You are on page 1of 357

NANOCRYSTALLINE

MATERIALS
A.I. Gusev, A.A. Rempel

CAMBRIDGE INTERNATIONAL SCIENCE PUBLISHING


iii

CONTENTS

Preface ..................................................................................................................... vii


List of Main Notations ........................................................................................ xiii

1.

INTRODUCTION ............................................................. 1
References ................................................................................................... 23

2.
2.1
2.2
2.3
2.4
2.5
2.6
2.7
2.8
2.9.

SYNTHESIS OF NANOCRYSTALLINE POWDERS ............. 27


GAS PHASE SYNTHESIS ....................................................................... 27
PLASMA CHEMICAL TECHNIQUE .................................................... 32
PRECIPITATION FROM COLLOID SOLUTIONS ............................. 44
THERMAL DECOMPOSITION AND REDUCTION ......................... 49
MILLING AND MECHANICAL ALLOYING ..................................... 52
SYNTHESIS BY DETONATION AND ELECTRIC EXPLOSION .. 59
ORDERING IN NON-STOICHIOMETRIC COMPOUNDS .............. 65
SYNTHESIS OF HIGH-DISPERSED OXIDES IN LIQUID
METALS ...................................................................................................... 76
SELF-PROPAGATING HIGH-TEMPERATURE SYNTHESIS ......... 78
References ................................................................................................... 79

3.

PREPARATION OF BULK NANOCRYSTALLINE


MATERIALS ................................................................. 89

3.1.
3.2.
3.3
3.4.
3.5.

COMPACTION OF NANOPOWDERS .................................................. 90


FILM AND COATING DEPOSITION ................................................. 101
CRYSTALLISATION OF AMORPHOUS ALLOYS .......................... 104
SEVERE PLASTIC DEFORMATION .................................................. 108
DISORDERORDER TRANSFORMATIONS .................................... 114
References ................................................................................................. 124

4.

EVALUATION OF THE SIZE

4.1.
4.2.
4.3.

OF SMALL PARTICLES .... 131


ELECTRON MICROSCOPY ................................................................. 132
DIFFRACTION ........................................................................................ 137
SUPERPARAMAGNETISM, SEDIMENTATION, PHOTON
CORRELATION SPECTROSCOPY AND GAS ADSORPTION .... 151
References ................................................................................................. 156
v

5.

PROPERTIES

ISOLATED NANOPARTICLES AND NANOCPOWDERS .............................................. 159

5.1
5.2
5.3
5.4
5.5

RYSTALLINE
STRUCTURAL AND PHASE TRANSFORMATIONS ..................... 159
CRYSTAL LATTICE CONSTANT ....................................................... 169
PHONON SPECTRUM AND HEAT CAPACITY .............................. 177
MAGNETIC PROPERTIES .................................................................... 190
OPTICAL PROPERTIES ........................................................................ 207
References ................................................................................................. 214

OF

6.

MICROSTRUCTURE OF COMPACTED AND BULK


NANOCRYSTALLINE MATERIALS ............................... 227

6.1
6.2.

INTERFACES IN COMPACTED MATERIALS ................................. 228


STUDY OF NANOCRYSTALLINE MATERIALS BY MEANS OF
POSITRON ANNIHILATION TECHNIQUE ...................................... 234
STRUCTURAL FEATURES OF SUBMICROCRYSTALLINE
METALS PREPARED BY SEVERE PLASTIC DEFORMATION .. 258
NANOSTRUCTURE OF DISORDERED SYSTEMS ........................ 268
References ................................................................................................. 275

6.3
6.4

7.

EFFECT OF THE GRAIN SIZE AND INTERFACES ON THE


PROPERTIES OF BULK NANOMATERIALS ................... 284

7.1
7.2
7.3

MECHANICAL PROPERTIES .............................................................. 284


THERMAL AND ELECTRIC PROPERTIES ..................................... 301
MAGNETIC PROPERTIES .................................................................... 311
References ................................................................................................. 330

8.

CONCLUSIONS ........................................................... 340


References ................................................................................................. 345

SUBJECT INDEX ................................................................. 347

vi

Tell all the Truth but tell it slant


Success in Circuit lies
Too bright for our infirm Delight
The Truths superb surprise
After Emily Dickinson

Preface
In 1998 the monograph Nanocrystalline Materials: Preparation and
Properties by A. I. Gusev was published by Ural Division of the
Russian Academy of Sciences Publishing House (Yekaterinburg).
The monograph was the first Russian and one of the first in the
world generalisation of experimental results and theoretical
considerations regarding the structure and properties of not only
dispersed but also bulk solids with the nanometer size of particles,
grains, crystallites and other elements of the structure. The
monograph was of considerable interest to readers and became,
almost immediately after publishing, a bibliographic rarity not only
for readers but also for the majority of scientific and technical
libraries. In more than 10 technical universities of Russia, this book
is used as a basis of a course of lectures Nanocrystalline
substances and materials for students, specialising in advanced
materials science. Therefore, already in the year 2000 and
subsequently in 2001, the Nauka Publishing House (Moscow)
published twice a supplemented edition of the book
Nanocrystalline Materials written by A. I. Gusev and A. A.
Rempel.
The English edition of the monograph by A. I. Gusev and A. A.
Rempel, presented here to the reader, has been greatly refreshed,
expanded and supplemented in comparison with the last Russian
edition. The monograph is concerned with one of the most
important current scientific problems, which is common for
materials science, solid state physics and solid state chemistry,
namely the nanocrystalline state of matter. It may be expected that
the publication of the monograph in its new, expanded version will
be available to a considerably larger number of investigators and
engineers concerned with the production and application of
nanocrystalline materials.
vii

Until recently, the main scientific data on the nanocrystalline


state of matter were published in various scientific journals,
conference proceedings and compilations of articles. The authors of
this monograph have taken the difficult task of presenting to the
reader information on hundreds of original investigations of the
nanocrystalline state, grouping these investigations in accordance
with the investigated materials and properties, describing the
general and special features in the results of these investigations,
and focusing attention on the most interesting and practically
important effects of the nanocrystalline state.
The term nano, which is derived from the Greek word nanos
which means dwarf, designates a milliardth (10 -9 ) fraction of a unit.
Thus, the science of nanostructures and nanomaterials deals with
objects in condensed matter physics on a size scale of 1 to 100 nm.
The special physical properties of small particles have been
utilised by peoples for a very long time, although this has been
carried out unknowingly. Suitable examples are ancient Egypt
glasses, colored with colloidal particles of metals, dye pigments
used in different historical periods. The first scientific mention of
the small particles is evidently the disordered movement of particles
of flower pollen, suspended in a liquid, discovered in 1827 by the
Scottish botanist R. Brown. This phenomenon is referred to as
Brownian motion. The article on this microscopic observation (R.
Brown, Phil. Mag. 4, 161 (1828)) laid foundations to many
investigations. The theory of Brownian motion, developed
independently by A. Einstein and M. Smoluchowski at the
beginning of the 20 th century, is the basis of one of experimental
methods of determining the size of small particles. The scattering
of light by colloid solutions and glasses was studied by M. Faraday
between 1850 and 1860.
The starting point of examination of the nanostructured state of
substance were the investigations in the area of colloid chemistry,
which were already quite extensive since the middle of the 19 th
century. At the beginning of the 20 th century, a significant
contribution to the experimental confirmation of the theory of
Brownian motion, to the development of colloid chemistry and
examination of dispersed substances, and to the determination of the
size of colloid particles was provided by the Swedish scientist T.
Svedberg. In 1919, he developed a method of separating colloid
particles from a solution using an ultracentrifuge. In 1926, he
received a Nobel Prize in chemistry for his work on dispersed
systems.
viii

The 20 th century is characterised by extensive investigations of


heterogeneous catalysis, ultrafine powders and thin films. These
investigations raise question about the effect of the small size of
particles (grains) on the properties of studied materials. At present,
the nanostructured materials include nanopowders of metals, alloys,
intermetallics, oxides, carbides, nitrides, borides and these
substances in the bulk state with the grains of the nanometer size,
together with nanopolymers, carbon nanostructures, nanoporous
materials, nanocomposites, and biological nanomaterials. The
development of nanomaterials is directly associated with the
development and application of nanotechnology. The examination
of nanomaterials has revealed a large number of grey areas in the
fundamental knowledge of the nature of the nanocrystalline state
and its stability under different conditions. On the whole, the field
of nanomaterials and nanotechnology is very wide and at present
has no distinctive contours.
The unique structure and properties of small atomic aggregations
are of considerable scientific and technical interest, because they
represent an intermediate state between the structure and properties
of isolated atoms and bulk solids. However, the problem at what
stage of atom agglomeration the properties of bulk crystals is
formed completely has not as yet been solved. It is not clear how
the contributions of surface (associated with the interfaces) and
bulk (associated with the size of the particles) effects to the
properties of the nanocrystalline materials can be separated. The
investigations in this field were carried out for a long period of
time on isolated clusters, consisting from two atoms up to hundreds
of atoms, small particles with a size of more than 1 nm, and
ultrafine powders. The transition from the properties of isolated
nanoparticles to the properties of bulk crystalline substances
remained a grey area because the intermediate member, i. e. a bulk
solid with the grains of the nanometer size, was not artificial
created. Only after 1985, when methods of preparation of bulk
nanocrystalline substances were developed, work was started to fill
this gap in the knowledge of solids.
The spectrum of properties of matter can be enormously
enhanced if nanometer-size particles are agglomerated to a bulk
material so that in addition to the crystallites with a nanometer-size
they consist of a large portion of interfaces with a disordered
structure and novel properties. In very small crystallites of the size
of a few nanometers, this is a few millionth of a millimeter, new
properties appear due to quantum size effects or scaling laws, which
ix

can be controlled by the size of crystallites or particles.


Indeed, the scientific interest to the nanocrystalline state of the
solids in the powdered and bulk form is associated mainly with
expectation of various size effects on the properties of the
nanoparticles and nanocrystallites whose sizes are comparable or
smaller than the characteristic correlation scale of a specific
physical phenomenon or the characteristic length, which are present
in the theoretical description of some property or process. These
characteristic lengths are the free path of the electrons, the length
of coherence in superconductors, the wavelength of elastic
oscillations, the size of the exciton in semiconductors, the size of
the magnetic domain in ferromagnetics, etc.
The industrial interest to the nanomaterials is caused by the
possibility of extensive modification and even principal changes in
the properties of the existing materials at transition to the
nanocrystalline state, and by new possibilities offered by
nanotechnology in the creation of materials and wares from
structural elements of the nanometer size. The essence of
nanotechnology is the possibility to work at the atomic and
molecular level, in the length scale between 1 and 100 nm, in order
to produce and use materials and devices characterised by new
properties and functions because of the small scale of their
structure. Thus, the term nanotechnology relates to the sizes of
structural elements in particular. Nanoproducts are already playing
an important role in almost all branches of industry. The range of
application of these products is huge: more efficient catalysts, films
for microelectronics, new magnetic materials, protective coatings on
metals, plastics and glasses. In the next couple of decades,
nanostructured objects will operate in biological systems and will
be used in medicine. The successes of nanotechnology may be
manifested most efficiently in electronics and computer technology
as a result of further miniaturisation electronic devices and the
development of nanotransistors.
In the content of this monograph, we have attempted to take into
account both purely scientific, fundamental interest in the problem
of the nanocrystalline state as a special non-equilibrium state of
matter, and also some technical aspects of this problem, which are
of considerable importance for materials science and the practical
application of nanomaterials.
The combined analysis of the structure and properties of isolated
nanoparticles and nanopowders presented in the book, on the one side,
and of bulk nanomaterials, on the other side, shows that the level of
x

LIST OF MAIN NOTATIONS

aB1
c
Cp, Cv
D, <D>
Ddiff
E
EF
FWHM
g( ), g( )
G
h, k, l
h, D = h/2
Hc
HV
I
kB
K hkl
m0
m*
NA
N(EF)
p
R( )
t
T
TC
Tmelt
T trans

lattice constant (period) of cubic unit cell


with B1 structure
concentration
heat capacity at constant pressure and at
constant volume
size or mean size (diameter) of particle
(grain, crystallite, domain)
diffusion coefficient
modulus of elasticity
Fermi energy
full linewidth at half maximum
frequency distribution function
shear modulus
Miller indices
Plancks constant
coercive force
microhardness
relative intensity
Boltzmann constant
Scherrers constant
free electron mass
effective electron mass
Avogadros number
density of electronic states on the Fermi level
pressure
angle resolution function
time
temperature
Curie temperature
melting temperature
phase transformation (transition) temperature
xiii

understanding and clarification of the structure and properties of


isolated nanoparticles is considerably higher in comparison with bulk
nanocrystalline materials. Evidently, this is a result of the considerably
longer (practically from the beginning of the 20 th century) study of
high-dispersed systems and nanoclusters in comparison with bulk
nanomaterials which became the object of investigations only in the
last 1015 years.
The monograph in the concentrated form includes a large part of
the most important data on the nanocrystalline state of solids. In
writing the monograph, we used a very large number of original
investigations, starting from 1828 up to the year 2003, inclusive.
It should be noted that more than 80 % of all references is made
to studies carried out since 1988. Thus, the monograph reflects
accurately the current state of investigations of the nanocrystalline
state of solids. We think it will be interesting and useful to
specialists in condensed state physics, solid state chemistry,
physical chemistry and materials science.

Yekaterinburg, March 2004

A. I. Gusev, A. A. Rempel

xi

d
h
s
e

D
T

B
p

relative content of interstitial atoms X in


nonstoichiometric compounds MX y
broadening of diffraction reflection
deformation broadening
inhomogeneity broadening
size broadening
electronic heat capacity coefficient
viscosity (liquid shear viscosity)
Bragg diffraction angle
characteristic Debye temperature
isotermal compressibility
radiation wavelength
normalized positron annihilation rate
chemical potential
Bohr magneton
magnetic permeability
surface tension
dispersion
positron lifetime
relaxation time
magnetic susceptibility

xiv

Introduction

...And freely men confess that this worlds spent,


When in the planets and the firmament
They seek so many new; they see that this
Is crumbled out again to his atomies.
This all in pieces, all coherence gone,
All just supply, and all relation...
After John Donne (An Anatomy of the World, 1611)

+D=FJAH
Introduction
The problem of production of ultrafine powders of metals, alloys and
compounds and submicrocrystalline materials for different areas of
technology, has been discussed in literature for many years now.
In the last couple of decades, the interest in this subject has greatly
increased because it was found (primarily, in metals) that a decrease
in the size of crystals below some threshold value may result in a
large change of the properties [116]. These effects form when
the mean size of crystalline grains does not exceed 100 nm, and are
most evident when the grain size is smaller than 10 nm. When
examining the properties of superfine materials, it is necessary to
take into account not only their structure and composition but also
dispersion. Polycrystalline superfine materials with a mean grain size
of 300 to 40 nm are referred to as submicrocrystalline, and those
with a mean grain size of less than 40 nm as nanocrystalline. The
conditional classification of materials on the basis of the size D of
particles (grains) is shown in Fig. 1.1. Nanomaterials can also be
classified on the basis of their geometrical form and the
dimensionality of structural elements from which they consist. The
main types of nanocrystalline materials as regards the dimensionality
are cluster materials, fibrous materials, films and multilayer
materials, and also polycrystalline materials whose grains have
1

Nanocrystalline Materials

&RDUVH JUDLQHG PDWHULDOV

3DUWLFOH JUDLQ VL]H 

 P

a[

6XEPLFURFU\VWDOOLQH PDWHULDOV

 QP

a

1DQRFU\VWDOOLQH PDWHULDOV
 QP

,FRVDKHGUDO PHWDOOLF FOXVWHUV


 QP

0ROHFXODU FOXVWHUV IXOOHUHQHV 09.


$WRPV DQG  DWRP PROHFXOHV



1XPEHU RI DWRPV LQ D SDUWLFOH JUDLQ 3

 P





 SP

Fig. 1.1. Classification of substances and materials on the basis of particle (grain)
size D.

Clusters

(0D)

Nanotubes, filaments
and rods

(1D)

Films and layers

Space network

(2D)

(3D)

Fig. 1.2. Types of nanocrystalline materials: 0D (zero-dimensional) clusters;


1D (one-dimensional) nanotubes, filaments and rods; 2D (two-dimensional) films
and layers; 3D (three-dimensional) polycrystals.

comparable size in all three mutually perpendicular directions (Fig.


1.2). In this book, attention will be given mainly to the structure and
properties of bulk and powdered substances and materials with a
particle size of 5 to 200300 nm, i.e. nanocrystalline and
submicrocrystalline.
2

Introduction

For the investigator and the engineer used to working with


traditional substances and materials in which the elements of the
microstructure have the size of approximately 1 m or more, the
first encounter with nanomaterials results in at least surprise. This
surprise is similar to that experienced by somebody who has seen
for the first time buildings constructed in the modern style: instead
of normal straight lines and angles, there are distorted planes and
complicated broken contours (Fig. 1.3). However, after some time
it becomes obvious that it is possible to live in such an unusual
building and that it may even be interesting. This also relates to the
nanocrystalline state, i.e. it is unusual but it may be interesting to
work with it.
The difference between the properties of small particles and the
properties of bulk materials has been known for a relatively long
period of time and has been utilised in different areas of technology.
Suitable examples are widely used aerosols, dye pigments, glass
colored by colloidal particles of metals.
Suspensions of metallic nanoparticles (usually iron or its alloys)
with the size from 30 nm to 12 m are used as additions to engine
oil during service for the restoration of worn components of
automobiles and other engines.
The small particles and nanosized elements are used for the
production of various aviation materials. For example, radiowave-

Fig. 1.3. The first encounter with nanomaterials causes the same surprise as these
buildings in the modern style, designed by F.OGehry in Dsseldorf (Germany).

Nanocrystalline Materials

absorbing ceramic materials, whose matrix is characterised by the


random distribution of fine-dispersion metallic particles, are used in
aviation industry. Whisker single crystals and polycrystals (fibres)
are characterised by very high strength, for example, graphite
whiskers have a strength of ~24.5 GPa or ten times higher than the
strength of steel wire. They are used as fillers for light composite
materials for aerospace applications. Carbon fibres and graphite
whiskers are relatively thick (approximately 110 m) and are not
nanomaterials, but their production and applications were the first
step on the path to the development of carbon nanomaterials. After
discovery in 19841985 of a new allotropic modification of carbon,
i.e. spherical fullerenes C 60 [17, 18], attempts were made to
produce other topological forms of carbon nanoparticles. One of the
proposed possible forms of carbon nanoparticles was, in particular,
a quasi-one-dimensional tubular structure [19], referred to as the
nanotube. The nanotubes form as a result of rotation of (0001)
basal planes of the hexagonal lattice of graphite and can be single
or multilayered. In fact, in 1991 and in the following years of the
20th century, researchers succeeded in detecting quasi-onedimensional tubular structures of carbon, i.e. carbon nanotubes)
[2022]. As an example, Fig. 1.4 shows a computer graphical model
of a double-shell carbon nanotube, and Fig. 1.5 shows experimentally
produced nanotubes [23]. Carbon nanotubes with a diameter of

Fig. 1.4. A computer graphic model for


a double-shell carbon nanotube showing
a helical arrangement of hexagons [23].
4

Introduction

Fig. 1.5. Multi-shell carbon nanotubes [23].

