Sie sind auf Seite 1von 162

MECHANICAL ALLOYING

i
ii
MECHANICAL ALLOYING
Fundamentals and Applications

P.R. Soni

Department of Metallurgical Engineering


Malaviya Regional Engineering College, Jaipur, India

CAMBRIDGE INTERNATIONAL SCIENCE PUBLISHING

iii
Published by
Cambridge International Science Publishing
7 Meadow Walk, Great Abington, Cambridge CB1 6AZ, UK
http://www.cisp-publishing.com

First published 2001

P.R. Soni
Cambridge International Science Publishing

Conditions of sale
All rights reserved. No part of this publication may be reproduced or trans-
mitted in any form or by any means, electronic or mechanical, including pho-
tocopy, recording, or any information storage and retrieval system, without
permission in writing from the publisher

British Library Cataloguing in Publication Data


A catalogue record for this book is available from the British Library

ISBN 1 898326568

Production Irina Stupak


Printed by Pear Tree Press Ltd, Stevenage, England

iv
TO MY FATHER

In memory of my father whom I loved most m


He expired in February 1996

About the Author

Dr. P.R. Soni received his M.Sc. (1975) degree in Physics from the
University of Udaipur; M.E. (1979) and Ph.D. (1991) from the Uni-
versity of Rajasthan, Jaipur. Dr. Soni joined the Department of Metal-
lurgical Engineering at Malaviya Regional Engineering College, Jaipur
in 1980, where presently he is a Reader. He had been a Research
Fellow at the Indian Institute of Technology, Bombay, in 1989. Dr.
Soni was selected as Manager (R&D) of Nappro Synthetics in 1990.
For the last fifteen years, he has been actively engaged in re-
search and development activities related to mechanical alloying,
composite materials and other P/M materials, and is closely associ-
ated with industries through consultancy. He has published thirty tech-
nical papers in various peer reviewed journals. He is a member of
the APMI International and many other professional bodies.

v
vi
Contents

1 INTRODUCTION ......................................................................................... 1
1.1 HISTORY ........................................................................................................................... 1
1.2 BENEFITS OF MECHANICAL ALLOYING ............................................................... 3

2 MECHANICAL ALLOYING ........................................................................... 6


2.1 ALLOYING MILLS ......................................................................................................... 6
2.1.1 Mills in Practice .............................................................................................................. 6
2.1.1.1 Szegvari attritor mill....................................................................................... 6
2.1.1.2 Spex vibratory mill ......................................................................................... 8
2.1.1.3 Planetary ball mill .......................................................................................... 9
2.1.1.4 Large diameter ball mills ............................................................................... 9
2.1.1.5 Grinding media ............................................................................................. 10
2.1.2 Improved Mills .............................................................................................................. 10
2.1.2.1 Modified attritor ........................................................................................... 11
2.1.2.2. Uni-ball mill ................................................................................................ 12
2.2 THE PROCESS ............................................................................................................... 15
2.2.1 Process Monitoring ....................................................................................................... 19
2.3 FACTORS AFFECTING ................................................................................................ 19
2.3.1 Mill Parameters ............................................................................................................ 20
2.3.1.1. Impact energy .............................................................................................. 20
2.3.1.2 Size of the grinding ball ............................................................................... 20
2.3.1.3 Balltopowder ratio ................................................................................... 20
2.3.1.4 Speed ............................................................................................................ 20
2.3.2 Temperature .................................................................................................................. 21
2.3.3 Atmosphere ................................................................................................................... 22
2.3.4 Contamination ............................................................................................................... 23

3 VARIATIONS OF MECHANICAL ALLOYING .............................................. 25


3.1 REACTION MILLING .................................................................................................. 25
3.2 CRYOMILLING ............................................................................................................. 26
3.3 REPEATED ROLLING .................................................................................................. 26
3.4 DOUBLE MECHANICAL ALLOYING ...................................................................... 28
3.5 REPEATED POWDER FORGING .............................................................................. 29

4 PROCESS CONTROL AGENTS IN MECHANICAL ALLOYING ..................... 31

5 MECHANISMS IN MECHANICAL ALLOYING ............................................ 35


5.1 ALLOYING ..................................................................................................................... 35
5.1.1 DuctileDuctile System ................................................................................................. 35
5.1.2 DuctileBrittle System .................................................................................................. 36
5.1.3 BrittleBrittle System ................................................................................................... 37
5.1.4 Idealness of MA Alloys ................................................................................................. 39
5.2 METASTABLE PHASE FORMATION ....................................................................... 40
5.2.1 Amorphization .............................................................................................................. 40

vii
5.2.2 Nanocrystallization ....................................................................................................... 49
5.2.3 Solid Solubility Extension (SSE) ................................................................................. 51
5.3 ACTIVATION OF SOLID STATE CHEMICAL REACTION .................................. 51

6 ENERGY TRANSFER AND ENERGY MAPS IN MECHANICAL


ALLOYING ................................................................................................. 55

7 CONSOLIDATION OF MECHANICALLY ALLOYED POWDERS .................. 58


7.1 CONSOLIDATION TECHNIQUES ............................................................................. 58
7.2 THERMOMECHANICAL TREATMENTS ................................................................ 62

8 MECHANICAL PROPERTIES OF MECHANICALLY ALLOYED


MATERIALS ................................................................................................. 65
8.1 TENSILE PROPERTIES ............................................................................................... 65
8.2 FRACTURE ..................................................................................................................... 69
8.3 CREEP ............................................................................................................................. 69
8.4 SCC SUSCEPTIBILITY ................................................................................................ 70

9 MODELLING MECHANICAL ALLOYING .................................................. 72


9.1 INTRODUCTION ........................................................................................................... 72
9.2 MECHANISTIC MODELS ........................................................................................... 72
9.2.1 Deformation, coalescence and Fracture ...................................................................... 73
9.2.2 Evolution of Particle Size ............................................................................................. 77
9.2.3 Milling Times ................................................................................................................ 79
9.2.4 Powder Heating ............................................................................................................ 85
9.2.5 Powder Cooling ............................................................................................................. 86
9.3 ATOMISTIC MODELS ................................................................................................. 87
9.4 THERMODYNAMIC AND KINETIC MODELS ...................................................... 89

10 JOINING OF MECHANICAL ALLOYING MATERIALS ............................... 91


10.1 WELDING ..................................................................................................................... 91
10.2 BRAZING ...................................................................................................................... 93
10.3 FORGE BONDING ....................................................................................................... 94

11 RAPID SOLIDIFICATION AND MECHANICAL ALLOYING ....................... 96


11.1 RAPID SOLIDIFICATION VERSUS MECHANICAL ALLOYING ..................... 96
11.2 MECHANICAL ALLOYING OF RAPIDLY SOLIDIFIED POWDERS ............... 96

12 APPLICATIONS ....................................................................................... 103


12.1 NICKEL-BASE SUPERALLOYS ............................................................................. 103
12.2 MA STEELS ................................................................................................................ 107
12.3 ALUMINIUM-BASE MATERIALS .......................................................................... 111
12.4 COPPER-BASE MATERIALS .................................................................................. 119
12.5 TITANIUM SYSTEM ................................................................................................. 121
12.6 MAGNESIUM-BASE MATERIALS ........................................................................ 123
12.7 SUPERSATURATED SOLUTIONS ......................................................................... 125
12.8 MAGNETIC MATERIALS ....................................................................................... 126
12.9 MA POWDERS FOR SPRAY-COATINGS ............................................................. 129
12.10 SUPERPLASTICITY ............................................................................................... 130
12.11 TRIBOLOGICAL MATERIALS ............................................................................ 130
12.12 COMPOSITES .......................................................................................................... 134

viii
12.13 AMORPHOUS SOLIDS ........................................................................................... 136
12.14 NANOCRYSTALLINE MATERIALS .................................................................... 137
12.15 MECHANICALLY ACTIVATED CHEMICAL REACTIONS ........................... 138
12.16 OTHERS .................................................................................................................... 140
CONCLUSIONS .................................................................................................................. 140
LIST OF SYMBOLS ........................................................................................................... 145
INDEX .................................................................................................................................. 147

ix
PREFACE
From the total initial laboratory success in 1968, the process of me-
chanical alloying (MA) has been developed into a well-controlled pro-
duction operation over the years, and applied to develop varieties of
materials. Being a new field, there is a wealth of recent scientific lit-
erature available, but it is all scattered The problem of a beginner
to get started in a practical way with the MA technique. This book
tries to address this problem and is aimed at the undergraduates, post-
graduates, materials scientists and engineers who want to have in-depth
knowledge in this field. The book is also designed to serve as an in-
troductory and refreshment reference tool for the manufacturing en-
gineers actively involved in MA or the allied industry but are in need
of detailed knowledge of metallurgical engineering or materials science.
A two year metallurgical engineering or materials science course should
provide the necessary basis for comprehension of the material discussed
in the book.
This book tries to put forward the fundamentals of MA recipes where
the technique has been successful and highlights the areas in technology
where it can provide benefits in developing high-tech materials. Not
only this, many secrets of the MA processing approach are still in the
embryonic stage and this book creates a brainstorm in the mind of ma-
terials scientists and engineers to reveal the same. The book comprises
of twelve chapters. Starting from the historical development of the MA
technique and highlighting its benefits in Chapter 1, Chapter 2 discusses
the basic process of MA devices used for, and factors affecting the
process. This chapter deals with the different variations of MA which
have been developed over the course of time. Chapter 4 deals with
the process control agents used in MA, while Chapter 5 deals with
mechanisms involved in basic processes of MA, metastable phases for-
mation and activation of solid state chemical reactions. Chapter 6 deals
with energy transfer and energy maps in MA. Chapters 7 and 8 deal
with basics of consolidation of MA powders and mechanical proper-
ties of MA products, respectively. Chapter 9 explains the basics of
models developed to predict the results of an MA process. Joining tech-
niques for MA materials are discussed in Chapter 10. The two non-
equilibrium processing techniques, MA and rapid solidification, are com-
pared and combinations of the two to enhance the properties of the
xi
Mechanical Alloying

rapidly solidified materials, have been discussed in Chapter 11. Chapter


12 is meant for highlighting the cases where the technique has been
applied to produce MA products at industrial levels and the potential
materials to find applications in the industrial scene.
My sincere thanks to the Minerals, Metals and Materials Society,
Warrendale, APMI (American Powder Metallurgy Institute) International,
New Jersey; Dr. J.S. Benjamin, Hitchiner Manufacturing Company, Inc,
Milford; Dr. H.K.D.H. Bhadeshia, University of Cambridge; Prof P.
Ramakrishnan, Indian Institute of Technology, Bombay; Prof. F.H. Froes,
University of Idaho and Prof. C.C. Koch, North Carolina State Uni-
versity, for their co-operation in completion of this task. I am grate-
ful to Prof. T.V. Rajan, Malaviya Regional Engineering College, Jaipur
for encouraging to pursue my interest in MA. I wish to acknowledge
my wife Pramila and my children Anshu and Ankit, for bearing my dis-
appearance when I was 'in the book'.

P.R. Soni

J.S. Benjamin, 'Father' of mechanical alloying (1968).

xii
Introduction

1 INTRODUCTION
1.1 HISTORY
The use of inert additions to improve elevated temperature mechani-
cal properties of metals was first exploited in 1910 by W.D. Coolidge
in thoriated tungsten [1]. The development of dispersion-strengthened
alloys by internal oxidation started in 1930 [2] and the invention of
dispersion-strengthened aluminium took place in 1949 [3]. However, the
relatively low melting point of aluminium was a severe limitation for
the use of SAP at elevated temperatures. This led to attempts in ap-
plying dispersion strengthening to higher melting point metals such as
copper and nickel.
In these metals, the self oxides cannot be used as they are not suf-
ficiently stable against Ostwald ripening at elevated temperatures. ThO 2-
dispersed nickel, having a finely distributed dispersoid, was produced
successfully by melting to improve the mechanical properties (1960)
[4]. The material has vastly developed elevated temperature properties.
However, the use of such materials was still limited due to their low
strength at intermediate temperatures and lack of corrosion resistance.
Hypothetically, these shortcomings could be solved by combining the
corrosion resistance and intermediate temperature strength of
precipitation-hardened, nickel-base superalloys with the high tempera-
ture strength and stability of oxide dispersion strengthening. One of the
primary hurdles to materialize this idea is the production of uniform
dispersion of fine oxide particles, less than 0.1 m in size, in alloy pow-
der particles in such a manner which leads to interparticle spacing of
less than 0.5 m in a consolidated product. The nickel-base superalloys
contain chromium, aluminium and titanium for effective precipitation
hardening, elements which are easily oxidizable. Aluminium and tita-
nium oxides are so stable that they cannot be reduced to a metallic
state, required in an alloy, without reducing deliberately the dispersed
oxide as well. Moreover, oxidation of these elements removes them as
a precipitation hardener in the alloy.
Four techniques, namely the simple mixing, ignition surface coat-
ing, internal oxidation and selective reduction, are available to com-
1
Mechanical Alloying

bine oxide dispersion strengthening and solid solution strengthening in


alloy systems containing relatively non-reactive elements. However, all
these four techniques were found to be unsuitable for the production
of oxide dispersion-strengthened precipitation-hardened nickel-base
superalloys [5], mainly due to reactiveness of the chromium, aluminium
and titanium present in the alloy.
The mixing technique requires interparticle spacing even with a large
mechanical reduction during consolidation and subsequent working
operations. Powders of this fineness containing formers such as alu-
minium and titanium are very reactive, leaving aside the health haz-
ards, because of their high specific surface area and complete or nearly
complete oxidation of aluminium and titanium that it can result in. The
problem can be overcome to some extent by employing slightly coarser
powders and grinding the mixture in a ball mill. To overcome the ten-
dency of fine particles to weld together during the milling, kerosene or
fatty acids are usually added. Although lubricants make fine grinding
possible, they may severely contaminate the powders and degrade the
alloy made from them.
The ignition surface coating technique involves mixing the matrix alloy
powders with a liquid solution of a salt of reactive metal. This mix-
ture is dried and pulverized, and the powders are heated in an inert
or reducing environment converting the salt to a refractory oxide. This
technique also produces oxide coated powders which have the same
disadvantages as powders made by the simple mechanical mixing tech-
nique. In addition, there is a greater contamination problem because
of the oxidizing potential of the reaction products of the salt decom-
position step.
The internal oxidation technique involves oxidation of the alloy pow-
der or thin strip containing a dilute solid solution of the reactive ele-
ment, in an oxidizing environment at elevated temperatures. It is found
experimentally that the particle size of the dispersoid increases with
increasing depth of penetration of internal oxidation. Therefore, very
fine powders or expensive ultra-thin strips are required to obtain suf-
ficiently fine dispersoid particles. In the case of nickel-based superalloys
containing reactive -forming elements, an additional problem would
arise. The oxygen potential could not be raised above the extremely
low values required to oxidize the forming elements, which could
prohibit the oxidation rate of the desired dispersion-forming elements.
The selective reduction process has been used to manufacture com-
mercial dispersion-strengthened materials. It involves production of an
intimate mixture of metal oxides and selective reduction of the oxides
of the matrix alloy, whilst leaving the dispersoid unreduced. Aluminium
and titanium present in the nickel-base superalloy make the reducing
2
Introduction

step impossible with gases, because of the stability of Al2O 3 and TiO 2.
These oxides can be reduced by the use of molten alkali and alkaline
earth metals. However, this introduces two major new problems, ex-
cessive growth of the dispersoid particles, and the necessity to remove
the reaction product oxides and carrier agent, usually salt.
To cope with these problems associated with the production of oxide
dispersion-strengthened nickel-base superalloys J.S. Benjamin (Father
of MA) and his colleagues at the Paul D. Merica Research Labora-
tory of the International Nickel Company processed a very high pu-
rity nickel powder, a fairly coarse chromium powder, a master alloy
powder of nickel, aluminium and titanium and a very fine powder of
Y 2O3 in a high energy attritor mill in the late 1960s [6]. As a historical
note, it should be recorded that at this stage, for the process of mixing/
milling the term mechanical alloying was coined by Ewan C.
MacQueen, a patent attorney for the International Nickel Company [7].
When such a mixture is mechanically alloyed, the degree to which its
various constituents maintain their form depends on their relative hard-
ness and their ability to withstand deformation. Nickel, which is the
softest constituent of the mixture, is the cemet that binds the other con-
stituents together. Chromium is somewhat harder and less ductile than
nickel, so it tends to form plate-like fragments that are embedded in
nickel. The Y 2O 3 disperses along the welds in the composite particles.
At the end of the process, the composite particle has a random dis-
tribution of oxide dispersoid in a metal-matrix composite. The powder
was then consolidated by hot extrusion. The grain size in the extruded
bar was very fine. However, to have maximum high temperature
strength, the grains should be coarser. For this, the extruded bar was
rolled and then annealed at about 1263C for 30 min, to recrystallize
to quite coarse grains.
As well as this, they could add elements such as tantalum, molyb-
denum and tungsten to the nickel-base superalloy, which gave added
strength at lower temperatures.

1.2 BENEFITS OF MECHANICAL ALLOYING


The MA process has several associated advantages:
1. The homogeneity in fine powder is independent of the initial pow-
der size, which avoids the hazards of fine powders.
2. Fine homogeneous dispersions can be obtained in a particle size
of 1 m or less at a high concentration of alloying elements without
occluding air, provided that enough ductile metal powder can be intro-
duced.
3. Grinding times are reduced to 1/10th or even less as compared
to that required in a conventional ball or pebble mill.
3
Mechanical Alloying

4. Liquid metal techniques are most convenient and cheaper to de-


velop an alloy. But for the case where it is not possible to get a ho-
mogeneous alloy by these techniques, powder metallurgy is adopted.
The value of MA becomes apparent when attempts to make an al-
loy cannot be made by these conventional routes. If the two metals
form a solid solution, MA can be used to accomplish the same at lower
temperatures. If the two metals are insoluble in solid state, i.e. im-
miscible solids (e.g. CuFe) or in liquid state i.e. immiscible liquids
(e.g. CuPb), an extremely fine dispersion of one of the metals in the
other can be achieved.
5. Mechanical alloying represents a cold alloying process, thus it
is suitable for hazardous operations. With proper precautions even vola-
tile inflammable materials can be handled safely.
Today, MA has been used for developing alloys from immiscible liq-
uids or solids, incongruent melting, intermetallics and metastable phases,
and has emerged and developed into a technology capable of providing
unique PM materials with consistent properties for high performance
applications over a wide range. Various application areas where the
MA technology has been utilized are illustrated in Fig.1.1.

Fig.1.1 Application areas of mechanical alloying.

4
Introduction

References
1. W.D. Coolidge, Proc. Am. Inst. Elec. Eng., 961 (1910).
2. C.S. Smith, Min. and Met., 11, 213 (1930).
3. R. Irman, Tech. Rundschen (Bern), 36, 9 (1949).
4. G.B. Alexander, U.S. Patent 2, 972, 529 (1961).
5. J.S. Benjamin, Met. Trans., 1, 2943 (1970).
6. J.S. Benjamin, Scientific American, 235, 40 (4) (1976).
7. J.S. Benjamin, MPR, 45 (2), 122 (1990).

Questions
1. Which are the various techniques available for preparing dispersion
strengthened materials.
2. How is MA helpful in the case of reactive materials?
3. Who invented MA, when and where?
4. Give in brief, a chronological development of MA.
5. Enlist various advantages associated with MA.

5
Mechanical Alloying

2 MECHANICAL ALLOYING
2.1 ALLOYING MILLS
A variety of milling equipment such as attritor mills, vibratory mills,
high speed blenders and shakers, planetary mill, and even large diameter
conventional ball mills have been used to carry out MA. The milling
machine stresses the maximum number of individual particles in a
powder mass to undergo plastic deformation or initiate fracture with
a minimum of energy. The motion of the milling medium and the charge
varies with respect to the movement and trajectories of individual balls,
the movement of the mass of balls, and the degree of energy applied
to impact, shear attrition and compression forces acting on powder par-
ticles. Impact is the instantaneous striking of one object by another.
Both objects may be moving or one may be stationary. Attrition is the
production of wear debris or particles created by rubbing action be-
tween two bodies. This type of milling force is preferred when the
materials are friable and exhibit minimal abrasiveness shear consisting
of cutting or cleaving of particles, and is usually combined with other
types of force. Shear contributes to fracturing by breaking particles into
individual pieces, with the creation of a minimal of fines. Compres-
sion is the slow application of compressibility forces to a body. A choice
between these is likely to be determined by the end results required,
as well as the chemical and physical properties of the powder.

2.1.1 Mills in Practice


2.1.1.1 Szegvari attritor mill
More than other ball mills, the Szegvari attritor is preferred by most
workers because of its operational flexibility. It was invented by Szegvari
originally for a comminution/blending machine meant for chemical in-
dustries [1]. It consists of a water-cooled stationary vessel with a cen-
trally mounted vertical shaft with impellers radiating from it (Fig.2.1).
The shaft is connected to a high-speed, geared motor. The vessel is
usually made gas-tight with rubber seals, especially when a control-
led atmosphere is required to be maintained. A laboratory attritor is
shown in Fig.2.2.
When the shaft rotates, arms or lifters stir the balls causing them
6
Mechanical Alloying

Fig.2.1 Attritor mill.

Fig.2.2 Laboratory attritor mill (courtesy Union Process).

7
Mechanical Alloying

to lift up and fall back. There is thus a differential movement between


the balls and materials being milled, giving a much higher degree of
surface contact. The rotating charge of balls and the powder form a
vortex at the upper end of the stirring shaft, into which the milling prod-
uct and balls are drawn. Attritors have many associated advantages as
listed below:
1. Milling is achieved by impact and shear forces, and is very in-
tensive because the force restoring the media downward is the weight
of all the media above it.
2. The high impact energy allows the use of smaller diameter balls
so that powder with a narrow size distribution may be produced.
3. As the greatest milling action is at 2/3 chamber-radius away from
the shaft, there is little contamination due to wear of the tank or shaft.
The minimal wear of chamber walls ensures long service life.
Today, these mills range from the laboratory scale where a few
grams of materials are processed to current production units where
1 ton of powder is processed in a 2 m diameter mill which contains
more than 1 million balls weighing approximately 10 tons. The alloying
process is both mill and powder specific, powders can be in almost any
form and can range in diameter from 11000 m [2].
However, these mills have a drawback of relatively low product out-
put (0.5 to 100 kg). They need more power and can be difficult to run
and maintain cost effectively.

2.1.1.2 Spex vibratory mill


The use of vibratory ball mills for MA has been promoted predomi-
nantly by Kuhn [3,4]. The vibratory ball mill is a long closed tube con-
taining grinding balls and powder. The mill agitates the charge in three
mutually perpendicular directions. A schematic of the vibratory mill is
shown in Fig.2.3. Impact forces acting on the powder in a vibratory
mill are a function of rates of milling, amplitude of vibration and mass
of the milling medium. High energy milling forces can be obtained by
using high frequency and high amplitude of vibrations. These mills nor-
mally operate at frequencies of the order of 200 Hz and a high am-
plitude of about 12 mm. It is estimated that the maximum accelera-

Fig.2.3 Vibratory mill.

8
Mechanical Alloying

tion of the grinding balls works out to be 24g, where g is the gravi-
tational force [2].
Maximum milling action takes place adjacent to the chamber wall
and minimum at the centre, as the motion of the medium decreases
from the chamber walls to the centre of the mill tube. The entire charge
slowly revolves counterclockwise to the oscillatory vibrations and the
high g-forces experienced by the agitated balls leads to intense MA.
Vibratory mills utilize small size grinding balls because of higher im-
pact forces, frequencies and acceleration. In fact, it is the most en-
ergetic milling device as compared to other mills. These mills are ex-
cellent for MA of amounts of up to 4.5 kg or more, depending on the
apparent density of the powder. The vibratory mill has, however, not
found wide acceptance for bulk production.
High speed Spex shaker mills or blenders are ideally suited for labo-
ratory experiment. In fact, some of Benjamins experiments were also
carried out in a high speed shaker mill.

2.1.1.3 Planetary ball mill


Bachin et al [5] carried out MA of dispersion-strengthened, nickel-base
superalloys in a centrifugal planetary ball mill. The mechanics of this
mill are characterized by the rotational speed of the plate p, that of
the container relative to the plate v, the mass of the charge, the size
of the ball, the ball to powder ratio and the radius of the container.
A schematic of the planetary ball mill is shown in Fig.2.4. Figure 2.5
shows a laboratory planetary mill.

2.1.1.4 Large diameter ball mills


For processing larger batches of powder, however, there has been a
trend recently to use conventional horizontal ball mills with larger

Fig.2.4 Schematic of a planetary ball mill.

9
Mechanical Alloying

Fig.2.5 Laboratory planetary mill (courtesy Fritsch GmbH).

diameters (0.5 to 2.5 m) to achieve high energy by rotating it just below


the critical speeds c (up to 0.9 c) [6]. Even though the time required
to accomplish MA by these mills is longer compared to attritor mills,
the overall economics are favourable.

2.1.1.5 Grinding media


As far as the grinding media are concerned, common practice is to use
hardened high carbonhigh chromium steel balls (4 to 12 mm diam-
eter), normally specified for ball bearings. Stainless steel balls have also
been used. When it is necessary to minimize iron contamination in the
charge, balls of tungsten carbide have also been used. When necessary,
the balls have been coated with the necessary oxide that was to be
dispersed in the composition to be mechanically alloyed [7].

2.1.2 Improved Mills


The commonly used MA mills just discussed are typically inefficient
due to the fact that only a small fraction of the expended energy re-
sults in the powder deformation, welding and fracture requisites for MA.

10
Mechanical Alloying

Fig.2.6 Regions of the attritor manifesting differing milling actions (Ref. 8, Met. Trans.,
24A, 175 (1993)).

These devices are not designed with specific needs of the MA proc-
ess in mind. This factor has become particularly obvious in recent
years, when the dependence of final products on the type of milling
device and energy regime used have been repeatedly reported. On these
premises a few improved milling devices have been proposed as dis-
cussed below.

2.1.2.1 Modified attritor


Attritor efficiency is limited by virtue of its geometrical characteristics
and dynamics. The region outside the volume containing the attritor
arms is relatively inactive (Fig.2.6) due to a radial ball velocity pro-
file in the mill (Fig.2.7). In fact, a dead zone, containing a dispropor-
tionately high volume fraction of powder, exists at the periphery of the
mill chamber bottom. However, some regions of the attritor outside the
volume containing the arms are characterized by ball velocity gradi-
ents which provide a mechanism for mechanical action by sliding and
a means for powder circulation within the attritor. The direct impacts,
rather than any rolling/sliding events, are primarily responsible for the
mechanisms effecting MA [8]. Elimination of the dead zone and a
means for facilitating powder circulation within the mill can reduce the
process time for an attritor. The following three approaches might lead
to improved attritor efficiency [8].

Fig.2.7 Radial ball velocity profile in an


operating attritor (Ref.8).

11
Mechanical Alloying

1. The first focuses on the geometrical characteristics of the grinding


media; milling with balls having the same diameter results in a closed
packed ball array in the attritor volume outside of the core. Such an
array results in the predominance of rolling, rather than impact events.
Use of balls with two (or more) different diameters would presumably
prevent the close packed array from forming, with an attendant increase
in mill agitation. The result could be more direct impact events and
a lesser relative number of rolling/sliding ones. In fact cinematography
has shown that ball mixing disrupts the close packed array assumed by
single sized balls in the region of the attritor outside its core. This
disruption ostensibly results in a more efficient milling action [9].
2. The other approaches could facilitate powder removal from the
dead zone and/or reduction of this zones volume. The bottom of the
attritor chamber could be contoured in such a way so as to steer balls
and powder up into the more active mill region (Fig.2.8a).
3. An additional arm could be placed on the attritor shaft, as in-
dicated in Fig.2.8b. This (the lowest) arm could be machined into a
wedge as opposed to the current round shape. This would force balls
up from the mill bottom into the more active regions. Other balls will
fall to fill the volume vacated by balls circulated in this way, thus
creating a localized convection of powder and grinding media.

2.1.2.2. Uni-ball mill


Since at the present stage of research very little is known about what
milling conditions are to be selected to produce an alloy with a prior

Fig. 2.8 Suggestions for improved attritor design


(Ref.8).

12
Mechanical Alloying

Fig.2.9 Schematic of the ball milling with controlled ball movement: 1) rotating cell; 2)
balls; 3) magnets (Ref.10, Mat. Sci. and Eng., A134, 1356 (1991)).

chosen structure, there is a need of a simple and versatile device. Ide-


ally, such a mill should offer:
A wide range of milling conditions;
Control of milling parameters.
Keeping this in view, the uni-ball mill (universal ball mill) was de-
veloped, which is close to these requirements [10]. A schematic of the
mill is shown in Fig.2.9. It is a conventional ball mill in which the ball
movement during the milling process is confined to the vertical plane
by the cell walls and is controlled by an external magnetic field. The
intensity and direction of the field can be externally adjusted. By ad-
justing the spatial and/or temporal profiles of the magnetic field, the
ball trajectories, the impact energy and the shearing energy can be var-
ied.
Figure 2.10 illustrates the three general patterns of ball movement
that can be achieved using this device. In the case illustrated in
Fig.2.10a, the magnetic field holds the balls in the bottom part of the
cell rotating with the frequency c. Friction causes the balls to rotate
R
in the same direction with the frequency w o = w c c , where Rc is the
Rb

Fig.2.10 Modes of operation of the ball mill: a) high energy mode, high rotation frequency;
b) low energy mode, low rotation frequency; c) high energy mode two point of equilibrium;
d,e) high energy mode, intermediate rotation frequency (Ref.10).

13
Mechanical Alloying

radius of the cell and Rb is the radius of the ball. Periodically, the outer
ball on the right-hand side gets released, completes most of the cir-
cle being pushed against the cell wall by the centrifugal force and hits
the left-most ball of the bottom. With this mode of operation, the
powder is worked by both by impact and shearing. The balls may be
confined to the bottom part of the cell for the whole time, either by
increasing the intensity of the magnetic field or by decreasing the fre-
quency c (Fig.2.10b). With this mode, the balls both rotate and os-
cillate around the equilibrium position of the bottom and the powder
is worked mostly by shearing.
In the third case, illustrated in Fig.2.10c, the ball movement caused
by the centrifugal force can be halted in two opposite directions: at the
lowest and highest point inside the cell. The ball trapped by the mag-
netic attraction in the upper position rotates with frequency b and can
be released to fall vertically on top of one of the bottom balls. The
two colliding balls rotate in opposite directions, which results in a com-
bination of shearing and uniaxial pressure at the surface of contact.
Two useful variations of mode (a) are shown in Figs.2.10c and 2.10d.
By slowing down the cell rotation frequency c a situation may be
achieved when the ball released from the bottom is not fully pinned
to the wall by the centrifugal force and can either hit one of the bottom
balls (Fig.2.10c) or the opposite cell wall (Fig.2.10d). The milling con-
ditions in situations (a), (d) and (c), (e) are similar, respectively.
In conventional ball mills, typical impact times are the order of
10 5 sec and the peak stress can reach 50 kbar. Depending on the col-
lision parameters, the orientation of the impact strain may be quite com-
plex and the temperature rise in the collision region may vary from a
few degrees to a few hundred degrees [11,12]. The mechanical energy
is utilized most efficiently if the balls collide head-on with maximum
possible velocities. Davice et al [12] reported eight head-on collisions
out of the total of 2132 collisions per second in a vibratory mill agi-
tated at 1200 rpm. There are only one to two collisions per second
in this device, but they are strictly head-on because of the guiding of
the ball movement by the magnetic field. The free-fall ball velocity dur-
ing an impact is 15 m/sec (it is increased in the presence of a mag-
netic field). For comparison, the average relative ball velocity in a vi-
brating mill varies from 3.9 to 6 m/sec [4], in a planetary mill [13]
it remains in the range of 2.54.7 m/sec and it is about 0.5 m/sec in
attritors [11]. The calculated values of collision time, Hertz radius (refer
to Section 9.2.3) and maximum impact stress have been found to be
6.510 5 sec, 4.610 4 and 37 kbar respectively for this mill. All these
values are close to the corresponding values quoted for commercial vi-
brating mills. Thus, the energy released per impact in this mill is not
14
Mechanical Alloying

much different from the characteristics of other devices. The unique


feature of the device is the specific ball movement pattern. With every
mode of operation this pattern is well defined and highly reproducible.
This contrasts with chaotic and unpredicted ball movement character-
istics of the other ball-milling devices.
Using this improved mill, it could be possible to reproduce both of
the amorphization paths: A+B(AB) am or A+B(AB) in(AB)am paths
for a Ni38 at% Zr mixture using two different modes of operation.
When the high-energy milling mode is used, the amorphous phase (AB)am
is formed directly from the mixture. In contrast to this, when the low
energy milling mode (b) is used, the intermetallic phase (AB) in form
first, which then slowly transforms into the amorphous phase. Using
such a device, it could be possible even to produce metalmetalloid high
melting point intermetallics which are believed to require very high
energies [14].
To produce fully amorphous Mg 70Zn 30 alloys from the crystalline
master alloy having a low crystallization temperature (95C), a two step
ball milling procedure has to be used using the high energy mode (d).
First, a steady state mixture of amorphous and crystalline phase is
generated, then full amorphization can be achieved by switching to the
low energy mode.
The high energy mode (d) is not suitable for MA of AlV, AlTi
and other aluminium-based and magnesium-based alloys. This mode fa-
vours the cold welding process and leads to the formation of large
lumps of the milled mixtures. The appropriate milling procedure is to
use the low energy mode (b) until the mean grain size decreases to about
0.1 m and then switch to the intermediate energy mode (e). Such a
procedure enables MA of the aluminium based alloys without a PCA
(refer to Chapter 4) to control the balance between fracturing and cold
welding [15].

2.2 THE PROCESS


Mechanical alloying is normally defined as a dry, high-energy, ball milling
process whereby two (or more) elemental powders are blended, cold
worked, welded and fragmented repeatedly, resulting in powders with
a uniform atom distribution, in stable or metastable phase in a finer
microstructure. While applying the technique for dispersoid distribution,
it is not really alloying due to insolubility of the dispersoids in the
matrix.
When an alloy powder instead of elemental powders is milled, the
process is termed as mechanical grinding (MG) [16]. In this case, the
function of MG is to introduce point and lattice defects such as va-
cancies, interstitials, dislocations, antiphase domain boundaries, etc. to
15
Mechanical Alloying

completely destroy the crystal structure and generate an amorphous


material. Thus, the MA process actually promotes particle welding in
contrast to the conventional milling practices in which welding is in-
hibited by the use of liquids and surfactants.
Every time two grinding balls collide they trap particles between them.
The force of the impact deforms the particles creating atomically clean
new surfaces. When the clean surfaces come into contact, they weld
together. Since such surfaces readily oxidize, the milling is carried out
in an inert atmosphere or vacuum. To facilitate interparticle welding,
there must be adequate compressive energy imparted to the grinding
medium during milling (hence the high energy mill) and usually the
presence of a malleable constituent that could act as a binder for the
other constituent and also readily bond with the balls. The other com-
ponents may include ductile metals, brittle metals, intermetallics, or non-
metals such as carbon, oxides and nitrides. This makes it necessary
that milling is done in a dry atmosphere especially for metals of high
melting point to promote cold welding.
At the early stages in the process, the metal powders are still rather
soft and the tendency for them to weld together into larger particles
predominates. A broad range of particles develops, with some parti-
cles being two to three times larger in diameter than the original ones.
As the process continues, the particles get harder and their ability to
withstand deformation without fracturing decreases. The smaller par-
ticles tend to weld into larger pieces. The large particles, on the other
hand, are more likely to incorporate flaws and to break apart when they
are struck by the balls. In time, the tendency to weld and the tendency
to fracture come into balance, and the size of particles becomes con-
stant within a narrow [17] range (Fig.2.11). The major factors contrib-

Fig.2.11 Balance between welding and fracturing (Ref.17).

