs for fragile ones (e.g. m = 81 and D = 6.8 for o-terphenyl).

Table I. Fragility parameters m, D and Tg/T0 for different glass-formers. Metallic glasses fall in between strong (SiO2) and fragile glasses (o-terphenyl)

Glass m D Tg/T0

SiO2 20 100 –

Zr46.75Ti8.25Cu7.5Ni10Be22.5 33 22.7 1.72

Mg65Cu25Y10 45 22.1 1.58

Pd40Ni40P20 46 18.1 1.44Fe40Ni40P14B6 69 10.1 1.28Al87Ni7Ce6 127 5.6 1.15o-terphenyl 81 6.8 –

The fragility concept can be also applied to metallic melts, e.g. Al87Ni7Ce6. This alloy is taken as a representative example of glass-forming liquids based on Al-rare earth elements-transition metals for which viscosity data are not available. They can be estimated by deriving suitable parameters for the VFT equation.

The relevant transformation temperatures and enthalpies are: Tm = 911 K. This is the eutectic melting. The liquidus occurs at higher temperature, but it is not considered here because it refers to the primary formation of an intermetallic which does not compete with glass formation, Tx = 523 K, Tg = 513 K (both data collected at the heating rate of 40 K/min); Hm = 9.8 kJ/mol and Hx = 5.3 kJ/mol [10]. Assuming that the specific heat difference between the liquid and crystal phases can be expressed by

Cp = C/T 2 in the temperature range from Tm to Tx, the condition of vanishing entropy difference between the liquid and crystal phases, Eq. (3), is imposed, and the T0

Page 5: [NATO Science Series] Properties and Applications of Nanocrystalline Alloys from Amorphous Precursors Volume 184 || Al-Rare Earth-Transition Metal Alloys: Fragility of Melts and Resistance

Al-Rare Earth-Transition Metal Alloys 271

temperature is found at 444 K. Cp at Tg is found as high as 21 J/mol K indicating thermodynamic fragility.

The pre-exponential factor A is taken as NAh/Vm = 3.8 10 5 Pa s, with NA the Avogadro constant, h the Planck constant and Vm the molar volume. Using the Andrade formula for the viscosity at the melting point [11]

3/2m

maCm V

TwAT , (5)

where AC is the proportionality coefficient and wa the average atomic weight, and assigning the AC coefficient the average value 6.5 10 7 (JK 1mol1/3)1/2 for metallic glass-formers at the eutectic temperature, we obtain: (Tm) = 7.8 10 3 Pa s. Finally (Tg) is taken as 1.0 1012 Pa s. With the latter two positions, the quantities B and T0. are obtained as B = 2474 K and T0. = 447 K, remarkably close to T0. The viscosity is then expressed as = 3.8 10 5 exp [2472/(T 447)] Pa s. The fragility parameters m and Dturn out to be 127 and 5.6 respectively, indicating that the ternary Al87Ni7Ce6 is a fragile melt. Although there are differences in the glass-forming tendency when substituting other transition metals and rare earth elements for Ni and Ce or when changing the alloy composition, the present result is indicative of the behaviour of this family of alloys. The above expression (5) for viscosity provides values in agreement with the experi-mental ones obtained above the liquidus for a number of Al-Ni-RE alloys [12].

The viscosity parameters of Al87Ni7Ce6 are compared with literature findings for some glassy substances and alloys in Tab. I. The general strong/fragile scheme is followed; note that strong melts need a critical cooling rate much lower than fragile ones to be vitrified [6]. The undercooled melts scale in strength in the order Zr46.75Ti8.25Cu7.5Ni10Be22.5> Mg65Cu25Y10> Pd40Ni40P20 >> Al87Ni7Ce6.

3. THE RELATIONSHIP BETWEEN VISCOSITY AND DIFFUSIVITY

When diffusion and viscous flow occur with the same brownian mechanism, the Stokes-Einstein equation (SE) is employed:

rkTD6

, (6)

where k is the Boltzmann constant and r is an ionic radius. The SE equation is valid at high temperatures near the melting point, but the correlation may fail when the liquid approaches the glass transition [6, 13].