D > 5 nm, consisting of 2 to 50 coaxial tubes, were detected for


the first time by transmission electron microscopy in a condensate
in an electric arc discharge between graphite electrodes [20]. The
results of modelling of the structure and electronic properties of
carbon nanotubes have been generalised in [24]. The carbon
nanotubes have high mechanical strength and may be used for
developing high-strength composites. Nanotubes have been used in
different mechanical nanodevices [25], like nanoindentors for
microhardness measurements. Depending on the type of helicoidal
ordering of the carbon atoms in the walls of the carbon nanotubes,
these nanotubes have semiconductor or metallic conductivity.
Consequently, they are used as conducting elements in electronic
nanotechnologies. In atomic force microscopes, the carbon
nanotubes have replaced the metallic probe [26].
By joining the carbon nanotubes it is possible to produce a large
number of structures with differing properties. Synthesis of these
structures is very important for electronic technology. T-connected
nanotubes, which may operate as a contact device, were produced
in [27]. The authors of [28, 29] grew Y-shaped carbon nanotubes
(Fig. 1.6); this structure is referred to as the Y-junction carbon
nanotube. Synthesis was carried out by chemical deposition from
the gas phase (CVD): pyrolysis of acetylene with subsequent
5

Nanocrystalline Materials

Fig. 1.6. Model of a carbon Y-nanotube.

growth of Y-nanotubes was carried out at a temperature of 920 K


in branching nanochannels of the aluminium matrix. Cobalt, being
a growth catalyst, was deposited on the walls and the bottom of the
nanochannels. The diameter of the stem of the produced Ynanotube was approximately 60 nm, the diameter of the branches
~40 nm. The authors of [30] produce carbon Y-nanotubes by
pyrolysis of organometallic precursors. As a result of a defective
structure in the area joining the prongs, the Y-nanotube passes
electric current only in one direction, i.e. it operates as a diode
[29]. If controlling voltage is additionally applied to one of the
prongs of the Y-nanotube, the nanotube operates as a current
stabiliser. The possibility of controlling the current leads to the
possibility of extensive application of Y-nanotubes in electronics.
Recently, a group of investigators [31] from the Department of
Materials Science and Engineering at the Rensselaer Polytechnic
Institute (Troy, USA) proposed a method of controlled growth of
carbon nanotubes on a substrate coated with a layer of SiO 2 .
Pattering of SiO 2 was generated by photolithography and then
subjected to combined wet and dry etching in order to produce
islands of SiO 2 distributed in a specific fashion. Subsequently,
bundles of nanotubes, forming a unique nanostructure (Fig. 1.7) were
grown in a gas mixture of xylene/ferrocene C 8 H 10 /Fe(C 5 H 5 ) 2 on
6

Introduction

Fig. 1.7. Controlled carbon nanotube growth on a silica-coated substrate [31].

the islands of SiO 2 by the CVD method. In this process, the iron
included in the composition of Fe(C 5 H 5 ) 2 plays the role of a
catalyst. Each bundle includes several tens of multiwalled nanotubes
with a diameter of 2030 nm. According to the authors of [31], such
nanostructures may be used in integrated systems of the next
generation and in microelectromechanical devices.
The first publications dealing with the production of boron nitride
nanotubes appeared in 19951996 [3234]. Intensive research is
being carried out into the synthesis of silicon carbide nanotubes.
The range of applications of these nanotubes is even wider because
of higher hardness and high melting point of silicon carbide.
Heterogeneous synthesis of silicon carbide fibres was described by
the authors of [35], the gas phase method of production of silicon
carbide nanofibres with a diameter of ~100 nm, produced from
silicon and carbon powders, was described in [36], and the authors
of [37] reported on hollow silicon carbide nanostructures. The
nanotubes and nanofilaments of silicon carbide and also of boron
carbide and SiO 2 , produced by these methods, were presented in
a lecture Elongated structures of silicon carbide: nanotubes,
nanofilaments and microfibres by A. I. Kharlamov at the NATO
Advanced Study Institute Synthesis, Functional Properties and
Applications of Nanostructures (July 26August 4, 2002, Heraklion,
Crete, Greece). At the same conference, in a lecture The stateof-art synthesis techniques for carbon nanotubes and nanotubesbased architecture P. Ajayan told about silicon carbide nanotubes
produced on an Al 2 O 3 substrate. At the 2nd NASA Advanced
7

Nanocrystalline Materials

Materials Symposium New Directions in Advanced Materials


Systems (May 2931, 2002, Cleveland, Ohio), D. Larkin presented
a report High temperature nano-technology: silicon carbide
nanotubes synthesis.
The results of investigations and application of various nanotubes
are presented in [38]. The book starts with the playful amateur
poem Material Ethereal by Peter Butzloff (University of North
Texas, Denton, USA) on a mystery nanotube and problems of
examining it. Only the first and final lines of this poem are given
here:
We speculate but underrate
what mystery we wrangle,
a Nanotube from carbon crude
that nature did entangle.
..............................................
Oh Nanotube of carbon crude
so cumbersome we trundle
through pass of phase, by time decays
and character to bundle!
The catalysis of chemical reactions is a very important and large
area of long-term and successful application of fine particles of
metals, alloys and semiconductors. Heterogeneous catalysis by
means of high-efficiency catalysts produced from ultrafine powders
or ceramics with grains of the nanometer size is an independent
and very large section of physical chemistry.
Various problems of catalysis have been discussed in hundreds
of books and reviews and tens of thousands of articles. The
extensive discussion of the problems of catalysis on fine particles
with respect to both the content and volume is outside the
framework of this book and, consequently, certain general
assumptions relating to the catalytic activity of fine particles will
be discussed only briefly.
Catalysis on fine particles plays an exceptionally important role
in industrial chemistry. Catalysed reactions usually take place at
lower temperatures than non-catalysed reactions and are more
selective. In most cases, the catalysts are represented by isolated
fine particles of metals or alloys deposited on a carrier with a
developed surface (zeolites, silicagel, silica, pumice, glass, etc.). The
main task of the carrier is to support obtaining the smallest size of

Introduction

deposited particles and prevent their spontaneous coalescence and


sintering.
The high catalytic activity of fine particles is explained by
electron and geometrical effects, although this division is very
conventional because both effects have the same source, i.e. the
small particles size. The number of atoms in an isolated metallic
particle is small and, consequently, the distance between the energy
levels E F /N (E F is Fermi energy, N is the number of atoms in
the particle) is comparable with thermal energy k B T. At > k B T
the levels are discrete and the particle loses its metallic properties.
The catalytic activity of the small metallic particles starts to be
evident when the value of is close to k B T. Consequently, it is
possible to evaluate the particle size at which the catalytic
properties become evident. For metals, Fermi energy E F is
approximately 10 eV, at room temperature 300 K the value = k B T
= 0.025 eV and, consequently N 400; a particle consisting of 400
atoms has a diameter of ~2 nm. In fact, the majority of data
confirm that the physical and catalytic properties start to change
markedly when the particles reached the size of 28 nm. In addition
to the discussed primary electron effect, there is a secondary
electron effect. This effect is caused by the fact that a large
fraction of atoms is situated on the surface of small particles. The
electron configuration of these atoms is different from one of the
atoms distributed inside the particle. The secondary electron effect
has a geometrical source and also leads to changes in the catalytic
properties.
The geometrical effect of catalysis depends on the number of
atoms distributed on the surface (on the faces), on the edges and
tops of the small particle because these atoms have a different
coordination. The atoms situated on the faces have a higher
coordination in comparison with the atoms on the tops and the
edges. If the atoms in the small coordination are catalytically most
active, the catalytic activity increases with decreasing particle size.
In another case, if the atoms located on the faces are catalytically
active, the rate of the catalysed reaction will be increased by larger
particles.
A specific role in catalysis is played by the carrier because the
atoms of the catalyst which are in direct contact with the carrier
may change their electronic structure because of the formation of
bonds with the carrier. It is evident that as the number of atoms
that are in contact with the carrier increases, the effect of the
carrier on catalytic activity becomes stronger. It is clear that the
9

Nanocrystalline Materials

effect of the carrier is relatively small for large particles but


increases and becomes quite strong with a decrease in the particle
size.
Metallic alloys (for example, alloys of catalytically inert metals
of group I with metals of group VIII) are used as a catalyst
because of the dilution of the metal-catalyst in the alloy increases
catalytic activity. This is similar to an increase in catalytic activity
with a decrease in the nanoparticle size. To a first approximation,
the similarity of the effects of a decrease in the particle size and
melting is caused by the fact that the valence electrons of every
metal in such alloys retain their affiliation and, consequently, a
catalytically inert metal (for example, copper) acts as a diluent for
the particles of the catalytically active metal.
Usually, the nanoparticles show catalytic activity in a very
narrow size range. For example, Rh catalysts, produced by the
dissociation of Rh 6 (CO) 16 clusters, fixed to the surface of disperse
silica, catalyse the reaction of hydrogenation of benzene only when
the particle size is 1.51.8 nm, i.e. only particles of Rh 12 are
catalytically active in relation to this reaction. The high selectivity
of the catalytic activity is also characteristic of nanoparticles of
widely used catalysts such as palladium and platinum. For example,
the hydrogenation of ethylene was studied at a temperature of
520 K and a hydrogen pressure of 1 atm on a platinum catalyst
deposited on SiO 2 or Al 2 O 3 . A distinctive maximum of the reaction
rate is observed when the size of Pt nanoparticles is about 0.6 nm.
This high sensitivity of catalytic activity to the size of small particles
confirms the importance of the development of selective methods
of production of nanoparticles with an accuracy to 12 atoms. The
very narrow size distribution of the nanoparticles is essential not
only for catalysis but also for microelectronics.
A new area of catalysis of small particles is photocatalysis using
semiconductor particles and nanostructured semiconductor films. For
example, this method is promising for photochemical purification of
effluents to remove various organic contaminants by means of their
photocatalytic oxidation and mineralisation.
Detailed analysis of the effect of the size of small particles of
metals and alloys, deposited on a carrier, can be found in [39] and
also in reviews [40, 41] concerned with catalysis using metallic
alloys and palladium.
Catalysis on small metallic particles may be regarded as the
chemical size effect. For example, nickel or palladium nanoparticles
on a SiO 2 substrate are used as catalysists for hydrogenation of
10

Introduction

benzene. Nanoparticles are produced by decomposition of


organometallic complexes. A decrease in the metallic particles size
is accompanied by an increase in specific catalytic activity, i.e. the
activity related to 1 surface atom of the metal. Let us consider the
reaction of hydrogenation of benzene at a temperature of 373 K
and at benzene C 6 H 6 and hydrogen H 2 pressures of 6700 and
46700 Pa, respectively. In this reaction, the specific catalytic
activity of nickel nanoparticles is increased 34 times when the
particle size become smaller than 1 nm and the dispersion tends to
unity. (Dispersion is the ratio of the number of atoms situated on
surface to the total number of the atoms in the particle.) In
catalysis on palladium nanoparticles with the dispersion close to
unity, the identical effect in the same reaction is observed at 300
K. A study of hydrogenolysis of ethane C 2 H 6 at a temperature of
473 K and a pressure of C 2 H 6 and H 2 equal to 6700 and 26700 Pa
showed that the very rapid increase in the specific catalytic activity
of the nickel nanoparticles is observed when their dispersion close
to unity.
The rate of the reaction of hydrogenolysis of cyclopentane and
methylcyclopentane, related to 1 surface atom of the metal-catalyst,
changes rapidly when the fraction of the surface atoms in the
nanoparticle of the metal-catalyst (Pt, Ir, Pd, Rh deposited on glass,
SiO 2 or Al 2 O 3 ) approaches unity [39].
Another chemical size effect is the shift of binding energy
3d 5/2 of the internal level of palladium in relation to the size of
palladium particles [39, 41]. For palladium particles larger than 4
5 nm, the binding energy of the 3d 5/2 level is ~335 eV, i.e. it is equal
to the value characteristic of bulk palladium. A decrease in the size
of palladium nanoparticles from 4 to 1 nm is accompanied
(irrespective of whether the material of the substrate is a conductor
(carbon) or and insulator (SiO 2 , Al 2 O 3 , zeolites)) by an increase of
the binding energy of the 3d 5/2 level. The most probable reason for
the positive shift is the size dependence of the electronic structure
of palladium, namely the decrease of the number of valence delectrons. An identical shift of the binding energy of Pt 4f 7/2 of the
internal level is recorded in the case of platinum nanoparticles [39].
The increase of the chemical activity of thin-film heterostructures is also a chemical size effect. For example, in two-layer
oxide heterostructures MgO/Nb 2 O 5 the reaction of the type

MgO + Nb 2O 5 MgNb 2O 6

(1.1)

11

Nanocrystalline Materials

takes place spontaneously at temperatures 8001000 K lower than


the temperature of the reaction between normal coarse-grained
oxides.
Hybrid nanocomposites of the metalpolymer type are produced
by forming nanoparticles in a specially prepared polymer matrix [42,
43]. Polymer composites with metallic nanoparticles are used as
electrically conducting film composite materials, and the amount of
the filler in the matrix may reach 90 vol %. The introduction of
metal ions into polymer fibers makes it possible to produce colored
lightguides suitable for application in computer equipment. The
optical properties of polymers with fillers of nanoparticles of metals,
alloys or semiconductors (CdS, CdSe, InP, InAs) are interesting.
Because of light machining and the possibility of producing films
from these polymer nanocomposites, they can be used for the
production of optical elements and light filters.
The nanoparticles and nanolayers are used widely in modern
technology. Multilayered nanostructures are used in the production
of microelectronic devices. A suitable example are layerheterogeneous nanostructures, i.e. superlattices in which superthin
layers (with a thickness from several to hundred of periods of the
crystal lattice or ~150 nm) of two different substances, for
example, oxides, alternate. The structure represents a crystal in
which in addition to the conventional lattice of periodically
distributed atoms, there is a superlattice of repeating layers of
different composition. Owing to the fact that the thickness of the
nanolayer is comparable with de Broglie wavelength of the electron,
the quantum size effect is realised in superlattices in electronic
properties. Utilising the effect of size quantisation in multilayered
nanostructures enables the production of electronic devices with
increased operating speed and information capacity. The simplest
electronic device of this type is, for example, the AlAs/GaAs/AlAs
two-barrier diode, consisting of a layer of gallium arsenide with a
thickness of 46 nm, distributed between two layers of aluminium
arsenide AlAs, with a thickness of 1.52.5 nm.
Of special interest are magnetic nanostructures characterised by
giant magnetoresistance. They are in the form of multilayered films
of alternating layers of ferromagnetic and non-magnetic metals, for
example, a ferromagnetic layer CoNiCu and a nonmagnetic
copper layer alternate in the CoNiCu/Cu nanostructure. The
thickness of the layers is of the order of the free path of the
electron, i.e. several tens of nanometers. Changing the strength of
the applied external magnetic field from 0 to some value of H it
12

Introduction

is possible to change the magnetic configuration of the multilayered


nanostructure in such a manner that the electrical resistance will
change in a very wide range. This makes it possible to utilise the
magnetic nanostructures as detectors of the magnetic field. The
highest value of giant magnetoresistance in the CoNiCu/Cu
nanostructure is obtained for very thin layers of copper, thickness
approximately 0.7 nm.
The development of electronics over a period of several decades
has also progressed along the path of miniaturisation. The first
jump in the development of electronic technology was the
transition from vacuum electron valves to the transistor. The second
jump is associated with the application of integrated microcircuits.
The transition to integrated microcircuits became possible after
understanding that all elements of the electronic circuit can be
produced from the same material of the semiconductor type, instead
of producing them from different materials. Silicon is such a
material. The application of the material of the same type enabled
the construction of all elements of the electronic circuit directly in
the same specimen of this material and, connecting the elements
together, produce an efficient microchip. The first necessity for
decrease in the size of the electronic circuits came from military
and space authorities of the USA, former USSR, European
countries, and Japan, who supported appropriate research projects.
If the first simplest chips (1959) consisted of tens of elements,
then in 1970, microcircuits included up to 10 000 elements. Advances
in electronics were accompanied by a rapid decrease of the cost
of electronic devices (Fig. 1.8). In 1958, the cost of one transistor

0LQLPDO W\SLFDO VL]H P



















6DOH YROXPH PLOOLDUGV GROODUV








Fig. 1.8. Decrease in the minimum characteristic size of electronic components and
growth of the volume of sales of electronic products [44].

13

Nanocrystalline Materials

was approximately 10 US dollars, and in the year 2000 this money


would purchase a microcircuit with tens of millions of transistors
[44]. In currently available mass-produced microcircuits,
approximately 1000 electrons are required for switching on/
switching off a transistor. At the end of the first decade of the 21st
century, the required number of electrons will decrease to 10 as a
result of miniaturisation [44] and work is already been carried out
to develop a single-electron transistor [45].
Semiconductor heterostructures, produced from two or more
different materials, are of special interest for electronics. In these
heterostructures, an important role is played by the transition layer,
i.e. the interface between two materials. In addition to this,
according to [46], the technical device in semiconductor
heterostructures is the interface itself.
Semiconductor heterostructures are fabricated from materials
containing such elements as Zn, Cd, Hg, Al, Ga, In, Si, Ge, P, As,
Sb, S, Se, Te. These elements belong to the groups IIVI of the
periodic table. Silicon occupies the most important place in the
technology of electronic materials, like steel in the production of
constructional materials. In addition to silicon, electronics require
semiconductor compounds A III B V and their solid solutions, and also
A II B VI compounds. Of the compounds of the A III B V type, gallium
arsenide GaAs is used most widely, and of the solid solutions it is
Al x Ga 1x As. The application of solid solutions makes it possible to
produce heterostructures with a continuous but not jump-like
variation of composition. The width of the forbidden band in these
heterostructures also changes continuously.
In the production of the heterostructures, it is important to match
the parameters of the crystal lattices of two contacting materials.
If the two materials with greatly differing lattice constants grow on
each other, then an increase in the thickness of the layers results
in the formation of high strains at the interface and mismatch
dislocations appear. Strains appear irrespective of whether the
transition between the two layers is smooth or not. To reduce the
strains, the lattice constants of the two materials should differ as
little as possible. Therefore, special attention in the study of
heterostructures is given to solid solutions of the AlAsGaAs
system because the arsenides of aluminium and gallium have almost
the same lattice constants. Single crystals of GaAs are an ideal
substrate for growing the heterostructures. Another natural substrate
is indium phosphide InP which is used in combination with solid
solutions GaAsInAs, AlAsAlSb, and others.
14

Introduction

A breakthrough in making thin-layer heterostructures took place


with the development of a technology for the growth of thin layers
by molecular beam epitaxy and liquid-phase epitaxy. It became
possible to growth heterostructures with a very sharp interface.
Consequently, it is possible to position the two heteroboundaries so
close to each other that the size quantum effects play the
controlling role in this intermediate space. The structures of this
type are referred to as quantum wells. In quantum wells, the mean
narrow-band layer has a thickness of several tens of nanometers
which results in splitting of the electronic levels because of the
size quantisation effect. This effect in the form of a characteristic
stepped structure of optical spectra of absorption of the GaAs
AlGaAs semiconductor heterostructure with the superthin GaAs
layer (quantum well) was detected for the first time by the authors
of [47]. They also found the shift of characteristic energies with
a decrease in the thickness of the quantum well (GaAs layer). In
quantum wells, superlattices and other structures with very thin
layers and high strains may form without the formation of
dislocations and, consequently, it is not necessary to match the
lattice parameters [48]. Heterostructures, especially double
heterostructures, including quantum wells, quantum wires and
quantum dots, enable the control of various fundamental parameters
of semiconductor crystals, such as the width of the forbidden band
(energy gap), the effective mass and mobility of charge carriers,
the electron energy spectrum.
The density of states N(E) in a three-dimensional (3D) semiconductor is a continuous function. A decrease in the dimensionality
of the electron gas results in a change of the energy spectrum from
continuous to discrete as a result of its splitting (Fig. 1.9). The
quantum well is a two-dimensional (2D) structure in which charge
carriers are restricted in the direction normal to layers, and may
move freely in the plane of the layer. In quantum wires, the charge
carriers are already restricted in two directions and move freely
only along the wire axis. The quantum dot is a zero-dimensional
(0D) structure where the charge carriers are already restricted in
all three directions and are characterised by a completely discrete
energy spectrum.
The size of the quantum dots produced by molecular beam
epitaxy and lithography ranges from 1000 to 10 nm; smaller
quantum dots (120 nm), whose surface is protected by organic
molecules preventing aggregation of the particles, may be produced
by colloidal chemistry methods [49].
15

'

G G a  

 

 

Nanocrystalline Materials

G G a 
 

G G a 

 

 '

 

'

G G a FRQVW

'

Fig. 1.9. Density of states N(E) for charge carriers as a function of the dimensionality
of the semiconductor: (3D) three-dimensional semiconductor, (2D) quantum well,
(1D) quantum wire, (0D) quantum dot.