16
Mechanical Alloying

uting to this grind limit [18] are:


Increasing resistance to fracture.
Increasing cohesion between particles with decreasing particle size
causing agglomeration.
Excessive clearance between impacting surfaces, which is mini-
mized as the ball diameter decreases.
Coating of fine particles on a grinding medium which cushions
the particles from impact.
Surface roughness of the grinding medium (highly polished, hard
mediums that retain a minimum root mean square surface roughness
during milling are most effective).
Bridging of large particles to protect smaller particles in the
microbed.
Generally, as the alloying proceeds over an extended time, the mean
applied stress needed for particle failure increases, while the magni-
tude of local stresses available to initiate fracture decreases.
Although there is little change in size of the particles after the equi-
librium is reached, the structure of the particles is steadily refined. The
alloying reaches a significant point when the welded layers of a par-
ticle can no longer be optically resolved. At this stage, two metals get
closely mixed on an atomic level. They have formed a true solid so-
lution rather than a mixture of fine fragments. The welds observed in
various systems studied have shown a maximum value of 0.7 m, but
in general this is seen to be far less. At this point, the powder is con-
sidered to be adequately processed.
It is found that the rate of refinement of the internal structure of

Fig.2.12 Reduction in thickness of the layers within a composite ironchromium particle


during MA (Ref.17).

17
Mechanical Alloying

the particles is roughly logarithmic with processing time (Fig.2.12). As


a result, the requirement of fine constituent starting powder is not criti-
cal. J.S. Benjamin [19] made the interesting observation that each time
a powder particle is trapped between colliding balls, it is plastically de-
formed sufficiently to reduce its thickness by a factor of two to three
times. A simple calculation then leads to the conclusion that for dis-
persion-strengthening superalloys, starting with a particle size of about
100 m, only six to eight MA events are needed to obtain a desired
lamellae thickness within the composite powders of about 0.2 m. To
reduce the inter-lamellae spacing down to five angstroms, or an atomic
diameter, requires between eleven and thirteen events. However, the
time necessary to cause the required number of collision events to occur
is generally in the tens of hours.
There is surprisingly little contamination (ppm level) of the powder
by the iron in the steel vessel and the steel grinding balls, due to the
fact that as the grinding proceeds, the balls and the inner walls of the
vessel get coated with a layer of the metals in the mixture [20]. The
layer is constantly flaked off the balls, fragmented and rewelded.
When a metal is plastically deformed by cold working, most of the
mechanical energy of the deformation gets converted into heat (about
5% is stored in the metal raising its internal energy). Heat is also gen-
erated by elastic deformation of metal grinding balls and mill cham-
ber walls. The energy expended to overcome the friction between the
particles is also translated into heat. Thus, if the temperature of the

Fig.2.13 Typical SEM layered structure of milled powders (Cr48% Nb elemental mixture)
indicating the degree of alloying as a function of milling time: a) 5 hrs; b) 15 hrs; c) 20
hrs; d) 25 hrs (A.H. Clauer and J.J. deBarbadillo (editors). Solid State Powder Processing
(1990), p.305 (The Minerals, Metals and Materials Society).

18
Mechanical Alloying

Fig.2.14 Microhardness as a function of processsing time for pure Al + 1.5%


NOPCOWAX-22. (Ref.24, Met. Trans., 24A, 513 (1983).

powder rises above a certain point, the cold worked metal particles may
undergo recovery and recrystallization. Therefore, water-jacketed milling
chambers are usually required for large, high energy vibratory and
attritor mills that reach temperatures as high as 200C [21].

2.2.1 Process Monitoring


The measurement of average flake size does not provide a meaning-
ful criterion for comparing the effect of changes in process param-
eters, competing alloying processes, and equipment. However, the
microstructural changes can reflect the effect of these changes (Fig.2.13)
[22,23]. Microhardness measurement is the most meaningful up to the
alloying time that produces maximum levels of cold work (Fig.2.14)
[24]. X-ray line broadening is sensitive to both the amount of cold work
and the refinement in crystallite structure that occurs with continuing
kneading and working of the metal well after saturation of cold work
[25]. Accordingly, the progress of alloying can be monitored by fol-
lowing microstructural changes, hardness changes and X-ray diffrac-
tion. As the absence of sharp X-ray peaks may not always be clear
evidence for amorphization, then high resolution electron microscopy
[26], scanning electron microscopy [27], Mssbauer spectroscopy [28],
and superconducting transition temperature [29] can be used to monitor
the process. If peak orientations and/or peak overlapping prevent ef-
fective exploration using X-rays, nuclear magnetic resonance spectra
[30] can be used.
In fact, any measurable property of the material which is affected
by MA processing can be chosen to monitor the process.

2.3 FACTORS AFFECTING


The progress and the end product of MA is greatly affected by a
number of processing parameters, such as mill parameters (impact en-
ergy, ball-to-powder ratio (BPR), mill speed, size and size of distri-

19
Mechanical Alloying

bution of balls, even the shape of impellers in the case of attritor mills),
temperature, atmosphere and contamination.

2.3.1 Mill Parameters


2.3.1.1. Impact energy
This depends on the specific mill, and the density and size of balls.
It is observed that microhardness developed in the MA microstructure
is dependent on the impact energy [31,32]. It has also been observed
that at high mill energies the degree of crystallization increases and with
low energies amorphization occurs [33].

2.3.1.2 Size of the grinding ball


The size of the ball affects size, morphology, recrystallization tempera-
ture and enthalpy of the powder produced [34]. As discussed in Section
2.1.2, welding/fracturing events can be enhanced by use of a range of
ball sizes, rather than using balls of the same size.

2.3.1.3 Balltopowder ratio


Increase in balltopowder ratio (BPR) reduces the mean free path of
the motion, whilst a low BPR minimizes collision frequency. Thus, im-
pact frequency and total energy consumption per second increase with
increasing BPR, while the average impact energy per collision decreases
with increasing BPR [31] and minimizes collision frequency. In gen-
eral, the effective BPR has been found to be in the range 5 to 30. For
amorphization cases, it has been found that as the BPR increases the
amorphization rate increases sharply, but the contamination, by iron from
the milling tools also increases [35]. In general, for amorphization a
BPR approaching 100 is frequently used.

2.3.1.4 Speed
The milling speed is one of the most important variables to be con-
sidered. Very low rotational speeds lead to very long periods of mill-
ing (>100 hrs) and a large inhomogeneity in the alloy because of in-
adequate kinetic energy input, resulting in insufficient localized attri-
bution heat input for alloying. Therefore, an extremely prolonged milling
time would presumably be required for homogeneous [20,36] alloying
(Fig.2.15). For speeds greater than the optimum, the milling time gets
reduced for the same number of revolutions and thus effectiveness of
alloying again decreases because of the decrease in the time available
for diffusion of the solute. However, very high speeds could lead to
excessive heating, high wear of the balls causing contamination from
the grinding medium and lower yields.

20
Mechanical Alloying

Fig.2.15 Effectiveness of alloying


as a function of attritor RPM for
constant number of revolutions
Ref.36, Trans. IIM, 39, 596 (1986),
The Indian Institute of Metals).

It is conjectured that there may also be an additional factor con-


tributing to the decreased effectiveness of alloying at very high speeds.
The horizontal, centrifugal component of the velocity vector is expected
to dominate, thus diminishing the vertical vector component of the agi-
tator velocities which is presumably an essential ingredient for homog-
enizing the alloying process, as is the case in a comparable situation
of mixing by agitators.
However, these milling parameters are interdependent. Therefore, it
should be adequate to optimize any one of the important parameters
for a given set of experimental conditions.

2.3.2 Temperature
The ambient temperature of MA and MG is an important parameter
which may influence the final structure. The amorphization of the sys-
tem, for which the heat of mixing is negative (H mix< 0) e.g. NiZr,
increase in ambient temperature increases amorphization due to the fast
diffusion rate of the constituent elements [27,37]. For the system, with
positive heat of mixing (H mix>0), the effect of ambient temperature
starts after a specific period of MA (e.g. CuTa, ambient temperature
effect starts after 30 hrs of MA). During this period, the alloying mostly
21
Mechanical Alloying

Fig.2.16 X-ray diffraction spectra


for a mixture of Cu30 Ta 70 powders
subjected to MA for 100 hrs. (Ref.
37, Mat. Sci. and Eng., A134,
1334 (1991)).

Fig.2.17 Total enthalpy H t vs milling time


for the Fe 2 B compound (smooth curve)
(Ref.37).

causes the reduction in grain size (100 ) and accumulation of strains.


Thus, the interdiffusion begins to act effectively when the average grain
size is reduced to a certain minimum. Hence, the retarded tempera-
ture effect [37] will be a unique feature observed in a system char-
acterized by a positive H mix (Fig.2.16).
The process of MG of intermetallic compounds starts from the low-
est free energy. Hence, it involves energetically the up-hill process due
to the accumulation of strains and defects, and a reduction in grain size.
The total stored energy Ht for the two systems: NiZr2 [38] compound
which has been confirmed to amorphize, and Fe 2B compound [37]
which does not amorphize, are shown in Fig.2.17. It can be clearly ob-
served that in the case of Fe 2B there is a negligibly small effect of tem-
perature which can be attributed to the lack of formation of the amor-
phous phase (H t levels off at 7.5 kJ/mol, which is smaller than the
calculated value of 19 kJ/mol).

2.3.3 Atmosphere
Mostly, the MA and MG processes are performed in an inert atmos-

22
Mechanical Alloying

phere, since a small amount of O 2 or H 2O may have a large influence


on the final product. Koch et al [39] have prepared amorphous Ni60Nb40
by MA of the elemental powders in air and helium. The crystalliza-
tion was found to be different in the two due to different amounts of
oxygen content in the alloy. The MA of Ni 60Ti 40 results in different
transformation paths when the process is carried out in high-purity ar-
gon, air and high-purity nitrogen atmospheres [40]. It is also pointed
out that in this case nitrogen has a small influence, while oxygen has
a greater effect on MA.

2.3.4 Contamination
The type and level of contamination from the wear of the mill chamber
and the grinding balls, though small (ppm level), may influence the
amorphization transformation path. For example, when MG of NiZr
is carried out in an attritor or vibratory mill, the compound transforms
directly to the amorphous state. But when the process is carried out
in a planetary mill, crystalline intermetallic compound appears as an
intermediate product which subsequently transforms to the amorphous
state. The difference is attributed to different levels of iron contami-
nation from the different mills [41]. There are cases when the change
of grinding medium balls (say from steel to WC) changes the end prod-
uct. The type of PCA used during MA influences the end product (refer
to Chapter 4), which may presumably be due to the level of carbon
added in the MA powder by the PCA.

References
1. W.E. Kuhn and M. Lucky, In: Fine Particles, W.E. Kuhn and J. Ehretsmann (eds).
The Electrochem. Soc. Inc. (1974), p.95.
2. G.B. Schaffer and P.G. McCormick, Mat. Forum, 16, 91 (1992).
3. W.E. Kuhn, In: Modern Developments in Powder Metallurgy, V.12, MPIF,
Princeton, N.J. (1980), p.195.
4. A.N. Patel and W.E. Kuhn, In: Modern Development in Powder Metallurgy,
V.13, MPIF, Princeton, N.J. (1980), p.2750.
5. B.N. Bachin, S.Ya. Kolupaeua, Yu.A. Kustor, A.J. Chernyok and B.V. Schetanov,
Poroshk. Metall., 235 (7) (1982), p.44
6. P.S. Gilman and W.E. Mattson, US Patent 4, 267, 959 (1986).
7. Anonymous, Attrition Mills, MPR, 51 (5), 58 (1996).
8. R.W. Rydin, D. Maurice and T.H. Courtney; Met. Trans., 24A, 175 (1993).
9. T.M. Cook and T.H. Courtney, Met. and Mat. Trans., 26A, 2389 (1995).
10. A. Calka and A.P. Radlinski, Mat. Sci. and Eng., A134, 1356 (1991).
11. D.R. Maurice and T.H. Courtney, Met. Trans., 21A, 289 (1990).
12. R.M. Davice, B. McDermott and C.C. Koch, Met. Trans., 19A, 2867 (1988).
13. J. Eckert, L. Schultz and E. Heustern, J. Appl. Phys., 64, 3224 (1988).
14. A.P. Radlinski and A. Calka, Mat. Sci. and Eng. A134, 1376 (1991).
15. A.P. Radlinski, A. Calka and B.W. Ninham and W.A. Kaczmarek. Mat. Sci. and
Eng. A134, 1346 (1991).

23
Mechanical Alloying

16. R.B. Schwarz and C.C. Koch, Appl. Phys. Lett., 49, 146 (1986).
17. J.S. Benjamin, Scientific American, 234, 40 (1976).
18. ASM Metals Handbook, Ninth Edition, V.7, 1990, p.56.
19. J.S. Benjamin, Mat. Sci. Forum, 88-90 (1992).
20. E.S. Rao, In: Powder Metallurgy - Recent Advances, V.S. Arunachalam
and O.V. Roman (eds), Oxford IBM, New Delhi (1989), p. 27.
21. H. Kimura, M. Kimura and F. Takada, J. Less-Common Metals, 140, 113 (1988).
22. J. S. Benjamin and T.S. Volin, Met. Trans, 5A, 1929 (1974).
23. K.H. Karmer, Powd. Metall. Int., 9, 105 (1977).
24. P.S. Gilman and W.D. Nix, Met. Trans., 12A, 513 (1983).
25. T.K. Wassel and L. Himmel, TACOM-TR-12571 (1981).
26. E. Bonetti, G. Cocco, S. Enzo and G. Valdre, Mat. Sci. and Tech., 6 (12), 1258
(1990).
27. L. Guoxian, W. Erde and W. Zhongren, J. of Mat. Proc. and Tech., 51, 122 (1995).
28. C. Michealsen and E. Hellstern, J. Appl. Phys., 62 (1), 117 (1987).
29. L.M. Di, PI. Loeff and H. Bakker, Mat. Sci. and Eng., A134, 1323 (1991).
30. B.Q. Li and Y.N. Wang, J. Appl. Phys., 75 (3), 1783 (1994).
31. H. Hashimoto and R. Watanabe, Met. Trans. JIP, 31 (3), 219 (1990).
32. E. Gaffet, Mat. Sci. and Eng., 132 (2), 181 (1991).
33. K.B. Gerasimov, J. Mat. Sci., 26 (9), 1296 (1991).
34. P. Yang, J. Jpn. Soc. Powd., Powd Met., 37 (5), 623 (1990).
35. A. Lasonna and M. Magini, Acta Met., 44 (3), 1109 (1996).
36. E.S.B. Rao, R.M. Mallya and D.H. Sastry, Trans. IIM, 39, 596 (1986).
37. C.H. Lee, T. Fukunage and U. Mizutani, Mat. Sci. and Eng.., A134 1334 (1991).
38. C.H. Lee, T. Fukunage and U. Mizutani, Jpn. J. Appl. Phys., 29, 540 (1990).
39. C.C. Koch, D.B. Cavin, C.G. McKamey and J.O. Scarborough, Appl. Phys. Lett.,
43 1017 (1983).
40. K.Y. Wang, T.D. Shen, M.X. Quan and J.T. Wang, J. Mat. Sci. Lett., 11, 129
(1992).
41. A.W. Weeber, W.J. Haag, A.J. H. Wester and H. Bakker, J. Less-Common Metals,
140, 119 (1988).

Questions
1 Which are the mills in practice for MA?
Which mill is most energetic amongst them?
2. Why, in general, are MA mills energy inefficient.
3. Describe an attritor mill, explaining its salient features.
4. Suggest measures to improve efficiency of an attritor mill.
5. What is a uni-ball mill? Explain specific advantages associated with
it?
6. Why is a vibratory mill preferred for amorphization.
7. Give schematic of a planetary mill.
8. Write a brief note on grinding medium used for MA.
9. What advantages are associated with large diameter horizontal ball
mills used for mechanical alloying?
10. How is the use of grinding balls of different diameter helpful in
improving efficiency of an attritor mill?
11. Explain how the use of a uni-ball mill can avoid use of PCAs in the
case of ductile-ductile system?
12. Define MA. How is it different from MG?

24
Variations of Mechanical Alloying

3 VARIATIONS OF MECHANICAL
ALLOYING
3.1 REACTION MILLING
While most MA is conducted in an inert atmosphere, there can be ad-
vantages to MA in a reactive atmosphere. In reaction milling, metal
powders react extensively with the milling fluid that is reactive dur-
ing milling. For example, processing powders in nitrogen atmospheres
can result in the formation of nitrides or nitrogen can be trapped in
the metal matrix [1,2,3,4], while processing in hydrogen atmospheres
may form metastable hydrides [5].
The metal powder is comminuted to a particle size much smaller than
the starting one. In general, chemical reactions between the fluid and
the powder assist comminution of metal powders by not allowing them
to weld together or to the balls/chamber walls.
Reaction milling can also be used to obtain the desired dispersoid
by introducing an element that reacts during milling or during subse-
quent heat treatment, or in part during both. For example, MA of Al
and C (lamp black or graphite) results in an AlAl4C 3 composite powder
[6], and of AlBgraphite in an AlBC composite powder [7].
The process can also be used to produce nanostructured and amor-
phous materials, if milled for prolonged hours [1], e.g. when iron pow-
der is mechanically alloyed with nitrogen at room temperature, it re-
sults in a highly disturbed, non-equilibrium, microfine microstructure
with an enhanced nitrogen concentration entrapped in both the metal
matrix and the highly disturbed regions. During MA, the impact deforms
the particles plastically, creating a new surface. The newly exposed sur-
face is highly reactive to the nitrogen atmosphere and gets adhered to
the fresh surface, dissociates and subsequently is incorporated into the
matrix when the particles become cold welded. The nitrogen infusion
by MA results in a linear increase in nitrogen concentration for up to
100 hrs and a nitrogen level of up to 1.0% in pure iron (the interstitial
nitrogen solubility in bcc iron is less than 0.05%) (Fig.3.1). During
subsequent consolidation, the matrix nitrogen diffuses rapidly and be-
come trapped in the heavily deformed microstructure and at the subgrain

25
Mechanical Alloying

Fig.3.1 Nitrogen pick-up as a function of


MA time (Ref.1).

boundaries. This entrapped nitrogen retards the grain growth and leads
to a nanostructural material.

3.2 CRYOMILLING
With the recognition of the fact that cryogenic temperatures (which are
used to control homologous temperatures) can be used to control the
MA process [8], the concept of cryomilling came into existence. In
cryomilling, liquid nitrogen is added directly to the milling vessel and
the technique has been applied to systems ranging from aluminium to
nickel-base superalloys [9,10]. In the case of aluminium, extremely fine
Al(ON) particles are reported to be formed in addition to the aluminium
oxide present on the surface of the starting powder particles. Cryomilling
of Al50 at% Ti in liquid nitrogen results in the formation and decom-
position of B2 TiAl, and the formation of AlN and Al 2O 3 , which
impede the grain growth [11].

3.3 REPEATED ROLLING


Mechanical alloying using the repeated rolling technique (Fig.3.2) has
been accomplished by several workers [1214]. The mixture of elemen-
tal powders is packed in a stainless tube. The stainless tube which con-
tains the sample powder is pressed to half of its original thickness by
a hydraulic press, then cold rolled to a thickness of about 1/20th of
the original. The stainless sheath is removed and the rolled sample is
packed again in the stainless steel tube of the same size, pressed and
rolled. The process is repeated about 30 times.
During the repeated rolling process, the thickness of the sample
particle d after the n-th cycle of repeated rolling should roughly obey

26
Variations of Mechanical Alloying

Fig.3.2 Schematic of the repeated rolling


process (Ref.14: A.H. Clauer and J.J. de
Barbadillo (editors). Solid State Powder
Processing (1990), p.21. (The Minerals,
Metals and Materials Society)).

the equation of kneading,

d = d 0 (1/a) n (3.1)

where d0 is the initial particle diameter and 1/a is the reduction by rolling
of one cycle. However, due to changes of the stainless tube after each
rolling cycle, the kneading of powders may deviate to some extent from
the relationship given in Eq. (3.1). The sharp decrease in reduction of
particle size after several initial rolling cycles (Fig.3.3) suggests a
change in the kneading mechanism, probably by clipping of particles.
P.H. Shingu et al [14,15] have employed this technique for four binary
systems; AgFe, CuFe, AgCu and AlFe. The heat of mixing for
these alloy systems varies from large positive (AgFe) to moderately
large negative (AlFe). It was shown that due to the simple kneading
of elemental powder mixtures by the repeated rolling, metastable al-
loy phases such as nanocrystalline structure, supersaturated solid so-
lution and amorphous phases can be formed. These results indicate the
impact force present in an attritor or ball mill may be effective but not
the essential factor for solid state alloying. Thus, this technique can
further aid in understanding of the MA process.

Fig.3.3 Plot of the maximum observable


thickness of Ag phase vs number of
repetitions of rolling process (Ref.19).

27
Mechanical Alloying

Fig.3.4 Flow sheet of the MA process, with two possible routes (Ref.17).

3.4 DOUBLE MECHANICAL ALLOYING


In contrast to conventional MA, an alternate process called double
mechanical alloying (DMA) has been developed [16,17]. Figure 3.4
illustrates the flow sheet of the DMA route.
This process production of MA powder involves three main steps.
The first mechanical alloying MA1 step is applied to blend the el-
emental powders leading to close mixing in a deformed lattice having
high dislocation density to facilitate diffusion process which takes place
during the second step [18].
In the second step of the process, the MA1 powder is subjected
to high temperature heat treatment under a protective atmosphere pro-
moting intermetallic formation.
In the third step, the heat treated powders are submitted to a sec-
ond MA process (MA2), where the grain size of the powder being
milled is further refined.
The technique has been found to be very effective in the case of
high temperature aluminium alloys like AlFeMn and AlFeCe. High
temperature precipitation-hardened Al alloys loose their strength above
200C due to dissolution of the precipitate and recrystallization of the
material. An intermetallic Al6(Fe/Mn) -strengthened AlFeMn alloy, de-
veloped by rapid solidification, was found to have a good combination
of strength and ductility [19]. However, direct MA of Al5% Fe4%
Mn hardly results in the formation of any intermetallic. It results only
in refined dispersion of Fe and Mn in the Al matrix [20].
In the DMA technique, a fine distribution of hard Fe and Mn in soft
Al results during MA1. During subsequent thermal treatment at 550C
of this powder, a reaction occurs between Al and Fe, and Al and Mn.
The first phase formed is metastable AlFe. This intermetallic is sta-
bilized by Mn diffusing readily in Al and substituting Fe in the
orthorhombic Al 6 (Fe,Mn) intermetallic. The process is limited by the
diffusion rate of Mn and Fe in the Al matrix [16]. The size of the
intermetallic particles formed ranges from far below 1 m to a few m.
The properties of the processed powders are shown in Table 3-1. Ex-

28
Variations of Mechanical Alloying

Table 3.1 Properties of Al5%Fe4%Mn powders

Thermally
MA1 MA2
treated

Mean size (m) 97 97 39

Microhardness (KHN) 1.47 1.31 3.17

Density (g cm3) 2.83 2.85

Carbon ( %) 1.18 2.27

Oxygen ( %) 1

cellent elevated temperature tensile properties are obtained in the con-


solidated product due to the distribution of high volume fraction of
intermetallics and the stabilization of the matrix by inert dispersoids
(Al 2 O 3, Al4 C 3). The ductility of the alloy is low, i.e. less than 2% due
to the high amount of intermetallics.

3.5 REPEATED POWDER FORGING


The highest powder forging has also been used for MA to prepare non-
equilibrium and nanostructured materials. Using a novel machine of 500
ton capacity for repeated powder forging, MA has been achieved in
CuAg and a range of Al alloys [21].

References
1. J.C. Rawers and R.C. Doan, Met. and Mater. Trans., 25A 381 (1994).
2. M. Miki, T. Yamasaki, Y. Ogino, Mater. Trans. JIM, 34 (10) (1993) p.952.
3. H. Yasuda, Mater. Sci and Eng., A159, 676 (1994).
4. T.D. Chem, and C.C. Koch, Nanostruct. Mater., 5 (6), 615 (1995).
5. M. Baricco, J. Non-Crystalline Solids, 155, 156 & 527 (1993).
6. H. Danninger, G. Jangg and J. Zbiral, In: Solid State Powder Processing, H. Clauer
and J.J. deBarbadillo (eds), The Minerals, Metals and Materials Society,
Warrendale, PA (1990), p.241.
7. T. Takahashi and M. Motoyama, Preparation of particle dispersion strengthened
aluminium by MA, Presented at Powder Metallurgy World Congress, San Francisco,
CA, 21-26 Jan, 1992.
8. J.S. Benjamin and M.J. Bomford, U.S. Patent 3, 816, 080 (1974).
9. R.P. Luton and J. Valone, U.S. Patent 4, 647, 304 (1987).
10. M.J. Luton, Symp. Proc., Mater. Res. Soc. L.E. McCandlish (ed), Pittsburgh, PA,
132 79 (1989).
11. M.J. Luton, Nanostruct. Mater., 5 (6), 631 (1995).
12. M. Atzmon, K.M. Unruh and W.L. Johnson, J. Appl. Phys., 58, 3865 (1985).
13. F. Bordeau, A.R. Yavari and P. Desre, Mat. Sci. and Eng., A197, 129 (1988).
14. P.H. Shingu, K.N. Ishihara, K. Venishi, J. Kuyama, B. Huang and S. Nasu., Ibid
ref.6, p.21.

29
Mechanical Alloying

15. K. Uenishi, K.F. Kobayashi, K.M. Ishihara and P.H. Singhu, Mat. Sci. and Eng.,
A134, 1342 (1991).
16. P. Le Brun, L. Froyen and L. Delaey, Mat. Sci. and Eng., A157, 79 (1992).
17. X. Niu, P. Le Brun, L. Froyen, C. Peytour and L. Delaey, Powd. Met. Int., 25,
(3), 119 (1993).
18. G.B. Schaffer and P.G. McCormick, Met. Trans., 22A 3019 (1991).
19. M.A. Zaidi, J.S. Robinson and T. Sheppard, Mat. Sci. and Tech., 1 737 (1985).
20. P. Le Brun, X. Niu, L. Froyen, B. Munar and L. Delaey, Ibid ref.6, p.273.
21. J. Kihara, Productive Mechanical Alloying by Repeated Powder Forging, Presented
at the 1996 World Congress on Powder Metallurgy, Washington DC, WA, June,
1996.

Questions
1. Why are PCAs required during MA?
What other benefits are associated with the use of PCAs?
2. What are the commonly used PCAs in MA?
3. How PCAs affect MA powder hardness?
4. Why is the role of PCA most critical during the earliest stage of
milling?
5. What factors decide the amount of PCAs being used?
6. Why is degassing necessary for ductile-ductile MA powders?
7. What is the role of oxygen to carbon ratio in organic PCAs.
8. What contamination can PCAs cause? How can this contamination
be avoided?
9. In what way can excess use of PCAs be harmful?
10. How use of PCAs affects morphology of powder produced.
11. Give two examples to cite the fact that use of PCAs may lead to
structural changes in the powder.

30
Process Control Agents in Mechanical Alloying

4 PROCESS CONTROL AGENTS IN


MECHANICAL ALLOYING
In the case of ductileductile systems (e.g. AlMg), it is difficult to
use the MA technique because of the excessive cold welding. In such
cases, MA results in lumping of the alloyed material, which in turn sup-
presses the process of alloying. Organic surfactants called process
control agents (PCAs) are used to achieve the critical balance between
cold welding and fracturing, and enhance the process efficiency [1].
The PCAs help in preventing fresh surface contact by giving a surface
coating on powders. During the MA process, the PCA is embedded and
finely distributed among the layers of flaky powder. The PCAs also
help in alleviating the tendency of ductile powder particles towards
powder-to-ball/vial welding. A common problem encountered with MA
the contamination with elements contained in the mill vial and/or balls.
The level of contamination depends on the type of ball mill used, the
powder being milled and the milling conditions. The use of a surfactant
can also reduce such contamination by at least a factor of 10 [2]. Thus,
use of a PCA may be of great practical significance when contami-
nation with surfactant itself does not pose a problem.
Benjamin & Bomford used a number of organic compounds including
acids, alcohols and ketones [3], for this purpose. Table 4-1 lists the
commonly used PCAs during MA.

Table 4.1 Commonly used process control agents

Composition ( %)
PCA Chemical Formula
C O H N

Oxalic Acid (COOH) 2 2H 2O 19.1 76.2 4.7

Methanol CH3OH 37.5 50.0 12.3

Stearic Acid CH3(CH2)16COOH 76.2 11.2 12.6

Ethylene bis disteramide C2H 22(C18H 36ON) 77.3 5.4 12.4 4.8
(Nopcowax-22 DSP)

31
Mechanical Alloying

Fig.4.1 Effect of PCA identity and


level on welding (elemental Al2% Cu
alloy) (Ref.4, A.H. Clauer and J.J. de
Barbadillo (editors). Solid State Powder
Processing (1990), p.227).

The following points regarding application of PCAs have been ob-


served experimentally [4].
1. Approximately the quantity necessary to cover the whole surface
area of the final MA mixture with a molecular monolayer of PCA should
be added (1%) for optimized results. A larger amount may reduce the
welding process, hence the powder may become pyrophoric. It is in-
teresting to note that the pyrophoric nature of the powder product with
stearic acid added is about 2% (Fig.4.1). As stearic acid has the highest
level of carbon of the three PCAs, the phenomenon of pyrophoric is
exhibited.
2. The PCA can be in a state of solid, liquid or gas. A fluid PCA
reaches the required sites relatively quickly.
3. The efficiency of the PCA is most critical in the early stages of
processing (Fig.4.2).
4. The size and shape of the powder produced depends on the ratio
of oxygen to carbon present in the PCA. As this increases, the powder
size increases and the powder shape becomes more equiaxed (Table
4.2). It appears that hydrogen present in the organic molecular PCAs
plays no significant role as a process facilitator. During MA of soft
magnetic material (CoFe) 75 Si 15B 10 , the use of sodium 1,2 bis
(dodecylcarbonyl) ethane-1-sulfonate (cationic) provides a circular shape
(100 nm) and ammonium dihexadecyl dimethyl-acetate (anionic) pro-
vides a needle-like shape (100 nm) of MA powders [5].

Table 4.2 Effect of PCA identity on powder size and shape (Ref.4)

PCA
PCA oxygen/carbon Powder size Powder shape
ratio

Stearic acid 0.15 Fine Flaky

Methanol 1.33 Medium Disc

Oxalic acid 4.00 Coarse Equiaxed

32
Process Control Agents in Mechanical Alloying

Fig.4.2 Effect of PCA identity and milling time on welding (elemental Al2% Cu alloy)
(Ref.4).
Fig.4.3 (right) Effect of PCA identity and level on PCA powder hardness (elemental
Al2% Cu alloy) (Ref.4).

5. The hardness of the MA powder is also dependent on the type


of PCA used (Fig.4.3) in the MA.
6. The use of the PCA may also lead to structural changes in the
MA powder, e.g. the use of sodium 1,2 bis (dodecyl carbonyl) ethane-
1-sulfonate increases the solid solubility of magnesium in aluminium to
13 at %, which is much more than the equilibrium solubility (about 1
at %) in a mixture of Al 70Mg30 (Fig.4.4) [2], (the alloying is judged
by the shift of aluminium peaks). The type of PCA used also affects
the structure produced, e.g. when Al50Mg 50 is mechanically alloyed for
160 hrs, it produces an amorphous-like structure. After annealing for
5 min at 350C, this structure transforms into a mixture of two equi-
librium phases, (Al 3Mg2 ) and (Al 12 Mg17 ). When different surfactants
are used for MA, the final product changes [2] drastically (Fig.4.5).
In general, the use of organic PCA contaminates the resulting powder
with residual carbon and oxygen [1,6], if a reactive element present

Fig.4.4 X-ray diffractograms of mechanically alloyed


elemental mixtures of AlMg (helium atmosphere;
80 hrs): a) alloyed without surfactant; b) alloyed with
the addition of sodium-1,2 bis (dodecyl carbonyl)
ethane-1-sulfonate (Ref.2, Mat.Sci. and Eng., A134,
1346 (1991)).

33
Mechanical Alloying

Fig.4.5 X-ray diffractograms of mechanically alloyed


elemental mixtures of Al 50-Mg 50 (helium atmosphere,
160 hrs: a) without surfacant; b) alloyed with addition
of sodium-1,2 bis (dodecyl carbonyl) ethane-1
sulfonate; c) dodecyloxycarbonyl sulfosuccinate; d)
alloyed with the addition of dodecyldimethylammonium
bromide (Ref.2).

in the composition being milled may react with such contaminates [7].
Organic PCAs also form hydrogen during MA which remains in the
granulates, necessitating a degassing step. In an effort to reproduce
high purity MA material, Gualandi [8] altered this practice by using sili-
con grease.

References
1. P.S. Gilman and W.D. Nix, Met. Trans., 12A, 813 (1981).
2. A.P. Radlinski, A. Calka, B.W. Ninham and W.A. Kaczmarek, Mat. Sci. and Eng.,
A134, 1346 (1991).
3. J.S. Benjamin and M.J. Bomford, U.S. Patent 3.816, 080 (1974).
4. J.H. Weber, In: Solid State Powder Processing, A.H. Clauer and J.J. deBarbadillo
(eds), The Minerals, Metals and Materials Society, Warrendale, PA (1990), p.227.
5. W.A. Kaczmarek, R. Bramley, A. Calka and B.W. Ninham, In: Conf. Proc. Intermag
90, Brighton, UK, April 1990.
6. F. Faudot, J. Mat. Sci., 28, 2669 (1993).
7. M. Nose, J. Jpn. Soc. Powd. Met., 42 (2), 166 (1995).
8. D. Gualandi and P. Johenson, In: Modern Developments in Powder Metallurgy,
H.H. Hausner (ed.) V.3, Plenum Press, New York (1966), p.36.

Questions
1. Describe MA by repeated rolling.
2. What is cryomilling?
3. For which various purposes can reaction milling be used?
4. How is reaction milling able to extend nitrogen solubility in iron?
5. What is DMA? How is it beneficial?

34
Mechanisms in Mechanical Alloying

5 MECHANISMS IN MECHANICAL
ALLOYING
The MA technique is being used mainly for three types of process-
ing: alloying, metastable phase formation and activation of chemical re-
actions. The underlying mechanisms involved in these processes to the
extent known today are discussed here.