Illustrative data are displayed in Fig. 2 for molecular glass formers: a clear break-down of the SE relationship can be noticed in correspondence of about 1.3 Tg.The measured self-diffusion coefficients are larger than those calculated by means of the SE equation using the experimental shear viscosity values. In correspondence to the SE breakdown de-coupling takes place between the translational diffusion and viscosity on the one hand, and rotational diffusion and viscosity on the other. In fact, while at high temperatures both Drot and Dtrans are inversely proportional to the viscosity, for high undercooling this relationship is lost; near Tg the molecules translate faster (two

Page 6: [NATO Science Series] Properties and Applications of Nanocrystalline Alloys from Amorphous Precursors Volume 184 || Al-Rare Earth-Transition Metal Alloys: Fragility of Melts and Resistance

272 L. Battezzati et al.

orders of magnitude) than suggested by the SE equation. Figure 3 also shows data for a

metallic alloy above Tg in the limited range where they could be collected. It is to be

remarked that the points for Fe40Ni40P14B6 refer to viscosity and an apparent diffusion

coefficient derived from the measurements of the rate of crystal growth, i.e. of the size

of eutectic colonies as a function of time at various temperatures. The points appear to

show a trend analogous to that of molecular glasses.

Figure 2. A plot showing the breakdown of the Stokes-Einstein equation for various glass-

formers. The logarithm of the quantity ( D/T) is plotted vs. a normalised temperature

(T/Tg). The points referring to salol (+) and o-terphenyl ( ) (translational diffusion

coefficients) conform to SE behaviour for T/Tg > 1.3. The points referring to o-terphenyl

( ) (rotational diffusion coefficients) conform to SE immediately above Tg. The points

referring to the Fe40Ni40P14B6 alloy (diffusion coefficients derived from the rate of crystal

growth) deviate from SE behaviour up to T/Tg = 1.17

The recent discovery of bulk metallic glasses having high thermal resistance to

crystallization allows us to verify the -D relationship in a larger temperature range,

above Tg. In the case of the Zr46.7Ti8.3Cu7.5Ni10Be27.5 bulk amorphous alloy, the SE

equation appears to lose its validity, since there is a discrepancy between the ex-

perimental translational diffusion coefficient of various elements and the one computed

via the SE equation. This is shown in Fig. 3 where the experimental trend of diffusion

coefficient of various elements [14] is shown together with that obtained by

means of the SE equation. The discrepancy is apparent. The -D behaviour found for

the Zr46.7Ti8.3Cu7.5Ni10Be27.5 amorphous alloy appears to be very similar to that of many

molecular glass-formers.

The failure of the Stokes-Einstein law near 1.2-1.3 Tg can be explained by assuming

a change in the diffusion mechanism. It has been speculated that the mechanism for

mass transport passes from a co-operative diffusion process, typical of viscous flow, to

activated “hopping” transitions (jump diffusion) involving a few molecular units.

The jump diffusion occurring at high undercooling resembles the vacancy diffusion in

crystals and does not contribute to viscous flow, so the measured diffusion coefficients

are higher than those calculated using the SE equation.

Page 7: [NATO Science Series] Properties and Applications of Nanocrystalline Alloys from Amorphous Precursors Volume 184 || Al-Rare Earth-Transition Metal Alloys: Fragility of Melts and Resistance

Al-Rare Earth-Transition Metal Alloys 273

Figure 3. Arrhenius plot giving the

tracer diffusion coefficient of various

elements in Zr46.7Ti8.3Cu7.5Ni10Be27.5

(solid lines) and a diffusion coef-

ficient computed as if the SE equation

was valid (dashed line). There is no

apparent correlation of diffusivity of

alloying or other elements, such as Al

with viscosity in this temperature

range

Contrary to molecular substances where the mobility behaviour has been studied

and modelled by investigating the liquid state, for metallic alloys the breakdown of the

relationship between diffusion and viscosity was discovered and discussed at first for

glasses below the glass transition. This is an outcome of studies on structural relaxation

of amorphous alloys where free volume is lost on annealing via defect annihilation, the

material increases its density and the ideal glassy state is approached. Measurements of

the equilibrium viscosity and the diffusion coefficient of Fe and Au tracer atoms in

Fe40Ni40B20 and Pd40Ni40P20 relaxed glasses, respectively, showed that at constant

temperature the D product is not constant [15]. It was proposed that the viscosity and

the diffusivity do not follow the same mechanism and it was assumed that the atomic

transport during a viscous flow involves a pair of defects, whereas, in the case of

diffusion, only one defect is required; formally this leads to

(7).const2D

It was also verified that this correlation holds for a molecular glass-former above Tg.