The application of nanostructures in electronics will lead to


further miniaturisation of electronic devices with transfer to
nanosized elements for producing processors of a new generation.
The i386 TM processor, produced by Intel Corporation in 1983,
contained 275000 transistors and performed more than 5 million
operations per second; the i486 TM processor, produced in 1989,
already contained 1 200000 transistors. The most widely used
processor at the end of the 20th century and the beginning of the
21st century, Pentium Pro, contains 5.5 million transistors and
carries 300 million operations per second. The size of the
transistors reached the smallest value available for current
technologies and, consequently, a further decrease in the size may
be achieved only by the application of nanotechnology. A practical
difficulty which must be overcome in making quantum dots and
single-electron transistors is the time instability of structures with
a small number of atoms. The stability of these quantum-electronic
elements is determined by the jump (diffusion) of already a small
number of atoms. Since the diffusion processes on the surface and
16

Introduction

at the interface of quantum-electronic elements are very fast,


processes of failure of the elements or even their movement on the
substrate as an integral unit are already detected at room
temperature [50]. A problem of the stability of nanoelectronic
circuits can be solved only using multicomponent materials,
including oxides, carbides and nitrides of metals. These compounds
have a high melting point and low mobility of atoms and,
consequently, have high thermal and time stability.
X-ray and ultraviolet optics uses special mirrors with
multilayered coatings of alternating thin layers of elements with
high and low density, for example, tungsten and carbon,
molybdenum and carbon, or nickel and carbon; the thickness of a
pair of layers is about 1 nm, and the layers should be atomically
smooth. The possibility of producing multilayered x-ray mirrors is
one of the factors determining the application in certain areas of
nanotechnology, such as x-ray lithography, on the one hand, and in
astronomical and astrophysical investigations, on the other hand.
The formation of x-ray mirrors has been described in sufficient
detail in [51], where multilayered nickelcarbon Ni/C nanostructures
with a period of ~4 nm were investigated. Other optical devices
with nanosized elements, intended for application in x-ray
microscopy, are Frenel zone sheets with the smallest width of the
zone of approximately 10 nm, and diffraction gratings with a period
smaller than 100 nm.
F/S heterostructures, formed by alterating thin layers of a
ferromagnetic and a superconductor, are very interesting. In the
F/S heterostructures, the superconducting and ferromagnetic regions
are divided in space but are linked together through the interface
between the layers. In most cases, ferromagnetic interlayers F are
produced using Fe, Co, Gd, Ni whose Curie temperature T C is
considerably higher than the superconducting transition temperature
T sc of metals (Nb, Pb, V), forming the layer S. Experimental
examination of these heterostructures started in [52] in which the
method of rf sputtering was used to produce two-layer sandwiches
F/Pb. Other methods of production of superlattices of the F/S type
are molecular-beam epitaxy, electron beam evaporation, and dc
magnetron sputtering [53]. Generally, superconductivity and
ferromagnetism are antagonistic phenomena. Primarily, this
antagonism is reflected in relation to the magnetic field. The
superconductor tends to push out the magnetic field (Meisner
effect), and the ferromagnetic concentrates the force lines of the
magnetic field in its volume (magnetic induction effect). From the
17

Nanocrystalline Materials

viewpoint of microscopic theory, the antagonism is reduced to the


following: In a superconductor, the attraction force between the
electron generates Cooper pairs, whereas the volume interaction in
the ferromagnetic tends to align the electronic forces paralelly.
Taking this into account, the coexistence of the superconducting
and ferromagnetic order in a homogeneous system is unlikely, but
it can easily be achieved in artificial multilayered systems consisting
of alternating ferromagnetic and superconducting layers (Fig. 1.10).
The F/S type heterostructures with the layers of atomic thickness
may be used in electronic devices of the next generation as logic
elements and switches of superconducting current [54, 55], and
superconductivity can be controlled by means of a weak external
magnetic field [56]. It should be mentioned that the properties of
the F/S multilayered systems, including the superconducting
transition temperature, depend on the thickness of the
ferromagnetic and superconducting layers. In most cases, the
thickness of the ferromagnetic layer is smaller than 1 nm, the
thickness of the superconducting layer is from 10 to 4050 nm [56].
It is interesting to note that the superconducting transition
temperature T sc in the F/S heterostructures may not only
monotonically decreases but also oscillates with increasing thickness
of the layer F. For example, in a Fe/Nb/Fe three-layer system, with
,

 $

$  $

 $ 

]
\

]
\

]
\

G1 G8
[

]
\
.

Fig. 1.10. Multilayer heterostructures ferromagnetic-superconductor F/S: (a) double


layers, (b) triple layers, (c) superlattices

18

Introduction

an increase of the thickness of the iron layer d Fe from 0.1 to


0.8 nm, the temperature T sc initially decreases from 7 to 4.5 K and,
subsequently, with an increase of d Fe to 1.01.2 nm, T sc increases
to 5 K and with a further increase of d Fe to 3 nm, the
superconducting transition temperature decreases to 3.23.4 K [57].
The layers of the metal and alloy, for example, Nb x Ti 1x /Co or
V/Fe x V 1x , may alternate in the F/S heterostructures. The heterostructures of the superconductorferromagnetic semi-conductor type
(for example, NbN/EuO/Pb or NbN/EuS/Pb) [58] with a Josephson
tunnelling transition are also interesting. In these heterostructures,
the thickness of the layer of the ferromagnetic semiconductor (EuO,
EuS) varies from 10 to 50 nm, and the thickness of the
superconducting layers is greater than 200 nm.
Engineering applications have no other parts, which are working
in complicated critical conditions, as blades of gas turbines in turbojet engines. The transfer to a new generation of gas-turbine engines
requires constructional materials, whose strength and hardness
would be 20% higher, fracture toughness 50% better, and wear
resistance twice as large. Actual tests showed that the use of
heat-resistant nanocrystalline alloys in gas turbines provides at least
one-half of the required improvement in the properties. Ceramic
nanomaterials are widely used for fabrication of parts working at
elevated temperatures, under nonuniform thermal loads, and in
aggressive environment. Thanks to their superplasticity, ceramic
nanomaterials serve as materials of intricately shaped highly
precision products used in the aerospace technology. Nanoceramics
based on hydroxy-apatite possesses biocompatibility and high
strength and therefore is used in orthopedy for the making of
artificial joints and in stomatology for the making of dentures. The
nanocrystalline ferromagnetic alloys of the FeCuMSiB systems
(M is the transition metal of the groups IVVI) are used as
excellent transformer soft magnetic materials with a very low
coercive force and high magnetic permeability.
The small grain size determines a long length of the intergranular
interfaces: At a grain size from 100 to 10 nm the interfaces
contained 1050% of the atoms of the nanocrystalline solid. Also,
grains may have various atomic defects (vacancies or their
complexes, disclinations, and dislocations), whose number and
distribution differ from those in coarse grains 5 to 30 m in size.
If the dimensions of a solid in one, two or three directions are
comparable with characteristic physical parameters having the
length dimensionality (the size of magnetic domains, the electron free
19

Nanocrystalline Materials

path, the size of excitons, the de Broglie wavelength, etc.), size


effects will be observed for the corresponding properties. Thus, size
effects imply a set of phenomena connected with changes in
properties of substance, which are caused by (1) a change in the
particle size, (2) the contribution of interfaces to properties of the
system, and (3) comparability of the particle size with some physical
parameters having the length dimensionality. Because of these
specific features of the structure, the properties of the
nanocrystalline materials greatly differ from those of usual
polycrystals. Therefore, the decrease in the grain size is viewed as
an efficient method for adjustment and modification of properties
of solids. In fact, there are reports on the effect of the
nanocrystalline state on the magnetic properties of ferromagnetics
(Curie temperature, coercive force, saturation magnetisation) and
magnetic susceptibility of weak para- and diamagnetics, on the
effect of shape memory on the elastic properties of metals and
large changes of their heat capacity and hardness, on the variation
of the optical and luminescence characteristics of semiconductors,
on the appearance of the plasticity of boride, carbide, nitride and
oxide materials which are relatively brittle in the normal coarsegrained state. In nanocrystalline materials, the combination of high
hardness and plasticity is usually explained by difficulties in the
activation of dislocation sources as a result of the small dimensions
of crystals, on the one hand, and by the presence of grain boundary
diffusional creep, on the other hand [13]. The nanomaterials are
characterised by very high diffusibility of the atoms at the grain
boundaries, which is 56 orders of magnitude higher than that in
the normal polycrystals. However, the mechanisms of diffusion
processes in nanocrystalline substances have not yet been
completely explained and the literature contains contradicting
explanations of this problem. The problem of the microstructure of
the nanocrystals, i.e. the structure of the interfaces and their atomic
density, the effect of the nanovoids and other free volumes on the
properties of nanocrystals requires solution.
Usually, when discussing the nonequilibrium metastable state, it
is assumed that this state may be compared with some equilibrium
state which exists in reality. For example, the metastable glass-like
(amorphous) state corresponds to the equilibrium liquid state (melt).
The specific feature of the nanocrystalline state in comparison with
other well-known nonequilibrium metastable state of matter is the
absence of the equilibrium state corresponding to this state in the
structure and a long length of the boundaries.
20

Introduction

The nanocrystalline materials represent a special state of


condensed matter, i.e. macroscopic ensembles of ultrafine particles
up to several nanometers in size. The unusual properties of these
materials are determined by both the specific features of separate
particles (crystallites) and by their collective behaviour, which
depends on the interaction between nanoparticles.
The main scope of investigations of the nanocrystalline state is
to answer the following questions:
(1) Is there a sharp, distinctive boundary between the bulk and
the nanocrystalline states?;
(2) Is there some critical grain (particle) size below which the
characteristic properties of nanocrystals become observable, and
above which the material behaves as a bulk one? In other words,
is the transition from the bulk to the nanocrystalline state the phase
transformation of the first order from the thermodynamic viewpoint?
The answer to this question is important for the correct procedure
of experimental investigations of the nanocrystalline state and for
understanding the results.
At the first sight, the transition to the nanocrystalline state is not
a phase transformation because the size effects increase gradually
with a decrease in the size of isolated particles or grain size in bulk
nanomaterials. However, all experimental investigations have been
carried out on materials with a high dispersion of the particle or
grain size. Obviously, one can assume that the dispersion of the
dimensions leads to bluring of the phase transformation, if such
a transformation exists. This could be confirmed by experiments
with the detection of the size effect. Such experiment should be
carried out on materials of the same chemical composition but
different dispersion in grain size. The important condition of this
experiment is equal size of the particles or grains of material
investigated. Only in this experiment it will be possible to exclude
completely the effect of the dispersion of the size of the particles
and determine whether the size dependence of a specific property
is a continuous and smooth function or whether it contains jumps,
inflection points and other special features. Unfortunately, at the
present time it is not possible to carry out such experiments.
In solid-state mechanics, successes have been achieved in
understanding of the nanocrystalline solid as an ensemble of
interacting grain boundary defects. This approach is most useful in
studying bulk nanomaterials. For analysis of the symmetric
properties of the polycrystals as a function of the changes of the
characteristics scale of structural heterogeneity, i.e. as a function
21

Nanocrystalline Materials

of the grain size, the authors of [59] used the gauge field theory
[60, 61]. This theory was developed to describe the structural and
physical properties of materials with defects. According to [59], a
decrease in the grain size is accompanied by a topological
transition from solitary waves of orientation-shear instability, which
are characteristic of the polycrystalline state, to spatial-periodic
structures of defects. The formation of these defects leads to the
transition to the nanocrystalline state. This topological transition in
an ensemble of grain boundary defects is accompanied by a large
change of the characteristics of connectivity and scaling parameters.
The main aim of the present monograph is to discuss the effects
of the nanocrystalline state, detected in the properties of metals and
compounds. The structure and dispersion (the size distribution of the
grains) and, consequently, the properties of nanomaterials depend
on the method of production of these materials. Therefore, in the
second and third chapters of the book we discuss briefly the main
methods of production of nanocrystalline powders and bulk
nanocrystalline materials. It should be mentioned that significant
advances in studying of the nanocrystalline state of solids were
achieved after 1985 as a result of improvement of the available and
development of new methods of production of both disperse and
bulk nanocrystalline materials.
The particle size has the strongest effect on the properties of
nanocrystalline substances. Therefore, the fourth chapter considers
the main methods of determination of the particle size. Special
attention is given to the diffraction method of determining the
particle size. This method is widely available and used.
The fifth chapter is concerned with the specific features of the
physical properties of isolated nanoparticles (nanoclusters) and
nanopowders. These properties are determined by the small particle
size. The methods of production of powdered nanomaterials have
been developed sufficiently and have been known for more than 50
years. A large amount of relatively reliable experimental material
has been collected for the properties of isolated particles (in most
cases, metallic particles), and an efficient theoretical base has been
developed for understanding their properties and structure. It should
be noted that the particles of the nanopowders occupy an
intermediate position between the nanoclusters and bulk solid.
The chapters 6 and 7, analysing the structure and properties of
bulk nanomaterials, contain the most recent experimental data.
Almost all the results described in these chapters have been
obtained after 1988. A great majority of the investigations of bulk
22

Introduction

nanocrystalline substances and materials have been concentrated on


several problems. One of this problems is the microstructure of the
bulk nanomaterials and its stability, the state and relaxation of grain
boundaries. The microstructure is studied by different electron
microscopy, diffraction and spectroscopic techniques. These
investigations are closed to the works on the study of the structure
of bulk nanomaterials by indirect techniques (investigation of phonon
spectra, calorimetry, measurement of the temperature dependence
of microhardness, elasticity moduli, and electrokinetic properties).
It is expected that the bulk nanomaterials will be used most widely
as constructional and functional materials in new technologies and
also as magnetic materials. Consequently, chapter 7 pays special
attention to the mechanical and magnetic properties of bulk
nanomaterials. Discussion of the structure and properties of isolated
nanoparticles and bulk nanomaterials should lead to united views on
the current state of the investigations of this special state of matter
and find common and specific features of isolated nanoparticles and
bulk nanomaterials.
References
1.
2.

3.
4.
5.

6.
7.
8.

9.
10.
11.

H. Gleiter. Materials with ultrafine microstructure: retrospectives and


perspectives. Nanostruct. Mater. 1, 1-19 (1992)
R. Birringer, H. Gleiter. Nanocrystalline materials. In: Encyclopedia of
Material Science and Engineering. Suppl. Vol.1. Ed. R. W. Cahn (Pergamon
Press, Oxford 1988) pp.339-349
R. W. Siegel. Cluster assembled nanophase materials. Ann. Rev. Mater. Sci.
21, 559-578 (1991)
R. W. Siegel. Nanostructured materials mind over matter. Nanostruct. Mater.
3, 1-18 (1993)
H.-E. Schaefer. Interfaces and physical properties of nanostructured solids.
In: Mechanical Properties and Deformation Behavior of Materials Having
Ultrafine Microstructure. Eds. M. A. Nastasi, D. M. Parkin, H. Gleiter.
(Kluwer Academic Press, Netherlands, Dordrecht 1993) pp.81-106
R. W. Siegel. What do we really know about the atomic-scale structures of
nanophase materials? J. Phys. Chem. Solids 55, 1097-1106 (1994)
I. D. Morokhov, L. I. Trusov, S. P. Chizhik. Ultra-Dispersed Metallic
Substances (Atomizdat, Moscow 1977) 264 pp. (in Russian)
I. D. Morokhov, V. I. Petinov, L. I. Trusov, V. F. Petrunin. Structure and
properties of small metallic particles. Uspekhi Fiz. Nauk 133, 653-692 (1981)
(in Russian)
I. D. Morokhov, L. I. Trusov, V. N. Lapovok. Physical Phenomena in UltraDispersed Substances (Energoatomizdat, Moscow 1984) 224 pp. (in Russian)
Yu. I. Petrov. Physics of Small Particles (Nauka, Moscow 1982) 360 pp. (in
Russian)
Yu. I. Petrov. Clusters and Small Particles (Nauka, Moscow 1986) 368 pp.
(in Russian)

23

Nanocrystalline Materials
12.
13.
14.
15.

16.

17.
18.
19.
20.
21.
22.
23.
24.
25.

26.

27.
28.
29.

30.
31.
32.
33.

34.