5.1 ALLOYING
To discuss the mechanism of MA, it is convenient to divide the powder
charge into three systems:

ductileductile
ductilebrittle
brittlebrittle

5.1.1 DuctileDuctile System


When both components are ductile, according to Benjamin and Volin
[1], a balance of plastic deformation, cold welding and fracture lead

Fig.5.1 Stages of MA.

35
Mechanical Alloying

to the final microstructure. The alloying occurs in five stages as shown


in Fig.5.1. In the first stage, when the particles start getting commi-
nuted, the malleable components are deformed into long lamellae by
the impact of the balls while the more friable components are com-
minuted. This is followed by growth in the number of lamellae due to
cold welding; these composite lamellae in the coarser particles have
a multilayered, oriented (plate-like) structure. The next stage is asso-
ciated with the reduction in the aspect ratio of the lamellae and the
plate-like coarse particles becoming equiaxed. In the fourth stage, the
welding orientation in the composite particles becomes random and con-
voluted. The final stage is characterized by a narrow particle size dis-
tribution and the composition becoming uniform. At this stage, the in-
dividual lamellae cannot be resolved in an optical microscope. A satu-
ration level of hardness of the particle is attained and the particles are
in the heavily cold worked state. Regarding the atomistic mechanisms
that could be responsible for MA, the presence of deformation bands
[2], which do not demonstrate in situ recrystallization in MA ductile
materials, suggest a collision induced temperature rise (at most a few
hundred degrees) is not high enough to account for the estimated dif-
fusion rate of the solute. It is therefore to be surmized that, by some
mechanism, the diffusivities are enhanced. One possibility is the role
of increased dislocation densities due to severe cold working result-
ing from MA, which enhance diffusion by providing many sites (tunnels)
for diffusion under low activation energy, through pipe diffusion as also
through higher vacancy concentrations. As the MA proceeds, plastic
deformation thins the lamellae of the powder particles being processed.
It decreases the diffusion distance, which further enhances the disso-
lution of the solute and homogenization. Thus, solute dissolution during
MA is facilitated by three factors, namely:
local heating
lattice defects
shortened diffusion distances

5.1.2 DuctileBrittle System


In ductilebrittle systems, as in the case of TiSi and Y 2O 3 in super-
alloys, MA usually results in a fine, homogeneous dispersion of brit-
tle phase in the ductile matrix. The process for this system also fol-
lows all five stages of ductileductile systems, with the only differ-
ence being that the dispersoid present in the constituent is believed to
be entrapped along the cold-welded interface [3,4] and its concentration
along the weld seams gradually decreases as the welds increase in
number and get randomized (Fig.5.2).
When the brittle constituent accounts for approximately half the
36
Mechanisms in Mechanical Alloying

Fig.5.2 Dispersoid distribution during


MA.

volume fraction, the characteristic layered structure does not develop,


instead both the constituents are reduced to nanometer sized crystallites
and are evenly distributed throughout the powder [5].

5.1.3 BrittleBrittle System


In the case of brittlebrittle systems like GeSi systems, intermetallic
compounds in the MnBi system and amorphous phases from
NiZr 2Ni 11Zr 9 intermetallic, the mechanism of MA is not well under-
stood yet. A granular morphology is observed in the process of MA
[6]. There is an initial period of processing which results in very lit-
tle change in the lattice parameters of the two components. Follow-
ing this initial period, the lattice parameters shift and eventually con-
verge, indicating complete alloy formation (Fig.5.3). The disappearance
of granular-type morphology and the appearance of approximately
equiaxed particles, embedded in the major constituent, with a homo-
geneous microstructure occur at the same milling time where alloying
is complete according to the lattice parameter results.
A phenomenon of interparticle necking (Fig.5.4) has been observed
in MA of such systems, suggesting diffusion under a sufficiently high
collision induced temperature rise in contrast to the ductileductile sys-
tem.
Such a large temperature rise may be possible in the light of findings
of Miller et al, who have showed that a local temperature ranging
between 402 to 502C can be attained in impacted non-metallic crys-
talline materials such as NaCl and ammonium perchlorate [7] (impact

37
Mechanical Alloying

Fig.5.3 Lattice parameter vs milling time for silicon and germanium powder for the
composition Ge72 at.% Si (Ref.6, Met. Trans., 19A, 2867 (1988). (The Minerals, Metals
& Materials Society)).
Fig.5.4 (right) SEM micrograph of interparticle 'necking' in Ge72 at.% Si after milling
for 8 hrs (Ref.6, Met. Trans., 19A, 2867 (1988)).

heating due to localized shear deformation, as in metals, may not be


possible in brittle systems).
Thus, MA of brittle components can be assumed as a thermally ac-
tivated process. The assumption has also been confirmed by the sup-
pression effect in the alloying process when carried out at cryogenic
temperatures using liquid nitrogen [5].
The presence of flat platelets in the mechanically alloyed powder
of such materials, in the absence of noticeable slip in these elements
at or near room temperature, indicates the temperature-enhanced de-
formation, provided the temperature approaches 0.5 Tm. However, large
scale deformation does not occur as evidenced by the absence of a
lamellar microstructure.
In the alternative approach to deformation, it is assumed that be-
low 0.5 T m, very little plastic deformation occurs and failure occurs due
to defects such as microcracks. This assumption is based on the finding
that the brittle materials show a brittle behaviour macroscopically but
plastic deformation at microscopic level [8]. Under these conditions
along with elevated temperatures, yielding could occur before fracture.
Numerous surface asperities are surely present in brittle materials.
Upon impact, these asperities of the respective powder particles would
penetrate the surface of an adjacent particle. In this way, plastic de-
formation and bonding might occur. The surface deformation could also
be aided by the presence of preferentially active surface dislocation
sources. With decreased critical shear stress for dislocation generation
at the surface and enhanced mobility provided by somewhat elevated
temperatures, plastic flow may be possible, which could lead to the ob-
served MA.

38
Mechanisms in Mechanical Alloying

Plastic flow during MA in the brittle materials can also be under-


stood in the light of the findings of Bridgman [9,10], who showed that
the application of superimposed shear or tensile stress in the presence
of hydrostatic stress can lead to plastic flow in brittle materials. During
the MA process, powder particles are subjected to impact (compress-
ive) forces when entrapped between the balls. Considering that many
randomly arranged particles in close contact are involved in a single
collision probably triaxial stress states may develop in certain regions
of the compact which may have significant hydrostatic components. This
could lead to MA via enhanced deformation induced by the hydrostatic
stress components.
Another possible mechanism involves alloying by material transfer
during frictional wear. This is qualitatively consistent with the relatively
long processing times observed in brittlebrittle systems compared to
ductileductile powders.
Thus, the mechanisms responsible for material transfer during MA
of brittle components may include plastic deformation induced by:

local temperature rise


microdeformation in defect-free volumes
surface deformation
hydrostatic stresses, or by a frictional wear mechanism.

These temperature induced microdiffusions or deformations may link


the mechanisms thought responsible for ductile systems to brittle sys-
tems.

5.1.4 Idealness of MA Alloys


To identify whether MA alloys are true alloys on an atomic scale or
merely a homogeneous mixture at a micron scale, J.S. Benjamin himself
and others examined these alloys with the help of magnetic measure-
ments [11,12] and the X-ray technique [13,14], and have demonstrated
the true alloying. In contrast to these facts, White and Nix [15] and
other investigators found in the NbSn systems that MA was merely
an intimate mixture of the constituents and not true alloying. Thus, re-
garding the true-alloying there is no complete agreement. Wassel and
Himmel [13] studied this problem in the CrMo system by X-ray dif-
fraction. However, they observed chemical inho-mogeneity in this com-
position. This could be attributed to their using non-optimized milling
parameters.
A detailed investigation of the problem was carried out by E.S.B.
Rao and his colleagues using nickelchromium as a model composi-
tion. Magnetic coercivity, X-ray line breadth analysis, electron probe
39
Mechanical Alloying

microanalysis (EPMA) and Auger electron spectroscopy (AES) have


been used to follow the state of alloying in the attritor milling of Ni-
20%Cr2%ThO 2. These studies have revealed that under the conditions
of effective milling, magnetic coercivity initially rises to the peak value
and drops down, while no such peak is observed if the milling con-
ditions are not conducive for effective alloying (Fig. 5.5). (This experi-
ment is a sensitive test as nickel is a ferromagnetic, but nickel alloy
with a chromium content of more than about 8% is paramagnetic). The
X-ray peak shifts, after substruction of the strain broadening, also sug-
gest that true alloying could be occurring. The EPMA of mechanically
alloyed particles clearly revealed the state of homogeneity of alloying
on a micron scale (Fig.5.6). Another interesting set of experiments using
AES showed that, with continuous etching of the surface atomic layers
of the mechanically alloyed particles, the composition remained homo-
geneous, confirming the atomic levels of mechanically alloyed parti-
cles (Fig.5.7) [16,17].
Thus, MA produces true alloys with a homogenous composition
provided that the milling conditions are properly optimized.

5.2 METASTABLE PHASE FORMATION


5.2.1 Amorphization
Since the discoveries by C.C. Koch et al (1983) of elemental alloy
Nb 40Ni 60 by MA [18] and later amorphization of intermetallic by MG
(1986) [19], extensive work has gone into the study of amorphous phase
formation by MA in the recent past (other methods for solid state
amorphization are hydrogen dissolution into a crystalline phase,
interdiffusion of two crystalline metals accompanied by a large negative
heat of mixing, and irradiation). In the case of MA, structural devel-
opment takes place due to materials transfer but in the case of MG
of elements or intermetallics no materials transfer accompanies the struc-
tural changes during ball milling.

Fig.5.5 Magnetic coercivity as a function


of milling time at low and high attritor
speeds (Ref.17).

40
Mechanisms in Mechanical Alloying

Fig.5.6 The homogeneity as


revealed by EPMA line scans of:
a) mechanically alloyed elemental
nickel and chromium powders; b)
atomised NiCr pre-alloyed powder
(Ref.17).

Fig.5.7 Homogeneity of composition of


mechanically alloyed particles as
revealed by the depth profile analysis
of AES by sputter etching (Ref.17).

Amorphization by MA
Several theories have been suggested for the mechanism of the amor-
phization phase formation by MA. These include:

1. Liquid quenching (LQ) model


This model assumes the occurrence of incipient melting under the im-
pact of colliding balls [20,21] and ultrarapid quenching of the partly
molten powder particles. It is natural to think that the severe plastic
deformation of the powder particles trapped in ball collisions may locally
raise the temperature of the particles sufficiently high to melt, which
subsequently solidifies rapidly by heat conduction into the less-deformed
and thus cooler regions of the particles. If this mechanism is present,
the amorphous phase would form by repeated rapid solidification (RS)
processing. However, two observations oppose this simple mechanism.
First, the composition range of the alloying does not agree with the
range obtained by RS at a cooling rate as high as 10 6 K/sec. Second,
although there is no direct measurement of the temperature change in
the particles during ball collisions, calculations [19] suggest that the
increase is at most a few hundred Kelvin, which is not sufficient for

41
Mechanical Alloying

partial melting of powder particles.

2. Solid state amorphization reaction (SSAR) model


This model assumes MA solid state amorphization as in diffusion couples
with a high negative heat of mixing and an anomalous high diffusion

Table 5.1 Typical list of amorphous alloys formed by MA (Ref.22 and 29).

Hmix Atomic
Alloy A 1xBx
(kJ/mole)
(kJ/mol) size ratio

HfAl 40 1.17

HfCu 23 1.30

HfNi 44 1.35

NbCuGe 2 1.14

NbCuSi 2 1.14

NbNi 32 1.18

PdSi 37

SnNi 4 1.31

SnNb 5 1.11

TiCu 18 1.15

TiNi 39 1.19

ZrCo 42 1.28

ZrFe 26 1.27

ZrNi 51 1.29

ZrV 4 1.19

CuW +33 0.908

CuTa +3 0.870

CuV +7 0.950

NiW 5 0.855

CuAl 2 0.892

TaAl 46 1.02

NbAl 44 1.03

TiNiCu

AlNiFeGd

42
Mechanisms in Mechanical Alloying

coefficient [19,21]. The metalmetal amorphous alloys that have been


formed by MA are given in Table 5.1.
These alloys can be put into two categories:
Metalmetal type (ductileductile) alloys consisting of early tran-
sitional metal (Ti, Hf, Zr, Nb) and late transitional metal (Fe, Co, Ni,
Cu);
Metalmetalloid type (ductilebrittle) alloys like CuNb(Si, Ge
or Sn), CuV(Si or Ge), Nb(Si or Sn), PdSi systems, PdNP and
PtNiP systems.
In most of the binary systems, the two elements have the follow-
ing two important characteristics [22]:
The two elements have a large negative heat of mixing in the amor-
phous (liquid) state,
The two elements have vastly different atomic sizes.
The difference in the atomic sizes is thought to be responsible for
the observation that the chemical diffusivity of the two elements in each
other and in the amorphous alloys in general differ by several orders
of magnitude. Schwarz and Johnson [23] proposed that these charac-
teristics are necessary for the formation of amorphous alloys by
interdiffusion at the boundary between two crystalline metals A and B
as in the thin film coupled systems, i.e. large negative driving force
of the reaction and kinetic suppression of intermetallic formation.
The formation of the metastable glassy phase then takes place if the
kinetics of transformation are such that the amorphous phase occurs
much more quickly than the equilibrium crystalline phase. The shear
strain spinoidal transformation makes the crystalline solid unstable and
leads to amorphous phase formation at a milling temperature below the
glass transition temperature [24].
The plastic deformation, cold welding and fracturing of the parti-
cles during MA generate a large concentration of clean metalmetal in-
terfaces, the negative heat of mixing provides a thermodynamic driv-
ing force for the reaction enhanced by the excess point, and lattice de-
fects generated by plastic deformation and by the momentary tempera-
ture increase in the powder particles trapped in ball collisions. A large
difference in chemical diffusivity of the elements in each other and in
the amorphous phase favours kinetically the formation of the amorphous
alloy in preference to the formation of more stable crystalline inter-
metallic compounds.
The operation of a SSAR during MA is also supported by the for-
mation of the reaction products of MA which are in good agreement
with the calculations based on the Phase Diagram method (CALPHAD)
[20,25,26]. Figure 5.8a shows a schematic phase diagram for a binary
system AB with a negative heat of mixing in the liquid state [21].
43
Mechanical Alloying

Phases and are crystalline primary solid solutions and phase is


the crystalline intermetallic. Figure 5.8b shows the free energy of ,
and , and of the amorphous phase () evaluated at the reaction tem-
perature T r . The free energy of the starting mixture of A and B is
located along the straight line joining the free energies of pure crys-
talline and . If the interdiffusion reaction of the A/B interfaces
reaches a state of thermodynamic equilibrium, the reaction products
would be determined by the common tangents between , and in
Fig.5.8b. However, by choosing a relatively low reaction temperature,
the nucleation and growth of the crystalline phase is prevented, while
still allowing atoms A and B to intermix. This is possible because, owing
to their atomic size difference, the chemical diffusivities of A and B
in each other and in the amorphous alloy are vastly different. In the
absence of -phase, the reaction products are determined by the com-
mon tangents between phases , and (Fig.5.8b). These tangents
predict five possible reaction products:
a crystalline solid solution, for 0 < x < x 1;
a two phase product of (x 1 ) and (x 2);
a single-phase amorphous alloy, for x 2 < x < x 3;
a two-phase product of (x 3) and (x 4 );
a crystalline solid solution, for x 4 < x <1.
The free energy diagram predicts that the composition range of the
single-phase amorphous alloy prepared by a solid state reaction is wide
and continuous and is located near the centre of the composition range,
which in fact has been found correct in most experimental studies.
In MA, deformation energies are usually small compared with the

Fig.5.8 Schematic phase diagram (a) for a


binary system, AB, with a negative heat of
mixing in the liquid state and a corresponding
free-energy diagram (b). At the temperature
T r, the bars at the bottom of the figure give
the homogeneity ranges of the amorphous
phase for an alloy prepared by MA and RS
(Ref.21, Metal Powder Report 43 (4), 231
(1988), (Elsevier Science)).

44
Mechanisms in Mechanical Alloying

heat of mixing [27] and therefore the stored energy of cold work is
believed to play only a minor role. It is well-known, however, that SSAR
does not take place in bi-layers made of defect free single crystals and
that the nucleation of amorphous phase usually starts at an interface
near a triple point defect [20]. The presence of structural defects which
are thermally stable at the temperature where SSAR takes place, is there-
fore a necessary condition in order for the amorphization to proceed
in mechanical alloying.
However, this SSAR model based on amorphization in thin films does
not explain the amorphization in systems with H mix>0. In the case of
AB thin film multilayers, it has been found experimentally that com-
pound nucleation and growth take place when the amorphous layer
thickness grows beyond a critical thickness x c. The growth in the critical
thickness x c has been correlated with the ratio of diffusivities of the
D
two species ( xc a B ). A.R. Yavari et al [28] reviewed this for the
DA
MA case in the light of the fact that the thermodynamic driving force
for crystallization disappears with the large concentration gradients is
shown to depend on the heat of mixing H mix . Neither H mix < 0 nor
D B >>D A are required for amorphization by MA. The critical thickness
x c is shown to be 2nm for the alloys with H mix=0, such that amor-
phization by solid-state remains insignificant (in the case of thin films).
It is shown, however, that in MA, amorphous layer growth is accom-
panied by a simultaneous thickness shrinkage due to mechanical de-
formation. This allows the amorphous interlayer thickness to remain
below the critical value x c whilst its volume fraction increases.
In another approach to explain the SSAR by MA, Y. Chakk et al
[29] developed an atomistic model based on the assumption that
amorphization is obtained when a solute atom penetrates into the in-
terstitial sites and distorts the lattice. When such distortions, even when
the chemical diffusion is of less significance (i.e. H mix>0 for systems
like CuTa), reach some critical value, the long-range order of the lattice
is destroyed and the amorphous phase is obtained. It was established
that for such a mechanism to be operative, lower and upper limits of
the atomic size ratio (ASR) and the minimum solute concentration can
be established (refer to Section 9.3).
This idea is demonstrated in Fig.5.9. It is divided into four regions.
The first region (I) includes systems which are characterized by a low
negative enthalpy of mixing and a small difference in atomic radii, as
well as pure metallic elements with H mix = 0 and r A/r B = 1. In this
region, the systems are not amenable to amorphization by MA. The
second region (II) corresponds to a low negative enthalpy of mixing

45
Mechanical Alloying

Fig.5.9 Two-parameter scheme ( H mix vs


ASR) to analyze the amorphization tendency
of bi-elemental metallic system by MA
(Ref.29, Acta Metall. and Mater., 42, 3679
(1994) (Elsevier Science)).

but a large deviation from unity of the ASR. In this region all the sys-
tems were successfully amorphized by MA and those systems having
low negative or positive heat of mixing could not be amorphized in dif-
fusion couples. The third region (III) contains systems which are char-
acterized by a high negative enthalpy of mixing and the ASR is close
to unity. In this region, the systems were amorphized by MA. The forth
region (IV) is characterised by a high negative enthalpy of mixing and
a large atomic size mismatch. In this region the systems were
amorphized by both MA and SSAR of diffusion couples.
The ASR criterion should be regarded as a preliminary step for
predicting the ability to amorphize a bi-elemental system by MA. The
next step was to determine the atomic concentration range where the
amorphization can be observed.
The maximum strain energy is found to be only about 2% of the
total energy stored and reveals that most of the energy of the cold work
must be stored in the grain boundaries (of less mobile species) and not
in the lattice strain [30]. However, the strain energy has been found
to represent about half of the enthalpy of crystallization of the amor-
phous alloy powder. It therefore can contribute significantly to the
amorphization reaction (however, it represents only 8% of that of fusion
and 4% of the heat of mixing). Also the rate of amorphization has been
found to be maximum when strain is maximum, it is believed that lattice
strain plays an important role in the amorphization reaction.

Amorphization by MG
Amorphous alloy powders have also been prepared by MG of single
phase intermetallic powder [30,31], mixtures of intermetallic powders
[19] (Ni 32TiNi 45Nb55, NiTi 2NiNb) and mixtures of intermetallic and
elemental powders [32]. (It is surprising that the same amorphous alloy

46
Mechanisms in Mechanical Alloying

powder can be synthesized by ball milling a mixture of elemental pow-


ders or by ball milling powders of an intermetallic compound). The MG
raises the free energy of the intermetallic to a value equal to or larger
than that of the amorphous phase (the reaction from point 3 to point
2, Fig.5.8b). This increase is thought to occur by the chemical disor-
dering of the crystalline lattice, and by the accumulation of point and
lattice defects [19,33].
Ball milling causes the lattice of the intermetallic to dilate, which
occurs as point and lattice defects are introduced and the lattice is
chemically disordered (the chemical disordering replaces strong AB
atomic bonds with weaker AA and BB bonds). If the dilation controls
the rate of dynamic recovery, then amorphization will be difficult to
achieve in crystalline compounds where dynamic recovery limits the
defect density and dilation to a value lower than that necessary for the
crystal-to-amorphous transformation [34]. (For example, in the MA of
nickel and titanium powders to form amorphous Ni 1-x Ti x with
(0.3<x<0.7), Schwarz et al [20] found no evidence of formation of
intermetallic compounds).
It has also been found that the amorphization reaction is associated
with a relaxation process of strain in less moving species of the couple
[30,35]. In Fig.5.10, the relative long-range order LRO parameter S
is plotted against the logarithm of the milling time for mechanical milling
of Ni 3Al. The hardness reaches a maximum value (1 hr milling) and
then due to relaxation progress drops and remains constant. The relative
LRO parameter is calculated after Carpenter and Schulson [36] as

Fig.5.10 Relative long-range


order parameter, S, and micro-
hardness as a function of milling
time (Ni3Al), (Ref.35, A.H. Clauer
and J.J. deBarbadillo (eds), Solid
State Powder Processing (1990),
p.49)).

47
Mechanical Alloying

LM bI / I g
s t t
p
OP 0. 5

S=
MN bI / I g
s t t =0
p
PQ (5.1)

where (I s /I t) is the ratio of intensity of the superlattice to the funda-


mental line at a given milling time tp .

Transformation path
Kim and Koch [37] reported that during the MA of niobium and tin
powders in a molar ratio 3:1, the Nb 3Sn intermetallic (A15 structure)
forms first and with continued ball milling, transforms to the amorphous
Nb 75 Sn 25 alloy. A direct amorphization reaction was also observed by
Tiainen and Schwarz during the ball milling of pure nickel and tin pow-
ders [38]. Likewise, several amorphous alloy powders have been pro-
duced by MA. It is thus logical to ask whether during the MA of a
mixture of pure powders A and B, the amorphous alloy (AB) am forms
by the direct reaction

( A+B) (AB) am

or through the indirect reaction

(A+B) (AB) cr (AB)am .

In the case of MA, the amorphization path is governed by milling


conditions, like the impact energy, atmosphere, temperature, type and
level of contamination and the PCA used (as discussed in Section 2.3
and Chapter 4).
It has been established that in the case of intermetallic phases (exists
in a significant range of compositions, e.g. Nb 3Sn, Nb 3Ge and Ni 3Al)
MG first results in formation of the equilibrium, or high defected
metastable crystalline compound, which is then destabilized by further
milling resulting in an amorphous phase, e.g. in compound Ni 3 Al, the
ordered L12 structure transforms to fcc Ni3Al solid solution which then
become partically amorphous. In the case of line compounds (negligible
range of composition beyond stoichiometry, e.g. NiZr 2 and Ni11 Zr9 ),
relatively short milling times result in crystalline intermetallic particles
embedded in an amorphous matrix. It appears that the growth of the
amorphous phase in this case occurs at the crystal/amorphous phase,
with a characteristic halo around the crystalline particles (Fig. 5.11)
[39].

48
Mechanisms in Mechanical Alloying

Fig.5.11 TEM micrograph of selected area exhibiting amorphous structure in Ni3 Al


powder, milled for 50 hrs. Inset shows corresponding diffraction pattern (Ref.35, A.H.
Clauer and J.J. deBarbadillo (editors)), Solid State Powder Processing (The Minerals,
Metals & Materials Society)).

5.2.2 Nanocrystallization
When nanocrystallization in pure metals is carried out by MA, the heat
enthalpy (fusion) (625%) and heat capacity (814%) are raised due
to the high density of grain boundaries and the incorporation of lat-
tice defects into the crystal, which results in lattice softening and in-
creased unharmonicity in the lattice [40]. The enthalpies stored due to
extensive ball-milling are considerably higher than for other known cold
working methods of metals and alloys, which are rarely found to ex-
ceed 1 to 2 kJ/mol and are never more than a small fraction of the
heat of fusion H f [41]. In the case of the ball milling process, the
enthalpy determined is considerably larger and can reach values typical
for crystallization enthalpies of a metallic glass (0.4 Hf ). The maxi-
mum dislocation densities measured in heavily deformed metals are less
than 10 13 cm 2, which would correspond to an energy of less than 1
kJ/mol. Typical energies determined experimentally by differential thermal
calorimetry for equilibrated high-angle grain boundaries of high melt-
ing point metals are 10 4 Jcm 2 [42]. Assuming that the two top
monolayers of the subgrains belong to a grain boundary, about 50% of
the atoms of a nanocrystalline material with an average grain size
boundary thickness of 15 to 20 corresponds to an energy of roughly
70 J/cm 3 or 1 kJ/mol. As a consequence, the grain boundary energy
of the ball-milled nanocrystalline solids seems to be considerably larger
than the grain boundary energy of fully equilibrated grain boundaries.
The process of the grain size reduction can be understood in the
following way. In general, plastic deformation proceeds by slip and

49
Mechanical Alloying

Fig.5.12 Atomic resolution TEM bright-field image including the corresponding diffraction
pattern of Fe powder after 24 hrs milling (Ref.40, Met. Trans., 21A, 2333 (1990)).

twinning at low and moderate strain rates, whereas at high strain rates,
the formation of shear bands, which consist of a dense network of dis-
locations, becomes the dominant deformation mechanism. These shear
bands in which the deformation is localized have a typical width of 0.1
to 1 m [2]. The local shear instability of the crystal lattice may be
triggered by material inhomogeneities and is probably enhanced by ther-
moplastic instabilities due to non-uniform heating in certain regions during
the deformation process. In the early stage of ball-milling, the aver-
age atomic level strain increases due to the increasing dislocation density.
At a certain dislocation density, within these heavily strained regions,
the crystal disintegrates into subgrains which are initially separated by
low-angle grain boundaries with offset angles of less than 20. Dur-
ing processing, deformation occurs in shear bands located in previously
unstrained parts of the material. The size of the subgrains in the ex-
isting band is further reduced to the final grain size, and the relative
orientation of the subgrains with respect to each other ultimately be-
comes completely random (Fig.5.12).
The nanocrystalline grains themselves are relatively dislocation free,
as suggested by several factors [43]. Once an entirely nanocrystalline
structure is achieved, further refinement seems to be impossible. The
very high stresses required for dislocation movement hinder plastic
deformation of very small crystallites. Hence, further deformation and
energy storage can only be accomplished by a glide along the grain
boundary, which results in a random rotation of the subgrains.
In the case of an alloy system, the presence of the alloying element
[44] plays an important role to obtain nanosize crystals. As shown in
Fig.5.10 the strain in the material increases considerably during the mill-

50
Mechanisms in Mechanical Alloying

ing. The major contribution to the lattice strain comes from the high
density of dislocation [27]. The final fracture of the crystal will probably
occur by a shear in this highly distorted region of the crystal.
The fact that even for extended milling the elemental powders ex-
hibit a larger grain size and a smaller r.m.s. strain than the interme-
tallic compound [45] suggests that nanocrystalline materials are more
easily obtained for alloys than for the pure metals.
Thus, it can be said that, if from the thermodynamics point of view
the amorphous phase is significantly less stable, the desired
nanocrystalline phase can be formed by MA. In particular, for inter-
metallic compounds, the rise in free energy due to chemical disorder
during milling has to be taken into account. Therefore, intermetallic
compounds exhibiting a large difference in the free energy with respect
to the amorphous phase are most favourable for the preparation of
nanocrystalline microstructures. A transformation into a nanocrystalline
metastable solid solution can occur if this phase has a higher stabil-
ity than the amorphous phase, as has been demonstrated in the case
of NbAl [45].
However, it should be realized clearly that nanocrystalline structures
do not form in alloy systems within the field of complete solid solu-
bility. In such cases, a fine lamellae structure is produced by MA. By
varying the composition and milling conditions, either an equiaxed crys-
talline structure with grain size less than 10 nm or an amorphous struc-
ture results. Nanocrystalline cermets can be produced by MA of metallic
powders and metalloids. In mutually immiscible, e.g. AgFe, systems
MA results in nanocrystallization with the extension of solid solubil-
ity rather than amorphous phase.

5.2.3 Solid Solubility Extension (SSE)


The achieving of the SSE metastable phase can be understood in the
light of discussions regarding Fig.5.8 as in Section 5.2.1.

5.3 ACTIVATION OF SOLID STATE CHEMICAL REACTION


Most solid state reactions (e.g. 4CuO+3 Fe4Cu+Fe 3O 4) involve the
formation of one or more product phases between the reactants, and
the reaction volume is continuously diminished as the reactants become
spatially separated. Reaction rates are therefore influenced by initial
contact areas and hence, particle size, and by the diffusion of the re-
actant species through the product phases. Factors which influence dif-
fusion rates, including defect structures and densities, local tempera-
tures and product morphology, have an important effect on reaction ki-
netics.
Mechanical alloying significantly increases solid state reaction rates
51
Mechanical Alloying

by dynamically maintaining high reaction interface areas [46] and si-


multaneously providing the conditions for rapid diffusion [47].
During MA, plastic deformation, welding and fracturing of powder
particles take place continuously. Plastic deformation and fracture of
powder particles create atomically clean surfaces. Particle welding can
occur when such surfaces are impacted during subsequent collision and
reactions can proceed across these new, internal interfaces. Hence the
chemical composition of powder particles changes during milling.
In the reduction of metal oxides by solid reducing agents [46-49],
which are highly exothermic redox reactions, MA causes an unstable
reaction which proceeds by the propagation of a combustion wave
through the partly reacted powder. In this reference, two critical tem-
peratures need to be defined:
critical reaction temperature T crit;
ignition temperature T ig.
T crit is the temperature at which the temperature difference between
the specimen and reference starts increasing due to the heat generated
by the reaction, i.e. when dT/dt > 0. T ig is the temperature that is re-
quired for the onset of combustion. Both T crit and T ig were found to
decrease with milling time [49], which can be attributed to an increase
in surface area of the powder system due to a decrease in crystal-
line size.
In addition to the energy dissipated by the collision, the reaction
enthalpy also contributes to the temperature rise during such milling,
and ultimately causes instability and combustion when the powder tem-
perature reaches the critical value. Figure 5.13 shows a schematic of
variation in the critical reaction temperature and local powder tempera-
ture maxima with milling time. Curve (a) shows the change in T crit with
increasing milling time, whilst curve (c) gives the variation in local pow-
der temperature. The combustion condition is reached after an incu-

Fig.5.13 Schematic showing variation in the critical reaction temperature maxima with
milling time (Ref.50, Met. Trans., 23A, 1285 (1992) (The Minerals, Metals and Materials
Society)).

52
Mechanisms in Mechanical Alloying

bation milling time tig required for the ignition temperature of the powder
mixture Tig to be reduced to a level equal to that achieved locally during
collision events. By changing the milling parameters (i.e. milling speed,
BPR, ball size), tig the incubation period can be changed but the critical
temperature and the ignition temperature remain unaffected for a system
[50].

References
1. J.S. Benjamin and T.E. Volin, Met. Trans., 5, 1930 (1974).
2. E. Hellstern, H.J. Fecht, C. Garland and W.L. Johnson, MRS Symp. Proc., 132,
(1989) 137.
3. J.S. Benjamin and M.J. Bomford, Met. Trans., 7A, 1301 (1977).
4. J.H. Weber, SAMPE Qly., 11 35 (1980).
5. G.B. Schaffer and P.G. McCormick, Met. Trans., 21A, 2787 (1990).
6. R.M. Davis, B. McDermott and C.C. Koch, Met. Trans., 19A, 2867 (1988).
7. P.J. Miller, C.S. Coffey and V.F. Devart, J. Appl. Phys., 54, 913 (1986).
8. H.M. McMillan and N. Gane, J. Appl. Phys., 41, 672 (1970).
9. P.W. Bridgman, Phys. Rev., 48, 825 (1935).
10. P.W. Bridgman, J. Appl. Phys., 18, 246 (1947).
11. J.S. Benjamin, Scientific American, 234, 40 (1976).
12. S.K. Kang and R.C. Benn, Met. Trans., 18A, 747 (1987).
13. T.K. Wassel and I. Himmel, Tr - 12571 (US Army Automotive Command, R and
D Control, Michigan) May 1981.
14. B.T. McDermott and C.C. Koch, Scripta Metall., 20, 669 (1986).
15. R.L. White and W. Nix, In: Proc Conf. New Developments and Applications in
Composites. K. Wilsdorf and W.C. Harrington Jr. (eds)., The Met. Sco. AIME,
N.Y. (1978).
16. E.S.B. Rao, R.M. Mallya. D.H. Sastry and V.S. Arunachalam, In: Proc. Int. Conf,
Horizons in Powder Metallurgy, W.A. Kaysser and W.J. Huppman (eds) Verlog
Schmid GAAH, Freiburg/Germany (1988), p.1275.
17. E.S.B. Rao, In: Powder Metallurgy - Recent Advances, V.S. Arunachalan and
O.V. Roman (eds), Oxford and IBM, New Delhi (1989), p.27.
18. C.C. Koch, O.B. Cavin, C.G. McKamey and Y.O. Scarborough, Appl. Phys. Lett.,
43, 1017 (1983).
19. R.B. Schwarz and C.C. Koch, Appl. Phys. Lett., 49, 146 (1986).
20. R.B. Schwarz, R.R. Petrich and C.K. Saw, J. Non-Crystalline Solids, 76, 281
(1985).
21. W. Krauss, C. Politis and P. Weimar, MPR., 43 (4), 231 (1988).
22. R.B. Schwarz and P. Nash, J. of Metals, 91, 27 (1989).
23. R.B. Schwarz and W.L. Johnson , Phys. Rev. Lett., 51, 415 (1983).
24. K.F. Kobayashi, N. Tachibana, P.H. Shingre, J. Mat. Sci., 25, 3149 (1990).
25. N. Saunders and A.P. Miodownik, J. Mater, Res., 1, 38 (1986).
26. Z.H. Yan, M. Oehring and R. Bormann, J. Appl. Phys., 76 (2), 2498 (1992).
27. E. Hellstern, H.J. Fecht, Z. Fu and W.L. Johnson, J. Appl. Phys., 65, 305 (1989).
28. A.R. Yavari and P.J. Derse, Mat. Sci. and Eng., A 134, 1315 (1991).
29 Y. Chakk, S. Berger, B.Z. Weiss and E.B. Levinson, Acta Metall.and Mater., 42
(11), 3679 (1994).
30 A.Ye Yermakov, Ye.Ye. Yurchikov and V.A. Barinov, Phys. Met. Metall., 52, 50
(1981).
31. Ibid ref. 30, 53, 935 (1982).
32. P.Y. Lee, J. Jang and C.C. Koch, In: Solid State Amorphization Transformations,

53
Mechanical Alloying

R.B. Schwarz and W.L. Johnson (eds), Elsevier, Lausanne (1988), p.28.
33. C.C. Koch and M.S. Kim, J. Phys. (Paris) Colloq., 46, 8-573 (1985).
34. R.B. Schwarz and R.R. Petrich, J. Less-Common Metals, 140, 99 (1988).
35. C.C. Koch, In: Solid State Powder Processing, A.H. Clauer and J.J. deBarbadillo
(eds), The Minerals, Metals and Materials Society, Warrendale, PA (1990), p.35.
36. G.J.C. Carpenter and E.M. Schulson, J. Nuclear Mat., 23, 180 (1978).
37. M.S. Kim and C.C. Koch, J. Appl. Phys., 62, 3450 (1987).
38. T.J. Tiainen and R.B. Schwarz, Ibid Ref. 32, p.69.
39. R. Bormann and R, Busch, In: Proc Conf, New Materials by Mechanical Alloying
Technique, E. Arzt and L. Schultz (eds), Deutsche Gesellschaft fr Metallkunde,
Germany (1989), p.73.
40. H.J. Fecht, E. Hellstern, Z.F, and W.L. Johnson, Met. Trans., 21A, 2333 (1990).
41. W.L. Johnson, Prog. Mater. Sci., 30, 81 (1986).
42. W. Gust, B. Predel and K.J. Stenzel, Z. Metallkd, 69, 721 (1978).
43 D.A. Rigney; Annu, Rev. Mater. Sci., 18, 141 (1988).
44. M.L. Trudeau and R, Schultz, Mat. Sci. and Eng., A134, 1361 (1991).
45. M. Oehring and R. Bormann Mat. Sci. and Eng., A134, 1330 (1991).
46. G.B. Schaffer and P.G. McCormick, Met. and Mater. Trans., 21A, 2789 (1990).
47 G.B. Schaffer and P.G. McCormick, Met. and Mater. Trans., 22A, 3019 (1990).
48. G.B. Schaffer and P.G. mcCormick, Scripta Metall, 23, 835 (1989).
49. G.B. Schaffer and P.G. McCormick, J. Mater. Sci. Lett., 9, 1014 (1990).
50. G.B. Schaffer and P.G. McCormick, Met. & Mater. Trans., 23A, 1285 (1992).