With the new findings on bulk metallic glasses illustrated in Fig. 3 it can be suggested

that this correlation is general. The above considerations would turn useful in estimating

diffusion coefficients from viscosity in the neighbourhood of Tg if the significance

and value of the above product D2 were elucidated. At present it has been ascertained

that the product in Eq. (7) is constant at constant temperature, but it changes with

temperature and diffusion couple studied. If various diffusing species in the same alloy,

Zr46.7Ti8.3Cu7.5Ni10Be27.5, are considered, then the product scales roughly with the atomic

mass of the diffusing atom.

4. CRYSTALLIZATION AND MOBILITY

If D is derived for an alloy from Eq. (7), its activation energy would be compatible

with that found in the diffusion and growth studies. Crystallization is very frequently

studied by thermal analysis and activation energies are easily obtained by means of non

isothermal scans. These values are often difficult to rationalise unless they can be

clearly referred to nucleation and growth processes. In most experiments a limited

temperature range is spanned, so the data can be fitted with an Arrhenius equation. If

Page 8: [NATO Science Series] Properties and Applications of Nanocrystalline Alloys from Amorphous Precursors Volume 184 || Al-Rare Earth-Transition Metal Alloys: Fragility of Melts and Resistance

274 L. Battezzati et al.

above Tg the mobility were determined by viscous flow, the activation energy would

approximate the curvature of the VFT function in the temperature range of interest.

Therefore, it should become the higher the closer the temperature range is to Tg. As an

example, the apparent activation energy for viscous flow in Al87Ni7Ce6, calculated from

the example in paragraph 2, just above Tg is of the order of 640 kJ/mol. Such high

activation energy would certainly be rate limiting for all processes involving atomic

mobility, but values of this order have not been found in any crystallization or growth

processes for this and related alloys. At the present stage Al-transition metal-rare earth

alloys the activation energies reported so far are widely scattered. They range from

140-160 kJ/mol for some Al-Ni-Gd glasses to values in excess of 400 kJ/mol for

Al86Ni5Co2Y8 according to their composition and heat treatment. Considering homo-

geneous nucleation, the activation barrier can be estimated of the order of 40-50 RT at

deep undercooling, again lower than the above value. In most instances nucleation is

heterogeneous, so the barrier will be even lower with respect to the value for SE viscous

flow [16].

In a series of experiments with Al87Ni7Ce6 and Al87Ni7Nd6 amorphous alloys [10], it

has been recently found that Tg becomes manifest at heating rates above a certain value

typical of each alloy when the onset of crystallization is displaced to sufficiently high

temperatures. The kinetics of the transformation was then studied and the Kissinger

plots showed a change in the mechanism as a function of the heating rate, i.e. of peak

temperature as implied by a deviation from linearity in the points for the first DSC

peaks. The Kissinger plots could be referred to definite processes. At low rates, the

crystallization must precede or overlap Tg so the activation energies of 340-370 kJ/mol

for the two alloys should be related to the motion of Al in a medium characterised by

short range order of different species, and therefore by the need of cooperative

displacements of the alloy constituents. Above Tg, the viscous flow in the highly

undercooled melt apparently has no effect since lower values of activation energies,

140-190 kJ/mol, are obtained. This suggests that the need for cooperative diffusion is

removed in the molten state and that the hopping of single Al atoms suffice for

promoting crystallization. It is apparent that estimating diffusivities via the SE equation

is not appropriate for these materials. Extending the work to an Al87Ni7La6 alloy

confirms these findings [17].

Concluding, a mobility different from that expressed by the SE relationship is rele-

vant for crystallization of Al based glasses. The SE supplant (cf. Eq. (7)) may be opera-

tive or crystallization may depend on single diffusion steps not related to viscous flow.