L. N. Larikov. Structure and properties of nanocrystaline metals and alloys.


Metallofizika 14, 3-9 (1992) (in Russian)
L. N. Larikov. Diffusion processes in nanocrystalline materials. Metallofizika
i Noveishie Tekhnologii 17, 3-29 (1995) (in Russian)
L. N. Larikov. Nanocrystalline compounds of metals. Metallofizika i
Noveishie Tekhnologii 17, 56-68 (1995) (in Russian)
R. A. Andrievski. The synthesis and properties of nanocrystalline refractory
compounds. Uspekhi Khimii 63, 431-448 (1994) (in Russan). (Engl. transl.:
Russ. Chem. Reviews 63, 411-428 (1994))
A. I. Gusev. Effects of the nanocrystalline state in solids. Uspekhi Fiz. Nauk
168, 55-83 (1998) (in Russian). (Engl. transl.: Physics - Uspekhi 41, 49-76
(1998))
E. A. Rohlfing, D. M. Cox, A. Kaldor. Production and characterization of
supersonic carbon cluster beams. J. Chem. Phys. 81, 3322-3330 (1984)
H. W. Kroto, J. R. Heath, S. C. OBrien, R. F. Curl, R. E. Smalley. C 60 :
buckminsterfullerene. Nature 318, 162-163 (1985)
J. W. Mintmire, B. I. Dunlap, C. T. White. Are fullerene tubules metallic?
Phys. Rev. Lett. 68, 631-634 (1992)
S. Iijima. Helical microtubules of graphitic carbon. Nature 354, 56-58 (1991)
S. Iijima, P. M. Ajayan, T. Ichihashi. Growth model for carbon nanotubes.
Phys. Rev. Lett. 69, 3100-3103 (1992)
P. Calvert. Strength in disunity. Nature 357, 365-366 (1992)
S. Iijima. Carbon nanotubes. MRS Bulletin 19, 43-49 (1994)
M. S. Dresselhaus. Future directions in carbon science. Ann. Rev. Mater. Sci.
27, 1-34 (1997)
K. Sohlberg, R. E. Tuzun, B. Sumpter, D. W. Noid. Application of rigidbody
dynamics and semiclassical mechanics to molecular bearings. Nanotechnology
8, 103-111 (1997)
L. Delzeit, C. V. Nguyen, R. M. Stevens, J. Han, M. Meyyappan. Growth
of carbon nanotubes by thermal and plasma chemical vapour deposition
processes and applications in microscopy. Nanotechnology 13, 280-284
(2002)
M. Menon, D. Srivastava. Carbon nanotube T junctions: Nanoscale metalsemiconductor-metal contact devices. Phys. Rev. Lett. 79, 4453-4456 (1997)
J. Li, C. Papadopoulos, J. Xu. Growing Y-junction carbon nanotubes. Nature
402, 253-254 (1999)
C. Papadopoulos, A. Rakitin, J. Li, A. S. Vedeneev, J. M. Xu. Electronic
transport in Y-junction carbon nanotubes. Phys. Rev. Lett. 85, 3476-3479
(2000)
B. C. Satishkumar, P. J. Thomas, A. Govindaraj, C. N. R. Rao. Y-junction
carbon nanotube. Appl. Phys. Lett. 77, 2530-2532 (2000)
B. Q. Wei, R. Vajtai, Y. Jung, J. Ward, Y. Zhang, G. Ramanath, P. M. Ajayan.
Organized assembly of carbon nanotubes. Nature 416, 495-496 (2002)
N. G. Chopra, R. J. Luyken, K. Cherrey, V. H. Crespi, M. L. Cohen, S. G.
Louie, A. Zettl. Boron nitride nanotubes. Science 269, 966-967 (1995)
A. Loiseau, F. Willaime, N. Demoncy, G. Hug, H. Pascard. Boron nitride
nanotubes with reduced numbers of layers synthesized by arc discharge.
Phys. Rev. Lett. 76, 4737-4740 (1996)
A. K. Cheetham, H. Terrones, M. Terrones, R. Castillo, J. P. Hare, H. W.
Kroto, D. R. M. Walton, K. Prassides, S. Ramos, W. K. Hsu, J. P. Zhang.
Metal particle catalysed production of nanoscale BN structures. Chem. Phys.
Lett. 259, 568-573 (1996)
24

Introduction
35.

36.

37.
38.

39.
40.
41.
42.

43.
44.

45.

46.
47.

48.

49.
50.
51.

52.
53.
54.

A. I. Kharlamov, S. V. Loichenko, N. V. Kirillova, V. V. Fomenko.


Heterogeneous synthesis of silicon carbide filaments. Teoret. i Eksper.
Khimiya 38, 49-53 (2002) (in Russian)
A. I. Kharlamov, N. V. Kirillova. Gas-phase reactions of forming of
nanothread-like silicon carbide from powdery silicon and carbon. Teoret. i
Eksper. Khimiya 38, 54-58 (2002) (in Russian)
A. I. Kharlamov, N. V. Kirillova, S. V. Kaverina. Hollow silicon carbide
nanostructures. Teoret. i Eksper. Khimiya 38, 232-237 (2002) (in Russian)
Science and Applications of Nanotubes. Eds. D. Tomanek, R. J. Enbody.
(Kluwer Academic Publishers: New York Dordrecht Moscow 2002) 398
pp .
M. Che, C. O. Bennet. The influence of particle size on the catalytic
properties of supported metals. Advances in Catalysis 36, 55-172 (1989)
V. Ponec. Catalysis by alloys in hydrocarbon reactions. Advances in Catalysis
32, 149-214 (1983)
Z. Karpinski. Catalysis by supported, unsupported, and electrondeficient
palladium. Advances in Catalysis 37, 45-100 (1990)
A. D. Pomogailo. Hybrid polymer-inorganic nanocomposites. Uspekhi Khimii
69, 60-89 (2000) (in Russian). (Engl. Transl.: Russ. Chem. Reviews 69, 5380 (2000))
A. D. Pomogailo, A. S. Rosenberg, U. E. Uflyand. Nanoscale Metal Particles
in Polymers (Khimiya, Moscow 2000) 672 pp. (in Russian)
J. S. Kilby. Turning potential into realities: The invention of the integrated
circuit (Nobel Lecture). Uspekhi Fiz. Nauk 172, 1103-1109 (2002) (in
Russian)
D. L. Klein, R. Roth, A. K. L. Lim, A. P. Alivisatos, P. L. McEuen. A
single-electron transistor made from a cadmium selenide nanocrystal. Nature
389, 699-701 (1997)
H. Kroemer. Nobel Lecture: Quasielectric fields and band offsets: teaching
electrons. Rev. Modern Phys. 73, 783-793 (2001)
R. Dingle, W. Wiegmann, C. H. Henry. Quantum states of confined carriers
in very thin Al x Ga 1x As - GaAs - Al xGa 1x As heterostructures. Phys. Rev.
Lett. 33, 827-830 (1974)
Zh. I. Alferov. Nobel Lecture: The double heterostructure: concept and its
applications in physics, electronics and technology. Rev. Modern Phys. 73,
767-782 (2001)
A. P. Alivisatos. Semiconductor clusters, nanocrystals and quantum dots.
Science 271, 933-937 (1996)
K. Morgenstern. Dynamische Mikroskopie von Nanostrukturen. Physik
Journal 1, 95-98 (2002)
Yu. P. Pershin, E. N. Zubarev, V. V. Kondratenko, O. V. Poltseva, A. G.
Ponomarenko, V. A. Sevryukova, J. Verhoeven. Features of formation of Ni/
C multilayer x-ray mirrors manufactured by electron-beam vaporization
completed with an ion-beam etching. Metallofizika i Noveishie Tekhnologii
24, 795-814 (2002) (in Russian)
J. J. Hauser, H. C. Theuerer, N. R. Werthamer. Proximity effects between
superconducting and magnetic films. Phys. Rev. 142, 118-126 (1966)
B. Y. Jin, J. B. Ketterson. Artificial metallic superlattices. Advances in
Physics 38, 189-366 (1989)
A. I. Buzdin, A. V. Vedyayev, N. V. Ryshanova. Spin-orientation-dependent
superconductivity in F/S/F structures. Europhysics Lett. 48, 686-691 (1999)

25

Nanocrystalline Materials
55.
56.

57.

58.

59.

60.

61.

L. R. Tagirov. Low-field superconducting spin switch based on a superconductor/ferromagnet multilayer. Phys. Rev. Lett. 83, 2058-2061 (1999)
Yu. A. Izyumov, Yu. N. Proshin, M. G. Khusainov. Competition between
superconductivity and magnetism in ferromagnet/superconductor
heterostructures. Uspekhi Fiz. Nauk 172, 113-154 (2002) (in Russian). (Engl.
transl.: Physics - Uspekhi 45, 109-148 (2002))
Th. Mhge, N. N. Garif yanov, Yu. V. Goryunov, G. G. Khaliullin, L. R.
Tagirov, K. Westerholt, I. A. Garifullin, H. Zabel. Possible origin for
oscillatory superconducting transition temperature in superconductor/
ferromagnet multilayers. Phys. Rev. Lett. 77, 1857-1860 (1996)
A. S. Borukhovich. Quantum tunneling multilayers and heterostructures with
ferromagnetic semiconductors? Uspekhi Fiz. Nauk 169, 737-751 (1999) (in
Russian). (Engl. transl.: Physics - Uspekhi 42, 653-667 (1999))
O. B. Naimark. Nanocrystalline state as a topological transition in an
ensemble of grain-boundary defects. Fiz. Metall. Metalloved. 84, 5-21 (1997)
(in Russian). (Engl. Transl.: Phys. Metal. Metallogr. 84, 327-337 (1997))
E. Krner. On gauge theory in defect mechanics trends in application of pure
mathematics to mechanics. In: Lecture Notes in Physics. Eds. E. Krner and
K. Kinchgassner (Springer, Heidelberg 1986) pp.281-296
A. Kadic, D. G. B. Edelen. A Gauge Theory of Dislocations and Disclinations
(Springer, Berlin 1983) 186 pp.

26

Synthesis of Nanocrystalline Powders

+D=FJAH

2. Synthesis of Nanocrystalline Powders


2.1 GAS PHASE SYNTHESIS
Isolated nanoparticles are usually produced by evaporation of metal,
alloy or semiconductor at a controlled temperature in the
atmosphere of a low-pressure inert gas with subsequent
condensation of the vapour in the vicinity or on the cold surface.
This is the simplest method of producing nanocrystalline powders.
In contrast to vacuum evaporation, the atoms of the substance,
evaporated in a rarefied inert atmosphere, lose their kinetic energy
more rapidly as a result of collisions with gas atoms and form
segregations (clusters).
The first studies in this topic were carried out in 1912 [1,2]:
examination of evaporation of Zn, Cd, Se and As in vacuum, and
also in hydrogen, nitrogen and CO 2 showed that the size of the
produced particles depends on the pressure and atomic weight of
the gas. The authors of [3] evaporated gold from a heated tungsten
filament at a nitrogen pressure of 0.3 mm Hg (40 Pa), and produced
in the condensates spherical particles with a diameter of 1.510 nm
(the mean diameter approximately 4 nm). They found that the
particle size depends on the gas pressure and, to a lesser degree,
on the rate of evaporation. The particle size was determined by the
high-resolution electron microscopy. The condensation of vapours
of aluminium in H 2, He and Ar at different gas pressures made it
possible to produce particles with a size of 20100 nm [4]. Later,
the method of combined condensation of metal vapours in Ar and
He was used to produce AuCu and FeCu highly dispersed alloys,
formed by spherical particles with a diameter of 1650 nm [5, 6].

27

Nanocrystalline Materials

A variant of the condensation of metal vapours in a gas atmosphere


is the method of dispersion of a metal by means of an electric arc
in a liquid with subsequent condensation of metallic vapours in liquid
vapours, proposed as early as in the 19 th century [7]; later, this
method was improved by the authors of [810]. The first extensive
review [11], concerned with detailed discussion of the condensation
method and the formation of highly dispersed metal particles by
condensation of metallic vapours, was published in 1969. Several
theoretical special features of condensation in supersaturated
vapours, which takes place by means of the formation and growth
of nuclei (clusters), were discussed in a review in [12].
The nanocrystalline particles with a size of 20 nm, produced
by evaporation and condensation, are spherical, and large particles
may be faceted. The size distribution of nanocrystals is
logarithmico-normal and is described by the function

F ( D) = ( 2 ln g )1 exp ln D ln < Dg >

2
) /(2ln 2 g )

(2.1)

where D is the particle diameter; <D g > is the mean diameter; g

2
is dispersion; ln g = ni ( ln Di ln < Dg > ) / ni

1/ 2

. Analysis

shows that the majority of distributions of nanoparticles of metals,


produced by the evaporation and condensation method, are described
by equation (2.1) with the values g = 1.4 0.2. Isolated
nanocrystals contain no dislocations, but disclinations, which are
energetically more advantageous in very small crystals than
dislocations, may form [13].
The systems using the principle of evaporation and condensation,
differ in the method of input of evaporated material; the method of
supplying energy for evaporation; the working medium; setup of the
condensation process; the system for collecting the produced
powder.
Evaporation of a metal may take place from a crucible or the metal
may be fed into the zone of heating and evaporation in the form of
wire, injected metallic powder or a liquid jet. The metal may also be
dispersed with a beam of argon ions. Energy may be supplied directly
by heating, the passage of electric current through a wire, electric arc
discharge in plasma, induction heating with high- and superhigh
frequency currents, laser radiation, and electron-beam heating.
28

Synthesis of Nanocrystalline Powders

Evaporation and condensation may take place in vacuum, in a


stationary inert gas, or in a gas flow, including in a plasma jet.
Condensation of the vapourgas mixture with a temperature of
500010000 K may take place during its travel into the chamber with
a large section and the volume filled with a cold inert gas; cooling
takes place both as a result of expansion and contact with the cold
inert atmosphere. There are systems in which two jets travel coaxially
into the condensation chamber: the vapourgas mixture is supplied
along the axis, and a circumferential jet of a cold inert gas travels
along its periphery. As a result of turbulent mixing, the temperature
of metal vapours decreases, supersaturation increases and rapid
condensation takes place. Favourable conditions for the condensation
of metallic vapours are generated in adiabatic expansion in a Laval
nozzle, when rapid expansion results in the formation of a steep
temperature gradient and almost instantaneous vapour condensation
takes place.
An independent task is the collection of the nanocrystalline powder
produced by condensation, because the individual particles of this
powder are so small that they are in constant Brownian motion and
remain suspended in the gas, not settling under the effect of the
forces of gravity. The produced powders are collected using special
filters and centrifugal deposition; in some cases, the particles are
trapped by a liquid film.
The main relationships of the formation of nanocrystalline
particles by the method of evaporation and condensation are as
follows, [11, 14]:
1. The nanoparticles form during cooling of the vapours in the condensation zone. The size of this zone increases with a decrease of the gas
pressure; the internal boundary of the condensation zone is in the vicinity of the evaporator, and its external boundary may, with a decrease of
gas pressure, extend outside the limits of the reaction vessel; at a pressure of several hundreds of Pa, the outer boundary of the condensation
zone is situated inside the reaction chamber with a diameter of 0.1 m
and convective gas flows play a significant role in the condensation
process;
2. When the gas pressure is increased to several hundreds of Pa, the mean
particle size initially rapidly increases and then slowly approaches the
limiting value in the pressure range greater than 2500 Pa;
3. At the same gas pressure, the transition from helium to xenon, i.e. from
a less dense inert gas to an inert gas with a higher density, is accompanied by a large increase in the particle size.
29

Nanocrystalline Materials

Depending on the evaporation conditions of the metal (gas pressure, the position and temperature of the substrate), its condensation
may take place either in the volume or on the surface of the
reaction chamber. Volume condensates are characterised by the
presence of spherical particles, whereas the particles of the surface
condensate are faceted. In the same evaporation and condensation
conditions, metals with high melting points form smaller particles.
If the gas pressure is lower than approximately 50 Pa, spherical
particles of metal with a mean diameter of D < 30 nm settle on the
walls of relatively large reaction chambers (diameter greater than
0.25 m). When the pressure is increased by several hundreds of Pa,
the formation of highly dispersed metallic particles is completed in
convective flows of the gas in the vicinity of the evaporator.
Gas-phase synthesis can be used to produce particles with a size
from 2 nm to several hundreds of nanometers. Smaller particles of
a controlled size are produced by means of the mass distribution
of clusters in a time-of-flight mass spectrometer. For example, metal
vapours are passed through a cell with helium with a pressure of
~10001500 Pa, and are then transferred into a vacuum chamber
(~10 5 Pa), where the mass of the cluster is stabilised during the
time of flight of a specific distance in the mass spectrometer. This
method is used to produce clusters of antimony, bismuth and lead,
containing 650, 270 and 400 atoms, respectively; the temperature
of gaseous helium in the case of Sb and Bi vapours is
80 K, and in the case of Pb vapours 280 K [15].
Recently, the gas-phase synthesis of nanoparticles has been
developed extensively as a result of the application of different
methods of heating the evaporated substance.
Highly dispersed deposits of silver and copper on glass were
produced by evaporation of metals in an inert atmosphere at a
pressure of 0.010.13 Pa [16]. The same method was used to
produce clusters of Li n , containing from fifteen to two lithium
atoms [17]; evaporation of lithium in high vacuum is accompanied
by the formation of only separate atoms of lithium, and clusters
form only in the atmosphere of a rarefied inert gas. Nanocrystalline
powders of the oxides Al 2 O 3 , ZrO 2 and Y 2 O 3 were produced by
evaporating oxide targets in a helium atmosphere [18], magnetron
sputtering of zirconium in a mixture of argon and oxygen [19], and
by the controlled evaporation of yttrium nanocrystals [20]. To
produce highly dispersed powders of transition metal nitrides, the
electron-beam heating of targets of appropriate metals, with
evaporation carried out in the atmosphere of nitrogen or ammonia
30

Synthesis of Nanocrystalline Powders

at a pressure of 130 Pa, is used [21].


Nanocrystalline powders have also been produced by plasma,
laser and arc heating methods. For example, the authors of [22, 23]
produced nanoparticles of carbides, oxides and nitrides by the
pulsed laser heating of metals in a rarefied atmosphere of methane
(in the case of carbides), oxygen (in the case of oxides), nitrogen
or ammonia (in the case of nitrides). The pulsed laser evaporation
of metals in the atmosphere of an inert gas (He or Ar) or a reagent
gas (O 2 , N 2 , NH 3 , CH 4 ) makes it possible to produce mixtures of
nanocrystalline oxides of different metals, oxidenitride or carbide
nitride mixtures. The composition and size of nanoparticles may be
controlled by the variation of the pressure and composition of the
atmosphere (inert gas or reagent gas), the power of the laser pulse,
the temperature gradient between the evaporated target and the
surface on which condensation takes place.
The method of condensation of vapours in an inert gas is used
most frequently in scientific investigations for producing small
amounts of nanopowders. The powders synthesised are not
agglomerated efficiently and are sintered at a relatively low
temperature.
The authors of [24] modified the condensation method for
producing ceramic nanopowders from organometallic precursors. In
the setup used in [24] (Fig. 2.1), the evaporator was a tubular
reactor in which the precursor was mixed with the carrier inert gas
Chamber pressure
(1~50 mbar)
Control
valve
To pump
Rotated
cold
cylinder
Carrier
gas
Heated tubular reactor
Gas
Precursor
source
Mass flow
Needle
controller
valve

Particles

Scraper

Funnel

Collector

Fig. 2.1. Schematic of the chemical vapour condensation (CVC) processing apparatus
for preparation of ceramic nanostructured powders from organometallic precursors
[24].

31

Nanocrystalline Materials

and dissociated. The resultant continuous flow of clusters or


nanoparticles travelled from the reactor in the working chamber and
condensed on a cold rotating cylinder. Successful realisation of the
process is ensured by a low concentration of precursor in inert gas,
rapid expansion and cooling of the gas which exits from the reactor
into the working medium, and a low pressure in the working
chamber. The characteristics (particle size distribution, agglomeration capacity, sintering temperature) of the nanopowders
produced by this method do not differ from those of the
nanopowders synthesised by the standard evaporation and
condensation method.
The properties of isolated nanocrystalline particles are greatly
determined by the contribution of the surface layer. For a spherical
particle with a diameter of D and the thickness of the surface layer
, the fraction of the surface layer in the volume of the particle
3
3
3
is V / V = 6 D 6 ( D 2) / 6 D 6 / D . At a thickness of the
surface layer equal to 34 atomic monolayers (0.51.5 nm), and
the mean size of the nanocrystal of 1020 nm, the surface layer
accounts for up to 50% of the substance. However, the highly
developed surface of nanocrystalline particles greatly increases their
reactivity and, in turn, this greatly complicates study of these
particles.