Questions
1. Explain the multi-layered structure of MA alloy-powder.
2. Explain the phenomenon of necking in brittle-brittle system dur-
ing MA.
3. How can microdeformation take place in a brittle-brittle system
during MA?
4. What are the various factors responsible for MA of a brittle-brittle
system?
5. Why was LQ model suitable to explain amorphization by MA?
6. Explain SSAR achieved by MA. How the MA amorphization in
systems with positive heat of enthalpy of mixing is caused.
7. Explain role of strain & plastic deformation in MA and MG.
8. Explain amorphization in intermetallic compounds.
Why is it not possible to have SSAR in a pure element?
9. Why are nanocrystalline grains considered to be defect free?
10. Explain the role of plastic strain during MA nanocrystallization.
11. Explain MA nanocrystallization in alloys.
12. Differentiate between critical temperature T crit and ignition tem-
perature T ig .
13. How MA helps in promulgating the solid state chemical reactions?
14. What is the effect of milling parameters on solid state chemical
reactions?
15. What condition must be satisfied for ignition to start during MA?

54
Energy Transfer and Energy Maps in Mechanical Alloying

6 ENERGY TRANSFER AND ENERGY


MAPS IN MECHANICAL ALLOYING
A mill is a device which transfers energy to the charge. The mecha-
nism of energy transfer from the milling tools to the milled powder is
certainly different with different apparatuses, e.g. in an attritor, an
important mechanism of energy transfer is just due to the trapping of
the powder between the colliding balls. In a vibratory or planetary mill,
the transfer is due to collision of the balls, charged with powder, against
the vial walls.
For consideration of the energy transfer during the MA process, let
us consider the process from fundamentals. What happens when a bare
ball falls from a given height (h) on a flat surface of the same ma-
terial? The ball rebounds to a certain height (h'), giving a rebound yield
(n b) = h'/h. Figure 6.1 illustrates a plot of rebound yield vs potential
energy of a ball of mass m b that is equal to the kinetic energy of a
ball E 0 at the instant of impact ( = m b gh = 1/2 m bv 2 = E 0 ). The yield
varies from 0.9 to 0.65 of increasing impact energies [1]. The energy
dissipated during an impact E b can be given as

E b = (1n b ) E 0 (6.1)

The dissipated energy gets transformed mainly into heat, resulting

Fig.6.1 Rebound yield of clean balls, b ,


vs kinetic energy of the balls, E0 (Ref.1).

55
Mechanical Alloying

in a temperature rise of both the ball and the plate. A small fraction
of it is stored in the material as structural disorder. During MA, af-
ter a milling period of 10 min to 1 hr, depending on the materials, the
balls and chamber walls get coated with a thin powder-layer, and the
balls remain glued after collision. The collision behaviour approaches
that of the elastic type and the energy dissipation becomes minimum
(Fig.6.2).
In Fig.6.2, the impact energy (x-axis) can be regarded as the intensive
milling variable which can be correlated with the impact pressure and
temperature rise required to overcome the activation energy barrier. A
solid state reaction can occur only if the impact energy is high enough
to force the atoms climbing over the activation energy of that reac-
tion. The y-axis represents the quantity of energy released for a col-
lision at any given impact energy. This extensive variable can be cor-
related to the total quantity of energy, i.e. total time, needed in or-
der to complete a given reaction.
Various investigators have developed energy maps for MA/MG for
the intermetallic formation (Pd3Si [2], NiAl [3]) and amorphization (NiAl
[3], TiNi [4], NiZr [5]). These maps give information about the minimum
energy required to start the process and the total energy required for
the process to complete. Figure 6.3 illustrates the energy map for the
formation of nanocrystalline NiAl intermetallic by MA of Ni and Al el-
emental powders [3]. It is very clear from Fig.6.3 that a minimum total
energy E t = 200 kJ/kg is essential for initiation of the phase forma-
tion and 450 kJ/kg for the formation to be complete. It also points out
that the completion of the reaction between Ni and Al leading to the
formation of single phase NiAl is governed by E t rather than the en-
ergy transferred in each impact E b .

Fig.6.2 Energy dissipation during collision, E, vs kinetic energy of the ball E0 (Ref.1).
Fig.6.3 (right) Energy map of NiAl formation reaction during MA (Ref.3).

56
Energy Transfer and Energy Maps in Mechanical Alloying

Fig.6.4 Energy map for disordered NiAl formation (Ref.3).

The energy map for the disordering of an ordered NiAl phase is


shown in Fig.6.4. It is clear from the energy map that complete dis-
ordering of NiAl formed by MA occurs only when the total energy E t
crosses 450 kJ/kg. Thus, the accumulation of a critical concentration
of defects leads to complete disordering, which is decided by E t not
E b. In addition to this, it is interesting to note that the E t required for
the completion of NiAl phase formation and for complete disordering
are quite close to each other.

References
1. M. Magini, Mat.Sci.Forum, 88-90, 121 (1992).
2. M. Magini and A. Iasomna, Mater. Trans. JIM, 36 123 (1995).
3. J. Joardar et al, In: Recent Advances in Metallurgical Processes, D.H. Sastry, et
al (eds), New Age International Publishers, New Delhi (1997), p.647.
4. B.S. Murty et al, Acta Metall., 43, 2443 (1995).
5. E. Gaffet, Mater. Sci. and Eng., A132 181 (1991).

Questions
1. What is the main mode of energy transfer in an attritor mill, a
vibratory mill and a planetary mill?
2. Describe phenomena of energy transfer during MA.
3. What are the energy maps used in MA?
4. What is the importance of energy maps in MA?
5. Draw energy maps for MA of (i) Al+Ni (ii) NiAl.

57
Mechanical Alloying

7 CONSOLIDATION OF
MECHANICALLY ALLOYED POWDERS
7.1 CONSOLIDATION TECHNIQUES
The MA powders have flake-like imperfectly packed layers and rough
surfaces which lower their flow rate due to extensive mechanical in-
terlocking. These particles are under heavily worked conditions, which
also lowers the shear induced bonding during cold compaction. Thus,
conventional methods of consolidation, cold compaction and sintering
are generally not suitable for their consolidation. Moreover, due to the
layered structure of these powder particles they have a high amount
of adsorbed and entrapped gases, which makes degassing, prior to con-
solidation, an essential step.
In the case of nanostructured and amorphous materials, the prob-
lems of consolidation are still severe. In the case of nanostructured
materials, there is rapid grain growth at elevated temperatures. Grain
size thermal stability in these materials can be given by a modified
Arrhenius equation [1]

d 2 = d 02 + K m t p 2 exp (Q/RT) (7.1)

where d0 is the initial grain diameter, K m is the material constant; p,


T and t are the pressure, absolute temperature and time used for
compaction, respectively, Q is the activation energy and R is the gas
constant.
In the case of amorphous MA powders, recrystallization at elevated
temperatures is a problem [2].
All compaction techniques make use of temperature, pressure and
shear. The temperature helps in diffusion bonding and in removing some
of the porosity. The pressure provides densification, but does not help
bonding. Shear helps to clean the powder surfaces and to bring about
interparticle bonding. The various techniques used for the consolida-
tion of MA powders have been compared in Fig.7.1 on the basis of
pressure used to achieve compaction and the timetemperature excursion
associated with the requirement of complete consolidation (time tem-

58
Consolidation of Mechanically Alloyed Powders

Fig.7.1 Schematic comparison of techniques for consolidiation of MA powder on the


basis of the time-temperature excursion and the stress or pressure used (Ref.3, E. Artz
and L. Schultz (editors)), New Materials by Mechanical Alloying (1989), p.143,
(Deutsche Geselschaft fr Metallkunde)).

Fig.7.2 Schematic comparison of techniques for consolidation of RS and MA powders


on the basis of whether shear or pressure forces are applied (Ref.3).

perature parameter = time (sec) temperature (C)). Long hold times


at high temperatures are useful for obtaining full consolidation by dif-
fusion bonding and by void removal, but leads to a significant loss of
metastability [3]. A comparison can also be made on the basis of
forces imposed, specifically on the pressure and shear components of
these forces (Fig.7.2).
The rapid omnidirectional compaction process (ROC) uses a forg-
ing press to create high pressures within a thick-walled, hot, solid die.
The temperature and forces are sufficient for the die to flow as a fluid
(i.e. fluid die), thus creating a nearly hydrostatic pressure [3,4]. Several
modifications of the basic process exist to produce shaped parts us-
ing simple and cheap fluid dies [4,5].
The choice of technique depends on the material as well as the
application as described below.
1. If the material is highly thermally stable, as for example oxide
dispersion-strengthened alloys, high temperature techniques can readily
be used.
2. For the medium thermally stable materials, for example alloys
dispersion strengthened by metallic particles, it is important to restrict
the consolidation to sufficiently low temperatures to avoid excessive
property degradation. Techniques of high shear, for example extrusion

59
Mechanical Alloying

or forging, are preferred for consolidation.


3. For the highly metastable materials, such as amorphous materi-
als, it is the avoidance of excessive loss of metastability by high tem-
perature excursions which is the most important criterion. Low tem-
perature consolidation techniques such as dynamic compaction, ROC
or forging seem to offer the best chances of consolidation.
These consolidation techniques provide compacts with nearly full
density and have associated advantages but limitations, also which are
as follows:
1. Being well characterised and the cheapest process as well as able
to provide an isotropic structure, hot extrusion of MA powder is gen-
erally preferred [6]. Temperature of the die, extrusion ratio, strain and
strain rate decide the grain growth. Higher temperature and lower strain
rates tend to allow grain sizes to enlarge gradually, which deteriorates
the mechanical properties.
2. Consolidation by forging presents an additional set of necessary
controls.
3. In hot isostatic pressing (HIP), the elevated temperatures used
during consolidation often result in significant grain growth [1,7]. In
addition, the configuration of the final compacts cannot be controlled
and is often irregularly shaped.
4. Compaction of MA powders by dynamic compaction techniques
like gas gun launchers [8,9] or explosives [10,11] results in minimum
grain growth, but due to the high shock pressure of relief waves (a pres-
sure two to three times of the material flow strength is required in these
techniques) the compacts often contain fine cracks. Secondly, the need
to utilize sufficient energy to bond powder particles which may be suf-

Fig.7.3 Optical micrographs showing dynamically compacted mechanically alloyed Fe-


40% Ni14% P6% B (at.%) amorphous powders: a) at 5 GPa pressure; b) at 4 GPa
pressure (Ref.3).

60
Consolidation of Mechanically Alloyed Powders

Fig.7.4 Optical micrographs of MA 760 superalloys:


a) as extruded, showing equiaxed grains; b)
recrystallized, showing texture and anisotropic grains
(Note that both the microstructures show depletion
of carbide particles as inhomogeneity) (Courtesy Dr.
H.K.D.H. Bhadeshia).

ficient for a temperature rise for localized melting (Fig.7.3a) and


recrystallization (Fig.7.3b).
Recently, consolidation techniques like cold/warm compaction and
sintering [12,13], plasma short activated sintering [14], and liquid phase
sintering [15,16] have also been used in the case of thermally stable
MA powders. These techniques have also provided compacts with nearly
full density and minimum grain growth. However, contamination due
to impurities present in the sintering atmosphere is a potential prob-
lem.
Thus, in consolidation of MA powders, there is difficulty in find-
ing the process window where good densification and bonding is ob-
tained, but where the extent of global as well as localised heating and
microstructural modification is limited to acceptable values.

61
Mechanical Alloying

7.2 THERMOMECHANICAL TREATMENTS


In the case of thermally stable materials, after consolidation several
thermomechanical processing steps are necessary to control the grain
size and achieve the desired properties in the as-fabricated condition.
Two structures are beneficial [17]:
1. Fine equiaxed grains, giving best room temperature strength, fa-
tigue strength and workability.
2. Coarse elongated grains, for high temperature stress-rupture
strength, thermal fatigue resistance and corrosion resistance.
The fine-grained condition exists in extruded material and, in gen-
eral, these alloys are not worked at relatively higher values of tempera-
ture compensated strain-rate, referred to as the ZenerHolloman param-
eter [18], which is defined by:

z = e& exp (Q/RT) (7.2)

where e& is the mean strain rate, Q is the activation energy of the rate-
controlling process in the deformation mechanism, R is the universal
gas constant and T is absolute temperature. Such strain rates lead to
increased substructural strengthening. For example, extruding at a rela-
tively low temperature and high speed produces a high value of z, which
can lead to a fine array of subgrains. Hot working at too high value
of z can result in sufficient stored energy to cause recrystallization.
Coarse elongated grains, e.g. in dispersion-strengthened superalloys,
are developed by secondary recrystallization of the materials.
Thermomechanical processing produces a high level of stored strain
energy [19]. It is a complex matter to control the processing history
of the material from the condition of milling, through the temperature,
ratio and speed of extrusion, to the details of post extrusion process-
ing, including hot or cold rolling or forging, and the grain coarsening
anneal. Achievement of the highest strength attainable by an MA
material demands the closest control of production parameters. The
recrystallized grains exhibit a texture ({100} <001> or {110} <001>)
and the grains are anisotropic (Fig.7.4). The texture is not only influ-
enced by the recrystallization temperature but also by the temperature
gradient conditions [2022]. Therefore, the desired recrystallized tex-
ture must be promoted by controlled nucleation and growth of these
very large grains during zone annealing, analogous to what occurs in
directional solidification. The development of anisotropic grains in these
alloys is attributed to the alignment of oxide particles along the extrusion
direction. The alignment is expected to be strongest along the surface
regions where the deformation imparted during extrusion is most intense.

62
Consolidation of Mechanically Alloyed Powders

Accordingly, the grains at edges tend to be more anisotropic and less


sensitive to heating rates. The core regions are much coarser and sen-
sitive to the heating rate. For randomly dispersed particles, the grain
boundary velocity v R is expected to be isotropic. For aligned particles,
it should be easier for grain boundary motion to occur parallel to the
extrusion direction (velocity v x) and more difficult for motion normal
to the extrusion direction v y,z, i.e. v y,z < v R < v x. Since growth along
the extrusion direction is less impeded, recrystallization can initiate more
readily, thereby explaining the faster kinetics at the surface regions. It
is also well established that particles can exert pinning forces. Thus
recrystallization in these materials is a complex process.

References
1. J. Ravers et al, Met. and Mater. Trans., 27(A), 3126 (1996).
2. C.P. Dogan et al, Nanostruc.Mater., 4(6), 631 (1994).
3. D.G. Morris and M.A. Morris, In: Proc.Conf. New Materials by Mechanical
Alloying, E.Artz and L. Schultz (eds), Deutsche Geselschaft fr Metallkunde,
Germany (1989), p.143.
4. W. Smarsly and W. Bunk, Powd.Met.Int., 17, 63 (1985).
5. R.L. Anderson and J. Groza, MPR, 43 (4), 272 (1988).
6. J.A. Pordoe, Powd. Metall., 22, 22 (1979).
7. C. Suryanarayana, Nanostruc. Mater., 2(5), 527 (1993),
8. D.G. Morris, Mat. Res. Soc. Symp. Proc., 28, 145 (1984).
9. D.G. Morris, Metal Sci., 14 215 (1980).
10. M.A. Meyers et al, J. Metals, 33, 21 (1981).
11. R.N. Wright et al, Adv. Mater. and Proc., 10, 56 (1987).
12. O. Dominguez and J. Bigot, Nanostruc. Mater., 6 (8), 877 (1995).
13. G.E. Fougere et al, Nanostruc. Mater., 5 (2) 127 (1995).
14. K. Kobayashi, J. Jpn. Soc. Powd. Powd. Met., 42 (2), 191 (1995).
15. G. Frommeyer, Paper presented at Euro 95, Birmingham (1995).
16. Y. Hashimoya, J. Jpn. Soc. Powd. Metall., 42 (2), 185 (1995).
17. M.J. Fleetwood, Mat. Sci. and Techn., 2, 1176 (1986).
18. J.R. Pickens: ASM Metals Handbook, Tenth Edition, 2 200 (1990).
19. R.C. Benn et al, Powd. Metall., 24, 191 (1981).
20. K. Murakani et al, Met. Trans., 24A, 1049 (1993).
21. W. Sha and H.K.D.H. Bhadeshia, Met. and Mater. Trans., 25A, 705 (1994).
22. T.S. Chou and H.K.D.H. Bhadeshia, Met. Trans., 24A, 773 (1993).

1. Why is conventional compaction & sintering not possible in the


case of MA powders?
2. What problems are associated in consolidation of MA materials?
3. What is the ROC process?
4. Which dynamic consolidation techniques are being used in case of
MA powders. What are basic advantages & limitations of these
techniques?
5. Compare, on the basis of pressure & time-temperature excursion
parameter, various consolidation techniques used for consolidation
of MA materials.

63
Mechanical Alloying

6. Why thermomechanical treatment is given to metastable MA


materials.
7. Why extrusion is normally preferred in the case of MA powder.
8. What disadvantages are associated with HIP when used to consoli-
date amorphous or nanocrystalline materials?

64
Mechanical Properties of Mechanically Alloyed Materials

8 MECHANICAL PROPERTIES OF
MECHANICALLY ALLOYED
MATERIALS
Mechanical alloying enables the effective superimposition of numerous
strengthening factors including:
oxide dispersion
carbide dispersion
fine grain
high dislocation density and substructure
solid solution strengthening
Both direct and indirect effects influence the mechanical properties
of these materials. The interaction of the second phase particles with
dislocations is an example of a direct effect, while grain morphology
and texture are both indirect effects.
Moreover, the aforementioned five strengthening contributions can
be augmented by precipitation strengthening as well as intermetallic
dispersion strengthening. Thus, the resulting enhancement in mechanical
properties of MA materials is far greater than can be achieved by
conventional methods or say by the rule of mixture.

8.1 TENSILE PROPERTIES


The tensile strength of MA materials is found to be relatively propor-
tional to the square of relative density [1] and is consistent with the
densitytensile strength relationship for PM materials (Fig.8.1):

u = k m ( measured / full C) 2 (8.1)

where u is the tensile strength, k m is the material constant, is the


density and C is a correction factor, equal to green density which rep-
resent the state of zero tensile strength.
The yield strength of MA materials is found to depend linearly on
the inverse of the square root of the grain size (Fig.8.2) (i.e. Hall
Petch relationship), even in the case of nanograin materials [1,2,3]:

65
Mechanical Alloying

Fig.8.1 Tensile stress vs corrected relative density square (Ref.1).

Fig.8.2 Grain size dependence of the yield strength of MA materials (The Int. J. Powd.
Met., 24(4), J.S.C. Wang et al, Microstructures and mechanical behaviour of mechanically
alloyed nickel aluminide (1988), p.315) (with permission from Prof. C.C. Koch)).

y = k m + k 0 d 1/2 (8.2)

where y is the yield strength, k m is a material constant, k 0 is the grain


size coefficient and d is the grain diameter.
Thus, the tensile properties of these materials depend strongly on
compact density, Eq.8.1, and, to a lesser degree, on grain size.
An approximately linear plot even in the case of oxide-dispersed
materials, (Fig.8.2), indicates that dispersoids influence the room tem-
perature yield strength through their capability to prevent grain growth
during powder consolidation.
This tensile behaviour is also consistent with the Orowan mecha-
nism in which the non-shearable oxide particles impede dislocation
66
Mechanical Properties of Mechanically Alloyed Materials

motion by requiring dislocations to bow between particles. The Orowan


Ashby expression [4]

0.13 Gb
s Or =
b
l ln rd / n g (8.3)

where rd is the dispersoid radius, l is the dispersoid spacing, b is Burg-


ers vector and G is the shear modulus, gives a value of about 15%
for the contribution to the yield strength, compared with the grain size
effect, which can be attributed to dispersoid strengthening. The appro-
priate dispersoid spacing l is given as:

LF p I
l = MG J
0. 5
-2
OP F 2 I 0 .5
rd
MNH f K PQ H 3 K (8.4)

where f is the dispersoid volume fraction and rd is the dispersoid particle


radius. Equations (8.3) and (8.4) suggest that fine and even distribu-
tion of the dispersoid has a stronger effect than its volume.
However, in the case of MA intermetallic nanograin materials, the
room temperature tensile strength is found to follow the inverse Hall
Petch relationship [5], as in HSLA steels and other nanocrystalline alu-
minides. This startling behaviour is attributed to the change in the con-
ventional deformation mechanism. Conventionally, in metals of normal
grain size the grain boundaries act as barriers to dislocation glide and
small grains prevent large dislocation pile-ups. The nanograins are sug-
gested to be essentially dislocation-free [6]. The grain size in these
solids is about three orders of magnitude smaller than in superplastic
alloys and diffusion rates are appropriately enhanced [7]. It is there-
fore probable that the mechanism of deformation at room temperature
in these materials is similar to the mechanism of superplasticity at el-
evated temperature like grain boundary sliding and diffusional creep,
dislocation glide and creep. However, not all nanocrystalline solids show
this inverse HallPetch effect [8]. Thus, further experimental work to
explain the deformation mechanism explicitly in these materials is re-
quired.
In general, a simplified relationship exists between ultimate tensile
strength and hardness of the metals and alloys [8]:

Hv
sy = km( m -2) (8.5)
(3.0 0.1)

67
Mechanical Alloying

where k m is a constant depending on the material, H v is the diamond


pyramid (Vickers) hardness (kg/mm 2 ) and m is the Meyers hardness
coefficient. The quantity (m2) is equivalent to the strain hardening co-
efficient.
The assumption of the strain hardening coefficient (m2) as zero
for simplicity in Eq.(8.5), allows an equation of the following form to
be used:

Hv
sy = (8.6)
(3.0 0.1)

During MA, powders are subjected to compressive impact forces


and undergo severe plastic deformation. During consolidation also these
alloys undergo a large plastic deformation (extrusion/rolling/forging, etc).
Thus, MA materials undergo extensive plastic deformation before they
get the shape, and the strain hardening coefficient can be assumed to
be zero [9]. Thus, MA alloys appear to be the ideal candidate for
applying Eq.(8.6). In fact, it has been applied for a range of aluminium
alloys by I.A. Hawk et al [10] and results have been plotted in Fig.8.3.
Equation (8.6) provides a good estimate (with 5% of the actual meas-
ured values) of the yield stress for MA alloys not only at room tem-
perature but also after elevated temperature heat treatment due to the
stability of fine grains.
Equation (8.6) provides a tool for a quick and relatively accurate
way to estimate the yield stress of MA alloys for a range of heat treated
conditions. It may be specially attractive in the cases of nanostructured
materials which are available in limited quantity [10].

Fig.8.3 Plot of yield stress as


calculated by Eq.8.6 using hardness
values vs the measured tensile yield
stress for a series of MA aluminium
alloys. (Ref.10, Met. Trans., 19A,
1363 (1988), (The Minerals, Metals
and Materials Society)).

68
Mechanical Properties of Mechanically Alloyed Materials

8.2 FRACTURE
In general, the ductility and toughness of MA materials is found to be
dependent on the morphology of prior particle boundaries [11,12]. The
main strengthening factor is found to be the grain size. Elongation is
shown to be greater at dynamic strain rates, which is attributed to the
initiation of cracks.
Dispersoids can harden the matrix by obstructing the motion of dis-
locations. The dislocation may circumvent the obstacles by passing,
cross-clipping, climbing, etc [13,14]. The fracture appears to be domi-
nated by the disaccomodation between the dispersoid particles and the
matrix. The fracture path presumably follows crack initiation at the
dispersoid oxide/matrix interface and to the extent that dispersoids deco-
rate grain boundaries as well as reside in grain interiors. The fracture
paths may be transgranular or intergranular leading to a mixed mode.
In the case of nanograin materials, factors such as:
development of secondary cracks;
development of cracks within the nanograin particles and subse-
quent crack arrest at the particle nanograinparticles and subsequent
crack arrest at the particle nanograinmicrograin interface;
crack propagation along the boundary between the nanocrystalline
and submicron grain regions increases the fracture toughness [1].

8.3 CREEP
Studies on creep behaviour of various fully dense compacts of MA
dispersion-strengthened materials [11,12,15,16] and aluminides [17] have
revealed that MA enhances creep resistance and microstructural sta-
bility. These improvements are attributed to the fine scale dispersion
(~ 30 m) throughout the matrix, at matrix intermetallic interfaces, and
on subgrain boundaries which inhibit coarsening, recovery and
recrystallization.
The improvement in creep strength, even in the presence of stable
fine grain microstructures, in the MA materials may be attributed to
the ability of the dispersoids to interact with mobile dislocations over
a much wider range of temperature compared to the precipitates present
in the conventional alloys [14]. Diffusional creep is generally consid-
ered to increase with decreasing grain size. Artz et al [18] proposed
that when diffusion distances are small, elevated temperature deformation
processes are controlled by the motion of dislocation along grain bounda-
ries rather than the lattice diffusion vacancies. The creep rate is then
determined by the resistance offered to the motion of the boundary dis-
locations by dispersoids situated along grain boundaries in much the
same way as Orowan strengthening. The fracture path is presumably
dominated by fracture initiation at the dispersoid/matrix interfaces in
69
Mechanical Alloying

these materials.
The low creep rates in the case of coarse-grained dispersion-strength-
ened materials (e.g. thermomechanically treated MA superalloys) can
be explained by the modified power-law diffusional creep, which is usu-
ally applicable for high temperature alloys, as given below [14]:

e& = km D LM s u - sth OP n
(8.7)
N E Q
where e& is the creep rate (strain rate), u is the tensile strength, th
is the threshold stress below which creep rates are considerably smaller,
k m is the material constant, D is the diffusivity, E is Youngs modu-
lus and n is the stress exponent.
The threshold stress th, proposed in modified power-law creep,
Eq.(8.7), is assumed to depend on a parameter k r, describing the re-
laxation of the dislocation and is proportional to the Orowan stress,

s th = sOr 1 - kr2 (8.8)

8.4 SCC SUSCEPTIBILITY


Mechanically alloyed alloys achieve an excellent combination of high
strength and superb SCC resistance because of the uniform introduction
of microstructural features that either improve SCC resistance, e.g.
intermetallic particles, or increase strength without degrading SCC re-
sistance, e.g. oxide and carbide particles, in aluminium alloys.

References
1. J. Rawers et al, Met. and Mater. Trans., 27A, 3126 (1996).
2. T.G. Neil and J. Wadsworth, Scripta Met. and Mater., 25, 955 (1991).
3. K.Y. Wang et al, J. Mat. Sci. Lett., 12 (23), 1818 (1993).
4. W.E. Frazier and M.J. Koczak, In: Dispersion Strengthened Aluminium Alloys,
Y.W. Kim and W.M. Griffith (eds), TMS, Warrendale, PA (1988), p.573.
5. A.J. Heron and G.B. Schaffer, In: Advances in Powder Metallurgy and Particulate
Materials, V.6, A. Lawley and A. Swanson (compilers), MPIF and APMI, NJ
(1993), p.77.
6. S.K. Ganapathi and D.A. Rigney, Scripta Met. and Mater., 24, 1675 (1990).
7. R. Birringer et al, Defect and Diffusion Forum, 59, 17 (1988).
8. J.S.C. Jang and C.C. Koch, Scripta Metall. and Mater., 24, 1599 (1990).
9. J.R. Cahoon et al, Met. Trans., 2, 1979 (1971).
10. I.A. Hawk et al, Met. Trans., 19A, 2363 (1988).
11. C. Zakine, Creep mechanism in mechanically alloyed oxide dispersion strengthened
alloys, Presented at 1994 Powder Metallurgy World Conference, Paris, June 1994.
12. H.R. Last and R.K. Garrett, Met. and Mater. Trans., 27A (3), 737 (1996).
13. J.S.C. Wang et al, Inter. J. Powd. Met., 24 (4) 315 (1988).

70
Mechanical Properties of Mechanically Alloyed Materials

14. R.C. Benn and P.K. Mirchandani, In: New materials by mechanical alloying, E.
Arzt and L. Schultz (eds), Deutsche Geselschaft fr Metallkunde (1989), p.19.
15. E.L. Erich, Report-AFML-TR-79-4210, US Air Force (1980).
16. A. Lawley and M.J. Koczak, Report No. AFOSR-TR-86-0567 (Oct. 1985).
17. J.W. Pyun and S.I. Kwun, J. Korean Inst. Metals, 33 (6), 814 (1996).
18. E. Arzt et al, Acta Metall, 13, 1977 (1983).

Questions
1. What are the contributing strengthening factors in MA materials?
Comment on them.
2. How do tensile properties depend on the grain size in MA materi-
als?
3. How can the inverse Hall-Petch relationship be explained in the
case of MA intermetallic nanograin materials?
4. Comment on relationship between yield strength and hardness of
MA alloys.
5. Comment on fracture toughness of
(i) MA alloys;
(ii) Nanograin materials.
6. Explain mechanism of creep in
(i) MA fine-grained materials;
(ii) MA coarse grained (thermomechanically treated) materials.
7. Why is SCC resistance high in the MA alloys?

71
Mechanical Alloying

9 MODELLING MECHANICAL
ALLOYING
9.1 INTRODUCTION
Modelling can be simply defined as the representation of a physical
entity or a process by another physical or conceptual entity or a process
with a view to obtain greater insights into the behaviour of the sub-
ject. Models developed for complex processes cannot be expected to
be absolutely precise. Rather, they are intended to identify important
parameters, define the functional dependence of the process output on
process variables and predict results with an acceptable level of pre-
cision. One useful result of such process modelling is a considerable
reduction in the empirical studies needed to refine a process into a
useful engineering tool.
Some aspects regarding what is occurring during the MA process
have been known qualitatively for some time, but the description of
the MA process is complex and manifested as it involves concepts of
mechanics, mechanical behaviour, heat flow, thermodynamics and ki-
netics. In spite of this, modelling of MA has been an avenue of re-
cent interest for further understanding.
Though the models for MA are in the early stage of development,
various physical/theoretical models developed can be subdivided into
mechanistic, atomistic, thermodynamic and kinetic. Mechanistic models
deal with powder deformation, coalescence and fracturing, and ball dy-
namics in a milling device. Atomistic models permit a deeper insight
into the physics of non-equilibrium phase formation using the CALPHAD
method of thermodynamics data fitting and phase diagram calculations.
Kinetic models consider the competitive diffusion-controlled growth
kinetics of crystalline phase [1]. In the following section, these are
discussed in brief.

9.2 MECHANISTIC MODELS


These models deal with the basic process of deformation, coalescence
and fragmentation, evaluation of steady state particle size distribution
and the milling kinetics. They can be used to predict the milling time
required for a particular system, under a given set of experimental con-

72
Modelling Mechanical Alloying

ditions. Mechanistic models can be classified into two types: local and
global:
1. Local modelling describes the various effects (thermal and me-
chanical) and events (deformation, fracture and welding) that transpire
when powder particles are entrapped between two colliding or sliding
surfaces. Thus, local modelling is generic in the sense that parameters
(relative impact velocity, angle of impact, charge ratio, etc.) affect-
ing the various events are common to all mills, although the values of
some of the parameters (e.g. relative impact velocity) are specific to
a particular type of mill and its operating conditions.
2. Global modelling is device specific. For example, this type of
modelling entails the study of factors such as distribution of impact
angles and heterogeneity of powder distribution within the mill, fac-
tors which clearly differ from one type of device to another. Several
efforts have been made at analysing the ductileductile system (to start
with) and these are described here.

9.2.1 Deformation, Coalescence and Fracture


Regardless of the mill used, the MA is characterised by collisions be-
tween tool and powder (ballpowderball, ballpowdervial and ball
powderimpeller arm, etc.) which result in powder fragmentation and
coalescence. Such collisions may take place over a range of impact
angles, and this geometrical feature might have an important effect on
the relative tendencies for coalescence and fragmentation. Figure 9.1
depicts the events possible during MA. Direct seizure (i.e. cold welding
taking place without sliding displacement) is favoured by normal im-
pacts, whereas if relative particle displacement precedes such weld-
ing, indirect seizure can occur. However, though only a small fraction

Fig.9.1 Various kinds of fragmentation and coalescence events that can be imagined
to occur during MA depending upon impact angle (Ref.4, Met. Trans., 21A, 289 (1990)).

73
Mechanical Alloying

of the total number of impacts [3], only the normal collisions are con-
sidered.
The structure developed in the MA powder depends on the defor-
mation, coalescence and fracture events taking place during the process.
Therefore, a model for the deformation, coalescence and fracture of
MA events has been developed under the following assumptions [4]:
only the ballpowderball collisions are considered as they are
greatest in number [2];
during the processing, a powder coating of about 100 m thickness
gets [2] coated uniformly;
it is assumed that the powder particle shape is a oblong spheroid,
though it varies from spherical to flake. This shape is characterised by
a ratio (shape factor) of minor and major axis of the spheroid;
it is assumed that these oblong spheroid shaped particles rest on
a grinding ball with their major axis parallel to the ball surface.