5. FORMATION OF PRIMARY NANOCRYSTALS

In most alloys where primary nanocrystals are obtained from an amorphous matrix

the transformation is reported to be fast and then to slow down progressively. This is

reflected in highly asymmetric peaks in DSC. The glass transition is seldom detected

before the onset of precipitation. It should be remarked that the primary phase is often,

but not always Al. In fact, there have been findings in various laboratories that

metastable intermetallics form as primary phases in Al-Ni-La and Al-Ni-Ce alloys [10,

18-21]. Examples of DSC traces are shown in Fig. 4 for Al87Ni7La6 and Al87Ni7La5Ti1

Page 9: [NATO Science Series] Properties and Applications of Nanocrystalline Alloys from Amorphous Precursors Volume 184 || Al-Rare Earth-Transition Metal Alloys: Fragility of Melts and Resistance

Al-Rare Earth-Transition Metal Alloys 275

Figure 4. DSC traces for the alloys listed

in the insert at the heating rate of 20 K min 1

amorphous alloys. The first crystallization step of the ternary alloy is due to the pre-

cipitation of a complex cubic phase, rich in the rare earth element. Substituting some Ti

for La causes destabilization of such a phase and precipitation is then due to fcc Al.

Nevertheless, both peaks are skewed and a substantial background signal occurs on their

high temperature side indicating a similar mechanism of transformation. It has been

shown that the background signal does not involve major precipitation of new crystals

but rather the coarsening of those already existing, and homogenization of the remain-

ing matrix where compositional gradients should occur because of solute rejection by

the precipitates. It has been shown earlier that rare earth atoms can accumulate ahead of

the nanocrystalline Al interface whereas transition metals diffuse faster in the matrix so

their concentration is more readily homogenized. The new finding that a homogeniza-

tion process occurs also when a primary La-rich phase forms, implies that a gradient in

the Ni content of the matrix will retard its transformation as well. So, the two phase

materials are resistant to crystallization because compositional gradients are established.

It has been also suggested that during primary crystallization stress builds up around

precipitates because of compositional gradients. This would be analogous to the stress

effect occurring in multilayered thin film made of early and late transition metals where

the interdiffusivity depends on the length scale of modulation, being fast for short dis-

tances, comparable to precipitate size, and dropping steadily for long range motion as

implied in homogenization [22].

6. NUCLEATION THERMODYNAMICS

Precipitation from the amorphous Al87Ni7Ce6 matrix involves only Al at low rates,

i.e. below Tg, but it implies the formation of a metastable intermetallic, besides Al,

when the transformation occurs above Tg. Therefore there is a clear change in

the transformation mechanism across the glass transition. It is likely that below Tg only

the growth of Al crystals occurs on pre-existing seeds whereas, in addition to Al

growth, the two phases nucleate and grow in the undercooled melt. In Al87Ni7Nd6

the first transformation does not imply the formation of an intermetallic compound at

any rate, so the change in the mechanism can be more clearly attributed to the formation

of Al by either a growth in a glassy matrix or nucleation and a growth in a liquid matrix

[10]. In addition, the tendency of the melt to undercooling shows that the nucleation of

Page 10: [NATO Science Series] Properties and Applications of Nanocrystalline Alloys from Amorphous Precursors Volume 184 || Al-Rare Earth-Transition Metal Alloys: Fragility of Melts and Resistance

276 L. Battezzati et al.

intermetallic compounds which is possible at high temperature, just below the liquidus,

is sluggish.