2.2 PLASMA CHEMICAL TECHNIQUE


One of the most widely used chemical methods of producing highly
dispersed powders of nitride, carbides, borides and oxides is plasma
chemical synthesis [2531].
The main conditions of producing highly dispersed powders by
this method is the occurrence of a reaction away from equilibrium
and the higher rate of formation of nuclei of a new phase at a low
growth rate of this phase. In the real conditions of plasma chemical
synthesis, the formation of nanoparticles can be carried efficiently
by increasing the cooling rate of the plasma flow in which
condensation from the gas phase takes place; this decreases the size
of particles and also suppresses the growth of particles by their
coalescence during collisions.
Plasma chemical synthesis is carried out with the use of lowtemperature (40008000 K) nitrogen, ammonia, hydrocarbon plasma,
argon plasma of arc, glow, high frequency or microwave
discharges. Starting materials are represented by elements, halides
32

Synthesis of Nanocrystalline Powders

and other compounds. The characteristics of the produced powders


depend on the starting materials used, synthesis technology and the
type of reactor. The particles of plasma chemical powders are
single crystals and their size varies from 10 to 100200 nm or
larger. Plasma chemical synthesis ensures high rates of formation
and condensation of the compound and is characterised by relatively
high productivity. The main disadvantages of plasma chemical
synthesis are the wide size distributions of particles and,
consequently, the presence of relatively large (up to 15 m)
particles, i. e. the low selectivity of the process. The other
disadvantage is a high content of the impurities in the powder. Up
to now, the plasma chemical technique has been used to produce
highly dispersed powders of nitrides of Ti, Zr, Hf, V, Nb, Ta, B, Al
and Si, carbides of Ta, Nb, Ti, W, B and Si, and oxides of Mg, Y
and Al [1529, 3237]. The plasma chemical technique is used most
widely for the synthesis of groups IV and V transition metal
nitrides; analysis of the structure and properties of ultrafine (with
a mean particle size smaller than 50 nm) nitride powders can be
found in a monograph in [38, section 1.4].
The plasma temperature, reaching up to 10000 K, determines the
presence in the plasma of ions, electrons, radicals and neutral
particles which are in the excited state. The presence of these
particles leads to high rates of interaction and a short time of
reactions (up to 10 3 10 6 seconds). High temperature ensures the
transfer of almost all starting substances into the gaseous state
followed by their interaction and condensation of the reaction
products.
Plasma chemical synthesis includes several stages. The first
stage is characterised by the formation of active particles in arc,
high frequency and microwave plasma reactors. The highest power
and coefficient of efficiency are shown by arc plasma reactors, but
materials produced are contaminated with products of electrode
erosion. High frequency and microwave plasma reactors do not
have this shortcoming. As a result of quenching, the next stage
results in the generation of reaction products. The selection of the
place and the rate of quenching makes it possible to produce
powders with the required composition, shape and size of the
particles.
The powders produced by the plasma chemical technique are
characterised by regular shapes and a particle size of 10100 nm
or larger.
The plasma chemical powders of carbides of metals, boron and
33

Nanocrystalline Materials

silicon are usually produced by the reaction of chlorides of the


appropriate elements with hydrogen and methane or other
hydrocarbons in the argon high-frequency or arc plasma; nitrides
are produced by reaction of chlorides with ammonia or a mixture
of nitrogen and hydrogen in low-temperature microwave plasma.
Plasma chemical synthesis may also be used for producing
multicomponent submicrocrystalline powders, representing mixtures
of carbides and nitrides, nitrides and borides, nitrides of different
elements, etc.
The synthesis of oxides in the plasma of electric arc discharges
is carried out by evaporation of metal with the subsequent oxidation
of vapours or oxidation of particles of a metal in an oxygencontaining plasma. The authors of [39] described the plasma
chemical synthesis of nanoparticles of aluminium oxide with a mean
size of 1030 nm. This study shows that the formation of
nanopowders of aluminium oxide with a smallest particle size is
achieved in the reaction of metal vapours with atmospheric oxygen
when an air is blowed intensively. It leads to a rapid decrease of
temperature. Rapid cooling not only decelerates the particle growth
but also increases the rate of formation of nuclei of the condensed
phase. Plasma chemical synthesis with the oxidation of aluminium
particles in the flow of oxygen-containing plasma leads to the
formation of larger oxide particles in comparison with the oxidation
of metal vapours produced in advance.
The plasma chemical synthesis is quite close to gas phase
synthesis using laser heating of a reacting gas mixture [4044].
Laser heating ensures controlled homogeneous nucleation and
prevents contamination. The size of nanocrystalline particles
decreases with increasing intensity (the power related to the unit
area) of laser radiation as a result of an increase of temperature
and the heating rate of gases-reagents. The authors [41] used this
method to produce silicon nitride Si 3 N 4 with a particle size of 10
20 nm from a gas mixture of silane SiH 4 and ammonia NH 3 .
The plasma chemical method is also used to produce metal
powders. For example, submicrocrystalline copper powders with a
particle size smaller than 100 nm and a comparatively narrow size
distribution of particles are produced by reduction of copper
chloride by hydrogen in argon electric arc plasma with a
temperature up to 1800 K.
Gas phase synthesis using laser radiation for producing the
plasma in which the chemical reaction takes place, is an efficient
method of producing molecular clusters. Molecular clusters are a
34

Synthesis of Nanocrystalline Powders

new structural modification of matter. We shall discuss in more


detail the achievements in the field of plasma chemical gas phase
synthesis and the new possibilities of producing previously unknown
polymorphous modifications of substances with the nanometer size
of structural elements.
The molecular clusters occupy a completely special position in
the group of substances with a nanostructure. The best known of
these are fullerenes [4547], i.e. a new allotropic modification of
carbon in addition to graphite and diamond.
Already in November 1966, the British journal New Scientist
published playful notes by D. E. H. Jones on the possibility of
producing solid materials with a low density (considerably lower
than the density of water). This material should consist of hollow
spherical molecules whose shell is constructed from graphite sheets.
A network of hexagonal rings C 6 should, for stability, also include
five-member rings. At that time, nobody noticed that a similar
design had already been proposed in 1951 by the well-known
American architect R. B. Fuller who patented the spherical
construction which was called the geodesic cupola. This
construction of the cupola was used, for example, in the US pavilion
at the Expo-67 World Exhibition in Montreal.
The fullerenes are produced by electric arc sputtering of graphite
in a helium atmosphere; gas pressure is 1.3310 4 Pa. Combustion
of the arc results in the formation of carbon black which condenses
on the cold surface. The collected carbon black is processed in
burning toluene or benzene. After evaporating a solution, a black
condensate appears which consists of approximately 1015% of a
mixture of fullerenes C 60 and C 70 . In addition to the electric arc,
fullerenes can be produced by electron beam evaporation and laser
heating.
The central position among the fullerenes belongs to the C 60
molecule with the highest symmetry and, consequently the highest
stability. As regards the shape, the molecule of fullerene C 60
resembles a soccer-ball and has the structure of a regular truncated
icosahedron (Fig. 2.2). In the molecule of fullerene C 60 the carbon
atoms form a closed hollow spherical surface consisting of 5- and
6-member rings, and each atom has a coordination number equal to
three and is situated in the tops of two hexagons and one pentagon.
The diameter of the molecule of fullerene C 60 is 0.72-0.75 nm.
Crystallisation of C 60 from a solution or gas phase is accompanied
by the formation of molecular crystals with an fcc lattice; the lattice
constant is 1.417 nm. Fullerene in the solid state is referred to as
35

Nanocrystalline Materials

Fig. 2.2. The structure of the most important fullerenes C 60 and C 70 . The C 60
molecule is built like a soccer-ball and its cage has a diameter of about 0.7 nm.
All fullerenes exhibit hexagonal and pentagonal rings of carbon atoms.

fullerite. Higher stability is also typical of fullerene C 70 having the


form of a closed spheroid. The fullerenes may be regarded as a
spherical form of graphite because mechanisms of interatomic
bonding in the fullerene and in bulk graphite are very similar. It is
interesting to note that the high stability of fullerene C 60 was
theoretically predicted already at the beginning of the 70s in
calculations of potentially possible frame structures constructed from
carbon atoms [48, 49].
The properties of fullerenes are very unusual. For example,
crystalline fullerenes represent semiconductors and are
characterised by photoconductivity in optical radiation. However,
crystals C 60 , alloyed with atoms of alkaline metals, have metallic
conductivity and change to the superconducting state at 30 K and
higher. Transformation of crystalline fullerene to diamond takes
place at room temperature at a pressure of 20 GPa. If a fullerene
is heated to 1500 K, a pressure of 7 GPa is sufficient for the
transition to diamond. Similar transformation of graphite to diamond
requires a temperature of 900 K and a pressure of 3050 GPa. The
solutions of fullerenes are characterised by non-linear optical
properties, i. e. a rapid decrease of the transparency of the solution
when exceeding some critical value of the intensity of optical
radiation.
Recently, ferromagnetic properties are found for the polymerised
form of fullerene C 60 at room temperature [50]. Investigation shows
that polymerised fullerene Rh-C 60 with a rhombohedral structure
has a Curie temperature of 500 K and its hysteresis curve is
typical of ferromagnetics. At heating and depolymerisation, the
36

Synthesis of Nanocrystalline Powders

specimen of Rh-C 60 lost its ferromagnetic properties.


At the beginning of 2001, a group of scientists [51] found a new
fullerene-like form C 48 N 12 in which in comparison with usual
fullerene C 60 a fifth of carbon atoms is substituted by nitrogen
(Fig. 2.3). If in the fullerene crystals the molecules C 60 are
connected by weak Van-der-Waals forces, then the presence of
nitrogen atoms results in the formation of strong covalent bonds.
For this reason, the fullerene-like crystalline material C 48 N 12 is
characterised by a unique combination of strength and elasticity.
The discovery of fullerenes C n (n = 6090) and subsequent
investigations show that clusters C n , containing less than 60 carbon
atoms, have low stability. One of the methods of stabilising carbon
fullerenes C 28 with a small number of atoms is the synthesis of
endohedral complexes M@C 28 in which the atom of the doping
element is introduced into the carbon sphere. A similar endohedral
complex Ti@C 28 has been synthesised in particular with titanium
[52].
Stabilisation of the unstable fullerene C 28 by means of
intercalation into its volume of atoms of non-metallic 2p-elements
(B, C, N and O) and metallic 3d-elements (Sc, Ti, V, Cr, Fe and
Cu) has been examined theoretically in [53]. When evaluating the
possibility of formation of endohedral complexes M@C 28 with 3dmetals, it is important to take into account the geometrical, chemical
and kinetic factors. For fullerene C 28 , the limiting value of the radius
of the metallic ion which can be intercalated in the internal cavity of
the fullerene is 0.090.10 nm [54]. Therefore, all 3d-metals satisfy the
geometrical criteria. The chemical factor is favourable if as a result

Fig. 2.3. The structure of fullerene C 48 N 12 [51]: the position of the symmetry
axis C 6 is shown by solid line.
37

Nanocrystalline Materials

of intercalation the transfer of electronic density enhances the binding


molecular orbitals. The kinetic factor takes into account the mechanism
of formation of endohedral complexes. They are produced by rolling
the graphen monolayer around the atom of the metal adsorbed on the
surface of graphite. This process takes place in the interaction of metal
when the surface of graphite strengthens the MC bond and weakens
the CC bond (especially interlayer bonds). The results published in
[53] show that intercalation of the atoms of metal M in the volume
of the fullerene C 28 is accompanied (1) by the transfer of an electron
charge from atom M to the atoms of the fullerene shell, (2) by the
change of the population of overlapping atomic orbitals of carbon, (3)
by the formation of a chemical bond of atom M with the carbon atoms
and (4) by a general variation of the electron energy spectrum.
Analysis of M@C 28 complexes with different metals M shows that
from the viewpoint of the chemical and kinetic factors the endohedral
complex Ti@C28 is most stable [53]. In this complex, the titanium atom
is situated in the centre of the polyhedron C 28 .
According to [55], the endohedral complex Ti@C 28 can be
considered as a possible compound of the type of molecular cluster
in the TiC system.
Fullerenes, as molecular clusters, have been examined in
thousands of original papers, tens of reviews and monographs.
Therefore, in this book, they are only mentioned in connection with
the synthesis of a new class of molecular clusters, having the
composition M 8 C 12 , where M is the metal atom.
After discovery of molecular clusters of carbon and initial
observations of the fullerene molecule C 60 [4547], and after
intensive and greatly differing investigations of synthesis, structure
and properties of fullerenes (see, for example [5658]), special
attention has been given to producing molecular clusters of other
substances. By analogy with fullerenes, it was expected that these
molecular clusters should have unique physical and chemical
properties, differing from the properties of the available
polymorphous modifications of the same substance.
The search for new molecular clusters was crowned with
success by the discovery in 1992 [59] of a new unusually stable
charged cluster Ti 8 C 1+2 corresponding to the molecule with
stoichiometric composition Ti 8C 12 in the form of a slightly distorted
pentagondodecahedron (Fig. 2.4). In the dodecahedral molecule,
all atoms are distributed on a sphere. The sphere consists of 12
pentagons including 2 atoms of titanium and 3 carbon atoms. In this
molecule, all atoms of titanium and carbon have the same (as in
38

Synthesis of Nanocrystalline Powders

%K

7L&
7L 

%G
7L 

7L 

7L 

7L 
7L 

7L

7L 
7L 

&

Fig. 2.4. Dodecahedral structure of the molecular cluster Ti 8 C 12 with the symmetries
T h and T d taking into account different length of TiC and CC bonds.

fullerene C 60 ) co-ordination equal to 3, occupy the same positions


and are distributed at the tops of the dodecahedron in such a
manner that titanium is bonded only with carbon, and 6 dimers C 12
alternate with 8 atoms of titanium. The dodecahedral structure of
Ti 8 C 12 can be regarded as a cube formed by 8 titanium atoms,
where each face is bonded with dimer C 2 . The distance between
the nearest atoms of carbon is 0.151 nm, the distance TiC is 0.196
nm, and the distance TiTi is 0.305 nm. The point symmetry group
T h of such a structure includes 24 elements of symmetry (rotations
and reflections). Because of high symmetry, the ideal molecule of
metallocarbohedren should be highly stable.
Another possible structure of the Ti 8 C 12 cluster has the point
symmetry group T d [60] (Fig. 2.4). In this configuration, the titanium
atoms occupy positions of two types, and the nodes, relating to the
positions of each type, form a tetrahedron. The smaller tetrahedron
is rotated by 90 in relation to the larger one. The difference in the
positions of the titanium atoms is due to the different position in
relation to dimers C 2 . In fact, 6 dimers C 2 are parallel to the edges
of the larger tetrahedron made of atoms of Ti(1) and are normal
to the edges of the smaller tetrahedron formed by 4 atoms Ti(2).
The atoms Ti(1) are linked with the 3 nearest carbon atoms, and
Ti(2) atoms with 6 carbon atoms. The Ti(1)C distance is 0.193
nm, Ti(2)C is 0.219 nm, Ti(1)Ti(2) is 0.286 nm and the distance
Ti(2)Ti(2) is 0.290 nm [56]. The linear size of the Ti 8 C 12 cluster
39

Nanocrystalline Materials

is approximately 0.5 nm.


The question of which of the two structures (with symmetry T h
or T d) is realised has not as yet been solved.
The Ti 8 C 12 clusters were produced by the plasma chemical gasphase synthesis. The inert gas was represented by helium, reagents
were hydrocarbons (methane, ethylene, acetylene, propylene, and
benzene) and titanium vapours, the pressure of the gas mixture in
the reactor was 93 Pa (0.7 mm Hg). A rotating titanium rod was
evaporated and the ionised beam of metal vapours was produced
by focused radiation of an Nd-laser with a wavelength of 532 nm.
Neutral and ionised clusters were separated from the reaction
products and analysed in a mass spectrometer. The mass spectra
of the reaction products contained a sharp peak corresponding to
the Ti 8C 12 molecule. In addition to the neutral molecules, stable ions
Ti 8 C 12 form in the mixture of ionised gases.
The author of [59] assumed that the Ti 8 C 12 cluster is a member
of a new class of molecular clusters and referred to this cluster as
metallocarbohedrene or Met-Car. In metallocarbohedrenes, the
atoms of a transition metal and carbon form a cage-like structure.
In fact, other clusters M 8 C 12 of transition metals such as Zr, Hf,
V [61, 62], Cr, Mo and Fe [63] were soon produced. The
metallocarbohedrenes and methods of producing them are described
in reviews [64, 65].
According to the authors of [59], the high stability of the Ti 8 C 12
cluster is a consequence of a special geometrical and electron
structure, typical of these clusters. Chemical bonds in the Ti 8 C 12
molecule are similar to those existing in carbon fullerenes.
However, in contrast to fullerene C 60 , the ionised or neutral
molecule of the type M 8 C 12 contains only five-member rings. As
regards shape of the surface, the highly stable cluster Ti 8 C 12
corresponds to a hypothetical unstable (and, therefore, not realised
in practice) fullerene C 20 . This comparison already shows that the
complete identity of the chemical bonds in the M 8 C 12 clusters and
in carbon fullerenes is unlikely.
In fact, the calculations of the equilibrium crystall and electron
structure of Ti 8C 12 [66] show that bonds of the titanium atoms with
3 adjacent carbon atoms are greatly differ from those in graphite
or in fullerene C 60 ; in particular, the lengths of the TiC and CC
bonds in Ti 8 C 12 differ by almost a factor of 1.5 and are equal to
3.76a 0 and 2.63a 0 (a 0 = 0.052918 nm is the radius of the first Bohr
orbit), respectively. According to [67], the length of the TiC bond
is approximately 70 % greater than the length of the CC bond. At
40

Synthesis of Nanocrystalline Powders

the same time, the atoms of carbon and titanium are situated at
almost the same distances from the centre of the cluster. This
means that the real dodecahedron Ti 8C 12 is greatly deformed and
distorted. According to [66], the binding states of the Ti 8C 12 cluster
are formed by the combination of d-orbitals of titanium and molecular
orbitals C 2 , and the filled level with the highest energy is situated
between the bonding and anti-bonding states of titanium. It ensures the
stability of the cluster. Identical conclusions according to which the
shape of the M 8C 12 is not ideal but they have the form of a distorted
pentagondodecahedron, were obtained in other theoretical calculations.
Slightly different results were obtained [68] in comparative
examination of the electronic structure of Met-Car Ti 8 C 12 with
symmetry T h and T d. According to [68], the structures of both types
contain the filled level with the highest energy at a sharp peak of
the density of states, formed mainly by C2p- and Ti3d-atom
orbitals. The high chemical stability of the Ti 8 C 12 compound is
determined by the combination of strong Ti3dC2p-interactions
between the titanium atoms and dimers C 2 , on the one hand, and
CC interactions in carbon dimers, on the other hand, in the
structures of both types, Ti 8 C 12 has an open electronic shell so that
it can play the role of a donor and also of the acceptor of electron
density. In calculations in [68], the parameters of the structure and
atomic spacing for symmetry T h were taken from [67] and for
symmetry T d from [60].
The atoms in molecules of metallocarbohedrenes form strong
bonds. For example, the binding energy per 1 atom of the molecule
of Ti 8 C 12 is 6.16.7 eV/atom [66, 67, 69]. For comparison, this
value in the fullerene molecule C 60 is 7.47.6 eV/atom [70, 71], and
in titanium carbide TiC with a cubic structure B1 it is 7.2 eV/atom
[66].
The considerations regarding the geometry and electron structure
of the molecular clusters Ti 8 C 12 , presented in [66, 68], explain
efficiently the peculiarities of reaction behaviour for these clusters
in relation to polar and non-polar substances.
+
Investigations of the interaction between Ti 8 C 12
clusters and polar
molecules of methanol CH 3OH, water H 2O and ammonia NH 3 [72]
show that at room temperature the reaction between them
+
+
Ti8C12
(P) n1 + P Ti8C12
(P) n