Deformation
During the collision, the kinetic energy of the balls is converted into
deformation energy during the approach of their centres. In the case
of alloying of two or more phase materials, it is taken as the resist-
ance offered by the softer component present. The stress homologous
to this energy conversion is applied on the deforming particles. Assuming
that the bulk powder particles undergoing collision constitute an in-
dividual particle, the deformation may be expressed as a function within
the contact area as

F r IJ
a (r ) = R v G
1/ 2
-
r2
(9.1)
HH K
b 0
V Rb

where r is the distance from the centre of contact, R b is the radius


of the balls, v 0 is the relative velocity of the balls at impact, is the
density of medium balls and H V is the powder hardness.
The deformation manifested strain (hardness) can be determined
from

e = - ln
LM h - a (r) OP
0
(9.2)
N h Q 0

where (r) is the deformation at a contact radius of r and h 0 is the


powder coating thickness. Thus, strain is a radial function of the contact
zone.

74
Modelling Mechanical Alloying

Coalescence
During plastic deformation, the brittle oxide layer present on the particle
surface breaks, the particles flatten in compression, their surface area
increases and the underlying metal is progressively exposed. The bonding
force F w, acting at the weld that is formed, can be given as

F w = A w u (9.3)

where u is the tensile strength of the weld and Aw is the effective new
surface area created. If the same metal welds together, u is the tensile
strength of the bulk material. When a cold weld is made between two
lamella particles, the weld strength is given by the rule of mixture.
Using the relationship, H v = 3 u, cf. Eqs. (8.6) and (9.2), one can
calculate the bonding force (F w) acting at the weld point which must
be more than this for fracture to take place.

Fracture
Considering only the forging fracture which may develop over several
impacts (chances of shear fracture and dynamic fracture are much less
during the MA process), the cracks formed will grow radially along
the major axes of the particles (Fig.9.2). A crack initiates when the
critical tensile strain is attained over a critical length and the crack
propagation occurs when the plastic energy release exceeds a value
characteristic of the material. Thus, if the particle is sufficiently small
so that this length is greater than the particle size, the particle is con-
sidered below its comminution limit and will not fracture.
The following equation can be used to predict the tensile true fracture
strain f of a particle

z (r) = 2 f (9.4)

where z is the axial strain.


When the crack reaches a critical length, determined by the criti-
cal value of the J integral, it propagates catastrophically. The value
of the J integral is approximately given by [5]

Fig.9.2 Schematic of forging fracture edge


cracks formed along the particle circumfer-
ence and their growth along the particle
axis.

75
Mechanical Alloying

J = bsu e a p
F 3s IJ
mG
m +1

(9.5)
H 2s K u

where = 1/ ( u /K) m , with K being the strength coefficient and


m = 1/n (n is the work-hardening coefficient). The critical J value (J Ic)
is related to the critical stress-intensity factor K Ic through J Ic = K 2Ic/
E. The critical crack length a c is obtained for J = J Ic or

ac = JIc
Km FG 2 IJ m+1

(9.6)
p m H 3 su K
Values of J Ic and ac are calculated based on the input values of the
material parameters K Ic, E, K, u and n, where E is Youngs modu-
lus of the powder material.
Using the expression for strain as a function of radial position within
the powder particle permits determination of the total approach between
balls needed in order to exceed some critical strain over a given length.
The condition for forging fracture can now be expressed as

a (r ) F a 2 f 2/ 3 I 0 .5

h0 GH
= 1 - 1 - c s2
4 Rw
JK exp( - e c ) (9.7)

where c is the critical strain to fracture, (see Eq.9.4), R w is the ra-


dius of the weld region and f s is the shape factor. The factor of 4 in
the denominator of the last term on the right-hand side of Eq. (9.7)
stems from the radial symmetry of the particles: to exceed the criti-
cal strain over a c requires that the strain be exceeded over a radial
distance one half of a c.
The particle shape may affect coalescence and fragmentation events.
The shape factor can be expressed in terms of deformation

FG
fsf = 1 -
a (r ) IJ 1.5
fsi (9.8)
H h0 K
where f sf and f si are the final and initial shape factors, respectively.

76
Modelling Mechanical Alloying

9.2.2 Evolution of Particle Size


The size and size distribution of the particles change as a result of
coalescence and fracture events taking place during the MA.
Assuming that respective fracture and welding probabilities are in-
dependent of particle size, an expression for the particle size distri-
bution during such fissionfusion can be written as [5,6,7]:

tc
dn ( v )
dt z
= - a f n ( v ) - a w n( v ) f (v' ) dv' +
0

+ 2a f z 1

v v'
a
z
2 0
v
m ( v' )dv' + w n ( v' ) f ( v - v' )dv'
(9.9)

where f and w represent the fracturing and welding probabilities re-


spectively, occurring in a specific impact, t c is the average time be-
tween impacts for a particle, n (v) dv is the number of particles having
a volume between v and v + dv, and f(v) dv is the corresponding frac-
tion of particles in this volume interval. The first term on the right-
hand side of Eq.(9.9) represents the removal of a particle by fracture
from the volume interval. The second right-hand side term represents
particles removal by welding. The third term represents introduction
of particles into the volume interval by the fracture of a particle having
a volume greater than the one considered. The factor of 2 in this term
takes into account that two particles form from one as a result of
particle fracturing. The term 1/v' within the integral reflects the assump-
tion that the particle fission distribution is uniform and thus the fission
probability is inversely proportional to the number of size intervals the
larger particle may break into. The last term represents introduction
of a particle into the volume interval by the welding of two particles
having the sum of their volume equal to the one considered. The term
of 1/2 in this term arises from the convolution integral and is needed
to prevent double counting.
There appears to be no analytical solution to such a type of equation.
Thus, application of Eq. (9.9) to these processes requires discretisation
of the equation. This is done by dividing the population into J equal
volume elements, each with an instantaneous population N J. In this for-
mulation, the equivalent of the first term on the right-hand side of
Eq.(9.9) is simply f N J. The second term is then represented as

n- J
-a w NJ fI (9.10)
I =1

77
Mechanical Alloying

Equation (9.10) does not allow particles to weld if the sum of their
volumes is larger than a maximum size class. The corresponding
discretised form of the third and fourth term on the right-hand side of
Eq.(9.9) are

n
NI
+2 a f I -1 (9.11)
I = J +1
J -1
aw
+
2
N J - I fI (9.12)
I =1

There are two special cases that must be considered during the
numerical analysis. The first deals with the smallest size class considered
(J = 1). Particles in this size class are not allowed to fracture. In ad-
dition, there are no smaller particles that could weld. The equation
applicable to this size class is thus written as

LMt dNJ OP = 2a f
n
NI
- aw
n- J
N J fI
N c
dt Q J =1 I = J +1
I -1 I =1
(9.13)

The second is for the largest size class (J = n). Particles of this size
range are not allowed to weld to any other particles. The equation for
this size class is

LMt dN J OP = -a f N J +
aw J -1
N J - I fI
N c
dt Q J =n 2 I =1
(9.14)

Fig.9.3 Cu, Nb and Cr particle size distributions obtained from model predictions are
compared to those found experimentally for 4 hrs milling (Ref. 8).

78
Modelling Mechanical Alloying

Fig.9.4 Total number of powder


particles vs milling time for Cu, Nb,
and Cr as observed experimentally
(data points) and as predicted by
the particle-size evolution model
(lines), (Ref.8, Met. Trans., 24A,
2465 (1993)).

The appropriate summation of Eq. (9.10) through (9.14) and stepping


through time provide an estimate of the number of particles in a given
size class during milling, as would be expected from integrating Eq.
(9.9).
A bimodel size distribution is predicted by the model and this is found
by experiment [8], also in the case of ductile metals like Cu and Nb
(Fig.9.3a and 9.3b). However, the model is less satisfactory in the case
of brittle metals like Cr (Fig.9.3c). It should be noted that the assump-
tion that fracture and welding probabilities are independent of particle
volume and time is not correct. The hardness of powder particles
changes concurrent with milling time, with a consequent reduction in
welding tendency. In addition, fracture tendencies are also size depend-
ent. Thus, the model is valid for the early stages of milling where frac-
turing and welding tendencies are approximately constant.
The time variation of the total number of particles in the mill can
also be predicted by the model and is compared with the experimen-
tal results. Such an exercise is illustrated in Fig.9.4, showing only fair
agreement between the model and the experimental results, which can
be attributed to the respective fracture and welding probabilities used
in the model, averaged over the milling time.

9.2.3 Milling Times


Since the powder particle aggregates are much smaller than the col-
liding bodies they are trapped between, it is reasonable to view the sur-
faces of the colliding bodies as having infinite curvature relative to the
powder. The collisions with powder may then be viewed as an upset

Fig.9.5 A cylindrical slug of initial height,


h 0 , Hertz radius, r h , entrapped between the
colliding surfaces (Ref.4).

79
Mechanical Alloying

forging process between the two parallel plates [9]. We consider a po-
rous metal cylinder, composed of a large number of individual pow-
der particle, of height h 0 and radius r h, as shown in Fig.9.5. The cyl-
inder is impacted between colliding workpieces having an average relative
collision velocity v 0. During an impact event, it is assumed that par-
ticles confined in the cylinder are subjected to a true deformation strain
. It is also assumed that the rate of this deformation decreases lin-
early during the impact. That is, this velocity is v at the initiation of
compaction and is zero at the time () at which the impacting
workpieces begin to rebound from each other.
This model can be used, among other purposes, for estimating the
time required for MA. For this first select a unit cell or the milling
device, i.e. in a SPEX mill, then the unit cell is taken as the mill vol-
ume divided by the number of balls within it. In the attritor, the unit
cell volume is the average volume belonging to each ball.
For such a unit cell, the powder volume strain per unit time is
r h2 h 0/t c, where is the deformation strain per impact and tc is the
time between impact events. To alloy the material, the critical defor-
mation strain c, Eq.(9.4), must be generated throughout the total pow-
der volume v p. The processing time t p is the product of V p and c di-
vided by the volume strain total per unit time [10]

t p = (V p c t c)/(r h2 h 0 ) (9.15)

Processing times thus can be estimated if the c, V p, tc, r h2 and h 0


are known.
V p is determined experimentally by mill conditions, and c can be
empirically estimated, assuming the number of collisions to be 35
would correspond to a lamellar structure with an initial spacing of 20
m having its spacing reduced to 1.00.1 m.
The estimation of other variables can be explained as follows:

Estimation of time between collisions


The time between collisions t c can be given by

t c = /v 0 (9.16)

where is the mean free path between collisions. The velocity of balls
in an attritor mill can be estimated by considering them as a fluid.
As the agitation speed increases, the balls rise higher against the chamber
wall (Fig.9.6). From elementary fluid mechanics, the height of a fluid
above the reference line is given by [10,11]:

80
Modelling Mechanical Alloying

Fig.9.6 Agitation of a ball charge in an


attritor may be likened to that of a
viscous liquid in a spinning tank (Ref.9).

h f = 2 R 2 /4g (9.17)

where is the rotational velocity, R is the chamber radius and g is


the gravitational constant. Equation (9.17) gives the ball rotational ve-
locity

2 0. 5
w=
R
d i
g hf (9.18)

The ball velocities calculated using Eq.(9.18) are found to be ap-


proximately equal to that of the impeller arms. Moreover, balls at the
periphery of the attritor mill have the maximum velocity, v = R, where
R is the chamber radius. The average velocity then can be obtained
by weighted average of the velocity of the number of balls. Since the
balls move generally in the same direction, the average collision ve-
locity (v 0) is half of this average velocity. On this basis, the collision
velocity in an attritor operating under typical conditions is 0.53 m/
sec. The mean free path between the collision surfaces is taken as the
mean free path between ball surfaces in a loose packed array of balls
(packing factor = 0.54).
In a Spex mill, it is assumed that the distance between collisions is
equal to the vial length. (The assumption is valid when the number of
balls in the mill is relatively low, so that ball collisions are infrequent.)
The vial accelerates over the first half of each trip and then it begins
to decelerate. The balls within the vial continue to travel at the maxi-
mum vial velocity. This velocity is twice the average velocity over this
range of motion, so the average impact velocity is taken as

81
Mechanical Alloying

2x
v0 = (9.19)
t

where x is the vial length. For a typical mill, the average velocity
is found to be 3.9 m/sec. These velocities are in good agreement with
the studies of Davis et al vis-a-vis a combination of computer simu-
lation and experiment [3].
The impact velocity in the case of a planetary mill can be calcu-
lated using the following relationship [12]

LM w F R - R I
3 b OP
v0 =M
H 2 K +w
v v
p wv
F RI
R P R - b
MM w p
p
PP H 2 K
v
(9.20)
N Q
where v and p are the angular velocity of the vial and the plate, re-
spectively, and R v and R b are the radius of the vial and the balls, re-
spectively. The impact velocity v 0 in planetary ball mills for typical
conditions is 11.14 m/sec [13]. This impact velocity is an order of mag-
nitude higher than the impact velocity in an attritor and is a primary
reason why processing times in a Spex mill or planetary mill are less
than in other MA devices. Estimated collision velocities, times between
impacts and mean free paths for common MA devices are given in Table
9.1.
For horizontal mills, the diameter of the drum varies from 1 to 2
m and the impact velocity ranges from 4.4 to 6.3 m/sec. These are
fairly high. The long processing times required by these mills are re-
lated to the longer time between collisions and/or the amount of ma-
terial impacted per collision.

Table 9.1 Typical impact velocity and impact frequency for common MA devices

Impact velocity, Impact frequency


Device
vo (m/sec) 1/tc (sec1)

Attritor mill 0.53 69

Vibratory mill 3.9 278

Planetary mill 11.24 90.7

Horizontal mill 6.3

82
Modelling Mechanical Alloying

Fig.9.7 Collision geometry at maximum


compression of a ball impacting on a
flat surface.

Estimation of material volume per impact


The material volume per impact is a cylinder with the volume r 2h h 0
(Fig.9.5). To know this, one must know the parameters rh and h 0. The
radius of the cylinder rh is estimated by considering collisions taking
place in the absence of powder. Such collisions can be approximated
using Hertzs theory of impact, which assumes that energy does not
dissipate during such impact. The kinetic energy utilized in elastic de-
formation of the colliding bodies is stored as elastic energy, which is
released as the bodies recover. This requirement is met if the relative
velocity of the impacting bodies is much less than the speed of sound
in the material. Sound velocities (v s ), which can be approximated as
(E/) 1/2, where E is the tensile modulus and is the density of the
impacting body, are of the order of kilometres per second, i.e. they
are about three orders of magnitude greater than the relative veloci-
ties of the colliding media.
Figure 9.7 shows the collision geometry of maximum compression
of a ball impacting on a flat plate having an infinite radius of curva-
ture and presumed mass much greater than that of the ball. If the ball
is compressed by a distance max, then the impact time (2) can be writ-
ten as
d max
2 t = k0 (9.21)
v0

where v 0 is the precollision velocity and k 0 the proportionality constant.


The distance, max can be expressed in terms of material and ma-
chine characteristics (i.e. milling machine, mill parameters) and the
relative collision velocity as follows:

b g
2 t = gt v0-0.2 r / E
0. 4
Rb (9.22)

where Rb is the ball radius, is the ball density, E is the tensile modulus
of the colliding media and gt is a parameter which depends on the col-
lision geometry, i.e. ball on ball, ball on plate (value is an order of
ten). If the ball and colliding surfaces are of different materials, Eeff
will be taken.
83
Mechanical Alloying

Table 9-2 Characteristics of Hertzian contacts for various mills (Ref.4)


Mill Material Rb(103m) 22 (105sec) rh(104m) Pmax(109 N/m2)

Attritor (ball on ball) Stainless steel 2.4 1.62 0.59 2.47

Tungsten steel 2.4 1.38 0.55 5.80

SPEX (ball on flat Stainless steel 2.4 1.25 1.99 4.16


surface)
Tungsten carbide on 5.6 3.12 4.81 6.18
stainless

SPEX (ball on curved Stainless steel 6.4 3.08 6.20 3.30


surface)
Stainless steel 2.4 1.22 2.10 3.85

The radius of contact between the colliding surfaces at maximum


compression rh, the Hertz radius, can be expressed as (using the Hertz
impact theory):

rh = gr v00.4 F rI 0. 2
Rb
H EK (9.23)

where g r is a geometrical proportionality constant (order of unity). The


average pressure developed across the contracting surfaces at maximum
compression can be used to determine the elastic strain energy involved
in the collision. The pressure is expressed as

Pmax = g p v00.4 F rI 0. 2
E
H EK (9.24)

where gp is also a geometrical constant, values of which range between


0.3 to 0.4. Actual values of impact time, Hertz radius and maximum
pressure are provided in Table 9.2 for some common MA devices. It
can be noted here that impact times are of the order of 10 5 sec, Hertz
radius of the order of 10 4 m and impact pressure of 10 9 N/m 2 .
When powder is entrapped between the colliding balls, we assume

Fig.9.8 The 'swept' volume model used to


determine the volume of powder impacted
during MA(The figure is appropriate to a
Spex mill) (Ref.9).

84
Modelling Mechanical Alloying

that Hertzian collision is mildly perturbed by the powder deformation,


fracture or coalescence processes.
Determination of h 0 is somewhat more uncertain. It has been de-
termined using a simple fluid mechanics in conjunction with the geo-
metrical sweeping mechanism (Fig.9.8). The cone height in Fig.9.8
is equal to the mean free path between colliding workpieces and has
a radius equal to the Hertzian one. It has been estimated that there
are of the order of one thousand particles entrapped between
workpieces in a single collision. If rh is taken as 210 4 m, a cylin-
der composed of 1000 spherical particles having a radius of 10 m and
with a 50% packing efficiency would have a height h 0 of 0.710 4 m.
Milling efficiency will clearly be greatest when the parameter
v 0t/2h 0 is close to unity.
The estimated times, Eq. (9.15), are the transient times during which
an appropriate steady state particle size distribution is being developed.
However, for chemical alloying, a transient time is also needed for
diffusional requirements during which alloy particles are developed.
These transient periods are not included in these models. Nevertheless,
the model provides milling times of the order of the observed ones.
Moreover, the predicted relative milling times for the different milling
devices are also in order.
The success of these mechanistic models depends on effective syn-
thesis of the local and global approaches. For this purpose, two com-
putational programs (MAP1 and MAP2) have been developed [14,15].
Programme MAP1 considers the behaviour of a single species with the
option of adding dispersoids. Program MAP2 considers two ductile spe-
cies (more pertinent), welded to form a third composite lamella (Fig.9.9).
However, the prediction of this program is accurate only with a fac-
tor of two or so.

9.2.4 Powder Heating


The temperature achieved during the collisions is of great significance
as the processes like alloying, intermediate phase formation, glass for-
mation, etc., which take place during MA, are profoundly influenced
by it. During conventional metal working operations, approximately 95%
of the work done is manifested in heat evolution and is so during the
MA as well. Assuming the plastic deformation work is entirely converted
into an (adiabatic) temperature rise in the impacted powder (per im-
pact), then

z
Vp s u de = Vp Cp DT (9.25)

85
Mechanical Alloying

where V p is the powder volume impacted and C p is the specific heat


of the powder.
Equation (9.25) gives the idea about the bulk temperature rise. De-
pending upon both milling conditions and the powder specific heat this
bulk temperature rise can vary from a few degrees to a few hundred
degrees [2,3,9]. Lim and Ashby have determined that the difference
between the surface temperature (T s) and the bulk temperature (Tb) dur-
ing such sliding conditions is of the order of 1 K [16].

9.2.5 Powder Cooling


Due to mechanical deformation, the powder temperature increases from
a base temperature. If the powders cool to this temperature prior to
experiencing an additional impact, the calculated temperature rise can
be used directly in analysis of diffusion taking place during MA.
As powder particles are small, temperature gradients within them can
be considered negligible and the cooling process can be analyzed in
terms of Newtonian cooling [9]. The rate of heat removal from the par-

Fig.9.9 MAP2 computional program (Ref.14).

86
Modelling Mechanical Alloying

ticle is the rate of heat transfer to the atmosphere:

-Vp Cp
F dT I = h AbT - T g
H dt K t a (9.26)

where h t is the heat transfer coefficient, A is the area of the particle


exposed to the atmosphere and T a is the ambient temperature. Although
particles are disc-like, we assume them to be spherical, and time (t)
can be given as

t = -8.53 10 -3 Cp
F rp2I ln bT - T g
a
GH k JK bT - T g
i a
(9.27)

where k is the particle thermal conductivity, r p the particle radius and


T i is the post impact temperature ( = T a + T). The values obtained are
of the order of 102 sec (for a particle of 50 m which has been heated
to several hundred degrees), which are much less than the time required
to have a subsequent collision in the MA mills (Table 9.1). Therefore,
it is safe to assume that particles will return to the ambient temperature
before being struck again.

9.3 ATOMISTIC MODELS


These models deal with the atomic approach to explain amorphization
achieved by MA. The structure of a crystalline solid is characterized
by a LRO and can be described in terms of the radii of the coordi-
nation spheres and lattice symmetry. As distinct from the crystal, an
amorphous solid has a short-range order and can be characterized by
the peak positions of the pair distribution function.
It is postulated that the relative change of the bond length be-
tween the amorphous and crystalline state of a pure metal can be chosen
as the criterion for the transition from the crystalline to amorphous
state. Thus, can be expressed in the following way:

FG R IJ - FG R IJ
i
a
i
c

h =
HR K HR K
l l
i
FG R IJi
c
(9.28)
HRK l

where the superscripts a and c refer to the amorphous and crystalline

87
Mechanical Alloying

Table 9.3 Relative changes in bond lengths for fcc, bcc and hcp lattices during
amorphization

Lattice 1 2 3

fcc 0.027 0.185 0.110


bcc 0.027 0.440 0.178
hcp 0.027 0.185 0.178

state respectively, i = 1,2,3 are the serial numbers of the co-ordina-


tion sphere, R ia is the i-th peak position of the pair distribution function
of the amorphous solid, Ric is the radius of the i-th co-ordination sphere
of the crystal, and R 1 is the nearest neighbour distance (assuming
R l a = R l c). Using Eq. (9.28), the change in bond length due to crys-
talline to amorphous transition can be calculated and this is given in
Table 9.3 [17].
The assumption that the high energy ball milling can result in changes
of the bond lengths equal to or greater than the value of i listed in
Table 9.3 would imply that the milling process can lead to crystalline-
to-amorphous transition of a pure metal. However, an amorphous phase
can not be produced by high energy ball milling of only one pure me-
tallic element as at least two elements are needed for successful
amorphization by MA [18].
When two elements, A and B, are mechanically alloyed, the impurity
atom A can occupy either a substitutional or one of the interstitial
positions in the B lattice. The lattice distortion made by an intersti-
tial, which is greater than that of a substitutional atom, can therefore
be used as a criterion for the ability to form an amorphous phase in
a bi-elemental crystalline system by MA. There should be two limits
to this criterion that consist of ASR of the constituent elements. A lower
limit below which the impurity atom can easily travel within the host
lattice and the distortion of the lattice is negligible, and an upper limit
above which the short range order of B atoms is destroyed.
Several kinds of interstices exist in lattices, keeping these in view
a lower limit of r A/r B, where r B is the radius of the matrix atom and
r A is the alloying (impurity) element. An upper limit can be calculated
which is used to predict the transformation.
The minimum solute concentration CBmin necessary for the amorphiz-
ation can be given by

vA 1
CBmin = l 0 = l0
vB - vB FG r IJ
B
3
-1 (9.29)
Hr K
A

88
Modelling Mechanical Alloying

where 0 is the atomic size factor, v A and v B are the atomic volumes,
respectively.
Such models have been successful in predicting the amorphization
even in the system with +H mix , as discussed in Section 5.2.1.

9.4 THERMODYNAMIC AND KINETIC MODELS


The basics of such models have been discussed in Section 5.2.1.
The study of these models developed clearly indicate a gap between
mechanistic and atomistic models. Thermodynamic and kinetic mod-
els are intended to fill this gap. Further progress in this area neces-
sitates the development of novel macrokinetic models linking the plastic
strain generation of defects to the mechanism and kinetics (i.e. solid
state diffusion and phase transformation) of a metastable phase for-
mation.

References
1. B.B. Khina and F.H. Froes, J. Metals, 7, 36 (1996).
2. R.W. Rydin et al, Met. Trans., 24A, 175 (1993).
3. R.M. Devis et al, Met. Trans., 19A, 2867 (1988).
4. D.R. Maurice and T.H. Courtney, Met.Trans., 21A, 289 (1990).
5. B.J.M. Aikin et al, Mat.Sci.and Engn., A147, 229 (1991).
6. J.W. Hilgers et al, Math.Mod., 6, 463 (1985).
7. K.L. Kuttler et al, Appl. Analysis, 19, 75 (1985).
8. B.J.M. Aikin and T.H. Courtney, Met.Trans., 24A, 2465 (1993).
9. D.R. Maurice and T.H. Courtney, Met.Trans., 21A, (2) 289 (1990).
10. T.H. Courtney and D.R. Maurice, In: Solid State Powder Processing, A.H. Clauer
and J.J. deBarbadillo (eds), The Minerals, Metals and Materials Society, Warrendale,
PA (1990), p.3.
11. V.L. Strater, In: Fluid mechanics, McGraw-Hill, New York (1951), p.25.
12. M. Magini, Mat.Sci.Forum, 88-90, 121 (1992).
13. M. Abdellaoui and E. Gaffet, Acta. Metall. and Mater., 43 (3), 1087 (1995).
14. D. Maurice and T.H. Courtney, Met. and Mater. Trans. A., 26A, 2431 (1995).
15. D. Maurice and T.H. Courtney, Met. and Mater. Trans. A., 26A, 2437 (1995).
16. S.C. Lim and M.F. Ashby, Acta. Metall., 35, 1 (1987).
17. Y. Chakk et al, Acta. Metall., 42 (11), 3676 (1994).
18. P.L. Brun et al, Scripta. Met. and Mater., 26, 1743 (1992).

Questions
1. With the help of a schematic show various possible fracture & coa-
lescence events which may take place during MA.
2. Give a typical account of impact velocity and impact frequency in
different MA devices; which mill is considered to be most effec-
tive?
3. Describe how the volume of powder impacted in single collision
can be estimated.

89
Mechanical Alloying

4. What is the significance of impact duration?


5. How strain per collision can be estimated.
6. How critical strain to fracture can be estimated.
7. Discuss computational model MAP2 to integrate local & global
models.
8. Discuss how impact velocity in the MA devices can be estimated.
9. Describe briefly the thermodynamic model based on CALPHAD
and ASR ratio for explaining amorphization by MA.
10. Describe the atomistic models to explain amorphization by MA.
11. Explain amorphization in intermetallic compounds; why is it not
possible to have SSAR in a pure element?

90
Joining of Mechanical Alloying Materials

10 JOINING OF MECHANICAL
ALLOYING MATERIALS
Despite the degassing of MA powders prior to their consolidation, MA
products still contain gaseous components which impair the joining of
these materials. Various potential joining techniques applicable to MA
materials are described here.

10.1 WELDING
In general, fusion welding is not suitable for joining MA materials. This
technique causes the dispersoid particles to be rejected from the molten
metal and the development of a high level of porosity in the weld zones
due to excessive degassing (Fig.10.1) [1]. Nevertheless, sound TIG
welds suitable for positioning or non load-bearing joints can be made.
Fusion welding processes, which minimize the size of the molten zone,
produce relatively sound welds. Spot and resistance seam welds with
excellent tensile strength can be made. However, it may never be pos-
sible to achieve the full stress rupture properties of the base material.
Solid state welding techniques, friction welding, diffusion welding,
explosive welding and magnetostrictive welding have demonstrated sound
weld joints [2]. Figure 10.2 shows the microstructure of a Dispal/Dispal
friction weld made by using a continuous drive type friction welding

1 mm

Fig.10.1 Optical micrograph of inert gas tungsten-arc melted Dispal showing porosity
in fusion region. (Ref.1, W.A. Kaysser and W.J. Huppmann (editors)), Horizons of Powder
Metallurgy, Part II, (1986), p.710 (Verlag-Schmid GmbH)).

91
Mechanical Alloying

500 m

30 m 0.9 m

Fig.10.2 Microstructure of Dispal/Dispal friction weld; (a) and (b) optical micrograph;
(c) TEM micrograph of the centre of the friction weld zone (Ref.1).

Fig.10.3 Continuous drive friction welding: basic principle (a); process variables (b).

machine [1]. During the process, one of the components is held sta-
tionary, while the other one is rotated at a constant speed (Fig.10.3).
As the two surfaces continue to rub against each other, heat is gen-
erated at the interface. After a certain weld-time, the rotating component
is stopped rapidly and the axial pressure is increased. Thus, the hot
material cools under pressure and the weld is consolidated. Due to
cooling of the molten pool under pressure, the welds have a fine micro-
structure without porosity (Fig.10.2), which leads to sound welds. The
variation of hardness within the weld zone is shown in Fig.10.4. Hard-
ness in the weld zone remains constant in the case of Dispal/Dispal,
but increases in the case of Dispal/steels. The bond strength in the
welds is given in Table 10.1.
Diffusion welding is performed by giving a suitable heat treatment
to the well-aligned polished surfaces. Diffusion welding of polished MA

92
Joining of Mechanical Alloying Materials

Fig.10.4 Hardness of a) Dispal/Dispal and b) Dispal/steel welds (100) (Ref.1).

Table 10.1 Ultimate tensile strength of friction welds

At room temperature At 300C At 400C


Material
(MPa) (MPa) (MPa)

Dispal/Dispal* 300 170 150

Dispal/stainless steel 260 175 146

Dispal/carbon steel 210 170 145

* For composition of Dispal refer to Section 12.3

6000 (see Section 12.1) surfaces have been performed by giving a heat
treatment at 1180C for 2 hrs [1]. The welded joint produced had good
tensile properties. However, the stress rupture properties in the bond
area were inferior to the base material due to fine grains. The prob-
lem of fine grains can be overcome by special surface treatment rec-
ommended by Helko [3]. However, the welding cycle used by Helko
is time consuming and requires well-aligned parallel surfaces as well
as a suitable atmosphere to get a sound weld everywhere.

10.2 BRAZING
Brazing is applied mainly for attachment of MA components in the
aerospace industry. The joint strength is of course limited by the strength
of available braze filler materials [2].

93
Mechanical Alloying

Fig.10.5 Optical micrograph of MA 6000, forged bonded and zone-annealed (Ref.5).

10.3 FORGED BONDING


A new method called forged bonding has been developed to join the
MA materials [4]. In forged bonding, the two surfaces are joined by
deforming the bond zone plastically in the fine-grained condition at the
same temperature and strain rate as used for bulk forging, followed by
subsequent recrystallization of the material to obtain the required grain
structure. The grains grow across the bond line, thereby creating a struc-
ture with very similar characteristics to the base material. The param-
eters (temperature, strain rate, strain) have to be controlled carefully.
Too low temperatures/too high strain rates lead to cracking and too high
temperatures/too low strain rates lead to loss of driving force and hence
insufficient recrystallization during the following treatment. Figure 10.5
shows a microstructure of a forged bond in MA 6000 [5]. Grain growth
over the bond line and microstructure of the bond region similar to the
base can be noticed. The yield stress, ductility, transverse stress rupture
strength and HCF properties of the forged bonded specimen are found
to be comparable with as-extruded and zone-annealed materials.

References
1. H. Kreve et al, In: Horizons of Powder Metallurgy, W.A. Kaysser and W.J.
Huppmann (eds), Verlag Schmid GmbH (1986), p.707.
2. J.J. Fisher and J.H. Weber, In: Proc. of Conf. on Structural Application of Mechanical
Alloying, F.H. Froes and J.J. deBarbadillo (eds), ASM International, Materials Park,
OH (1990).
3. K.H. Helko, US Patent 3, 787, 748 (1974).
4. R.F. Singer and G.H. Gessinger, Met. Trans., 13A, 1463 (1982).
5. H. Rydstad and R.F. Singer, Ibid Ref.1, p.713.

Questions
1. Why fusion welding techniques are not suitable for MA materials.
2. Give one example where diffusion welding has been used success-
fully for MA materials.

94
Joining of Mechanical Alloying Materials

3. What is forged bonding? In which case of MA alloy has it been


used successfully?

95
Mechanical Alloying

11 RAPID SOLIDIFICATION AND


MECHANICAL ALLOYING
Two far from equilibrium processing techniques, such as RS and MA
have been available for the last three decades to produce materials with
properties much superior to those obtained by conventional ingot met-
allurgy (IM) techniques.
The RS technique refers to cooling of metallic melts (most com-
monly using melt spinning or atomisation techniques) at rates of about
10 410 6 K. It is well established that RS can result in one or more of
the following constitutional and microstructural modifications [1]:
refinement of grain size and second phase particles or segregations;
extension of solid solubility;
formation of non-equilibrium crystalline or quasi-crystalline inter-
mediate phases;
production of amorphous phases.
On the other hand, MA involves repeated welding, fracturing, and
rewelding of powder particles in a dry high-energy mill. The process
can also lead to the above-mentioned metastable effects, but entirely
in the solid state. With either of the techniques, the product obtained
(powder only in MA and powder, foil, wire or ribbon in RS) has to
be consolidated into larger, more stable forms. To retain all the advan-
tages gained by RS or MA, consolidation should be done at relatively
low temperatures. Thus, the challenge is to achieve the necessary con-
solidation within a temperature window, which is often quite narrow.
The consolidation methods described in Chapter 7 are applicable to
either of these two types of materials.

11.1 RAPID SOLIDIFICATION VERSUS MECHANICAL


ALLOYING
A comparison of the RS and MA techniques is given in Table 11.1.

11.2 MECHANICAL ALLOYING OF RAPIDLY SOLIDIFIED


POWDERS
It is also pertinent to point out that MA of RS powders can provide

96
Rapid Solidification and Mechanical Alloying

Table 11.1 Comparison of RS and MA


A COMPARISON OF RAPID SOLIDIFICATION AND MECHANICAL ALLOYING

1. REFINEMENT OF MICROSTRUCTURES
High strength alloys can be developed through compositional modification and suitable
heat treatments. refined microstructures and higher solute contents, in conjunction with a
uniform distribution of fine precipitates during subsequent ageing of the RS alloy, have
been found useful for this purpose. For example, addition of about 1% of Fe, Ni, Zr or
Mn to conventional 7xxx series aluminium alloys improve their mechanical properties.
RS of these alloys results in about 20% increase in YS and UTS without loss of ductility
(2). These alloys also possess improved corrosion resistance and stress corrosion
cracking resistance. Similarly, addition of eutectoid formers (e.g. Co,Cr, Cu, Fe, Ni etc.)
to titanium using RS leads to much higher strengths than normally possible through IM
techniques. Table 11.2 lists the room temperature mechanical properties of RS alloys
showing the significant improvement achieved through this route. A similar but better
improvement in the properties of MA Alloys can be achieved using MA techniques
(Table 113).