Figure 5. The driving force for

nucleation computed by means of

the common tangent construction

for fcc-Al and Al11Nd3 in the al-

loy Al91Nd9

The nucleation scenario in these alloys can be described thermodynamically by

means of free energy curves representing the driving force for nucleation of either Al or

Al11Nd3 computed according to the parallel tangent construction as a function of tem-

perature (Fig. 5). The free energy of the liquid and crystal phases was optimized by

means of a CALPHAD software and an appropriate account was made of the excess

specific heat which is specific of glass-forming melts as well as of the occurrence of

a glass transition [23]. At the glass-forming composition of 9 at.% Nd, the driving force

for nucleation of the intermetallic builds up below the liquidus but levels off at high

undercooling because of the progressive stabilization of the liquid phase due to the

excess specific heat. Meanwhile, the driving force for nucleation of Al increases

steadily and becomes close to that of Al11Nd3 at temperatures close to the glass

transition. The two phases can actually compete for nucleation and factors such as glass

(liquid)-crystal interfacial tension can become decisive in phase selection. It is expected

that the simple fcc structure of Al will involve a lower interfacial tension with respect to

the complex Al11Nd3 compound. On the other hand, metastable compounds may display

short range order closer to that of the glass (liquid) and, therefore, nucleate preferen-

tially.

7. CONCLUSIONS

Al-transition metal-rare earth melts have been suggested to be fragile in the Al-rich

composition range where amorphisation can occur by rapid quenching. This is mostly

determined by the low value of the ideal glass transition temperature, T0, corresponding

to the point of vanishing entropy difference between liquid and crystal phases which

was estimated from thermodynamic data. Although there is limited information on

the viscosity of such alloys, referring to temperature ranges above the liquidus, the T0,

parameter entering the VFT for viscosity can be recovered. This stems from the as-

sumption that the viscosity behaviour at the eutectic and glass transition temperature

should be typical of glass forming alloys. The T0, is close to the thermodynamic T0 and

Page 11: [NATO Science Series] Properties and Applications of Nanocrystalline Alloys from Amorphous Precursors Volume 184 || Al-Rare Earth-Transition Metal Alloys: Fragility of Melts and Resistance

Al-Rare Earth-Transition Metal Alloys 277

the overall viscosity in the undercooling regime conforms to the fragile type in Angell

classification.

Evidences have then been reviewed of the relationship between viscosity and dif-

fusivity. At high temperature the proportionality expressed by the Stokes-Einstein

equation appears valid within the scatter of experimental data but it breaks down close

to Tg. In fact, the data obtained on the diffusion and the rate of crystal growth (including

activation energies for crystallization) show that mobility is due to either single hopping

jumps (vacancy-like mechanism) or the movement of groups of atoms (cooperative

mechanism) instead of viscous flow.

In Al-Ni-RE amorphous alloys, primary crystallization occurs frequently involving

either fcc Al or an intermetallic compound. Such two-phase materials are rather stable

kinetically in that they resist crystallization in a large temperature range. This has been

attributed to sluggishness in nucleation of stable and metastable intermetallics

depending on the existence of compositional gradients in the amorphous matrix which

is enriched in the slow diffusing elements, and on the lowering of the driving force for

nucleation due to the shape of the liquid free energy curve.

ACKNOWLEDGMENTS

This work is funded by the Research Training Network of the European Commission “Nano

Al” HPRN-CT-2000-00038.

REFERENCES

1. Inoue. A., (1998) Amorphous, nanoquasicrystalline and nanocrystalline alloys in Al-based

systems, Progr. Mater. Sci., 43, 365-520.

2. Battezzati, L., Pozzovivo, S., and Rizzi, P., (2003) Nanocrystalline Aluminium Alloys, in:

H.S. Nalwa (ed.) Nanoclusters and Nanocrystals, American Scientific Publishers, 25650

North Lewis Way, Stevenson Ranch, California 91381-1439, USA.

3. Guo, F.Q., Poon, S.J., and Shiflet, G.J., (2000) Glass formability in Al-based multinary

alloys, Mater. Sci. Forum, 331-337, 31-42.

4. Perepezko J.H., Hebert R.J., and Tong W.S., (2002) Amorphisation and nanostructure

synthesis in Al alloys, Intermetallics, 10, 1079-1088.

5. Rizzi, P., Baricco, M., Borace, S., and Battezzati, L., (2001) Phase selection in Al-TM-RE

alloys: nanocrystalline Al versus intermetallics, Mater. Sci. Eng. A, 304-306, 574-578.

6. Battezzati, L. and Castellero, A., (2002) Nucleation and the Properties of undercooled

Melts, M. Magini and F. H. Wohlbier (eds.), Materials Science Foundations, Trans Tech

Publications Ltd., Uetikon-Zurich, Switzerland, 2002, vol. 15, 1-80.