(2.2)

takes place through eight consecutive steps of annexation of the


41

Nanocrystalline Materials

polar molecule P. This means that the first solvation shell of the
+
Ti 8 C 12
ion is formed by eight polar molecules. In reactions with
benzene and ethylene, the resultant first solvation shell includes
only four hydrocarbon molecules with -bonds. For example, at
room temperature the Ti 8 C +12 cluster is fully inert in relation to nonpolar molecules of oxygen and methane. According to the authors
of [72], if the clusters Ti 8 C 12 can be held together by
Van-der-Waals forces and form large crystals, like fullerene C 60 ,
then the bulk material Ti 8C 12 will be very stable in air. It could also
+
be mentioned that, regardless of reactivity of Ti 8 C 12
in relation to
many substances, interaction between them takes place only as
association of ligands, without rupture of any chemical bonds in a
cluster. This confirms the high stability of metallocarbohedrenes.
It is very interesting to note that in plasma chemical gas phase
synthesis [59, 61, 62] there was preferential formation of cluster
particles of M 8 C 12 and M mC n (M Ti, Zr, Hf, V) with the ratio M:C
1.52.0, and not of nanoparticles of carbides TiC, ZrC, HfC, VC
with the fcc crystal structure. In identical synthesis, systems TaC and NbC were characterised by the formation, in addition to
clusters Ta mC n and Nb mC n with the compositions similar to M 8 C 12 ,
of small amounts of nanocrystalline particles M m C n with m n,
with a cubic structure. At the same time, conventional plasma
chemical synthesis (without laser heating of plasma) makes it
possible to produce only carbide nanoparticles. Thus, in gas phase
synthesis, two structures cubic and the metallocarbohedrene type
can form in the transition metalcarbon systems. Because each
MC system usually should initially contain clusters (or particles)
of only one structural type, it could be assumed that the selective
formation of a specific structure is determined by its
thermodynamic stability.
However, the author of [73] reported that as a result of
synthesis in the experimental conditions used in the study, the Ti
C and VC systems are characterised by the simultaneous
formation of cubic MC (M 14 C 13 ) and dodecahedral M 8 C 12
structures. The more extensive formation of cubic nanoparticles in
comparison with M 8C 12 took place at a relatively low laser radiation
power. In [73] it is assumed that the formation of
metallocarbohedrenes may take place by photo-dissociation of cubic
nanoparticles, indicating high stability of M 8C 12 clusters.
Taking into account the results in [59, 6163, 73], it is natural
to assume that possible reasons for the preferential formation of
carbide fcc nanoparticles or molecular clusters M 8 C 12 may also be
42

Synthesis of Nanocrystalline Powders

of a kinetic nature. The correct answer to the question of the


reasons for the preferential formation of a specific structure is
important for practice because it makes it possible to produce the
crystalline modification of the nanostructural material which is
required, i.e. carry out in practice the directional synthesis of the
nanomaterial.
To solve this problem, investigations were carried out into the
formation of Nb m C n nanoclusters in the NbC system [74] in
relation to the synthesis conditions (the concentration of a carboncontaining reagent in a gas atmosphere, laser radiation power). To
evaporate a rod of metallic niobium, heat and sustain the plasma,
the authors used the radiation of an Nd-laser with a wavelength of
532 nm. The buffer gas was helium, the total pressure of the gas
mixture was 0.40.65 MPa. The mass spectra of ionised clusters
Nb m C +n were recorded directly from the plasma using a quadrupole
mass spectrometer.
Analysis of the mass spectra shows that the nanoparticles with
a cubic structure and with the ratio Nb:C 1:1 (Nb 14 C 13 ) form at
a relatively low (4%) concentration of methane CH 4 in helium and
a radiation power of 1015 mJ pulse 1 . At a methane concentration
from 8 to 20% and the radiation power not lower than 15 mJ
pulse 1 , dodecahedral particles with the Nb:C ratio close to 1:2 (for
example, Nb 11 C 21 , Nb 13 C 22 ) form preferentially. It is interesting that
an increase in the concentration of hydrocarbons is accompanied
by the growth of clusters: for example, at a methane concentration
of 8%, the largest clusters are Nb 8 C 12 , and a concentration of 12%
the clusters Nb 11 C 21 , Nb 12 C 22 , Nb 13 C 22 and Nb 14 C 25 appear. The
identical growth of clusters is detected in laser gas phase synthesis
in the ZrC system, where clusters Zr 13 C 22 , Zr 14 C 21 and Zr 13 C 23
appear [75]. It indicates the formation of structure representing
doubled dodecahedrals. The authors of [74] concluded that the
metallo-carbohedrenes (in particular, large clusters consisting of two
or more connected dodecahedrals) are formed at a high
concentration of hydrocarbon and a high laser radiation power,
causing dehydrogenation of carbon. Thus, the metallocarbohedrenes
are formed at a higher carbon content in the plasma. A decrease
in the hydrocarbon concentration or in radiation power decreases
the carbon content in the plasma. As a result, a relative decrease
in carbon content leads to the formation of carbide nanoparticles
MC with the cubic structure B1, in which the carbon content is
lower than in molecular clusters M m C n . This really shows that in
the gas-phase synthesis conditions, the formation of cubic or
43

Nanocrystalline Materials

dodecahedral structures in the MC systems is determined to a


greater extent by kinetic factors instead of thermodynamic ones.
On the whole, plasma chemical synthesis with different methods
of formation of plasma is one of the most promising methods of
producing different nanostructured materials.
2.3 PRECIPITATION FROM COLLOID SOLUTIONS
Precipitation from colloid solutions is evidently the first method of
producing nanoparticles. Visitors to the Royal Institutions Faraday
Museum in London, opened in 1973 by Queen Elizabeth II, may see
two glass vessels with a colloid solution of gold (Fig. 2.5) produced
by M. Faraday in the first half of the 19 th century. These solutions
have retained their stability for almost 200 years. The production
and optical properties of colloid solutions of gold were described by
Faraday in 1857 [76].
The conventional method of producing nanoparticles from colloid
solutions is based on a chemical reaction between the components
of the solution and interrupting the reaction at a specific moment
in time [7781]. Subsequently, the dispersed system is transferred
from the liquid colloidal state to the nanocrystalline solid state. For
example, nanocrystalline powders of sulfides are produced by
reaction of hydrosulfuric acid H 2 S or sulfide Na 2 S with the water-

Fig. 2.5. Colloid solutions of gold, produced by M. Faraday (the Museum of the
Royal Institute in Great Britain, London).

44

Synthesis of Nanocrystalline Powders

soluble salt of a metal: Nanocrystalline cadmium sulfide CdS is


produced by precipitation from a mixture of solutions of cadmium
perchlorate Cd(ClO 4) 2 and sodium sulfide Na 2 S

Cd (ClO 4 ) 2 + Na 2S = CdS +2NaClO 4

(2.3)

The growth of nanoparticles of CdS is interrupted by a sudden


increase of the pH of the solution.
The colloid particles of metal oxides are produced by hydrolysis
of salts. For example, TiO 2 particles form easily in hydrolysis of
titanium tetrachloride

TiCl4 +2H 2 O = TiO 2 +4HCl

(2.4)

The formation of metallic or semiconductor clusters with a very low


dispersion of the dimensions (or even monodispersed clusters) is
possible inside pores of a molecular sieve (zeolite). Isolation of the
clusters inside the pores is maintained at heating to very high
temperatures. For example, semiconductor clusters (CdS) 4 are
synthesised inside cavities of zeolites [82]. Analysis of the
properties of clusters, produced in ultrafine channels and, in
particular, in pores of zeolites, has been the subject of a review in
[83]. Large semiconductor nanoparticles are synthesised by
annexation of additional molecules to the initial small cluster which
is stabilised in advance in a colloid solution by organic ligands. This
synthesis of large nanoparticles can be regarded as polymerization
of inorganic compounds.
Nanoparticles can also be produced by means of ultrasound
treatment of colloid solutions, containing large particles.
Precipitation from colloid solutions makes it possible to synthesise
nanoparticles of a mixed composition, i.e. nanocrystalline
heterostructures. In this case, the core and the shell of the mixed
nanoparticle are produced from semiconductor substances with
different electron structures. Formation of heterostructures, for
example, CdSe/ZnS or ZnS/CdSe, HgS/CdS, ZnS/ZnO, TiO 2 /SnO 2
takes place as a result of controlled precipitation of molecules of
a semiconductor of one type on pre-synthesised nanoparticles of a
semiconductor of another type [8487]. These heteroparticles may
be coated with a layer of another semiconductor. Nanocrystalline
heterostructures are used in photocatalysis.

45

Nanocrystalline Materials

In the group of all the methods of producing isolated


nanoparticles and other powders, the method of precipitation from
colloid solutions is characterised by high selectivity and makes it
possible to produce stabilised nanoclusters with a narrow size
distribution which is very important for the application of
nanoparticles as catalysts or in nanoelectronic devices. The main
problem of precipitation from colloid solutions is how to avoid
coalescence of the produced nanoparticles.
The chemical synthesis of large metallic clusters using colloid
solutions has been discussed in detail in [88]. There are different
chemical methods of producing nanoparticles in colloid solutions but
in any case it is necessary to protect particles in order to prevent
their coalescence. Stabilisation of the colloid particles and clusters
is possible using ligand molecules. Various polymers are used as
ligands. The schematic reaction of production of a ligand-stabilised
metallic cluster M n has the following form:
mL
nM + + ne M n
M n L m ,

(2.5)

where L is a ligand molecule. The metallic clusters of gold,


platinum and palladium, produced by this method, may contain from
300 to 2000 atoms. The metallic clusters have a cubic or hexagonal
close-packed structure. In these clusters, the central atom is
surrounded by several shells in which the number of atoms is
10k 2 + 2 (k is the number of the shell), i.e. the first shell contains
12, the second 42, the third 92 atoms, etc. The total number of
atoms in a cluster is n =  2 N + 1 + 10

k
k =1

, where N is the number

of atomic shells (layers). In the clusters, stabilised with ligands,


there is a metallic core where the nearest neighbours of the metal
atom are only metallic atoms, and the external shell of the metallic
atoms, partially bonded with the ligand molecules. Protection of the
clusters by a means of an outer shell is shown in Fig. 2.6: The
surface of a dark colloid nanoparticle of gold with a size of
approximately 12 nm is coated with a light shell of trisulfonated
triphenylphosphine P(m-C 6 H 4 SO 3 Na) 3 ligand molecules.
The metallic clusters, consisting of 55 atoms, distributed in two
shells, are evidently particles with a smaller size still retaining some
of the properties of the metal; however, examination by scanning
tunneling spectroscopy at room temperature already indicates the
46

Synthesis of Nanocrystalline Powders

Fig. 2.6. High resolution microscopic image of a single gold colloid of about
1113 nm, covered by a shell of P(m-C 6 H 4 SO 3 Na) 3 ligands (image obtained by
J.-O. Bovin and A. Carlsson, University of Lund) [88].

splitting of electron levels in these particles.


The hydrolysis of metal salts is carried out to produce colloid
oxide particles [8991]. For example, the nanocrystalline oxides of
titanium, zirconium, aluminium and yttrium may be produced by
hydrolysis of appropriate chlorides or hypochlorides. Ultrafine
titanium oxide is also produced by the hydrolysis of titanyl-sulfate
with subsequent heating of the amorphous deposit at 10001300 K.
Colloid solutions are stabilised to prevent coalescence of the
nanoparticles using polyphosphates, amines, and hydroxyl ions.
The colloid solutions of semiconductor oxides and sulfide
nanoparticles are used directly (without precipitation) in
photocatalytic processes of synthesis and destruction of organic
compounds, and dissociation of water. To produce highly dispersed
powders, the deposits of colloid solutions, consisting of
agglomerated nanoparticles, are heated at 12001500 K. For
example, a highly dispersed powder of silicon carbide (D ~
40 nm) is produced by hydrolysis of organic salts of silicon with
subsequent heating in argon at 1800 K [92]. Highly dispersed
powders of oxides of titanium and zirconium are often produced by
precipitation by means of oxalates.
Cryogenic drying is also used to produce highly dispersed powders
from colloid solutions. A solution is pulverised in a chamber with
47

Nanocrystalline Materials

a cryogenic medium and, consequently, freezes up in the form of


fine particles. Subsequently, the pressure of the gas medium is
reduced in such a manner that it becomes lower than the equilibrium
pressure above the frozen solvent and the material is heated with
continuous pumping for sublimation of the solvent. This leads to the
formation of the finest porous granules of the same composition,
and heating these granules gives powders.
The precipitation methods also include the production of
nanocrystalline composites from tungsten carbide and cobalt used
for producing hard alloys [93, 94]. Colloid solutions of tungsten and
cobalt salts are dried by spraying. The produced powder is
subjected to low-temperature carbothermal reduction in a suspended
layer thus retaining high dispersion. To decelerate the growth of
grains and decrease the solubility of tungsten carbide in cobalt, nonstoichiometric vanadium carbide in the amount of up to 1 wt.% is
added to the mixture. The hard alloy, produced from this
nanocrystalline composite, is characterised by the optimum
combination of high hardness and strength [9395].
It is shown [96] that every nanocomposite particle of WCCo,
with a size of ~75 m, consists of several millions of nanocrystalline
WC grains smaller than 50 nm, distributed in the cobalt matrix.
Sintering of the nanocomposite mixture of tungsten carbide with 6.8
wt.% Co and 1 wt.% VC leads to the formation of alloys in which
60% of WC grains were smaller than 250 nm and 20% smaller than
170 nm. An even finer grain structure was found in the alloy
containing, in addition to tungsten carbide, 9.4 wt.% Co, 0.8 wt.%
Cr 3 C 2 and 0.4 wt.% VC. After sintering at 1670 K 60% of
tungsten carbide grains in this alloy are smaller than 140 nm and
20% smaller than 80 nm. Comparison of the nanocrystalline and the
conventional polycrystalline alloy, having the same hardness, shows
that the fracture toughness of the nanocrystalline alloy is 1.21.4
times higher than the fracture toughness of the conventional coarsegrained alloy [96]. According to [97], the hard WCCo alloy,
produced from the nanopowder of tungsten carbide with a particle
size of 3050 nm, has a more uniform fine-grain structure and high
hardness and strength than the conventional hard alloy of the same
composition. In [97] it is also reported that the addition to the
coarse-grained charge of the standard WCCo alloy of 35 wt.%
of tungsten carbide nanopowder decreases the scatter of the
hardness and strength values, i. e. stabilises the properties of the
hard alloy.

48

Synthesis of Nanocrystalline Powders

2.4 THERMAL DECOMPOSITION AND REDUCTION


Thermal decomposition is usually carried out using complex
element-organic and organometallic compounds, hydroxides,
carbonyls, formiates, nitrates, oxalates, amides and imides of metals
which at a specific temperature decompose with the formation of
a synthesised substance and generation of the gas phase. The
production of highly dispersed metallic powders by thermal
decomposition of different salts has been described in detail in [98].
For example, the pyrolysis of formiates or iron, cobalt, nickel, and
copper in vacuum or an inert gas at 470530 K produces the
metallic powders with a mean particle size of 100300 nm.
A variant of pyrolysis is the decomposition of organometallic
compounds in a shock tube. Subsequently, free metal atoms are
condensed from the supersaturated vapours [14]. A long steel tube,
closed on both ends, is sectioned into two different parts with a thin
diaphragm made of mylar film or aluminium foil. The longer part
of the tube is filled with argon under a pressure of 10002500 Pa
with an addition of 0.12.0 mol.% of a organometallic compound.
The other part of the tube is filled with helium or a mixture of
helium with nitrogen until the membrane is ruptured. Rupture of the
membrane leads to the formation of a shock wave. The temperature
at the front of this wave may reach 10002000 K. Impact heating
of the gas results in the decomposition of the organometallic
compound several microseconds after the passage of the wave
front, and free metal atoms form a strongly supersaturated vapour
capable of rapid condensation. This method is used to produce
ultrafine powders of iron, bismuth and lead.
A combination of thermal decomposition and condensation is the
supersonic discharge of gases from a chamber, in which increased
constant pressure and temperature is maintained, through a nozzle
into vacuum [14]. In this case, the thermal energy of gas molecules
is transformed into the kinetic energy of the supersonic flow. During
expulsion the gas cools down and transforms into supersaturated
vapour in which clusters, containing from 2 to a million of atoms
may form. An increase of the initial pressure in the chamber at a
constant temperature increases supersaturation. The authors of [99]
describe the production of submicrocrystalline powder
(Si 3 N 4 + SiC) by pyrolysis of liquid polysilazane [CH 3 SiHNH] n ,
discharged in the form of an aerosol through an ultrasound nozzle.
The aerosol was heated by the radiation of a continuous-wave CO 2 laser.
49

Nanocrystalline Materials

The highly dispersed powders of silicon carbide and nitride are


produced by pyrolysis of polycarbosilanes, polycarbosiloxanes and
polysilazanes [100102]. Initial heating is carried out by a means
of low temperature plasma or laser radiation and the products of
pyrolysis are subsequently annealed at a temperature of ~1600 K
to stabilise the structure and composition.
Poly[2-(vinyl)pentaborane] is used to produce nanocrystalline
boron carbide. Boron-containing polymers of the type of
polyborazine, polyborazole and poly(B-vinylborazine) are proposed
for the production of highly dispersed powders of boron nitride and
also as additions to the titanium powder for synthesis of
nanocrystalline compositions TiN + TiB 2 [102]. The nanocrystalline
AlN nitride powder with a mean particle size of 8 nm is produced
by decomposition of ammonium polyamide imide [Al(NH 2) 3NH] n in
ammonia at 900 K [103]. The submicrocrystalline titanium nitride
powder with a mean particle size of 100300 nm is prepared by
decomposition of polytitanimide Ti[N(CH 3 ) 2 ] 4 [100].
The transition metal borides may be produced by pyrolysis of
borohydrides at 600700 K, i. e. at a temperature which is
considerably lower than the conventional temperatures of solid
phase synthesis. For example, high-dispersion powders of zirconium
boride with a specific surface of 40125 m 3 g 1 are produced by
thermal decomposition of zirconium tetraborohydride Zr(BH 4) 4 under
the effect of pulsed laser radiation [104].
The powders, produced by thermal dissociation of monomer and
polymer compounds, must be additionally annealed to stabilise the
composition and structure; the annealing temperature of nitrides and
borides is 9001300 K, and for oxides and carbides it is 1200
1800 K.
Section 2.1 describes the method of producing nanocrystalline
powders in which the thermal decomposition of the organometallic
precursor is combined with the condensation of nanoparticles on the
cold surface in the atmosphere of a rarefied inert gas.
The main disadvantage of the thermal decomposition method is
the relatively low selectivity of the process, because the reaction
product is usually a mixture of the target product and other
compounds.
A widely used method of producing highly dispersed metallic
powders is the reduction of compounds of metals (hydroxides,
chlorides, nitrates, carbonates) in a hydrogen flow at a temperature
of <500 K. Advantages of this method are the low content of
impurities and a narrow size distribution of the powder particles;
50

Synthesis of Nanocrystalline Powders

typical size distributions of the particles for several metal


nanopowders, produced by reduction in a hydrogen flow, are shown
in Fig. 2.7.
Metallic catalysts are usually produced by saturation of a porous
material (silica gel, zeolite, etc.) by a solution of hydroxide or other
compound of the required metal. The saturated porous carrier is dried

1XPEHU RI SDUWLFOHV 



1L




























QP

1XPEHU RI SDUWLFOHV 



&X
























QP

1XPEHU RI SDUWLFOHV 



)H


























QP

Fig. 2.7. Typical distributions of metal particles by their size D. The particles are
synthesized by reduction of metals from compounds in a hydrogen flow.