2. The addition of 1% (about 3.5 at %) Li decreases the density of aluminium by about 3%


and increases the elasticity modulus by about 6%. A similar effect is also observed on
addition of B to titanium and Li to magnesium. Problem of segregation associated with
IM technique of AlLi can easily be overcome through the RS route.
Mechanical properties of the MA AL-Li alloys have an edge over the RS or other PM
alloys. however, the Li content should be limited to about 1.3% (13). If added more, it
has been observed that fabrication becomes difficult due to high strength of the alloy.
Thus, the RS approach allows increased density reduction while MA optimizes
mechanical properties.

3. DISPERSION STRENGTHENING
One of the serious difficulties encountered in the precipitation hardened alloys is the
coarsening of the precipate on elevated temperature exposure and consequent degradation
of the mechanical properties. Thus, development of high-temperature alloys requires very
low equilibrium solid solubility of the solute elements and a very low diffusion rate.
Addition of Fe, Cr and V transition metals such as to aluminium (4,5), and of rare-earth
elements like Er and Nd to titanium (6) by RS techniques are found to be useful in
retaining the reasonably high temperature strength levels in these alloys. The transition
elements form fine, thermically stable intermetallic dispersoid particles whereas the
rare-earth elements form uniformly dispersed oxides in situ. However, sometimes regions
near the grain boundaries are devoid in these materials (Fig.11.1).
It may be emphasized that MA is a much more powerful technique for dispersion
strengthening. Because of the heavy working involved, the matrix grain size is very fine,
occasionally reaching the nanometer levels, and further the distribution of dispersoids is
much finer and uniform throughout the matrix, at the interface, and on the subgrain
boundaries. MA enhances microstructural stability at elevated temperatures (Fig.11.2) ,
is attributed to the presence of fine scale dispersoids in the microstructure which inhibit
coarsening, recovery and recrystallization (7).

4. EXTENSION OF SOLID SOLUBILITY


One of the early advantages realized by RS is the extension of sollid solubility in several
alloy systems. This extended solid solubility coupled with the under cooling experience
by the melt lead to enhanced solid-solution and fine-grain strengthening. RS can also
avoid formation of coarse inter metallics and insolubles in some of the systems.
Solid solubility extensions has been achieved in AlMg, AlSi, AlCu and MgAl
alloys by mechanical alloying. Although, these values are significantly higher than those
in equilibrium state but lower than values achieved in RS alloys (Table 11.3) (8,9). On
the other hand, since MA is carried out completely in the solid state, alloying
capabilities are much higher through this route. Addition of magnesium to titanium by
conventional methods is difficult becuase magnesium boils before titanium melts; but MA

97
Mechanical Alloying
conventional methods is difficult becuase magnesium boils before titanium melts; but MA
allow about 3% Mg dissolution in titanium (10). further, MA can produce homogeneous
solid solutions in liquid-immiscible (CuPb) and solidimmiscible (CuFe) systems
(11).

5. AMORPHIZATION
The free-energy diagrams perdict that the homogeneity range of amorphous alloys formed
by RS is usually divided into relatively narrow regimes located near deep eutectics in
the liquidus (refer Fig.5.8).
In contrast, the composition range of the single phase amorphous alloy prepared by solid-
state reaction is wide and continuous, and is located near the centre of the composition
range. (Fig.11.3) (12). Mechanical alloying prevents intermetallic phase formation during
interdiffusion because of extremely thin layer, thus resulting in a wider glass forming
range(13).

6. MA amorphous phases have a relatively more relaxed disordered structure with a more
developed short range ordering and contains a large milling induced stored energy.(14)

7. Amorphization in many systems like Ni-Pd, and Ni-Ti intermetallic is not possible by RS
technique.
But MA made the amorphization possible in such cases.(12)

The MA amorphouse alloys and rapidly quenched alloys, both exhibit very similar
properties like crystallization temperatures and atomic structures.

Table 11.2 Mechanical properties of RS and MA aluminium alloys (Ref.1)

0.2% YS UTS K1c


Temper %El.
(MPa) (MPa) (MPam)

1M 7075
T6 503 572 10
RS
Extruded 634 717 9
7075(+1/Ni+1/Fe)
T7 Extruded 580 614 12 47
RS CW67
T6 595 645 3.5
MA 7075
848 986 13
1M Ti6A/4V
1090 1110 2.5
RS Ti6A/4V+1B
T4 277 464 22 32
1M 2024
326 542 24
RS 2024
T4 500 570 12 30
MA IN 9021
T6 560 600 12 44

CW67: Al9.0 Zn2.5% Mg1.5% Cu0.14% Zr0.1%Ni


MA IN 9021: Refer to Section 12.3

synergetic effects and remove many limitations of RS materials as dis-


cussed above.

Aluminium system
Rapidly solidified plus mechanically alloyed (RSMA) aluminium
titanium alloys have been shown to have elevated temperature mechani-
98
Rapid Solidification and Mechanical Alloying

Table 11.3 Extension of solid solubility by MA/RS (Ref.8)


Solid solubility (at%)
Solvent Solute
element element Mechanical Rapid
Equilibrium
alloying solidification

Mangesium Aluminum 1 3.7 22.4


Aluminum Magnesium 1 14.1 36.8
Aluminum Copper < 1 5.6 17.3
Aluminum Silicon 1 4.5 11.1

Fig.11.1 TEM micrographs showing the size and distribution of dispersoids in: a) RS
Ti3Al2% Er alloy;(b) MA Ti3Al20%Nb2% Mo3%V2% Er (Ref.1, Light Metal Age,
6, 18 (1989)).

Fig.11.2 Grain growth in heat treated ribbon, extruded ribbon-powder, and extruded
and heat treated MA powder of Cu5% Cr (Ref.7, E.Artz and L. Schultz (editors)),
New Materials by Mechanical Alloying (1989), p.3 (Deutsche Gesellschaft fr
Metallkunde).

Fig.11.3 Comparison of the homogeneity ranges of the amorphous phase of NiZr, Co


Zr, and FeZr alloys prepared by MA and RS (Ref.12).

99
Mechanical Alloying

Fig.11.4 Effect of temperature on yield strength of several types of AlTiX alloys


(Ref.15).

Fig.11.5 Effect of temperature on tensile elongation of several types of AlTiX alloys


(Ref.15).

cal properties superior to those of RS aluminium alloys [15]. The size


and distribution of intermetallics are refined through RS and MA in-
troduces a fine distribution of oxide and carbide particles which con-
trols the grain size and enhances grain boundary strength at elevated
temperatures.
The effect of the RSMA process on the yield strength and elongation
for these alloys is shown in Figs.11.4 and 11.5, respectively. The
ambient temperature yield strength of melt spun (MS) Al4%Ti (224
MPa) is 28% greater (Fig.11.4) than that of atomised (AT) Al4%Ti
(178 MPa) due to a faster cooling rate achieved in the case of MS (106
K/sec) than AT (10 4 K/sec), which is believed to result in a higher vol-
ume fraction of fine intermetallic particles. Mechanical alloying enhanced
both the ambient and elevated temperature strengths of these alloys pro-
duced by RS, i.e. gas atomization and melt spinning. AM6 alloy has
a room temperature yield strength (325 MPa) 80% greater than that
of AT6 alloy (180 MPa). At 300C, the yield strength of AM6 alloy
(196 MPa) is 105% greater than that of AT6 alloy (96 MPa). The

100
Rapid Solidification and Mechanical Alloying

MSMA 185 alloy has an ambient temperature yield strength (526 MPa)
106% greater than that of the RS Al3%Ti3% Ce alloy (255 MPa).
At 300C, the yield strength of MSMA 185 alloy (172 MPa) is 25%
greater than that of MS Al3%Ti3%Ce alloy (137 MPa). Elongation
in the MS alloys is found to be 22% (Fig.11.5). The MA, however,
reduces ductility. AM6 alloy has an elongation of 9% vis-a-vis alloy
AT6 of 22%. Similarly, MA Al3%Ti3% Ce has a tensile elongation
of 5%, but the MS Al3%Ti3% Ce alloy has elongation of 21%. Os-
tensibly, the reduction in ductility can be attributed to the introduction
of oxide and carbide dispersoids during MA. These dispersoids pref-
erentially locate themselves on the grain boundaries and inhibit defor-
mation and limit plastic accommodation.
Ternary RS AlFeCe alloys can be used up to 327C. However,
these alloys lose strength rapidly at high temperatures (see Fig.12.12).
To overcome this deficiency, the RS Al8.4%Fe3.4%Ce alloy has been
mechanically alloyed in a Spex mill to uniformly disperse carbides,
oxides and intermetallic phases (metastable Al 10 Fe 2 Ce, and stable
Al13Fe3Ce and Al13Fe4 intermetallic crystalline) in the aluminium, which
results in the increase in strength and stiffness [16,17].
In another investigation using the RSMA process for Al4%Cu
1%Mg1.5%Fe0.75%Ce alloy, the strength level could be increased
to 435 MPa [18].

Titanium system
Two advanced titanium-based alloys, Ti1% Al8%V5% Fe1% Er and
Ti24%V10%Cr5% Er, both with extensive rare earth additions have
been developed using the RSMA process. The Ti1%Al8%V5%Fe
1% Er alloy consolidated directly from a gas atomised powder had
coarser beta grains of about 30 m. A MA of up to 40 hours [19] did
not yield a totally uniform structure, but the majority of the structure
in the consolidated alloy consisted of submicron grains with dispersion
of 30 to 50 nm in size. The fine-grain structure was found to be stable
with little grain coarsening, and a little drop in hardness was observed
on solution treatment at 675C.
The Ti24%V10%Cr5%Er alloy in the as-atomised condition was
badly segregated with large chunks of free Er. The microstructure of
the consolidated RSMA alloy consisted of both very fine grained re-
gions and some coarser grains, with dispersion in fine grained regions
of less than 10 nm in size. The RSMA alloy had higher hardness (437
in comparison with 304 KHN) due to the grain refinement and disper-
soids.
Thus, fine dispersions and stable grain refinements are possible by

101
Mechanical Alloying

adopting a RSMA process which leads to improvement in the properties


of RS materials.

References
1. C. Suryanarayana and F.H. Froes, Light Metal Age, 6, 18 (1989).
2. P.K. Domalavage et al, Met. Trans., 14, 1599 (1983).
3. W.E. Quist et al, In: Aluminium-Lithium Alloys, C. Baker et al (eds), The Institution
of London (1986), p.625.
4. P.S. Gilman et al, Industrial Heating, 56 (2), 30 (1989).
5. Y.M. Kim, In: Proc. Conf. on Dispersion Strengthened Aluminium Alloys, Y.M.
Kim and W.M. Griffith (eds), The Minerals, Metals and Materials Society,
Warrendale, PA (1988), p.157.
6. S.M.L. Sastry et al, Met. Trans., A15, 1451 (1984).
7. D.G. Morris and M.A. Morris, In: Prof. Conf. on New Materials by Mechanical
Alloying, E. Artz and L. Schultz (eds), Deutsche Geselschaft fr Metallkunde,
(1989), p.3.
8. C.R. Clark et al, In: Proc. Int. Conf. PM 2 Tech 95, M.A. Phillips and J. Porter
(eds), MPIF, Princeton (1995).
9. Z.A. Zhang et al, Met. and Mater. Trans., 25A, 73 (1994).
10. R. Sundresan and F.H. Froes, Key Eng. Mater., 29-31, 199 (1989).
11. R. Sundresan and F.H. Froes, J. Metals, 39 (8), 22 (1987).
12. E. Hellstern et al, J. Less-common Metals, 140, 1 (1988).
13. J. Eckert and L. Schultz, Mat. Sci. and Engn., A134 1389 (1991).
14. A. Inoue et al, Mat. Trans. JIM, 32 (2) 148 (1990).
15. W.E. Frazier and J. Cook, In: Solid State Powder Processing, A.H. Clauer and
J.J. deBarbadillo (eds), The Minerals, Metals and Materials Society, Warrendale,
PA (1990), p.257.
16. M.L. Ovecoglu and W.D. Nix, In: New Materials by Mechanical Alloying Techniques,
E. Artz and L. Schultz (eds), Deutsche Gesellschaft fur Metalkunde, Oberusel,
Germany (1989), p.287.
17. M.L. Ovecoglu et al, Met. and Mat. Trans., 27A, 1033 (1996).
18. P.S. Gilman and K.K. Sankaran, Ibid Ref.5.
19. R. Sundares and and F.H. Froes, 6th World Conference on Titanium, Cannes, France
(1988).

Questions
1. In what way are the two techniques MA and RS similar?
2. In what do mechanically alloyed materials differ from rapidly solidi-
fied ones?
3. Give two examples of amorphization reaction which can be achieved
by MA but not by RS.
4. Explain how magnesium can be added to titanium?
5. How AlLi alloys produced by RS techniques from from that of MA.
6. What advantages are associated with MA of rapidly solidified mate-
rials? Explain the phenomenon with the help of suitable examples.

102
Applications

12 APPLICATIONS
Since its advent, the MA technique has been applied to develop various
novel compositions and improve the performance of existing materials.
Many of these are produced at industrial level and find applications in
the commercial sector, while many are potential candidates to find an
application. This is discussed in the following section.

12.1 NICKEL-BASE SUPERALLOYS


Four nickel-base superalloys are now produced commercially using the
MA technique [1]. These are INCONEL MA 754, MA 758, MA 6000
and MA 760. The nominal composition of these alloys is given in Table
12.1. MA 754 has a solid solution strengthened Ni20% Cr matrix with
only low levels of titanium (0.5%) and aluminium (0.3%) along with
0.6% Y 2O 3 as the dispersoid. It can withstand up to 1100C. MA 758
has been developed for resistance to molten glass corrosion. In MA
6000 alloy, tungsten and molybdenum provide solid solution strength-
ening. Chromium and aluminium with titanium, tantalum and tungsten
improve oxidation and sulphidation resistance. MA 760 is a relatively
new variant of commercial alloys prepared by MA. It has been designed
for the purpose of achieving a balance of high-temperature strength,
long-term microstability and oxidation resistance.
Batches of mechanically alloyed superalloy powders are produced
in a commercial size attritor mill under controlled conditions. A mix-
ture of powders of nickel (particle size = 5 m), chromium, molyb-
denum, tungsten, tantalum and a master nickel-base alloy powder con-
taining the reactive elements aluminium, titanium, boron and zirconium

Table 12.1 Composition of MA nickel-base superalloys*


Alloy Ni Cr Al Ti Mo W Y 2O 3

MA 754 Bal. 20 0.3 0.5 0.6


MA 758 Bal. 30 0.3 0.5 0.6
MA 6000 Bal. 15 4.5 2.5 2.0 4.0 1.1
MA 760 Bal. 20 6.0 2.0 3.5 0.95

* Products of Inco Alloys International Inc

103
Mechanical Alloying

Fig.12.1 Basic process of consolidation and recrystallization for MA nickel superalloys.

(particle size 150 m) is used. Yttrium oxide is introduced into the mix-
ture in the form of 1 m aggregates, each consisting of numerous par-
ticles of 20 to 40 nm diameter. The MA powder is then degassed in
vacuum at 538C and sealed in mild steel cans. These cans are then
extruded using an extrusion press at temperatures in the range 1010
1066C and at an extrusion ratio of 13:1 (Fig.12.1). The microstruc-
ture at this stage consists of incredibly fine (0.4 m) equiaxed grains
resulting from dynamic recrystallization during the hot deformation. Hot
rolling is carried out at 1025C with 20% thickness reductions per pass
with a total thickness reduction ranging up to 90% [2]. Afterwards,
zone annealing is carried out at 12301290C, which increases the grain
aspect ratio to more than 15 by a process of secondary recrystallization,
giving a structure reminiscent of directional solidification [3] and hence
the high-temperature strength of the alloy. In general, 4060% thick-
ness reductions are required to produce excellent directional
recrystallization. The recrystallized grains tend to assume plate-shaped
morphologies (see Fig.7.4).

Fig.12.2 TEM micrograph of recrystallized MA 760 superalloy, cooling rate 10C per
min (courtesy Dr H.K.D.H. Bhadeshia).

104
Applications

Microstructural observation shows that precipitates when fully


recrystallized appear as cubes in TEM (Fig.12.2). In general, three
types of particles Y 2 O 3, Al 2O 3Y 2O 3 garnets and chromium carbides
form during recrystallization rather than during extrusion [4]. Chromium
carbides are found to be present in large quantities in the extruded ma-
terial. These carbides get dissolved during heating to recrystallization
and can reprecipitate during subsequent cooling if the alloy sample is
not recrystallized. In recrystallized samples, carbides do not reprecipitate
rapidly because of the dearth of grain boundary nucleation sites.
Inhomogeneous distribution of particles appears in forms of stringers
and also as carbide depletion zones (Fig.7.4). However, particle depletion
has a small influence on the local microhardness.
The properties of these MA superalloys are shown in Table 12.2.
They possess excellent high- and low-cycle fatigue resistance strength
(~1300 MPa) and ductility (~3%). Comparison of 1000 hrs specific
rupture strength with three established directionally solidified and single
crystal alloys (Fig.12.3) shows that these materials are superior above
900C. The failure mechanism at elevated temperature is intergranular
fracture along transverse grain boundaries nucleated by cavities that
form during grain boundary sliding. Nucleation of voids is retarded due
to diffusional accommodation of grain boundary sliding. Depletion of
chromium, aluminium and titanium in the surface zones, due to pref-

Table 12.2 Typical properties of MA nickel-base superalloys (Ref.5)

Alloy
Property
MA 754 MA 758 MA 6000 MA 760

Alloy type NiCr NiCr NiCr ' NiCr '


3
Density, g/cm 8.3 8.14 8.11 7.88

Young's modulus at 20C, 149 203


GPa

Yield strength, 0.2% 134 147 192 140


offset, MPa (at 1095C)

Tensile strength, MPa (at 148 153 222 141


1095C)

Elongation, % (at 12.5 9 9 15


1095C)

Stress to rupture, MPa (at 1095C)

100 hr 102 50 131 115

1000 hr 94 127 105

105
Mechanical Alloying

Fig.12.3 Stress for 1000 hours life to rupture as a function of temperature for alloys
MA 6000, and other reference material.

Fig.12.4 (right) Comparison of oxidation and sulphidation resistance of MA 6000 with


other superalloys (Ref.2).

erential evaporation, also contributes to initiation of the failure. Fig-


ure 12.4 shows that MA 6000 alloy possesses excellent oxidation re-
sistance and sulphidation resistance equal to IN-100 and IN-792, re-
spectively, which can be attributed to the homogeneous distribution of
the alloying elements and the improved scale adherence due to the
dispersoid itself. The mechanical properties of the MA materials com-
bined with their resistance to high-temperature corrosion/oxidation attack
offer a level of performance that cannot be obtained in conventional
alloys.
106
Applications

The MA nickel-base superalloys are difficult to form. However, good


progress has been made in forging and forged components which are
in commercial use. Other forming operations such as hot shear forming,
ring rolling and hot upsetting have been done at the stage when alloy
has low strength and high ductility in fine grain conditions.
The MA 754 alloy has been used as components in the hot section
of military jet engines and gas turbine vanes at operating temperatures
above 1000C (Fig.12.5). Many of the recently developed single crystal
alloys are unable to perform in these type of applications unless they
are plasma sprayed with MCrAl(Y) type materials to improve the oxi-
dation resistance.
The MA 760 alloy is mainly for the use in gas turbine components.
The alloy combines excellent hot corrosion resistance with long term,
high-temperature strength that is beyond the capabilities of conventional
superalloys.

Fig.12.5 Brazed aircraf, gas-turbine vane assembly fabrication of MA 754 superalloy.

12.2 MA STEELS
Many MA materials are now commercially available, but steels produced
using MA show particular promise in a variety of applications. For
example, MA 956 (Fe20% Cr4.5% Al0.5% Ti0.5% Y2O3) is a chro-
mium-rich, ferritic stainless steel containing aluminium for oxidation re-
sistance, (aluminium forms an Al 2O 3 protective layer at high tempera-
tures) together with a dispersion of Y 2O 3 particles for creep resistance.
An alternative ferritic stainless steel variant MA 957 (Fe13.5%Cr
0.3%Mo1%Ti0.3%Y2O3) contains titanium rather than aluminium, and
is designed for application in the nuclear industry. MA 956 can be used
at operating temperatures of over 1300C in corrosive atmospheres. The

107
Mechanical Alloying

b
Fig.12.6 Components made of MA 956: a) burner nozzles machined from forgings; b)
air-stream swirlers fabricated from sheet.

Fig.12.7 Use of MA 956 in heat treatment equipment: a) vacuum furnace hearth; b)


heat treatment furnace basket fabricated from rods.

alloy is used as sheet material for aircraft and industrial gas turbine
combustors, in burners, swirlers and heat exchangers of power generation
equipment (Fig.12.6), and in heat treatment equipment (Fig.12.7) [5,1].

108
Applications

INCONEL MA 957 alloy has been developed for an application as


nuclear fuel cladding material in a fast breeder reactor. Conventional
austenitic alloys can not be used in this application due to swelling
caused by the high neutron fluxes, while conventional ferritic steels
generally have inadequate creep strength at the service temperature of
700C. These steels are produced basically to meet these requirements.
The steels after MA and consolidation have an ultrafine microstructure
containing submicron grains of ferrite. The hardness in this condition
is unacceptably high (in contrast to the MA nickel-base superalloys),
so the steels are usually recrystallized into a coarse directionally
recrystallized grain structure which is also ideal for elevated temperature
applications where creep resistance is of prime importance. The MA
956 steel exhibits a strong {110} <100> fibre texture in the extruded
condition, while MA 957 has a relatively poor texture. However, af-
ter recrystallization there was a significant difference in texture which
cannot be explained on the basis of deformation or recrystallization
theory. Chou and Bhadeshia [6] revealed that the ferrite present in MA
957 partially transforms to austenite at the extrusion temperature (970
to 1010C) because of its relatively low chromium concentration and
the absence of aluminium. The austenite then transforms to martensite
on cooling, giving a new set of orientations and, consequently, a more
random crystallographic texture. During annealing, recrystallization nu-
cleates via both a subgrain coalescence mechanism and grain-bound-
ary migration. The existence of austenite phase enhances the develop-
ment of cubic texture.
Dour Metal have developed a series of oxide dispersion microforged
material (ODM) iron base FeCrAl alloys, by MA, which meet the re-
quirements of strength and oxidation resistance at high temperatures [7].
The composition of these alloys is given in Table 12.3, while a com-
parison of their properties is given in Table 12.4. These materials have
good hot forming properties in the microcrystalline condition as com-
pared to conventional FeCrAl alloys. The ODM has been found suit-

Table 12.3 Composition of Fex%Cry%Al1.5%Mg0.6%Ti0.5%Y2 O 3 alloys

Alloy Cr% Al%

ODM 331 13 3

ODM 361 13 6

ODM 031 20 3

ODM 061 20 6

ODM 751 16.5 4.5

109
Mechanical Alloying

Table 12.4 Hot tensile properties of ODM alloys

Microcrystal 900C Macrograin 900C

Alloy YS, 0.2% YS, 0.2%


UTS Elongation UTS Elongation
Offset Offset
(MPa) (%) (MPa) (%)
(MPa) (MPa)

331 28 41 157 169 176 10

361 42 61 86 166 177 12

061 49 71 88 159 173 16

751 32 47 134

FeCrAl 35

MA956 108 115 8

F10R 165 16

able even at temperatures above 1100C, which is the upper limit for
the use of superalloys for both the mechanical properties and the oxi-
dation resistance. A suitable heat treatment of ODM has also been
developed to favour the transgranular rupture in these materials, which
improves ductility. When compared with MA 956, the time to rupture
is multiplied by a factor of 100. The superior mechanical properties
of these ODMs above 1000C is due to the strengthening effect of
dispersoids. Thus, it is envisaged that these ODMs may find applications
in energy conversion systems (particularly of heat exchanger tubes) as
well as other high temperature applications in corrosive atmospheres.
For the requirements of modern combustion engines, high wear re-
sistance steel having a tough matrix and hard, fine dispersoids is re-
quired. In this respect, four hard phases can be used, namely: NbC,
TiC, TiN and Al 2O 3 . All other hard phases are either too soft or un-
stable within the steel matrix at sintering temperatures. Sintered steels
containing hard phase in the range of (1025 vol.%) have been pro-
duced by MA of elemental composition in an attritor mill followed by
liquid phase sintering at 12801320C in a hydrogen atmosphere. A high
phosphorous content leads to a harmful embrittlement of the grain
boundaries, as a consequence the phosphorous content is chosen very
carefully [8,9]. The influence of this hard phase on the mechanical
properties of the composite material is given in Table 12.5. It can be
observed that the effect of wettabilities of different hard phases with
the matrix on sintered densities of the composite material is much less.
However, stiffness of the composite increases as a function of Youngs
modulus of the reinforcing component. The strength of Al2O3-containing

110
Applications

Table 12.5 Properties of the used hard phases and the composite materials with a
composition of Fe10 vol% hard phase 0.6% P0.9% C

NbC TiC TiN Al2O3

Hard phase density (g/cm 3) 7.78 4.98 5.40 3.98

Young's modulus (kN/mm 2) 580 470 590 410

Composite material theoretical


7.85 7.57 7.61 7.47
density (g/cm 3)

Sintered density achieved (% TD) 97.9 98.2 97.4 97.46

Tensile strength (N/mm 2) 1008+88 995+66 806+64 742+21

Elongation (%) 3.7 3.2 1.8 1.6

Young's modulus (kN/mm 2) 229 218 205 215

material is 25% less compared to that of NbC, and the elongation frac-
ture is 50% less. Surprisingly, sintering densities are found to be in-
creased with the amount of an inert second phase in the composite
which does not take part in the sintering process. It has been attrib-
uted to the fact that sintering in the composite materials takes place
by means of intergranular diffusion for a longer period of time while
the grain size remains more or less stable. The wear behaviour of these
materials has been found to be the best when the hard phase content
is about 10 vol.%. It is also reported that dry wear is reduced using
coarser hard phase. For oil lubrication at higher surface pressures (800
N/mm 2) and longer times (100 hrs), finer hard phase particles lead to
less wear. Thus, for oil lubricated tribosystems, these hard phase con-
taining sintered steels offer a better alternative to conventional wear
resistance PM steels like sintered and forged T15 MSS. Moreover, the
matrix hardness can be controlled within a wide range by heat treat-
ment and, hence, can be matched to the respective application.

12.3 ALUMINIUM-BASE MATERIALS


The MA process can be applied to aluminium-base systems provided
that organic process control agents are used to control the extreme ten-
dency of aluminium to weld to itself during high energy milling. All MA
aluminium products have highly uniform dispersion of equiaxed oxide
(Al2O 3) and carbide (Al 4C 3) particles, formed due to the reaction of alu-
minium with oxygen and carbon (formed as residue of PCAs) parti-
cles during the consolidation. These dispersoids of extremely fine sized
oxide and carbide particles (3040 nm) stabilise a very fine equiaxed
grain (0.2 to 0.5 nm) and dislocation substructure during thermomech-

111
Mechanical Alloying

Fig.12.8 Flow sheet for processing of MA aluminium alloys (Ref.11).

anical processing [10]. In addition to the strengthening caused by these


finely dispersed oxides and carbides, a significant part of the strength-
ening in these alloys is due to the fine grain size and high dislocation
density resulting from the severe working of the powders. Further ben-
efits can be obtained in these alloys by adding small amounts of carbide-
forming elements such as titanium. A typical flow sheet for process-
ing MA aluminium alloys is shown in Fig.12.8 [11]. The development
of MA aluminium alloys can be put into three classes: high strength
alloys, low density alloys and high temperature alloys.

High strength alloys


The major application of MA aluminium alloys is at ambient tempera-
tures rather than elevated temperatures. One example is IN 9052
(Al4% Mg1.1% C0.8% O) (Fig.12.9). Although a solid solution alloy,
IN 9052 gives a combination of tensile strength, fatigue strength and
toughness normally seen only in precipitation hardened 7000 series
alloys. The high strength of this alloy arises from magnesium solid
solution strengthening, ultrafine grain size, finely dispersed MgO, Al2O3
and Al 4 C 3 and substructural strengthening [12]. Strengthening due to
precipitation can be superimposed on the above mechanisms when
higher strength is essential but corrosion resistance is of secondary im-
portance. Because it is a solid solution material, IN 9052 also possesses

112
Applications

Fig.12.9 TEM micrograph showing fine-grain structure of IN 9052.

Fig.12.10 Strength and corrosion resistance of IN 9052 compared with conventionally


made alloys (Ref.11).

excellent resistance to corrosion and stress corrosion cracking


(Fig.12.10). The fatigue crack growth rates of IN 9052 and another
mechanically alloyed aluminium alloy IN 9021-T4 (Al4%Cu1.5%Mg
1.1%C0.8%O) are comparable with that of 7075 alloy. These alloys
are being considered suitable for marine and metal matrix composites.

Low-density alloys
For the aerospace industry, an alloy with properties equivalent to or
better than those of the widely used 7075-T73 alloy and with lower
density is desired. For density reduction in aluminium, there are generally
only two choices of alloying additions that merit practical considera-
tion lithium and magnesium. The proportion of these elements which
can be tolerated is limited by the possible occurrence of stress-corrosion

113
Mechanical Alloying

cracking at levels above the solubility limits. The possibility of achieving


up to 15% reduction in density along with up to 10% increase in elastic
modulus (amounting to a combined increase of approximately 30% in
specific strength) by alloying aluminium with lithium is possible. The
aluminiumlithium system seems to hold high promise in the aerospace
industry and because of its plasma-confinement properties, it also has
prospects in nuclear engineering [13]. The existing production meth-
ods of casting and powder-making by atomisation lead to a material
of inhomogeneous composition structure. Lithium or lithium compounds
(Al 3Li) are segregated at the grain boundaries, causing excessive brit-
tleness of the material and reducing its resistance to fatigue.
The MA process has specific advantages in the case of highly re-
active constituents (like Al, Mg and Li) under atmospheric conditions
as well as for constituents which have high partial vapour pressure at
their liquid temperatures, as is the case with lithium. The difficulties
inherent in the production process are due to the wide disparity be-
tween the three constituent metals. Pure aluminium is ductile, while mag-
nesium is relatively brittle. Fine magnesium powder is pyrophoric and
tends to ignite spontaneously under ambient conditions. Pure lithium
metal cannot be handled directly, and must always be kept under a
protective liquid (e.g. paraffin oil). It is very soft and has a high rate
of sublimation at moderately high temperatures.
Al3% Li2.2% Mg has been produced by carrying out MA in a vi-
bratory mill in a vacuum of 10 5 torr. The master alloy powder is first
produced by milling a mixture of Mg powder and AlLi powder. The
aluminium powder is used as a diluent. The difficulty presented by the
tendency of the soft aluminium particles to agglomerate and adhere to
the grinding medium of the ball-mill is removed by the presence of a
brittle (both magnesium and AlLi) constituents, which interposes itself
between the aluminium particles and keeps them apart [14]. Flow-sheet
of the process is shown in Fig.12.11. The X-ray diffraction (Fig.12.12.)
showed that AlLi undergoes partial solution after grinding and full
solution following hot pressing. The Mg is found to distribute uniformly
in the alloy. In the fine grains, Li diffuses to the grain boundaries form-
ing a (AlLi) phase which causes brittleness because of its incoher-
ence. However, in the case of larger grains, coherent precipitates of
-(Al 3Li) are coherent with the matrix. Thus, MA provides a micro-
structure with coarse grains having a precipitate of hard -(Al3Li) and
very fine grains, which improves the mechanical properties at room tem-
perature. After solution treatment at 450C and ageing at 150C, the
alloy achieves the highest value of specific modulus of elasticity E/
(28). This makes the alloy suitable for aviation application which can
be ascribed mainly to the coherent precipitation of (Al3 Li) and to
114
Applications

Fig.12.11 Flow sheet of the production process for AlLiMg alloy (Ref.14, Powd.
Met. Int., 22(3), 22 (1990) (Verlag Schmid)).

Fig.12.12 X-ray diffraction pattern of an Al2% Li4.4% Mg powder: a) before MA;


b) after MA; c) after hot pressing (Ref.14).

115
Mechanical Alloying

the low density of the lithium, the lightest existing metallic element.
Based on the work of S.J. Domachie et al [15,16], IncoMAP has
developed the IN-905 XL alloy (Al4% Mg1.5% Li1.2% C0.4%O)
which matches 7075-T73 in strength but is 8% less dense and 15% more
stiff. The outstanding feature of the alloy is the minimal degradation
of properties when stressed in the transverse direction. This charac-
teristic is of great importance in forgings which are uniaxially stressed.
Its general corrosion resistance is about 100 fold better than that of
7075-T73. Heating cycles up to 450C do not alter the strength. This
characteristic can be important in service as well in manufacturing.
Figure 12.13 shows a prototype aircraft wheel forged of this alloy.

Fig.12.13 Prototype aircraft wheel forged of IN 905 XL.

High-temperature alloys
The development of MA high-temperature aluminium alloy can be put
in three classes:
DISPAL
AlTi alloys
AlFeCe alloys
The MA process has lead to the development of Dispersion Strength-
ened Aluminium (DISPAL) high-temperature aluminium alloys [17]. The
alloy powders are processed directly by reaction milling with carbon
black and controlled additions of oxygen from the milling atmosphere
and with this independent control of the dispersoid content, the alloys
can suitably be tailored for applications. Already, forged automotive pis-
tons, high-temperature electrical conductors, interferometers and structural
parts in aerospace cryogenic fields have been made from various
DISPAL alloys.