7. Angell, C.A., (1991) Relaxation in liquids, polymers and plastic crystals – strong/fragile

patterns and problems, J. non-crystalline Solids, 131-133, 13-31.

8. Debenedetti and P.G., Stillinger, F.H., (2001) Supercooled liquids and the glass transition,

Nature, 410, 257-267.

9. Battezzati, L., (1999) Metallic Glasses, T. M. Letcher (ed.), in Chemical Thermodynamics, a

“Chemistry for the 21st Century” monograph, Blackwell Science, 239-249.

10. Battezzati, L., Pozzovivo, and S., Rizzi, P., (2002) Phase Transformations in Al87Ni7Ce6

and Al87Ni7Nd6 amorphous alloys, Mater. Trans., 43, 2593-2599.

Page 12: [NATO Science Series] Properties and Applications of Nanocrystalline Alloys from Amorphous Precursors Volume 184 || Al-Rare Earth-Transition Metal Alloys: Fragility of Melts and Resistance

278 L. Battezzati et al.

11. Battezzati, L. and Greer, A.L., (1989) On the viscosity of liquid metals and alloys, Acta

Metall., 37, 1791-1802.

12. Yamasaki, T., Kanatani, S., Ogino, Y. and Inoue, A., (1993) Viscosity measurements for liquid

Al-Ni-La and Al-Ni-Mm (Mm: Mischmetal) alloys by an oscillating crucible method, J. Non-

cryst. Solids, 156-158, 441-444.

13. Chang, I., Sillescu, H., (1997) Heterogeneity at the glass transition: translational and rotational

self-diffusion, J. Phys. Chem. B, 101, 8794-8801.

14. Zumkley, T., Naundorf, V. Macht, M.-P., (2000) Diffusion in the Zr46.75Ti8.25Cu7.5Ni10Be27.5

bulk glass: on the diffusion mechanism in supercooled liquids, Z. Metallkd., 91, 901-908.

15. Duine, P.A., Sietsma, J., van der Beukel, A., (1993) Atomic transport in amorphous

Pd40Ni40P20 near the glass transition temperature: Au diffusivity and viscosity, Phys. Rev. B 48,

6957-6965.

16. Battezzati, L., Interplay of process kinetics in the undercooled melt in the proximity of

the glass transition, Mater. Sci. Eng. A (in press)

17. Battezzati, L., Rizzi, P., Ronto, V., The difference in devitrification paths in Al87Ni7Sm6

and Al87Ni7La6 amorphous alloys, Mater. Sci. Eng. A, (in press)

18. Zhuang, Y.X, Jiang, J.Z., Lin, Z.G., Mezour M., Crichton, W. and Inoue, A., (2001) Evi-

dence of eutectic crystallization and transient nucleation in Al89La6Ni5 amorphous alloy,

Appl. Phys. Lett., 79, 743-745.

19. Ye, F. and Lu, K., (1999) Pressure effect on crystallization kinetics of an Al-La-Ni amor-

phous alloy, Acta mater., 47, 2449-2454.

20. Gangopadhyay, A.K. and Kelton, K.F., (2000) Effect of rare-earth atomic radius on

the devitrification of Al88RE8Ni4 amorphous alloys, Phil. Mag. B, 80, 1193-1206.

21. Abrosimova, G. E., Aronin, A. S., Zver'kova I. I., Kir'yanov, Y. V., (2002) Phase trans-

formations upon crystallization of amorphous Al-Ni-RE alloys, Phys. Met. Metallov., 94,

113-118.

22. Karpe, N., Krog J. P., Bottiger, J., Chechenin, N. G., Somekh, R. E., Greer, A. L., The ther-

modynamic factor in interdiffusion – A strong effect in amorphous Ni-Zr, Acta Met.

Mater., 43, 551-558.

23. Baricco, M., Gaertner, F., Rizzi, P., Battezzati, L., Cacciamani, G., Greer, A.L., (1998)

Thermodynamics of homogeneous crystal nucleation in Al-RE metallic glasses, Mater. Sci.

Forum, 269-272, 553-558.


Recommended