51

Nanocrystalline Materials

and then heated in a hydrogen flow to reduce the metal. As a result,


catalytically active fine metallic particles form in the pores of the
carrier.
2.5 MILLING AND MECHANICAL ALLOYING
The basis of mechanical alloying is the mechanical treatment of
solid mixtures leading to refining and plastic deformation of
substances, acceleration of mass transfer and mixing of the
components of the mixture on the atomic level, together with the
activation of chemical reaction of the solid reagent [105107].
A stress field forms in the near-contact regions of the solid
substance as a result of the mechanical treatment. A relaxation of
the stress field may take place by the generation of heat, formation
of a new surface, formation of different defects in crystals,
excitation of chemical reactions in the solid phase. The preferential
direction of relaxation depends on the properties of the substance,
the loading conditions (power of supplied energy, the relationship
between pressure and shear), the sizes and shape of the particles.
An increase of the power of the mechanical pulse and the duration
of the mechanical treatment is accompanied by a gradual transition
from relaxation by the generation of heat to the relaxation
associated with the failure, disintegration and plastic deformation of
the material and the appearance of amorphous structures of
different nature. Finally, the chemical reaction, initiated by different
mechanisms, may be the channel for relaxation of the stress field.
In the group of these mechanisms there are direct excitation and
fracture of a bond which may take place at the crack tip; local
thermal heating; emission-free breakdown of excitons, etc.
The mechanical effect during refining of materials is pulsed and,
consequently, the formation of a stress field and its subsequent
relaxation are observed only at the moment of collision of the
particles and for a short time after this collision. For this reason,
in mechanochemical synthesis it is necessary to take into account
the formation of the stress field with time and the kinetics of
subsequent relaxation processes. The mechanical effect is not only
pulsed but also local because it does not take place in the entire
mass of the solid substance but take place in areas where the stress
field forms and subsequently relaxes.
Mechanical milling is the most productive method of producing
large quantities of nanocrystalline powders of different materials:
metals, alloys, intermetallics, ceramics, composites. Mechanical
52

Synthesis of Nanocrystalline Powders

milling and mechanical alloying may lead to the complete solubility


in the solid state of elements characterised by very low mutual
solubility in the equilibrium condition [108, 109].
In mechanical milling of powders, deformation is initially localized
in shear bands containing a large number of dislocations. At a
certain strain level, these dislocations annihilate and recombine into
low-angle boundaries separating the individual grains. This stage of
milling is already characterised by the formation of grains with a
diameter of 2030 nm and their number increases with increasing
milling time. In the following stage of milling the orientation of the
individual crystals in relation to each other becomes random
because of grain boundary sliding. This behaviour during milling is
typical of bcc metals and intermetallics [109].
Milling and mechanical alloying are carried out using high-energy
planetary, ball and vibration mills, where the mean size of the
produced powders may vary from 200 to 510 nm. For example,
when milling boride -FeB in a ball mill, it is possible to produce
a powder of -FeB with a mean crystallite size of ~8 nm [110].
Mechanical treatment of barium titanate BaTiO 3 in a planetary mill
is applied to produce nanocrystalline powder with a mean particle
size of 525 nm [106].
Mechanical alloying of the powders of borides, carbides, silicides,
oxides and sulfides of transition metals has been carried out by the
explosive method in vibration mills [111, 112]; the high rate of
reaction is initiated by mechanical activation of the powders of
initial components (metal and carbon, boron or silicon) over a
period of several minutes. Examination of the powders of carbides
of B, Ti, Zr, Hf, V, Ta, and W, produced by mechanical alloying in
mills, shows that the mean particle size is 620 nm [113].
The powders of the transition metal nitrides with a particle size
of several nanometers are synthesised by milling metallic powders
in a vibromill in a nitrogen atmosphere [114].
Mechanochemical synthesis of nanocrystalline TiC, ZrC, VC and
NbC carbides from a mixture of powders of metal and carbon is
described in [115]. The mixture is produced by ball milling. Carbides
form after milling for 412 hours; the size of the powder particles
after 48 hours of milling is 71 nm. In the group of the produced
carbide nanopowders, the highest stability at heating is found for
the niobium carbide: with an increase of temperature from 300 to
1300 K, the size of the NbC grains increases from 710 to 30 nm.
The lowest stability at heating is exhibited by the vanadium carbide.
The intensive recrystallisation of this carbide at 10001200 K leads
53

Nanocrystalline Materials

to grain growth to a size of 90 nm.


The nanocrystalline bcc alloys FeNi and FeAl with a grain
size of 515 nm are synthesised by ball milling of metal powders
in a vibration mill for 300 h [116].
The ideal and practically applicable variant of mechanochemical
synthesis, combined with the production of a nanocomposite mixture,
is described in [117]. A mixture of coarse-grained (~75m) powders
of tungsten, graphite and cobalt is milled for 100 h in an argon-filled
ball mill. This leads to the formation of a nanocomposite mixture
WCCo of grains of tungsten carbide and cobalt with a mean size
of 1112 nm.
In the hard alloy, produced by cold pressing and subsequent
sintering of this mixture at 1310 K, the majority of the WC grains
are smaller than 200 nm, i.e. several times smaller than in
conventional alloy of the same composition. The hardness of the
sintered specimens of the hard alloy is ~18 GPa, and the relative
density is equal to 80% of theoretical density. The possibility of
sintering of the nanocomposite powder mixture WCCo at a lower
temperature in comparison with that used for the coarse-grained
mixture, is the result of a lower melting point of nanocrystalline
cobalt in comparison with coarse-grained cobalt.
A detailed description of the mechanical synthesis of
nanocrystalline powder of titanium carbide is presented in [118].
Powders of titanium and graphite were taken in a ratio 44:56
resulting to the formation of the composition Ti 44 C 56 . A powder
mixture was milled in a sapphire ball mill. The ratio of the mass of
the balls to the mass of the powder was 10:1. Milling was carried
out at room temperature in argon.
After milling for 2000 s, the X-ray diffraction pattern of the
reacting mixture (Fig. 2.8) showed only wide reflections
corresponding to Ti and C. After milling for 1110 3 s, the
reflections, corresponding to graphite, disappeared almost
completely, and after milling for 1510 3 s the reflections of the new
cubic phase with the B1 structure appeared. The lattice constant
of the produced titanium carbide is 0.4326 nm. An increase in the
milling time to 410 4 s resulted in the complete disappearance of
the diffraction reflections of metallic titanium and increased the
intensity of reflections of titanium carbide. An increase in the milling
time to 810 4 s was accompanied by an increase of the degree of
mechanical deformation of the powder particles and by a large
decrease of the grain size, as indicates by the large broadening of
the diffraction reflections. Milling for 7210 4 s leads to the
54

Synthesis of Nanocrystalline Powders

7L
7L




7L








7L

7L
&

7L


7L
7L






&

&X.

  V

  V

&RXQWV DUELWUDU\ XQLWV

  V

  V

  V

  V

  V

  V

 7L&  7L&

  V







 7L&





 7L&  7L&





 GHJUHHV
Fig. 2.8. X-ray diffraction patterns of mechanically alloyed Ti 44 C 56 powders as a
function of the ball-milling time [118].

formation of nanocrystalline titanium carbide. A further increase of


the milling time to 10 6 s did not lead to any observable changes in
the produced carbide.
Figure 2.9 shows the change of the grain size of the powder
during milling according to the results of X-ray diffraction and
electron microscopy. According to Fig. 2.9, four stages are evident
in the process of formation of nanocrystalline titanium carbide.
Figure 2.10 shows scanning electron images of powder particles in
different milling stages. The starting powder (Fig. 2.10a) consists
of randomly distributed particles of different sizes and shapes. The
first stage (milling time up to 1010 3 s) is characterised by the
formation of Ti/C composite particles with a mean size of
approximately 1000 nm (Fig. 2.10b). Metallographic examination
55

Nanocrystalline Materials

SXUH 7L

3DUWLFOH VL]H ! QP

FRDUVHJUDLQHG 7L&

SXUH &
FRPSRVLWH 7L&
SDUWLFOHV

7(0
;5'

QDQR7L&

 

 

 

 

 

 

9 V

Fig. 2.9. Dependence of the particle size <D> of mechanically alloyed Ti 44 C 56


powders on the ball-milling time t [118]: (open circle) particle size which is determined
by TEM techniques, (closed circle) particle size which is determined by X-ray
diffraction analysis from broadening of diffraction reflections.

shows that these particles consist of a large number of titanium and


carbon layers. The second stage of milling, lasting from 1110 3 to
210 4 s, is a mechanical solid-phase reaction. During this reaction
titanium and carbon almost completely react together and form
large grains of titanium carbide with a size of 8001000 nm. The
third stage, lasting from 210 4 to 810 4 s, is characterised by
extensive refining of titanium carbide grains and by the formation
of ultrafine powder with a relatively wide size distribution of the
grains, from 5 to 100 nm; the titanium carbide grains merged into
particles with a size of 510 m (Fig. 2.10c). The last stage, from
810 4 to 110 6 s, is the stage of homogenising of the
nanocrystalline powder with respect to the grain size the produced
powder of titanium carbide is characterised by a narrow size
distribution of the grains and consists of particles with a size of
approximately 21 nm; the grains are agglomerated into spherical
particles not larger than 300 nm (Fig. 2.10d).
The titanium carbide powder produced by milling with different
durations (2.210 4 , 410 4 , 810 4 and 7.210 5 s) was sintered in
activated plasma [118]. This sintering is a version of hot pressing
but differs from the latter by the application of pulsating voltage for
56

Synthesis of Nanocrystalline Powders

Fig. 2.10. SEM micrographs of mechanically alloyed Ti 44C 56 powders after (a) 0 s,
(b) 1.110 4 s, (c) 410 4 s and (d) 7.210 5 s of the ball-milling time [118].

the activation of plasma and the non-axial application of pressing


pressure. The sintering time was less than 8 min. Sintering resulted
in the formation of dense specimens of titanium carbide with a high
density (up to 5.2 g cm 3 ) whilst retaining the mean grain size
below 70 nm. The dependence of the hardness of the titanium
carbide specimens, sintered from powders with different mean grain
sizes, is shown in Fig. 2.11. Hardness was measured at a load of
50 kg.
The results published in [118] demonstrate the high efficiency of
mechanosynthesis of nanocrystalline titanium carbide. At the same
time, the study [118] has a number of important disadvantages. First
of all, it does not contain any data on the chemical composition of
the produced titanium carbide. The implicit assumption of the
authors of [118] on the coincidence of the composition of the
Ti 44 C 56 charge with the composition of the synthesised titanium
57

























+'













3DUWLFOH VL]H G QP

+DUGQHVV +' *3D

Nanocrystalline Materials





 V

Fig. 2.11. Correlationship between ball-milling time t, Vickers hardness H V (closed


circle) and the grain size d (open circle) of sintered mechanically alloyed Ti 44 C 56
titanium carbide [118].

carbide creates doubts. The reported lattice constant of the titanium


carbide is slightly smaller and the measured density higher than the
lattice constant (0.4328 nm) and density (4.91 g cm 3 ) of
stoichiometric carbide TiC 1.0 . In particular, this high density of the
produced titanium carbide is difficult to understand. Even if it is
assumed that the synthesised carbide with structure B1 has the
same composition as the charge, i. e. Ti 0.44 C 0.56 (or TiC 1.27 !?), the
metallic sublattice of such a carbide should contain defects because
the number of sites of the non-metallic and metallic sublattices in the
structure type B1 is the same and there are no sites for distributing
the excess number of atoms of any type. This relative excess of the
atoms of some type indicates that the sublattice of the atoms of
another type contains structural vacancies. In the examined case, the
possible composition of the titanium carbide with the defective metallic
sublattice is Ti 0.78C 1.00, but the titanium carbide with this composition
should have the density 22% lower than the theoretical value.
Regardless of special investigations, it was not possible to find
vacancies in the metallic sublattice of carbides and it is therefore
assumed that the metallic sublattice of the transition metal carbides
does not contain structural vacancies. Taking this into account, the
results [118] require careful verification. At first, it is necessary to
determine the chemical composition of synthesised titanium carbide, i.e.
58

Synthesis of Nanocrystalline Powders

content of titanium, bonded and free carbon, and also metallic


impurities. It is necessary to determine the lattice constant and the
composition of sintered titanium carbide because it is evident that this
carbide contains a large amount of a metallic impurity so that the
density is higher than the theoretical density.
2.6 SYNTHESIS BY DETONATION AND ELECTRIC EXPLOSION
There is another type of mechanical treatment which creates
simultaneously the conditions for both the synthesis of the final
product and its dispersion. It is the shockwave. The shockwave
treatment of mixtures of graphite with metals at a pressure in the
shockwave of up to several tens of GPa is used to produce
nanocrystalline diamond powders with a mean particle size of 4 nm.
It is more efficient to produce diamond powders by explosion of
organic substances with a high carbon content and a relatively low
oxygen content.
The detonation of explosive substances, i.e. explosion energy, is
used quite widely for ensuring phase transitions in substances and
synthesis by detonation. Synthesis by detonation as a high-rate
process makes it possible to produce ultrafine powders in dynamic
conditions where the important role is played by kinetic processes.
Detonation synthesis of diamond was carried out for the first
time by authors of [119] by shockwave loading rhombohedral
graphite up to 30 GPa. Grains of explosion diamond were
agglomerates of individual particles and it was not possible to
determine the size of these diamond particles by optical microscopy
[119]. In [120], diamond powders were produced by shockwave
treatment of mixtures of graphite and metal; the duration of the
shockwave was 1020 ms, and the pressure generated by the wave
2040 GPa. Later, it was shown that the diamond powder, produced
in these conditions, contains single crystals not larger than 50 nm,
and also tightly bonded agglomerates with a size of up to 5 m or
larger. These agglomerates consist of individual crystals with sizes
of 14 and 10160 nm.
After 1983, studies were publisher (for example [121, 122])
which discussed the possibility of the formation of ultrafine diamond
particles by detonation of condensed explosive substances with a
negative oxygen balance. As a result of detonation a free carbon
is generated with subsequent formation of the diamond. This process
of production of diamond particles with their subsequent cooling in
the gas phase (the so-called dry synthesis) is realised by the
59

Nanocrystalline Materials

authors of [123, 124] in detonation decomposition of carboncontaining explosive substances with subsequent expansion of
explosive products in an inert atmosphere. At present, this process
is used for industrial production of ultrafine diamond particles for
various applications. The volume of explosion chambers is not
smaller than 23 m 3 . In the second version of detonation synthesis
of diamond powders from condensed explosive substances with a
negative oxygen balance, referred to as water synthesis, a water
cooler of the produced diamond particles is used.
The synthesised diamond powder forms in the zone of chemical
decomposition during a period not longer than 0.4 s and consists
of compact cubic particles with a mean size of approximately 4 nm.
The application of more powerful explosive substances makes it
possible to produce larger, up to 1 m, diamond particles [125].
Pressures of hundreds of thousands of atmospheres and
temperatures of up to several thousand degrees, characterising the
detonation process, correspond to the range of thermodynamic
stability of the diamond phase on the pT diagram of possible
states of carbon (Fig. 2.12). Therefore, the application of the
detonation method for synthesis of diamond in the dynamic conditions
is fully logical. At the same time, it must be taken into account that
in synthesis by detonation at a short duration of high pressures and
temperatures, required for the formation of diamond, a significant
role is played by the kinetics of formation and growth of nuclei of
the diamond phase.
This is confirmed by, for example, explosion decomposition of
trinitrotoluene characterised by the generation of the maximum
amount of free carbon. The parameters of the detonation wave in
the ChapmanJouguet plane (p 18 GPa, T = 3500 K) when the
chemical reaction has already been almost completed, correspond
to the region of stability of the diamond phase (Fig. 2.12). However,
detonation of trinitrotoluene is not characterised by a large yield of
the diamond phase [124].
To obtain a large yield of the diamond powder it is necessary to
use more powerful explosive substances. It makes possible to
increase the pressure and temperature generated by the
shockwave. Usually, ultrafine diamond powders are produced using
mixtures of trinitrotoluene and hexogene in a weight ratio of 50:50
or 60:40 [124, 126]. For these mixtures, the pressure and
temperature in the detonation wave are p15 GPa and T3000 K.
In dry detonation synthesis, the process takes place in special
explosion chambers, filled with an inert gas or carbon dioxide, which
60

Synthesis of Nanocrystalline Powders




'LDPRQG




*3D




*UDSKLWH

Fig. 2.12. Equilibrium phase pT diagram of carbon indicating the regions of synthesis
of diamond by different methods [119]: 1) detonation (impact wave) synthesis using
graphite, 2) static transformation using a catalyst, 3) static transformation without
a catalyst, 4) detonation synthesis using a mixture of trinitrotoluene and hexogene
at a ratio of ~1:1.

prevents oxidation of the produced diamond particles and their


transformation to graphite. The formation of diamond nanoparticles
takes place up to reaching the ChapmanJouget plane and is
completed within 0.20.5 s. It corresponds to the duration of
chemical reactions for the trinitrotoluenehexogene mixtures. In the
chemical reaction zone, the pressure and temperature may be
considerably higher than in the ChapmanJouget plane and,
consequently, the calculations of the formation of the diamond phase,
based on the parameters of the ChapmanJouget detonation wave,
should be regarded as a rough estimate. It should also be mentioned
that in detonation synthesis at a very short time of formation of the
diamond particles, the rate of growth of these particles is several
orders of magnitude higher than that in the static conditions. After
explosion, the condensed products of synthesis are collected and
processed in hot perchloric acid HClO 4 and in mineral acids under
pressure to remove carbon black and other impurities. This is followed
by multiple rinsing in water with drying. The yield of the diamond
powder is 89% of the initial mass of explosive substances which in
different systems may vary from tens of grams to several kilograms
[124].
Russian industry has mastered a conversion method of producing
61

Nanocrystalline Materials

diamond nanopowders by explosion of ammunition in special


chambers. High pressures and temperatures, developed during
explosions, lead to the synthesis of diamond from carbon-containing
explosive substances. Synthesis is catalysed by particles and
vapours of the metal from the shells of ammunition.
The characteristic special feature of diamond nanopowders,
produced by detonation synthesis, is the very narrow distribution of
the sizes of the nanoparticles: the main fraction of the particles has
the size of 45 nm [123128]. In fact, determination of the size of
nanoparticles by the Raman scattering method and from the
broadening of X-ray diffraction reflections shows that the diamond
particles are the nanocrystals with a characteristic size of 4.3 nm,
irrespective of the method and kinetics of cooling [126]. The narrow
range of the sizes of diamond nanocrystals is detected in different
investigations. According to [126], it is the consequence of the fact
that for small nanoparticle sizes it is the diamond and not graphite
which is the thermodynamically stable form of carbon. This
assumption has been confirmed by numerical calculations [129].
A detailed description of the formation of ultrafine diamonds in
the detonation wave, and the methods of their extraction from the
explosion products may be found in a monograph [130].
Another detonation method of synthesis of different
morphological forms of carbon and nanopowders of oxides of Al,
Mg, Ti, Zr, Zn was described by the authors of [131, 132]. A layer
of starting substance (high-porosity metallic medium, chemical
compound, sol or gel of a metal hydroxide) is subjected to the
shockwave treatment from a contact charge of an explosive
substance. The compression and heating of the high-porosity metal
or dissociation of the starting compound to the oxide with
subsequent stabilisation of the oxide phases take place in a
shockwave. After exit of the shockwave to the free surface of the
starting substance, the material is dispersed into the gas atmosphere
of the explosion chamber.
If the starting substance is represented by metals, an active
oxygen-containing medium (for example, O 2 + N 2 ) is used. In this
case, the scattering stage is characterised by the combustion of
metal with the formation of an ultrafine oxide. When using a
carbon-containing atmosphere CO 2 , it is possible to synthesise
nanotubes and spherical particles of carbon (Fig.2.13) and also
filament crystals of MgO. The mean size of filament crystals of
MgO is 60 nm, and the ratio of length to diameter reaches 100.
When using chemical compounds as starting materials, a gas or
62