116
Applications

To replace titanium in aerospace substructures and the cooler parts


of advanced jet engines, Al(7.515)% Ti alloys, e.g. E5(Al11.6% Ti
1.9% C0.7% O), have been developed. These alloys consist of a
matrix of aluminium containing a second phase of Al3Ti particles. Several
characteristics of AlTi alloys make them potentially attractive for en-
gineering use. The high melting point of Al 3Ti and the low solubility
and low diffusion rate of titanium in aluminium make them suitable for
elevated temperatures. In addition, increased stiffness can be derived
from the high elastic modulus of Al 3Ti. When adding more than 0.15%
(0.15% amount is sufficient for grain refinement), Al 3Ti particles pre-
cipitate directly from the melt and grow to large sizes, which decreases
ductility. Even rapid solidification techniques have failed to control the
phenomenon as the addition of Ti greatly increases the liquidus tem-
perature. The MA of elemental aluminium and titanium powders dis-
tributes titanium uniformly in the matrix of aluminium. Alloying is car-
ried out in a chilled water-cooled attritor mill in a flowing argon at-
mosphere. The MA powder is degassed at 550C and extruded into bars
at the same temperature with an extrusion ratio of 25:1. During heating
for consolidation, the titanium and aluminium react in situ to form Al3Ti
particles about 20250 nm in diameter, providing a structure not pos-
sible with conventional manufacturing methods. A MA Al12.5%Ti (E6)
alloy has an elastic modulus of 50% greater than that of conventional
aluminium alloys, accompanied by a density increase of 4% with good
ductility and high strength at elevated temperatures [18]. The high
strength of the MA-processed alloy (grain size 100300 nm) is obtained
by dispersion strengthening through the Al 2O 3 (30100 nm), Al4C 3 (30-
100 nm) and Al3 Ti (20250 nm), by the addition of alloying elements
and due to work hardening [12]. It is reported that Al 3Ti forms above
260C during degassing. These precipitates are stable thermally up to
1000 hrs at 300C and 500 hrs at 400C [19]. The tensile strength of
Al10%Ti (E4) alloy has been reported to be almost the same as the
conventional precipitation-hardened aluminium alloys (2014-T6, 2219-
T18) up to 100C, but it is three times higher above 200C (Fig.12.14)
due to the presence of the dispersoids.
The stress rupture strength of the alloy at 300C is three times as
much as the conventional precipitation hardened alloy. The activation
energy for creep is 228.34 kJ/mol, indicating the strong interaction be-
tween the precipitates and the dislocations. The fatigue strength of MA
Al10% Ti at room temperature is somewhat lower than that of 7075
which also has high tensile strength, but higher than that of 2024. The
fatigue strength is found to decrease with an increase in temperature,
but much slower than the decrease of tensile strength. In elevated tem-
perature use (472C), MA alloys are most promising as is a combination
117
Mechanical Alloying

Fig.12.14 Strength at elevated temperatures of


mechanically alloyed Al10% Ti (Ref.11).

of RS with MA of alloys such as AlFeCe (as discussed in Section


11.2). At the same time, ductility levels are significantly higher than
AlSiC p composites (see Section 12.12). The good ductility is attrib-
uted to good binding between in situ formed Al3 Ti and the matrix in
situ contrast to the artificial incorporation in metal-matrix composites
(MMCs).
Another high temperature alloy Al8% Fe4% Ce is prepared by MA
using stearic acid as the PCA, degassing at 425C and finally hot ex-
trusion at 425C using an extrusion ratio of 6:1 [20]. The MA proc-
ess introduces a fine-scale uniform dispersion of oxides and carbides.
The MA material microstructure was stable even after exposure for 1
hr at 610C. A comparison of the creep response of the material with
the RS one, under the same conditions of stress and temperature, shows
(Fig.12.15) that MA material is superior. However, ductility values are
low due to cleavage type of fracture.
A summary of the composition and mechanical properties of MA alu-
minium alloys developed is given in Tables 12.6 and 12.7, respectively.
For other MA aluminium materials, refer to Sections 12.7, 12.10, 12.11
and 12.12.

Table 12.6 Composition of MA aluminium-base materials

Alloy Al Mg Cu Li C O

MA 9021* Bal. 4 1.1 0.8

MA 9052* Bal. 1.5 4 1.1 0.8

MA 905XL* Bal. 4 1.5 1.2 0.4

DISPAL* * Bal. 2.0 2.5

* Product of Inco Alloys International Inc.


** Product of
**Product of Sintermetall
Sintermetall Werke,
work, Krebsge
Krebsge (Germany)
(Germany)

118
Applications

Table 12.7 Properties of MA aluminium alloys and other reference materials

Fracture Elastic Corrosion Fatigue


YS UTS Density
Alloy El.(%) Toughness Modulus Rate Strength
(MPa) (MPa) (gm/cm 3)
(MPa/m) (GPa)` (mm/year) (MPa)

IN 9052 542 577 8 44 76 2.68 0.009 >380


IN 9021 613 634 12
RS 2024 326 542 24
IN2024 279 464 22 32
IN 905 XL 462 525 13 30 80 2.58 0.003 345
IN 7075T73 503 572 10 28.6 69 2.80

RS7075 (+1%Ni+1%Fe)
634 717 9
595 645 3.5
MA 7075

193 345 18 16 2.55


IM(Al5%Mg1.5%Li)
Dispal2 340 370 10 80 2.70 110

Dispal1 (Al-12%Si) 340 380 8 85 2.66 150

Dispal
(Al12%SiCuMgNi) 340 370 3 81 2.70 120

Fig.12.15 Comparison of creep rate for non-MA (RS) and MA Al8% Fe4% Ce alloy
(Ref.20).

12.4 COPPER-BASE MATERIALS


Heat-resistant copper alloys having high electrical and/or thermal con-
ductivities are required in a variety of applications, including resistance
welding electrode contacts, electrical switches, microwave and X-ray
tube components, incandescent lamp lead wires and neutron irradiation
targets [21].
The wrought CuCr and CuZr alloys commonly used for elevated
temperature applications are strengthened through thermomechanical
treatments, which cause precipitation and stabilization of a strain-hard-
ened structure. In such alloys, phenomena such as recovery and
recrystallization start taking place at a temperature of about 450C. Using

119
Mechanical Alloying

the MA technique, it has been possible to extend the amount of chro-


mium in the copper matrix. The use of PCAs is avoided in order to
have high electrical and thermal conductivity in the final material by
minimising impurities. Cu5% Cr powder could be prepared by MA of
elemental powders. The MA powder was consolidated by hot extru-
sion. A strength level of 75 MPa could be achieved, which is attrib-
uted to fine grain structure and dispersion strengthening. Ductility of
8% and conductivity of 60% of the copper were found after the heat
treatment [22,23].
Alumina dispersion-strengthened copper has been produced by MA
of copper powder and an amorphous Al(OH) 3 powder [24]. The Al2O 3
is created in situ in the copper matrix at 254C during heating of the
green compacts to the sintering temperatures. Consolidation of the MA
powder was carried out by cold compaction followed by sintering at
940C in a nitrogen atmosphere. The highest hardness (110 VHN) was
obtained for the optimized value of 0.65% of Al 2O3 in the matrix show-
ing a significant improvement in comparison with the copper matrix,
hardness (48 VHN). It was also interesting to note that use of a crys-
talline powder as a dispersoid results in hardness lower than in the case
of amorphous Al2O 3. The hardness variation of the MA matrix with ex-
posure to high temperature at 940C for different periods of time is
shown in Fig.12.16. With an initial minor drop in hardness due to the
recovery of the cold worked state, the material is found capable of re-
taining its hardness at high temperature exposure. Thus, ODM is ca-
pable of retaining hardness/strength up to 0.9 of the absolute melting
temperature of the copper. This approach can be used to develop dis-
persion-strengthened, high-temperature copper alloys.
An elemental alloy (Cu8% Ti4% B) has been mechanically alloyed
[25]. The solubility of Ti in Cu is reported to be 2 at.%. TiB2 is formed

Fig.12.16 Effect of annealing time on softening behaviour of MA Cu0.98% Al2 O 3,


annealing temperature 940C (Ref.24, Powd. Met. Int., 22 (5), 15 (1990) (Verlag Schmid)).

120
Applications

in the ternary mixture. It is reported that refinement is accelerated by


TiB 2 or B. The MA powder was consolidated by sintering at tempera-
tures between 400 to 900C. The hardness was found to increase with
sintering time. High temperature sintering enhanced the densification
and formation of TiB 2 but reduced the hardness. The MA process has
also resulted in copper-rich solid solutions in the prealloyed CuPb [26],
CuC [27] and CuAlC [27] systems. For other copper-base mate-
rials, refer to Section 12.11.

12.5 TITANIUM SYSTEM


The MA titanium base alloys are still in the research state, but some
interesting advances have been made in the following:
TiMg immiscible;
dispersion strengthened TixAl and Ti5 Si3 .
Though TiMg forms an immiscible system and solubility obtained
of Mg in titanium by conventional methods is negligible, it is an at-
tractive alloying element to titanium due to its low density. Using MA,
solid solution of Mg up to 3.1% in titanium has been achieved [28].
Magnesium also appears to be a very strong stabiliser, which depresses
the transus in binary TiMg as well as in ternary TiMgGd sys-
tems.
Titanium aluminides based on -TiAl (L1 structure) are potentially
very attractive low-density materials for use at elevated temperatures
in aerospace applications. However, these alloys, like most intermetallics
suffer from low ductility and fracture toughness at ambient tempera-
tures due to the presence of a superlattice, and low corrosion resist-
ance even at the moderately high temperature. However, the inherent
brittleness of these alloys limits their fabricality. Alloying additions such
as Mn for ductility and Cr and Nb for oxidation resistance improve these
properties appreciably. The nanocrystalline equiaxed -grains and the
fine lamella / 2 microstructure also results in the same effects. The
MG process has been applied to two such alloys [29,30], Ti25% Al
10% Nb-3% V2% Mo(at%) and Ti48% Al2% Mn2% Nb(at%). The
microstructure of the consolidated gas atomised powder shows a mainly
uniform two-phase structure consisting of equiaxed 2 with a grain size
of about 4 m, with intergranular B2 and some occasional patches of
lenticular 2 grains. However, MG of powder prior to consolidation
results in a structure consisting mainly of very fine grains of about 1
m, with some coarser grains of about 3 m retained (Fig.12.17).
Dispersoids in the consolidated alloy are coarse, about 0.1 m in size.
An appreciable improvement in the hardness of the MA material (720
KHN) is noticed as compared to atomised powder (340 KHN), which
is attributed to the refinement of the microstructure and the transfor-
121
Mechanical Alloying

Fig.12.17 Optical micrographs of compacts of MA Ti25% Al10% Nb3% V2%


Mo: a) fine grained structure; b) coarser grains, containing a larger portion of lenticular
grains (Ref.29, Metal Powder Report, 44 (3), 206 (1989) (Elsevier Science)).

mation of 2 phase to B2 phase. The microstructure of atomised powder


has completely B2 structure, which turns to a mixture of both 2 and
B2 phases on consolidation. The proportion of B2, a more ductile phase,
has been found to be more in the case of MA powder as compared
to atomised powder.
The Ti5Si 3 intermetallic compound is also a promising high tempera-
ture material for aerospace application. However, the inherent brittleness
of this alloy limits its formability. The MA process has been found to
be suitable to produce this alloy in an amorphous and/or nanocrystalline
form to improve its mechanical properties [31]. The MA of elemen-
tal powders was carried out in a Spex 8000 mixer mill in argon atmos-
phere using hexane as a PCA medium, until a completely amorphised
phase was produced (25 hrs). The DTA results show that the powder
has a crystallisation temperature of 640C. The MA powder is then con-
solidated using a modified shock consolidated method. The modification
includes a cylindrical gap provided around the powder sample region
of the capsules to prevent radial focusing of shockwaves to minimize
generation of very high and non-uniform pressures within the axial re-
gions of the sample. Thus, the average calculated peak pressure in the
powder mixtures packed to 64% theoretical density, for an impact
velocity of 0.9 km/sec, is in the range of 20 to 40 GPa, with the one-
dimensional loading pressure being 5 GPa. During compaction, a sig-
nificant amount of crystallization, forming 30 to 40 nm crystals of
122
Applications

orthorhombic TiSi 2 and Ti 5Si 3 intermetallic compounds, has been re-


ported. This partial crystallisation of the amorphous powder during shock
compact occurs due to mean-bulk shock temperatures exceeding 1000C
(recrystallization temperature = 640C). The compacts are subsequently
annealed above the crystallization temperature at 800C for 1 hr, which
results in limited growth to 50 nm crystallite size. The microhardness
of the shock-compact was 1100 KHN, which increases to 1250 KHN
upon subsequent annealing with the formation of a more homogene-
ous nanocrystalline microstructure.
Nanocrystalline TiN powder has also been prepared by reaction
milling in a nitrogen atmosphere. The formation of TiN was found to
depend on TiN solid solution. A strong solid solution hardening has
been detected [32] both in the as-milled and the annealed (at 350 to
540C) nanocrystalline TiN alloys. The approach may probably be used
in improving the titanium aluminides.
Two advanced alloys Ti1% Al8% V5% Fe1% Er and Ti24%
V10% Cr5% Er, both with extensive rare earth additions, are being
developed for subsequent oxide dispersions by a combination of RS and
MA, as discussed in Section 11.2.

12.6 MAGNESIUM-BASE MATERIALS


The MA process has been used to alloy magnesium with Fe/Cu/Cr/C/
Ti for developing supercorroding alloys, which are used for making
links with precisely controllable corrosion rates for releasing deep sea
equipment at specific depths [17]. Such supercorroding alloys have also
been of use in other submarine applications, such as a heat source in
diver suits and as a hydrogen gas generator. The advantage of MA in
producing a supercorroding alloy is the extent of control afforded by
the closeness of the anode/cathode metals achieved. By adjusting the
alloy composition, the reaction rates can be precisely controlled in the
device (Fig.12.18).
MgM is a promising system for hydrogen storage. One of the ways
to kinetically improve magnesium-based hydrogen storage material is
by the addition of metals and alloys whose hydrides have a high dis-
sociation tension. By alloying with catalyst elements/intermetallic com-
pounds, the hydrogenation capability of magnesium is substantially en-
hanced [17]. The additions include elements such as:
Ni, Ce, La which can be alloyed with Mg conventionally. These
hydride-forming elements can function as 'hydrogen pumps' owing to
stoichiometric variation.
Elements such as Fe, Co, Ti, Nb that do not normally alloy but
form intermetallics with magnesium which can absorb and deabsorb hy-
drogen.
123
Mechanical Alloying

Fig.12.18 Corrosion rates in 'supercorroding'


MA magnesium alloys (Ref.17, Metal
Powder Report, 44(3), 195 (1989) (Elsevier
Science)).

Intermetallics such as LaNi5, Mg2Cu, CeMg12 and even oxides such


as TiO 2.
These are generally prepared by argon melting. The MA technique
has been used to produce the first and second type materials with im-
proved performance (Fig.12.19). The MA is carried out in a planetary
ball mill in an argon or hydrogen atmosphere. Severe cold work re-
sulting from MA aids diffusion by providing many sites with low ac-
tivation energy of diffusion. In addition, the close mixture of the powder
constituents decreases the diffusion distances to the micrometer range.
Precipitation may occur or metastable phase may form throughout the
powder particles, or the dispersoids may provide nucleation sites for
the formation of hydrides. For example, the MA of an elemental mag-
nesium and nickel alloy results in dispersion of the MgO film present
on the surface of magnesium particles, and disordering of the surface
of magnesium and nickel particles in the matrix. The XPS results [33]
show that the nickel atoms present in the subsurface of MA powder

Fig.12.19 Kinetics of five-cycle hydriding and dehydriding: 1) MA Mg; 2) MA (Mg-


5% Nb); 3) MA (Mg5% Fe)/Mg5% Co)/(Mg5% Ni); 4) MA (Mg5% Ti); 5) MG
(Mg 2 Ni); 6) Mg (P=0.7 MPa, T=85C) (20 m); 7) MA (Mg50% C); 8) Mg25% Ni
(conventional) (Ref.33).

124
Applications

have a binding energy of Ni 2p 3/2 less by 0.5 eV, which is the rate-
determining step for dissociative adsorption of hydrogen on nickel atoms.
The formation of Mg 2Ni is found to take place in the hydriding reaction
of the MA alloy, and practically all the nickel converts into the
intermetallic compound within 23 cycles. Mg 2 Ni is shown to form
mainly by

2MgH 2 + Ni Mg 2Ni + 2H 2

Thus, the presence of these alloying elements in the MA powder


surface, hydrogen adsorption and magnesium hydride nucleation are
essentially accelerated, and the reaction begins at maximum rate. How-
ever, the formation of a ternary hydride like MgFeHx (x = 6) decreases
the hydrogenation and the formation of a stable ternary hydride like
Mg 2CoH x (4 < x < 5) is not favourable for dehydriding kinetics.

12.7 SUPERSATURATED SOLUTIONS


The use of the SSE to produce supersaturated solutions without pro-
ducing a second phase can be achieved by various techniques like: solid
state quenching, rapid solidification, condensation from the vapour state
and irradiation/ion implementation. Supersaturated solid solutions are
stronger and harder than equilibrium solid solutions, and on decompo-
sition can produce a higher volume fraction of fine second phase par-
ticles. The actual values for the SSE are calculated using Vegards law,
which states that the lattice parameter and the amount of solute added
to the solvent lattice have a direct, linear relationship. The technique
of MA has been applied for the SSE in a variety of systems as dis-
cussed below.
A non-equilibrium supersaturated fcc solid solution is formed in an
AgCu system in which the entire composition range has positive heat
of mixing and is immiscible to each other by MA [34]. The
supersaturated solid solution formation in this system means the eleva-
tion of free energy caused not only by the stored energy as defects
but also by the effect of mixing of immiscible materials in dimensions
as fine as that of a nanometer order. Decomposition of the
supersaturated solid solution in MA alloy takes place at above 157C
which can be identified by the accompanied sharp decrease in electrical
resistivity (Fig.12.20). At the initial stage of decomposition, the grain
size of the decomposed phases remains at nanometer size. Upon heating
above about 327C grain growth takes place.
The SSE has also been achieved of Mg, Cu and Si in aluminium,
as well as of Al in magnesium as discussed in Section 11.1. A
supersaturated nanocrystalline Al8% Ti0.3% Fe (at.%) alloy powder
125
Mechanical Alloying

Fig.12.20 Changes of the electrical resistivity with


temperature in AgCu alloys (Ref.34).

has also been prepared by MA [35].


The equilibrium solid solubilities of VCu and VFe alloys are low
but V30 at% Cu and V50 at.% Fe alloys are reported to form solid
solution after MA [36].

12.8 MAGNETIC MATERIALS


NdFeB magnets
The advent of high-performance permanent magnets based on NdFe-
B in 1983 has generated a huge scientific and technological interest.
The main reason for this trend is the cheapness and abundance of the
involved elements. The origin of the best magnetic properties in these
magnets is the tetragonal structure of Nd 2 Fe 14B phase. The Curie tem-
perature (T c) of these magnets is 314C which can be improved by a
small addition of cobalt. An anisotropic field H A of about 6000 kA/m
enables high coercivities (800 to 200 kA/m) and a saturation magneti-
zation of 5120 mA/m.
An MA technique using a planetary mill has been recently used to
produce isotropic and anisotropic NdFeB magnets [37]. The MA pro-
duces a closely mixed powder of Nd, Fe and B, which has low coercivity
due to the presence of -Fe magnetic phase which is transformed to
hard magnetic Nd 2Fe14B phase by a solid state reaction by an annealing
treatment typically at 700C for 30 min. These isotropic microcrystalline
powders are used for making bonded type isotropic and compacted
anisotropic magnets (Fig.12.21). Compaction is carried out by hot press-
ing at 700C for 5 min under a pressure of 1 kbar. Density levels of
98% of theoretical values could be obtained using the process of die
upsetting where a hot pressed sample is uniaxially compressed to half
its height in a die whose diameter is 2 times the sample. The
crystallographic texture present in the consolidated sample is attributed
to stress-induced grain growth and grain rotation during the compression.

126
Applications

Fig.12.21 Preparation of NdFeB


magnets from mechanically alloyed
powder (Ref.37).

The ratio of remanences (measured parallel and perpendicular to the


press direction) critically depends on the deformation rate and the
number of deformation steps. A value of 2.6 could be obtained for this
ratio. Thus, flake geometry, as obtained in rapidly quenched material,
is not essential for the formation of anisotropic samples, since the me-
chanically alloyed powder has a highly irregular shape. Figure 12.22
shows that the coercivity of the hot pressed samples is slightly lower,
but shows the same composition dependence. With an increase in neo-
dymium content, a higher coercivity of up to 1400 kA/m at x = 9 is
obtained with an increase of the second phase after x = 6.

SmFe magnets
Magnetic alloy systems like Sm (Fe,X)12 and (Sm,Zr)Fe3 have also been
prepared successfully using an MA technique [38,39]. The MA of el-
emental samarium and iron powders leads to a two-phase mixture of
amorphous SmFe and -Fe. Heat treatment at 650800C for sev-
eral hours produces Sm 2 Fe 17, with a Th 2Zn17 crystal structure, which
has a Curie temperature of only 116C and an in-plane anisotropy. In
this state, the material is still soft magnetic because of the in-plane
anisotropy of Sm2Fe17. During the subsequent nitriding treatment at 400-
550C in a N 2 atmosphere, nitrogen atoms are introduced into its
Th 2Zn 17 crystal structure, resulting in an extension of the lattice pa-
rameter as calculated by the X-ray diffraction patterns (Fig.12.23b and

Fig.12.22 Demagnitization curves of resin-bonded


(MM1), compacted isotropic (MM2) and
anisostropic (MM3) Nd16Fe76B8 magnets (Ref.37).

127
Mechanical Alloying

12.23c), to form Sm 2Fe 17N x phase [40]. The amount of residual -Fe
depends on the Sm content, and its optimum amount has been found
to be 12.5 at.%. The grain size after milling was typically 1030 nm
and increased to 50 nm after annealing. The magnetic domain size was
found to be 300 nm. Oxygen was found to enhance precipitation of
-Fe and a maximum of 9000 ppm is suggested [41]. Low oxygen is
found to suppress precipitation of -Fe which deteriorates the magnetic
properties. This new phase, Sm 2Fe 17N x, has coercivities up to 2400 kA/
m and uniaxial anisotropy. The phase has also improved the Curie
temperature T c (470C compared to 320C), the saturation magnetization
M s, (1.6 T compared to 1.54 T) and the anisotropy field H A (14 T com-
pared to 8 T) as compared to NdFeB magnets. Therefore, magnets
of SmFeN should have the potential to compare favourably with well-
established NdFeB magnets, if the high anisotropy field can be ex-
plained to create a high coercivity and a processing route can be es-
tablished. An isotropic resin-bonded (Sm 12.5Fe 87.5) 1xN x magnet shows
remanence M r = 0.71, which is the same for mechanically alloyed
Nd 16 Fe 76 B 8 . The temperature dependence of the coercivity and
remanence of this phase is much smaller than that for Nd-Fe-B. At
150C, the coercivity is still 1450 kA/m. The major drawback of this
new hard magnetic material is its limited stability at elevated temperature.
It has been found that 2:17 nitride is stable up to 650C, thereafter
it decomposes into samarium nitride and -Fe. This instability of the
2:17 nitride prevents the application of sintering technique used for Nd-
FeB or SmCo for the production of anisotropic magnets.

SmCo magnets
The MA process has also been used to produce conventional SmCo 5
and Sm 2 Co 17 magnetic materials. To produce these materials, MA of

Fig.12.23 X-ray diffraction diagrams of


Sm 12.5 Fe 87.5 : a) after MA; b) after the
formation of Sm2 Fe17 by annealing in vacuum;
c) after the nitriding reaction (Ref.39).

128
Applications

Sm 16.7 Co 83.3 and Sm 10.5Co 89.5 powders, corresponding to 1:5 and 2:17
stoichiometry, respectively, is performed in a planetary ball mill under
an argon atmosphere [42]. The MA in the case of Sm 16.7Co 83.3 pro-
duces an amorphous phase possibly with traces of free Co, and for
Sm 10.5 Co 89.5 a two-phase microstructure of an amorphous matrix and
Co solid solution. The two-phase region in the case of Sm 10.5Co 89.5
results from a metastable equilibrium between the amorphous phase and
the Co solid solution. The crystallization of amorphous state to the 1:5
structure, SmCo 5 , occurs at about 550C for both alloys, which has
a high coercivity of about 2200 kA/m and a remanence of 0.52T. The
degradation of the remanence M r results at a high annealing temperature
due to Sm losses during annealing accompanied by the precipitation of
Co. The coercivity of Sm 2Co 17 is found to be low due to the lower
anisotropy field of 2:17 phase. To increase the coercivity, the MA pow-
der is annealed at 600C during which Sm 2Co 17 transforms to SmCo 5.
The order of grain size in the annealed material is found to be less
than 200 nm and the single domain particle size is of about 1 m. As
the coercivity strongly depends on grain size when it is less than the
single domain particle size, MA provides hard magnetic powder at low
temperatures without any sophisticated annealing programme. Conven-
tionally, the preparation of SmCo5 magnets is based on high-temperature
heat treatment above 800C followed by a rapid cooling to room tem-
perature. However, a suitable technique for consolidation of these MA
powders is to be evolved. For MA soft magnetic material refer to
Section 12.13.

12.9 MA POWDERS FOR SPRAY-COATINGS


The MA process is suitable for the production of MCrAl (Y) (M=Fe,
Ni or Co) types of powder like MA-754 superalloy which can be
plasma-sprayed in a low pressure chamber to provide oxidation and
corrosion-resistant coatings on gas turbine blades and vanes [43]. By
employing a fine mixture of mechanically alloyed Cu/Al 2O3 composite
powder for low pressure plasma spraying, a fine structure uniform coat-
ing has also been applied [44]. However, such coatings have been found
consisting of two layers: matrix-rich layer and dispersoid layer. This
separation presumably takes place due to the lack of wettability and
to some extent by the acceleration caused in the plasma jet.
A uniform single phase intermetallic coating of Ti 3Al has been ob-
tained by employing MA TiAl powder for RF plasma spraying [44].
The interdiffusion of Ti/Al takes place quite quickly during the proc-
ess of spraying, which results in the formation of Ti3 Al intermetallic
phase. Moreover, an intermetallic matrix coating has also been obtained
using MA TiAl10vol.% Si 3N4 powder. A fine distribution of Si3N 4 is
129
Mechanical Alloying

obtained along with the other phases like Ti5Si 3 and TiN exhibiting de-
composition of the ceramic phase [44]. Using DC plasma spraying, a
Cu/Ti/C MA powder coating of uniformly dispersed TiC in a copper
matrix has been obtained [45]. These coatings have slightly higher hard-
ness due to homogeneously dispersed structures.
Thus, the proper selection of the ceramic phase both from the re-
activity and wettability point of view is essential in order to have the
uniform desired coating [44].

12.10 SUPERPLASTICITY
The MA 6000 oxide dispersion-strengthened, nickel-base superalloy has
been found to possess superplasticity at 1000C [46]. In the hot-
extruded and hot-rolled condition, i.e. prior to zone annealing, the ma-
terial has a grain size of 0.26 m and a dislocation density of 310 9
cm2. It can be believed that at least some of the is evenly distributed
as very fine (30 nm) particles (Fig.12.2, a similar microstructure, may
be referred). The alloy exhibits superplasticity, having a maximum elon-
gation of over 300% and a maximum strain rate sensitivity of 0.47.
Transmission electron microscopy shows no evidence of ordinary
recrystallization and grain growth is slight. However, strain rates of less
than about 1.0 sec 1 alter the initial microstructure and prevent grain
coarsening on subsequent annealing at higher temperatures. Deformation
of the fine grained MA 6000 can be described as a combination of
power law creep and diffusional (Coble) creep [47] (Fig.12.24). With
a threshold stress caused by the presence of particles, it exhibits only
the diffusional creep process. Threshold stresses for dislocation creep
are not observed.
Superplasticity has also been observed in MA IN 90211 aluminium
alloy (Al2%Mg4.4%Cu1.1%C0.8%O) at 447C and a strain rate
of 1.0 sec 1. The alloy has a grain size of 0.5 m. The deformation
at a high strain rate is found to be governed by co-operative grain
boundary sliding, i.e. the movement of grain groups as a unit [48]. Such
a process proceeds as shear of grain groups along two intersecting sys-
tems, having grain boundary surfaces oriented at 45 to 60 with re-
spect to the tensile axis. Superplasticity has also been observed in IN
9021-SiC p composite material (see Section 12.12).

12.11 TRIBOLOGICAL MATERIALS


Aluminium alloys
It is common practice to use solid lubricants such as graphite, tung-
sten disulphide or hexagonal boron nitride in sleeve-bearing materials
to reduce friction in the boundary lubrication region. The MA process
has proved to be a powerful tool in producing an ultrafine microstructure
130
Applications

Fig.12.24 Comparison of the flow data for MA


6000 at 1000C with the predicted dislocation
and Coble creep behaviour of Ni (Ref.47, Met.
Trans., 16A, 777 (1985) (The Minerals, Metals
and Materials Society)).

containing dispersed solid lubricants such as graphite and tungsten


disulphide [49,50], which has improved tribological properties.
Al4.5%Cugraphite has been produced by MA of the elemental mix-
ture in an attritor mill in an inert atmosphere for a period of 3 hrs fol-
lowed by consolidation by hot pressing. The MA results in a homo-
geneous distribution of graphite particles in the matrix, which leads to
a dramatic decrease in the friction coefficient in the boundary region
(Fig.12.25) [49]. It is also interesting to note the slope of the curves
in the boundary region, which is a measure of the rate of decrease in
the friction coefficient f with respect to the bearing parameter ZN/
P, where Z is the viscosity of the lubricating oil, N is the rpm of the
shaft and P is the pressure on the bearing, which is faster in the case
of MA materials. It is observed that an addition of approximately 3%
graphite by MA is sufficient to improve the bearing performance of the
alloy. Even in the presence of 3% graphite, the k-factor of the ma-
trix remains unchanged, which is attributed to the refining of the grains
and the strengthening effect of oxide and carbide dispersoids. The MA
process also results in more than a 100% improvement in the wear re-
sistance of the material.
The MA technique has also been used successfully to improve the
tribological performance of aluminiumlead alloys. Aluminium base-bear-
ing alloys invariably contain tin as an alloying element which is costly.
Lead is an attractive alternative to tin because it is much cheaper and
more freely available. However, the AlPb system is immiscible with

131
Mechanical Alloying

Fig.12.25 Coefficient of friction, f, vs


bearing parameter, ZN/P.

a wide miscibility gap. In addition, the two metals are virtually insoluble
in each other at room temperature. The situation is aggravated further
by a large solidification range of their alloys and a wide density dif-
ference, which greatly increases the kinetics of lead segregation dur-
ing melting and solidification. As a result of all these difficulties, it is
virtually impossible to produce AlPb castings having a uniform dis-
tribution of lead. The MA of elemental aluminium and lead powders
was carried out in an attritor mill for 2 hrs using 1% stearic acid as
a PCA. After degassing at 200C in a vacuum of 10 2 torr, the MA
powder was consolidated by hot pressing at 500C under a pressure
of 60 MPa [50]. During MA, lead particles undergo plastic deforma-
tion without becoming work-hardened as its recrystallization temperature
(3C) is even below the ambient temperature. Thus, the lead parti-
cles are lamellar in shape and remain soft, whilst the aluminium par-
ticles which undergo plastic deformation, are work-hardened and be-
come fragmented. Consequently, these fragments stick to the surface
of the lead lamellae. The consolidated MA alloy has an improved micro-
structure, which is also reflected in an improvement of approximately
250% in the wear resistance of the alloy.
High energy milling of Al, Fe and ferrochromium powders has been
carried out to develop Al8%Fe2% Cr alloy. The reason for using
chromium is that, being a carbide-forming element, it is likely to combine
with the carbon of the PCA, therefore increasing the wear resistance
of the alloy. After degassing, the powder was consolidated by hot press-
ing [51]. The MA material was found to possess a hardness of 162
BHN, which is comparable to several rapidly solidified AlFex-alloys.
The MA alloy has high wear resistance compared to PM and wrought
alloy, and EDAX analysis has shown the presence of Al4 (FeCr) in-
132
Applications

termetallic compound. The alloy has the potential to be used in a wide


variety of components for aerospace applications because of its high
specific strength.

Brasses
The MA technique has also been used to improve the tribological prop-
erties of brasses. The MA of Cu40%Zn blended alloy and solid lu-
bricant WS 2 or graphite (20 and 40 vol.%) has been carried out in a
planetary mill in an argon atmosphere for a period of approximately
525 hrs. It was observed that MA continuously refines the microstruc-
ture with milling hours (Fig.12.26) but in the case of brass WS2, milling
beyond 13 hrs results in the chemical reaction of zinc with WS 2 to form
ZnS due to activation by the energy input of MA. The graphite par-
ticles retain their hexagonal lattice, as confirmed by SAD patterns of
the composite powder, even after milling for a period of 25 hrs. The
MA powder containing WS 2 was consolidated by hot pressing at 620C
at a pressure of 28 MPa, and the powder containing graphite was con-
solidated by cold compaction under a pressure of 700 MPa followed
by sintering at 810C for one hour [52]. The microstructures of these
consolidated MA composites have uniform fine scale distribution of WS2
graphite, as shown in Fig.12.27.
Friction coefficient f (in unlubricated conditions) and wear rate meas-
urements for these MA materials have shown that it could be possi-
ble to decrease the friction coefficient as low as that of the pure solid
lubricants (f = 0.15), indicating that the sliding surface is completely
covered by a thin layer of the solid lubricant.

Fig.12.26 TEM micrograph of Cu40% Zn brass powder particle, milled for 13 hrs.
Note the lattice planes of WS 2 (Ref.52, Metal Powder Report, 48 (11), 20 (1993)
(Elsevier Science)).
Fig.12.27 (right) Micrograph of sintered MA brassgraphite composite (Ref.52).

133
Mechanical Alloying

Steels
Wear resistant steels for tribosystems have also been developed using
MA techniques as discussed in Section 12.2.

12.12 COMPOSITES
Aluminium-base MMCs, having particulate reinforcements such as SiC,
Al2O 3, spinel MgAl2O4 and BN, are being developed. However, a major
problem with these composites is the excessive contamination by iron
from the alloying mill abrasion. Unlike other processing routes for similar
composites, the particulate reinforcement does not enhance significantly
the strength of the matrix that has already been strengthened by the
oxide and carbide dispersions. The matrix (IN 9021) is thermally un-
stable and loses its strength at temperatures above 150C due to dis-
location annihilation, and the coarse SiC particles do not appear to result
in any significant strengthening at high temperatures [53]. Whilst the
particlematrix interface strength is substantial (well over 1000 MPa),
coarse particles of SiC break during fast fracture and the fracture tough-
ness of the composite is low [54]. However, at high temperatures in
the range 425450C and at high strain rates (>0.7 sec 1), the com-
posite material with the IN 9021 matrix shows extended ductility, up
to 300% total elongation [53,55].
Elemental 7010 alloy has been mechanically alloyed with 10% SiC
particles [56]. After degassing at 200C for 2 hrs in vacuum (10 2 torr),
the composite is consolidated by hot pressing at 565C under a pressure
of 110 MPa and hot extrusion at 585C with an 11:1 extrusion ratio.
The consolidated composite was solution treated at 500C and quenched
in ice-water followed by ageing at 175C for 16 hrs. The composite
had a clean SiCPmatrix interface (Fig.12.28) showing minimum interface
reaction. It should be noted that SiC P appears to inhibit the alloying

Fig.12.28 TEM micrograph showing SiCp interface in 7010 matrix (courtesy Prof. P.
Ramakrishnan).

134
Applications

process. It is also revealed that Zn dissolves with greater difficulty than


Cu and Mg [57]. The matrix in the consolidated composite had dynami-
cally recrystallized nanograins (2050 nm) with a fine dispersion of
oxides and carbides (515 nm) prior to recrystallization, resulting in high-
strength properties in the composite material. The composite had a
tensile strength of 100 MPa, almost double that of the matrix at 350C.
A composite with a 2014 aluminium alloy matrix having 8% SiC P
has also been prepared successfully using the MA technique. The com-
posite powder was consolidated by hot pressing at 535C under a pres-
sure of 125 MPa. It was observed that the MA process increases wear
resistance of the composite by a factor of two in comparison with the
conventionally blended composite [58]. Aerospace Metal Composites
(AMC), UK has developed a technique to produce a 2124-T4 aluminium
alloy matrix with 17.8% SiC P using MA. Blending as well as MA is
carried out in an argon atmosphere to control oxidation, maintain product
purity and reduce the explosion risk. Consolidation of the MA pow-
der is carried out by HIP to produce a billet which is extruded into
tubes to make MMC frames of road racing bicycles (Fig.12.29) [59]
Al 2O 3 and spinel MgAl 2 O 4 have also been tried as the reinforcement
with Al4%Mg blended alloy matrix using the MA technique, result-
ing in composites with improved creep rates [60]. The spinel MgAl2O 4
has a stable structure up to 2000C and is shown to minimize metal/
ceramic reaction.
The MA aluminium-based, copper-based and iron-based tribo-com-
posites were discussed previously in Section 12.11.
In situ TiNiTiC and NiTiC composites were prepared by MA of
elemental powders. The powder was found to contain granulates of 3-
10 nm after 35 hrs of MA due to a melt and quench mechanism taking

Fig.12.29 Raleigh's Cronos-MMC frame made of 2124-T4 aluminium alloy - 17.8%


SiCp composite (XMT-100).

135
Mechanical Alloying

place during the process. Spherical TiC grains with martensite and B2
structures also form during processing, which causes self-propagating
synthesis initiated by the heat of formation of TiC [61].
Surprisingly, the MA process has been applied to polymer-based
composites. The MA of powders of polytetrafluoroethylene and cop-
per or nickel has been carried out. Amorphization through the SSAR
has been achieved in the composite [62].
During MA, in general the softer material tends to form the matrix
and the harder material disperses within it. This tendency, together with
the tendency of mechanically alloyed materials to become harder with
increasing processing, can be utilized to produce complex, multi-level
metal composites. For example, if tungsten, a relatively ductile metal,
is mechanically alloyed with a very fine ZrO2 nickel powder, it results
in the dispersion of ZrO 2 in tungsten. If nickel powder is added and
processed with tungsten ZrO 2 composite powder, a two-level composite
is formed. The hard and brittle tungstenZrO 2 is broken up and dis-
persed in the matrix of more ductile nickel. Contact between the zir-
conium oxide and nickel is minimal in such a composite. Thus, a
number of levels and the relative degree of dispersion of different in-
gredients that can be obtained by this technique are almost limitless.

12.13 AMORPHOUS SOLIDS


Using the MA/MG technique, a large number of binary, ternary and
even quarternary systems have been amorphized. A typical list can be
seen in Table 5-1. Here, the few systems which have practical signifi-
cance are discussed.
At present, there is considerable interest in the formation of amor-
phous phases in soft magnetic alloy systems because of their poten-
tial properties and processing advantages. These materials have tradi-
tionally been prepared by the RS of molten alloys. FeB and FeB
Si containing 75% of the amorphous phase was prepared by MA [63].
FeSi alloy is used for the construction of transformers due to a com-
bination of low magnetostriction and high saturation magnetization.
Amorphous FeSi alloys, potentially, add to these characteristics the
possibility of controlling magnetic anisotropies by stress and/or mag-
netic field annealing, facilitated by the atomic mobility of the amor-
phous state. The MA technique has been successfully used to amorphize
the alloy [64].
Aluminiumtransition metal (TM)rare-earth (RE), quarternary sys-
tems (Al-TM1-TM2-RE) have yielded the best glass-formers to date.
Optimal glass-forming compositions found in the AlNiFeGd system
are located around Al 85Ni 6Fe3 Gd6 and Al85Ni 5Fe 2Gd 8. These amorphous
alloys have a strength of 1280 MPa, which is 2-3 times higher than
136
Applications

the tensile strength of commercially available crystalline aluminium


alloys. They are comparable to the strength of common high-strength
steels which have low densities and an elastic modulus comparable with
standard aluminium alloys, and have satisfactory corrosion resistance.
Al 80Ni 8Fe 4Gd 8 alloy powder has recently been produced in the amor-
phous state by the MA technique (Spex mill, 160 hrs), which suggests
the possibility of forming the bulk material through the powder met-
allurgical MA technique [65]. The MA amorphous powder was found
to have good stability with a crystallization temperature of 620650
KHN. K c is found to be in the range 15 MPam, identifying the pro-
duced powder as the brittle type amorphous phase. Dispersed ductile
nanocrystals are found to improve the fracture toughness at the expense
of hardness.
The TiSi system is of particular interest for semiconductor devices,
and TiSi intermetallic compounds such as Ti5 Si3 and TiSi2 are con-
sidered as high-temperature structural materials with extremely low spe-
cific weight, favouring their utilization in aerospace applications. The
system has been amorphized by a planetary ball mill in an argon at-
mosphere [66]. In the case of a Ti 67.5 Si 32.5 powder mixture, an almost
amorphous phase forms when milling takes place with intervals and con-
tinuous milling results in the Ti 5 Si 3 stable phase. In the case of
Ti 33.3Si 66.7, the metastable C49 structure is formed. Both intermetallic
phases exhibit a crystalline size typically between 10 and 15 nm.

12.14 NANOCRYSTALLINE MATERIALS


In general, nanocrystalline solids are exclusively prepared by conden-
sation from the vapour phase to a fine powder or by the solgel method
(ceramics). In these materials, 50% of the actual volume consists of
grain boundaries. Because of this large number of crystalline interfaces,
an important fraction of the material has a disordered microstructure
with no short-range order similar to a gas phase. As a result,
nanocrystalline solids exhibit physical and chemical properties differ-
ing from those found in normal crystalline or amorphous materials, e.g.
the diffusion coefficient in these materials can be as high as 10 9 times
of the volume diffusion in a single crystal [67], which leads to enhanced
solid solubility (even in the immiscible system with positive enthalpy
of mixing) and, in certain cases, an increase of the catalytic activity.
Carrying out MA for 40 hrs, the Mo solubility in the nickel lattice,
which is approximately 8 at.% at room temperature equilibrium limit,
could be increased by a factor of 3 (24 at.%) with an fcc nanocrystalline
structure. The average particles size was, as studied by high-resolu-
tion transmission electron microscopy and image processing, 4 nm. The
solubility limit was found to increase with a decrease in crystallite size.
137
Mechanical Alloying

The electrocatalytic activity, for the hydrogen evolution reaction in al-


kaline solutions, of the cold-pressed MA NiMo nanocrystalline cath-
odes is among the highest ever reported for this type of alloy [68]. This
high activity appears to be directly related to the size of the crystallites
[69]. It is reported that elemental Nb and Al powders have been me-
chanically alloyed for 5hrs in an argon atmosphere to synthesize
nanocrystalline Nb 3Al. The MA powder has an average particle size of
4 nm. The mixture becomes two-phase, Nb 2Al (19%) and Nb3Al (81%),
after heat treating for 1 hr at 600C but has a stable grain size (5 nm).
Grain growth is found to occur above 900C [70]. Cu16 at.% Fe and
Fe40 at.% Cu have also been produced in nanograin size using the
MA technique [71]. Fe50 at.% V alloy powder has also been
nanosized (9 nm) by ball milling for 180 hrs [72]. High-temperature
alloys such as AlTi, TiAl, TiSi and NbAl (T m 1850C) have also
been prepared in nanosize using MG with the aim of improving their
mechanical properties as discussed in Sections 12.3 and 12.5.
Reaction milling in the case of titanium (as discussed in Section 12.5)
and FeAl [73] has been employed for nanocrystallization.

12.15 MECHANICALLY ACTIVATED CHEMICAL REACTIONS


As discussed in Section 5, MA can be used effectively to reduce the
T crit and a concomitant increase in reaction rate. These processes were
applied to reduce a number of metal oxides and halides to pure met-
als [74]. Typically, copper metal forms when CuO is milled with cal-
cium, and brass forms when CuO and ZnO are simultaneously milled
with calcium. A break-through in the process came with the advent of
a separating and extracting technique for the nanosized powder particles
produced [75]. The nanosized particles have good chemical purity and
surface oxidation less than 1015%. So far, iron, nickel, cobalt and
copper powders, as well as some ceramic powders such as alumina and
zirconia, having a particle size as small as 5 nm, could be produced
using the MA technique. These ultrafine powders have significant po-
tential for high technology applications such as cutting tools, advanced
ceramics, high density magnetic fluids and catalysts. Thus, the MA
process allows direct production of metal powder without the need to
manufacture bulk alloy and then convert it to powder form. A number
of high-temperature processes can be combined into one single room
temperature process with the potential for significant cost savings. It
is envisaged that the process may find application in the production of
reactive elements and alloys such as rare-earth and titanium, which are
difficult to produce by conventional metallurgical processing techniques
that require recycling of a process reductant.

138
Applications

Recently, using a uni-ball mill for milling a mixture of ilmenite and


carbon, it was possible to produce Ti powder at a low temperature
reduction of rutile [76]. The mineral ilmenite is a naturally occurring
iron titanate (nominally FeTiO3 ) and is abundant in nature. Commer-
cial grades of ilmenite contain 4565.8% TiO2 and are regarded as a
huge resource for the production of rutile (TiO 2), which can be used
directly as a pigment or for the manufacture of titanium). The reduction
rate was found to increase with milling hours and milling intensity. How-
ever, the carbothermic reduction does not take place during milling (even
after 400 hrs of milling), but during subsequent low-temperature an-
nealing at 760C. Under normal conditions, this reduction usually takes
place at temperatures between 8601000C. Thus, high-energy ball
milling produces a mechanical activation effect, as shown in Fig.12.30.
The gaseous reduction rate is enhanced due to the small grain/parti-
cle size and lattice defects created in ilmenite crystals, enabling com-
plete ilmenite reduction of rutile to occur before an appreciable fur-
ther reduction to TiO 2-x (where 2 < x < 0) occurs. Thus, MA leads
to a high degree of selectivity also for the reduction of rutile.
The effect of MA on elemental mixtures of Ti, Ni and C powders
has also been investigated [77]. It is observed that an explosive reaction
takes place during milling in the range 3 hrs 30 min3 hrs 35 min, and
large agglomerates with a size of 310 mm form abruptly, suggesting
melting and subsequent quenching of the powders during MA. The
phases present are indentified as spherical TiC grains, lath twin
martensite, B2 phase and Ni depending on the initial composition. The
self-sustained, high-temperature synthesis (SHS) is believed to be trig-
gered by the release of the heat of formation of TiC and ignited by
mechanical collisions. Thus, MA can be a versatile method for inducing
SHS reactions in highly exothermic systems.

Fig.12.30 XRD patterns for the: a) 200 hrs milled sample isothermal annealed at 760C
for 30 min, and b) for an as-mixed sample heated continuously up to 1100C without
premilling, i; FeTiO 3 , r; TiO 2, F; Fe, A; -Fe (C) (austenite) (Ref.76).

139
Mechanical Alloying

12.16 OTHERS
Other potential applications include dental prostheses [78] and super-
conductors [7981]. The MA process is being used to produce precursor
powders by INCO alloys based on BaCu and BiSrCaCu alloys for
producing superconducting wires [82]. The use of ceramics for super-
conducting applications is not suitable because they are very brittle.

CONCLUSIONS
At present, MA materials are used not only in the aerospace industry,
but also in the thermal processing industry (furnace fixtures, muffle
tubes, furnace racks, furnace transport rollers, skid rails, mesh belts,
electrical resistance windings, burner nozzles, etc.), the glass industry
(molten glass stirring rods, furnace hearths, tiles, tableware, bushing
used to produce single and multistrand fibres), the energy production
industry (flame stabilisers, fuel cladding in fast neutron breeder reactors,
heat exchanger components in high-temperature gas-cooled reactors,
components for industrial gas turbines), the recreational industry (bi-
cycle frames and fork brakes) and the spray-coating industry. However,
the progress of industrial acceptance of MA materials has been slow,
mainly due to the following reasons:
high cost (though high performance/cost ratio),
reluctance to try new materials,
lack of in-service experience,
non-availability of the necessary product forms (bar, plate, sheet,
wire, tube, forging stock, etc.),
size range; the upper limit usually depends on the casting pro-
duction facilities (at present 500 kg MA 956 plates represent the largest
made), whereas the lower limit (thickness or diameter) may depend on
the working characteristics of the alloy.
The outlook for the growth of MA materials in industrial markets
appears to be a good one. Efforts to increase the availability of the
product forms and lower cost production methods will certainly improve
the market scenario of these materials. The RSMA technique will prob-
ably result in the advent of a new class of extraordinary materials.

References
1. R. Sundaresan and F.H. Froes, J.Metals., 39 (1), 22 (1987).
2. R.C. Benn, L.R. Curwick and G.A.J. Hack, Powder Metal., (4), 191 (1981).
3. K. Murakami et al, Met. Trans., 24A, 1049 (1993).
4. W. Sha and H.K.D.H. Bhadeshiam Met. and Mater. Trans., 25A, 705 (1994).
5. J.J. Fischer and J.H. Weber, Adv. Mat. Proc., (10), 43 (1990).
6. T.S. Chou and H.K.D.H. Bhadeshia, Met. Trans., 24A, (4), 773 (1993).
7. B. Kazimierzak, M. Prigon and D. Coutsouradis, MPR, 45 (10), 699 (1990).
8. E. Kohler, C. Gutsfeld and F. Thummler, Powd. Met. Int., 22 (3), 11 (1990).

140
Applications

9. C. Gutsfeld and F. Thummler, MPR, 45 (11), 769 (1990).


10. J.F. Faure and L. Ackermann, Comparative evaluation of aluminium base
materials and processing routes for elevated temperature applications, Presented
at the symposium on Dispersion Strengthened Aluminium Alloys, TMS, Phoenix,
Arizona (1988).
11. R.D. Schelleng, J. Metals., 41 (1), 32 (1989).
12. J.R. Pickens, J. Mat. Sci., 10, 1437 (1981).
13. A. Layyous, S. Nadiv and I.J. Lin, Powd. Met. Int., 19 (1), 11 (1987).
14. A. Layyous, S. Nadiv and I.J. Lin, Powd. Met. Int., 22 (3), 22 (1990).
15. S.J. Donachie and P.S. Gilman, In: Aluminium-lithium alloys II, T.H. Sanders and
E.A. Strake (eds.), TMS, Pennsylvania, PA (1984), p.507
16. R.D. Schelleng, P.S. Gilman and S.J. Donachie, In: SAMPE technical conference
series (1985), p.106.
17. R. Sundaresan and F.H. Froes, MPR., 44 (3), 195 (1989).
18. P.K. Mirchandani and R.C. Benn, In: Spage Age Metals Technology, V.2, SAMPE
(1988), p.188.
19. K.S. Lee, S.J. Kim and K.B. Shin, In: Advances in Powder Metallurgy and
Particulate Materials, V.4, A. Lawley and Swanson (compilers), MPIF and APMI,
Princeton, N.J., USA.
20. S.S. Ezz, A. Lawley and M.J. Koczak, Dispersion strengthened Al-Fe-Ce, a dual
rapid solidification-mechanical alloying approach, Presented at the symposium on
Dispersion Strengthened Aluminium Alloys, TMS, Phoenix, Arizona (1988).
21. H. Baker, ASM Metals Handbook, V.2, Metals Park, OH (1979), p.269.
22. A.N. Patel and S. Diamond, Mat. Sci. and Eng., 98, 329, (1988).
23. W. Erde, PM Technology, 14, (3), 175 (1966).
24. M.L. Mehta et al, Powd. Met. Int., 22 (5), 15 (1990).
25. T. Morooka, J. Jpn. Soc. Powd. Metall., 41 (3), 277 (1994).
26. J.L. Estrada, Mechanical alloying of prealloyed copper-lead powders, Presented
at the 1996 World Congress on Powder Metallurgy, Washington DC (1996).
27. J. Palacios, Production of copper-carbon and copper-aluminium-carbon alloys by
mechanical alloying, Presented at the 1996 World Congress on Powder metallurgy,
Washington DC (1996).
28. R. Sundaresan and F.H. Fores, In: Proc. Int. Conf. PM and Related High-
Temperature Materials, P. Ramakrishnan (ed), Oxford and IBH (1987), p.215.
29. R. Sundaresan and F.H. Froes, MPR, 44 (3), 206 (1989).
30. N.Hoo et al, J. Metals, 48 (7) 40 (1996).
31. S.C. Glade amd N.N. Thadhani, Met. and Mater. Trans., 26A, 2565 (1995).
32. T.D. Chan and C.C. Koch, Nanostruc. Mater., 5 (6), 615 (1995).
33. E.Ivanov et al, J. Less Common Metals., 132, 25 (1987).
34. K. Uenishi et al, Mat. Sci. & Eng., 134A, 1342 (1991).
35. T. Yamane, Mater. Trans. JIM, 37 (2), 130 (1996).
36. M. Mori, J. Jpn. Soc. Powd. Powd. Met., 38 (1), 71 (1991).
37. W. Heisz and L. Schultz, Appl. Phys. Lett., 53 (4), 342 (1988).
38. L. Schultz, K. Schnitzke and J. Wecker, Appl. Phys. Lett., 56, 868 (1990).
39. K. Schnitzke, et al, Appl. Phys. Lett., 57 (26), 2853 (1990).
40. M.Umemoto, J. Jpn. Soc. Powd. Powd. Met., 38 (7), 957 (1991).
41. M. Ito, J. Jpn. Inst. Met., 59 (6), 666 (1995).
42. J. Wecker, M.Katter and L. Schultz, J. Appl. Phys., 69 (8), 6058 (1991).
43. P.S. Gilman and J.S. Benjamin, Ann. Rev. Mater. Sci., 13, 279 (1983).
44. F. Fukumoto and I. Okane, In: Proc. Int. Conf. Thermal Spray Conference and
Exposition, Orlando, Florida (1992), p.595.
45. M. Umemoto and I. Okane, J. Jpn Inst. Met., 58 (1), 43 (1994).

141
Mechanical Alloying

46. R. Singer and G.H. Gessinger, In: Proc. Int. Symp. Deformation of Polycrystals,
N. Hansen et al (eds), Riso National Laboratory, Roskilde (1981), p.365.
47. J.K. Gregory, J.C. Gibeling and W.D. Nix, Met. Trans., 16A, 777 (1985).
48. M.C. Zelin, T.R. Bieler and A.K. Mukherjee, Met. Trans., 24A, 1208 (1993).
49. A. Sharma, P.R. Soni and T.V. Rajan, MPR 43 (1), 37 (1988).
50. V. Kumar et al, Unpublished work.
51. G.H. Borhani, A.N. Tiwari and P. Ramakrishnan, Trans. PMAI., 20, 7 (1993).
52. B. Gnther et al, MPR., 48 (11), 20 (1993).
53. Wadsworth, In: High-Strength Powder Metallurgy Aluminium Alloys II, G.J.
Hildeman and M.J. Koczak (eds), TMS, Warrendale, PA (1986), p.137.
54. D.L. Davidson, Met. Trans., 18A, 2115 (1987).
55. R.G. Nieh, Scripta Metall., 18, 1405 (1984).
56. A. Bhaduri et al, Trans. PMAI, 21, 1 (1994).
57. A.P. Miodownik, Met. and Mater. Trans., 27 (10), 3718 (1996).
58. N.R. Rudrappa et al, PMAI Newsletter, 11 (4), 23 (1985).
59. S. Pickering, MPR, 50 (6), 30 (1995).
60. J.W. Pyun and S.I. Kwun, J. Korean Inst. Metals,33 (6), 814 (1995).
61. H.Q. Ye, Acta Mater., 44 (5), 1781 (1996).
62. T. Ishida and S. Tamasu, J. Mater. Sci. Lett., 12 (23), 1851 (1993).
63. S. Surimach, et al, Mat. Sci. and Eng., A134, 1368 (1991).
64. A.G. Escorial et al, Mat. Sci. and Eng., A134, 1394 (1991).
65. G.M. Dougherty, G.J. Shiflet and S.J. Poon, Acta Metall. Mater., 47 (2), 2275
(1994).
66. Z.H. Yan, M. Oehring and R. Bormann, J. Appl. Phys., 72 (6), 2478 (1992).
67. Y. Minamino et al, Met. Trans. JIM, 37 (2), 130 (1996).
68. M.L. Trudean and R. Schultz, Mat. Sci. and Eng., A134, 1361 (1991).
69. J. Kuyama, J. Jpn. Appl. Phys. Lett., 30 (5a), 854 (1991).
70. Y. Tanaka, Mater. Trans. JIM, 37 (3), 265 (1996).
71. J.Y. Huang et al, Acta Metall. and Mater., 45, (1), 113 (1997).
72. H. Kuwano and T. Hamaguchi, J. Jpn. Soc. Powd. Powd. Met., 39 (3), 216 (1992).
73. J. Rawers et al, Met. Mater. Trans., 27A, 3126 (1996).
74. G.B. Schaffer and P.G. McCormick, Mat. Sci. Forum, 88-90, 779 (1990).
75. P.G. McCormick, MPR, 51 (9), 5 (1996).
76. Y. Chen et al, Met. and Mat. Trans., 28A (5), 1115 (1997).
77. J.Y. Huang et al, Acta Metall. and Mater., 44 (5), 1781 (1996).
78. E. Ivanov, Int. symp. Mechanical Alloying, Kyoto, Japan (1991).
79. C. Politis, Physica, 135B, 286 (1985).
80. M. Stubicar et al, Int. symp. mechanical alloying, Kyoto, Japan (1991).
81. V. Mizutani et al, Mat. Sci. Forum, 88-90, 415 (1992).
82. Anonymous, MPR, 43 (10), 698 (1988).

Questions
1. Give a general flow sheet for the production MA aluminium alloys.
2. What are DISPAL alloys? How are they produced?
3. Discuss development of MA high temperature aluminium alloys.
4. Discuss with the help of a flow sheet the production of Al-Li MA
alloys.
5. Name the MA alloys produced at commercial level, discuss their
properties and uses:
(i)Nickel-base superalloys;

142
Applications

(ii)Aluminium-base alloys.
6. Which MA steels are produced on an industrial scale? What are
their applications?
7. What are ODMs? What are their possible uses?
8. Discuss MA steels useful for tribological purposes.
9. How MA techniques can be helpful in developing the following:
(i)High temperature copper alloys;
(ii)Ti-Mg alloys.
10. How MG is helpful in improving ductility of titanium aluminades.
11. Discuss how Ti 5Si 5 can be improved with the help of MG.
12. How MA produces supercorroding materials.
13. Discuss development of MA hydrogen-storage materials.
14. Compare SSE achieved by MA with that by RS.
15. How MA technique is helpful in developing NdFeB magnets.
16. Write a note on MA powders useful for spray coatings.
17. Why MA techniques are helpful in achieving superplastic
behaviour.
18. Give an account of development of tribological materials using MA
techniques.
19. Give an account of development of particulate composites using
MA techniques.
20. What are the advantages of nanocrystalline materials? Discuss one
example where nanocrystallization has been achieved using MA
techniques.
21. Discuss how MA techniques are advantageous in production of tita-
nium powder from ilmenite.
22. List the reasons for the slow acceptance of MA products at indus-
trial level. How can this state of affairs can improved?

143
Mechanical Alloying

144
List of Symbols
Applications

LIST OF SYMBOLS

Aw Effective surface created during welding events; the weld area


b Burgers vector
C Bmin Minimum concentration of solute atoms B
Cp Specific heat of powder
D Diffusivity
D A , D B Self-diffusion coefficient of atoms A and B, respectively.
d Particle diameter; grain diameter
d0 Initial grain diameter
E Youngs modulus
Eb Energy dissipated during an impact
E0 Pre-collision kinetic energy of the grinding medium ball
Et Total energy dissipated during MA
f Shape factor; coefficient of friction
fw Force acting at the particles weld
f(v)dv Fraction of particles having volume between v and v+dv
f sf Final shape factor
f si Initial shape factor
G Shear modulus
g Acceleration due to gravity
g p , g r ,gt Geometrical constant with respect to Hertzian-impact pressure, radius and
time, respectively
H Powder hardness
Hk Knoop hardness
Hv Vickers hardness
H mix Enthalpy of mixing
Ht Total enthalpy input by cold work
h Height
h Rebound height
h0 Powder coating thickness
ht Heat transfer coefficient
I Intensity of a superlattice line
k Particle thermal conductivity; pre-collection velocity constant with respect
to impact time
k0 Grain size coefficient
km Materials constant
kr Dislocation relaxation parameter with respect to applied tensile stresses
l Dispersoid particles spacing
N Revolutions per minute (rpm)
n Number of rolling cycles; work hardening coefficient
n(v)(dv) Number of particles having volume between v and v+dv
Q Activation energy
R Universal gas constant
Rb Radius of the grinding ball
Rc Radius of the mill-cell
R ia i th peak position of the pair distribution function of the amorphous solid
R ci Radius of the i th co-ordination sphere of the crystal
R al, Rcl Nearest neighbour distance in amorphous and crystalline states respectively
Rw Radius of weld region
r Radius of atom
r A ,r B Radius of atoms A and B, respectively

145
Mechanical Alloying

rd Dispersoid radius
rh Hertz radius
rp Particles radius
S Relative long range parameter
T Absolute temperature
Ta Ambient temperature
Tb Bulk temperature
Tc Curie temperature
T crit Critical temperature
Ti Post-impact temperature
T ig Ignition temperature
Ts Powder surface temperature
T Change in temperature
t Mechanical alloying processing time
tc Time between collisions
t ig Incubation time for combustion to start
V Volume
Vp Volume of powder impacted per collision
x Vial length in Spex mill
xc Critical thickness of amorphous layer
Z Zener Holloman parameter; viscosity of lubricant
f Particle fracture probability
w Particles welding probability
Deformation manifested strains
e& Particles strain rate during collision
c, f Critical strain to fracture
Relative change in bond length
b Rebound yield
Mean free path between collisions
0 Atomic size factor
0 Relative colliding velocity of the ball
r Grain boundary velocity during recrystallization
s Velocity of sound
Density of ball
Stresses experienced by powder during collision
th Threshold stress below which there is negligible creep
u Tensile strength (ultimate)
y Yield strength (0.2 per offset)
Half of impact duration
Angular velocity of horizontal ball mill
c Critical rotational velocity of horizontal ball mill
p Angular velocity of plate in a planetary mill
v Angular velocity of vial with respect to plate in a planetary mill

146
Index
Applications

Index

A CALPHAD method 72
acids 31 carbide dispersion 65
AgCu 27, 125 chemical reactions 51, 133
AgFe 27 chromium carbides 102
Al(7.515)% Ti 117 (CoFe)75Si15B10 32
Al2% Li4.4% Mg 115 closed packed array 12
Al2%Mg4.4%Cu1.1%C0.8%O 130 Coble creep 130
Al3% Li2.2% Mg 114 consolidation 58
Al4% Mg1.1% C0.8% 112 co-operative grain boundary sliding
Al4.5%Cugraphite 131 (CGBS) 128
Al8% Fe4% Ce 118 critical reaction temperature 52
Al8%Fe2% Cr 132 cryomilling 26
AlAl4C3 composite powder 25 crystal-to-amorphous transformation 47
AlFe 27 CuAg 29
AlFeCe 28, 116, 118 CuAl 42
AlFeMn 28 CuAlC 116
AlMg 31 CuC 116
AlTi 116 Cu40%Zn 133
Al2O3Y2O3 105 Cu8% Ti4% B 120
Al50Mg50 33 CuCr 119
aluminium alloys (see dispersion strengthened) CuFe 4, 27
amorphization 21 CuPb 4
amorphization by MA 41 CuV 42
amorphization reaction 48 CuW 42
amorphous solids 136 CuZr 119
Arrhenius equation 58
atomistic models 72, 87 D
attrition 6 dead zone 11
attritor 6 degree of crystallization 20
attritor efficiency 11 differential thermal calorimetry (DTC) 49
attritor mill 3 diffusion welding 92
average flake size 19 diffusivity 53, 45
B Dispal 91, 112, 4, 115
dispersion strengthening (DS) 2
B2 TiAl 26 materials 2
ball mill 2 commercial materials 2
ball rotational velocity 81 superalloys 1, 2, 3, 18, 100, 105
ball to powder ratio 9 iron-base materials 4
ball velocity gradient 11 aluminium alloys 107
balltopowder ratio (BPR) 19, 20 copper 115
brazing 93 dispersoid particle radius 67
brittlebrittle system 37 dispersoid radius 67
bilk temperature rise 85 dispersoid spacing 67
dispersoid strengthening 67
C dispersoid volume fraction 67
CALPHAD 43 double mechanical alloying 28
ductilebrittle system 36

147
Mechanical Alloying

ductileductile system 35 H
dynamic compaction techniques 60
HallPetch relationship 65
E Hall-Petch effect 67
head-on collisions 14
EDAX 128
heat treatment 25, 28, 105, 106, 123
elongated grains 61, 62
Hertz impact theory 84
equation of kneading 27
Hertz radius 14, 84
EPMA 39, 41
Hertzian collision 85
Energy maps 55, 56
HfAl 42
ethylene bis disteramide 31
HfCu 42
exothermic redox reactions 51
HfNi 42
explosive compaction 89
high-angle grain boundaries 49
extrusion ratio 60, 101, 113
high-energy milling mode 15
F high-speed blenders and shakers 7
horizontal mills 81
fatty acids 2 hot isostatic pressing (HIP) 60
Fe13.5%Cr0.3%Mo1%Ti0.3%Y2O3 hot upsetting 104
107 HSLA steels 67
Fe20% Cr4.5% Al0.5% Ti0.5% Y2O3 hydrogen storage materials 4, 123
107 hydrostatic stress 39
Fe2B 22, 23
FeCrAl 109 I
fluid dies 59
ignition surface coating technique 2
fussionfusion 77
ignition temperature 52
forged bonding 94
ilmenite 139
friction coefficient 133
immiscible liquids 4
fracture
immiscible solids 4
dynamic 72
impact energy 20
forging 72,74
impact stress 14
shear 72
impact time 83
free-ball velocity 14
IN 9021-T4 113
forging 60,62
IN 9052 112
IN-905 XL 116
G IncoMAP 111
INCONEL MA 754 103
gas gun launcher 60 incubation period 52
GeSi 37 initial grain diameter 58
glass transition temperature 43 internal oxidation 2
glide 50, 66 International Nickel Company 3
global modelling 73 interparticle necking 37
mechanistic 71 interparticle spacing 2
atomistic 71, 86 iron titanate 139
thermodynamic 71, 87
kinetic 71, 87 K
grind limit 17
ketones 31
grinding balls 9, 23
kerosene 2
grinding media 10
kinetic models 72

148
Index
Applications

kneading technique 27 MSMA 100, 101

L N
laboratory planetary mill 9 nanocrystalline cermets 51
large diameter ball mills 9 (see also horizontal nanocrystalline solids 137
mills) nanocrystalline TiN powder 123
lattice strain 46 nanocrystallization 49
lattice defects 15 Nb3Ge 48
liquid phase sintering 61 NbAl 42
liquid quenching model 41 NbCuSi 42
local modelling 73 NbNi 40, 42
localized melting 61 NbSn 48
long-range order (also see LRO parameter) NbC 110
low-angle grain boundaries 50 Nb40Ni60 40
low density alloys 113 NdFeB magnets 126
low energy milling mode 15 Nd16Fe76B8 128
low pressure plasma spraying (LPPS) 125 NiZr 21
LRO parameter 47 NiZr2 48
Ni20%Cr2%ThO2 40
M Ni32TiNi45Nb55 46
MA 6000 alloy 91, 92, 103, 106 Ni60Nb40 23
MA 754 103, 107 Ni60Ti40 23
MA 758 103 nickel-base superalloys 3, 103
MA 760 103, 104, 107 NiTi2NiNb 46
MA 956 107 NiZr2 22
MA 957 107 NiZr2Ni11Zr9 37
magnesium-base materials 123 O
magnetic materials 126
mean free path between collisions 80 occluding air 5
mechanically activated chemical reactions ordering 56
138 Orowan mechanism 66
mechanical grinding 15 Orowan strengthening 69
mechanical interlocking 58 Orowan stress 70
mechanistic model 85 OrowanAshby expression 67
mechanistic models 72 Ostwald ripening 1
metastable hydrides 25 oxidation 1
Meyers hardness coefficient 68 oxidation resistance 103
MgAl2O4 134 oxide coated balls 10
MgO 112 oxide dispersion microforged material
MgZn 16 (ODM) 109, 110
Mg70Zn30 15 oxide dispersion 65
microcracks 38 oxygen/carbon ratio 32
microhardness 18, 29
mill parameters 19 P
mill speed 19, 21 packing factor 81
mixing technique 2 pair distribution function 86
modelling mechanical alloying 72 particle thermal conductivity 87
modified attritor 11 particle welding 15
morphology 20, 37, 51, 101 pebble mill 3
Mssbauer spectroscopy 19 pipe diffusion 36
MS 100, 101 planetary ball mill 9

149
Mechanical Alloying

plasma short activated sintering 61 Spex vibratory mill 8


polytetrafluoroethylene (PTFE) 136 strain hardening coefficient 68
post impact temperature 87 superconducting transition temperature 19
potential energy of a ball 55 superconductors 4
powder cooling 86 superplasticity 4,66,125,130
powder hardness 74 supersaturated solutions 4, 125
powder heating 85 Szegvari attritor 6
precollision velocity 83 surfactants 31
process control agents 31
processing time 80 T
R thermomechanical processing 62
Ti1% Al8% V5% Fe1% Er 123
radial ball velocity profile 11 Ti1% Al8%V5% Fe1% Er 101
rapid omnidirectional compaction Ti24% V10% Cr5% Er 123
(ROM) 59, 60 Ti24%V10%Cr5% Er 101
rapid solidification 28, 96 Ti24%V10%Cr5%Er 101
reaction milling 25 TiCu 42
rebound yield 55 TiMgGd 117
relative long-range order 47 TiMg 121
relative velocity of the balls at impact 74 TiN 122
repeated powder forging 29 TiSi 36,136
repeated rapid solidification 41 TiNiTiC 135
repeated rolling 26 TiC 110
retarded temperature effect 22 timetemperature excursion 58
RSMA process 100 time between collisions 80
rutile 139 TiN 110
titanium 1
S titanium aluminides 121
SSAR (see solid state amorphization reaction) titanium oxide 1
SAD 133 titanium systems 98
SAP 1 total stored energy 22
SCC resistance 69 transformation path (see also amorphization
selective reduction process 2 path) 22,23,48
self-sustained high-performance synthesis U
(SHS) 139
size of the grinding ball 20 uni-ball mill 12
SmCo magnets 128
SmFe magnets 127 V
SmFeN 128 vacancies 15
Sm-Co magnets 128 Vegards law 125
Sm-Fe magnets 127 vial 81
Sm10.5Co89.5 129 vial length 82
Sm16.7Co83.3 129 vibratory ball mill 8
solgel method 137
solid solubility extension 51 W
solid solution strengthening 2, 65
solid state amorphization reaction model 42 water-jacketed milling chambers 19
solid state welding techniques 91 wear 20
Spex 8000 mixer mill 122 welding 89,91
SPEX mill 80 wettability 129
Spex shaker mills 9 WS2 133

150
Index
Applications

X
X-ray technique 39
X-ray line broadening 19 (see also X-ray
technique)

Y
Y2O3 13, 36, 102
Yield strength 64,65,67

Z
ZenerHolloman parameter 62
zone annealing 62,100
ZnFe 42
ZnNi 42

151

Das könnte Ihnen auch gefallen