Synthesis of Nanocrystalline Powders

Fig. 2.13. Carbon nanotubes prepared by detonation [127, 128].

liquid medium, chemically neutral in relation to the produced


material is used. This results in rapid cooling of the substance and
stabilisation of high-temperature and metastable crystalline
modifications. For example, the authors of [132] used the detonation
method to produce a nanocrystalline powder of cubic modification
of ZrO 2 , stabilised by yttrium oxide. The mean size of the powder
particles is 30 nm.
A rapidly developed method of producing ultrafine powders is the
electric explosion of a conductor during the passage of a powerful
current pulse with a duration of 10 5 10 7 s and a density of 10 4
10 6 A mm 2 [133, 134]. Wires with a diameter of 0.11.0 mm are
used for this purpose. The electric explosion of a conductor is a
large change of the physical state of the metal as a result of the
intensive generation of energy during the passage of a high-density
current pulse [135]. Electric explosion is accompanied by the
generation of shockwaves and makes it possible to heat rapidly
metals at a rate greater than 110 7 K s 1 to high temperatures of
T > 10 4 K. The capability of electrically exploded conductors to
rapidly change properties and efficiently transform the primary
electric or magnetic energy of accumulators to other types of
energy (thermal energy, the energy of radiation of the plasma,
energy of the shockwaves) is used in particular for producing
ultrafine powders.
In the initial stage of electric explosion, the Joule heating of a
conductor is accompanied by its linear expansion at a relatively low
63

Nanocrystalline Materials

rate of 13 m s 1 . In the stage of an explosion the passage of a


current pulse superheats the metal above the melting point, the
expansion of the material of the exploded conductor takes place at
a rate of up to 510 3 m s 1 and the superheated metal is dispersed
in an explosion-like manner [136]. The pressure and temperature
at the front of the shockwave reach several hundreds of MPa
(thousands of atmospheres) and ~10 4 K, respectively. Particles of
very small sizes form as a result of condensation in the flow of the
rapidly expanding vapour. Regulating the explosion conditions, it is
possible produce powders with a particle size from 100 m to
50 nm. The mean size of the particles monotonically decreases with
increasing current density and shortening of the pulse time. Electric
explosion in an inert atmosphere makes it possible to produce
powders of metals and alloys. Adding of additional reagents into the
reactor (air, mixture of oxygen and inert gas, nitrogen, distilled
water, decane C 10 H 22 , paraffin, commercial oil) allows to produce
ultrafine powders of oxides, nitrides, carbides or their mixtures
(Table 2.1). The authors of [134] described copper powders
produced by electric explosion in an inert gas at a pressure of
200 Pa with the maximum size distribution corresponding to
~20 nm, and aluminium powders with a mean particle size of
approximately 50 nm.
According to the experimental data [137], the submicrocrystalline
powders, produced by electric explosion of wires, have a very large
excess energy. For example, aluminium powders with a mean
particle size of 500800 nm have an excess energy of 100200
kJ mol 1 , and silver powders with a mean particle size of ~120 nm
Table 2.1 Some nanopowders produced by electrical explosion in a vacuum and
various media
Metal

Vacuum
<1.3106
(Pa)

Air

Nitrogen

Water

Decane

N2

H2O

1022

n-AlN

n-Al(OH)3
or -Al2O3

n-Al4C3

Carbides

n-FeO

n-FeO

Carbide
mixture

n-Ti2O3

n-TiCy

Carbide
mixture

n-WCy

n-WC

Al

n-Al

n-Al, covered by
oxide

Fe

n-Fe

n-Fe, covered by
oxide

Ti

n-Ti

n-Ti, covered by
oxide

n-W

n-W, covered by
oxide

n-WO2

Cu

n-Cu

n-Cu, covered by
oxide

n-Cu2O

n-TiNy

64

Paraffin

Synthesis of Nanocrystalline Powders

have an excess energy of 4080 kJ mol 1 , which is several times


higher than the melting heat of the bulk substance. This excess
energy cannot be determined only by the contribution of surface
energy. The storage of a large amount of excess energy by
ultrafine powders, produced by electric explosion, has not as yet
been completely explained. In [138] it is suggested that the excess
energy is stored in the form of surface energy, energy of internal
defects and charge states.
The size distribution of the powders is logarithmico-normal and
the maximum of the distribution is in the range 10500 nm, g =
1.31.8. The particles of the powders of metals and alloys,
produced by electric explosion, are spherical, and the particles of
nitride powders have faceting.
Recently, special attention has been given to the development of
an electric erosion method of producing submicrocrystalline and
nanocrystalline powders of metals and alloys.
2.7 ORDERING IN NON-STOICHIOMETRIC COMPOUNDS
Nanocrystalline ceramic materials have been studied intensively
because of the need to develop hard and, at the same time, nonbrittle materials with high fracture toughness. Promising materials
in this respect are non-stoichiometric carbides of groups IV and V
transition metals whose hardness is inferior only to that of diamond
and cubic boron nitride [139].
The monocarbides of transition metals MC y are included in a
group of strongly non-stoichiometric compounds [140142]. In the
disordered state, the monocarbides MC y have the B1 cubic structure
and may contain up to 50% of structural vacancies in the non-metall
sublattice [38, 140147]. At a temperature below 1300 K, the B1
structure is unstable and disorderorder phase transformations take
place in non-stoichiometric carbides. Ordering leads to the
formation of ordered phases with complicated superstructures [140
147]. The orderdisorder transformations in the carbides are phase
transitions of the first kind [38, 140146] with a sudden change of
volume at the disorderorder phase transformation temperature T trans
[140, 141, 146, 148, 149]. The ordering process is of the diffusion
type and, consequently, transformation does not take place instantaneously but over a period of several tens of minutes. The carbides
are synthesised at temperatures of 14001800 K which are higher
than the temperature T trans . In cooling from the synthesis
temperature to room temperature, the non-stoichiometric carbide
65

Nanocrystalline Materials

passes through the ordering temperature and tends to the ordered


state. If the cooling rate is high, the ordering process has not time
to complete and the nonstoichiometric carbide remains in a
metastable disordered state. Because of the difference in the lattice
spacings or distances between atoms of the disordered and ordered
phases, stresses arise in a specimen. With time these stresses lead
to cracking of crystallites at the interfaces between the disordered
and ordered phases. Controlling the size of domains of the ordered
phase, it is possible to produce nanostructured powders of nonstoichiometric carbides.
The concept of formation of the nanostructure by means of
atomic-vacancy ordering of the non-stoichiometric compounds is
realised for the first time by the authors [150] on an example of
non-stoichiometric vanadium carbide. Previously, the ordering
phenomenon was not used for producing the nanostructured state.
The vanadium carbide was chosen among all non-stoichiometric
carbides for investigation because ordering is pronounced most in
this carbide [149, 151153]. In nanocrystalline solids and
nanopowders, the important role is played not only by volume
(associated with particle sizes) but also surface (associated with the
state and structure of the interface) effects [154156]. Therefore,
the authors of [150] and subsequent investigations [157159] in
examination of the physicalchemical properties of vanadium carbide
paid special attention to the surface of the investigated substance.
The starting powder of vanadium carbide VC 0.875 with particles
12 m in size was produced by carbothermal reduction of V 2 O 5
oxide and then subjected to long-term ageing at 300 K in a closed
vessel preventing the penetration of water vapours from air.
The aged powder of vanadium carbide proved to be very
hygroscopic. Immediately after the powder was removed from the
vessel, it contained not more than 0.2 wt.% of physically adsorbed
water. After the powder was kept in the ambient atmosphere for
several months, the water content reached 2.0 wt.% and came to
saturation. High hygroscopicity is an indirect indication of the
catalytic activity of the aged VC powder. It should be mentioned
that the conventional coarse-grained vanadium carbide powder does
not possess such hygroscopicity.
Chemical composition of aged VC powder (wt%)

Component

Cbond

Concentration 77.50.1 16.00.1


(mass %)

Cfree

H2O

Ochemisorb

Olattice

0.90.1

2.00.2

3.10.1

0.1

0.20.02

66

Synthesis of Nanocrystalline Powders

According to the results of chemical analysis, the aged vanadium


carbide powder had the composition VC 0.875 , corresponding to the
upper boundary of the homogeneity interval of the cubic phase B1.
The impurities Ti, Nb, Ta and W with a total content of
approximately 1 wt.%, were detected by the EDX method only in
the surface layer of the sintered specimen; the content of these
impurities in the volume of the specimen was less than 0.1 wt.%.
The total oxygen content (5.0 wt.%) of the powder is so high
that it could indicate complete filling of the free non-metallic sites
(structural vacancies) by oxygen atoms and the formation of the
vanadium oxycarbide. Therefore, detailed analysis was carried out
to determine the type in which oxygen is present in the specimen.
It was shown that 1.8 wt.% of oxygen relates to the fraction of
adsorbed water (water content in the amount of 2 wt.% was
determined by heating in vacuum to 500 K, and also by gas
chromatography and thermogravimetric analysis). Oxygen in the
amount of 0.1 wt.% is dissolved in the lattice of vanadium carbide
resulting in filling a small fraction of structural sites of the carbon
sublattice approximately 0.4%. According to the results of
quantitative gas chromatography, the main portion of oxygen in the
amount of 3.1 wt.% is in the chemisorbed form. A very small
amount of oxygen forms a surface oxide film with a thickness of
23 atomic monolayers; the film contains oxide V 2 O 5 and
homologous oxides V n O 2n1 located between V 2 O 3 and VO 2 . The
nitrogen content, determined without pre-heating the powder, was
only 0.2 wt.% and could not have any significant effect on the
powder properties. To analyse the oxide film, the vanadium carbide
powder was heated in a diluted solution of HCl (concentration
0.36 wt.%, pH = 1) at 330 K for several minutes. On the basis of
the light blue colour of the solution it may be concluded that it
contains ions with trivalent and four-valent vanadium. This confirms
the presence of an oxide film on the surface of the vanadium
carbide powder.
In examination of the vanadium carbide powder under a
microscope at a magnification of approximately 100 times (Fig.
2.14) there are individual agglomerates of irregular shape with a
size of 550 m and consisting of particles with a size of
approximately 1 m. However, at a high magnification, it becomes
evident that these particles have a complicated structure and, in
reality, are a set of a large number of very small particles of the
nanometer size (they will be referred to as nanoparticles or
nanocrystallites). Figure 2.15 shows a micrograph produced at a
67

Nanocrystalline Materials

Fig. 2.14. Scanning electron micrograph of aged vanadium carbide VC 0.875 powder:
agglomerated particles with irregular shape have size from to 5 to 50 m and consist
of small particles with a size of approximately 1 m or less [136, 137, 155].

Fig. 2.15. Microstructure of the powder of vanadium carbide VC 0.875 subjected to


long-term ageing at ambient temperature (magnification 10 000) [136, 137, 146].

magnification of up to 10000 in high-resolution scanning electron


microscope DSM 982 Gemini. It may easily be seen that each of
the particles with a size of ~1m looks like an open rose bud or
a loose cabbage head and consists of nanocrystallites. At a
magnification of up to 50000 times it is evident that the observed
nanocrystallites have the form of distorted petals. Joining together,
the nanocrystallites form a structure resembling corals (Fig. 2.16).
The nanostructure of vanadium carbide is unique and has not been
detected in any of the substances investigated. To a first
68

Synthesis of Nanocrystalline Powders

Fig. 2.16. Morphology of particles of aged powder of VC 0.875 carbide at a


magnification of 50 000 [136, 127, 146, 153155]: particles with a size of
approximately 1 m represent a set of nanocrystallites in the form of distorted
petalsdiscs with a diameter from 400 to 600 nm and a thickness of approximately
1520 nm.

approximation, the nanocrystals may be simulated with a disc with


a diameter 400600 nm, thickness approximately 1520 nm. Despite
the fact that the volume of this disc corresponds to a relatively large
spherical particle with a diameter of 150220 nm, the ratio between
the surface area S and the volume V is S/V = 0.1070.143 nm 1
and is not large thanks to the small thickness of the disc. This
S/V value corresponds to a specific surface of the powder from 19
to 26 m 2 g 1 . Electron microscopic examination also indicates a
narrow size distribution of nanocrystallites in the powder.
The crystal structure of the vanadium carbide was examined by
X-ray diffraction analysis in CuKa 1,2 radiation [150, 159]. Figure
2.17 shows X-ray diffraction patterns of the disordered carbide
VC 0.875 , annealed coarse-grained ordered carbide VC 0.875 (V 8 C 7 )
and nanostructured carbide VC 0.875 . In addition to the structural
reflections of the basic phase B1, the X-ray diffraction pattern of
the VC 0.875 nanopowder (Fig. 2.17c) contains superstructure
reflections, corresponding to the cubic ordered V 8C 7 with a space
group P4 3 32. The lattice constant of the ordered phase is 0.8337
0.0001 nm. The ideal cubic superstructure of the type M 8 C 7 with
a space group P4 332 (Fig. 2.18) has a double (in comparison with
the disordered basic phase B1) lattice constant [140, 141, 147] and,
consequently, for the investigated vanadium carbide, the lattice
constant of the basic phase is a B1 = 0.41685 nm. This is greatly
69

Nanocrystalline Materials
GLVRUGHUHG FDUELGH 9&

,



 QP

&RXQWV DUELWUDU\ XQLWV


DQQHDOHG FRDUVHJUDLQHG RUGHUHG FDUELGH 9&

,

 QP




QDQRFU\VWDOOLQH RUGHUHG FDUELGH 9&

,



 QP
















 GHJUHHV

Fig. 2.17. X-ray diffraction patterns of powders of disordered carbide VC 0.875 (a),
ordered coarse-grained carbide VC 0.875 (V 8 C 7 ) (b) and nanostructured carbide VC 0.875
(c) (examination in CuK 1,2 , number of counts is shown on the logarithmic scale)
[146, 155]. Superstructure reflections, observed in the X-ray diffraction patterns
(b) and (c), correspond to an ordered cubic (space group P4 3 32) phase V 8 C 7 . The
high intensity Isuper of the superstructure reflections, detected for the VC0.875 nanopowder,
and the anomalous increase of I super of the nanopowder in the range 2 >100 may
be the result of large static displacements of vanadium atoms.

larger (by 0.00047 nm) than the lattice constant of the disordered
carbide VC 0.875 . According to [140, 141, 148, 152] this large
difference of the lattice constants of the ordered and disordered
carbides VC 0.875 may be observed only at maximum or nearly
maximum degree of ordering.
The ratio of the intensities of the structural and superstructure
reflections confirms that the degree of long-range order is nearly
at a maximum in nanostructured vanadium carbide. In addition, this
relationships shows that the ordered phase occupies the entire volume
70

Synthesis of Nanocrystalline Powders

of the substance, i.e. the nanopowder consists of a single ordered


phase. Analysis of the ratio of the intensities yields another important
conclusion regarding the almost complete absence of oxygen in the
carbon sublattice; this is in agreement with the chemical analysis data
according to which the lattice contains only 0.1 wt.% of oxygen.
Attention should be given to the fact that the intensity of
superstructure reflections I super of the coarse-grained ordered
carbide VC 0.875 decreases with increasing diffraction angle 2
(Fig. 2.17b), whereas the intensity of superstructure reflections of
the VC 0.875 nanopowder in the range 2 > 100 not only does not
decrease but even increases (Fig. 2.17c). A possible reason for this
is the presence of large relaxation static displacements of vanadium
atoms in the vicinity of carbon vacancies. Earlier, these
displacements were actually detected in the ordered carbide V 8 C 7
[153].
Regardless of the nanometer thickness of the nanocrystals,
analysis of the width of diffraction reflections did not reveal any
large deviations from the instrumental width. Since all atoms inside
the crystal are scattered coherently, the absence of broadening of
the diffraction reflections is in agreement with the presence of a
relatively large number of atoms in the nanocrystallites because of
their large size in two other dimensions.
The pycnometric density of nanostructured VC 0.875 powder is
considerably lower than the theoretical density and is 5.15 g cm 3 .
This could have been caused by the presence in the specimen of
a substance with a low density (for example, adsorbed water) or
by the high imperfection of the metal sublattice determined by the
presence of vacant metal sites (metal structural vacancies). To
explain this, the pycnometric density was measured after heating
the carbide in vacuum. The removal of water at a temperature of
470 K resulted in a pycnometric density of 5.48 g cm 3 which is also
slightly lower than the theoretical value and is associated with the
presence of chemisorbed oxygen in the specimen. Only after
vacuum annealing at 900 K it is possible to achieve a pycnometric
density of 5.62 g cm 3 coinciding with the theoretical density. The
establishment of theoretical density as a result of annealing
indicates that the metal sublattice did not contain structural
vacancies and the reduced initial density is determined by adsorbed
impurities of water and oxygen. Although after annealing at 900 K
12 atomic layers of oxides remained on the surface of
nanocrystallites, the difference between the theoretical and
pycnometric densities cannot be detected because the content of the
71

Nanocrystalline Materials

Fig. 2.18. Position of vanadium V and carbon C atoms in the unit cell of the ordered
cubic (space group P4 3 32) phase V 8 C 7 : vacant (unfilled by carbon atoms) octahedral
interstitials of the metallic sublattice are indicated).

surface oxide phase is less than 0.1 wt.%, and the densities of the
oxides and vanadium carbide are similar.
The technique of magnetic susceptibility has been used
successfully for analysis of the structural state of weak magnetics
[160, 161]. This technique is highly sensitive to phase
transformations in non-stoichiometric carbides [140, 141, 144146]
and, consequently, it was used in the study of VC 0.875 nanopowder.
The magnetic susceptibility of the nanopowder of vanadium
carbide VC 0.875 was measured in the temperature range from 300
to 1200 K in a vacuum of 110 3 Pa after treatment in hydrochloric
acid for removal of vanadium oxide [157, 159].
The temperature dependence of the magnetic susceptibility of
the vanadium carbide (Fig. 2.19) is in good agreement with the data
[162, 163] for coarse-grained specimens. This confirms the slight
effect of the nanostructure of vanadium carbide on its electronic
properties.
The most effective and sensitive method of study of defects at
the interfaces and surfaces of nanoparticles is electron-positron
annihilation. Trapping of positrons by defects such as vacancies or
nanovoids leads to an increase in the positron lifetime in
72

Synthesis of Nanocrystalline Powders







 FP J


QDQR9&















Fig. 2.19. Temperature dependence of magnetic susceptibility of VC 0.875


nanopowder [153, 154].

comparison with the lifetime for a defect-free material [164]. The


value of the lifetime may be used to determine the type of defect.
In [150, 157159] the lifetime of positrons was measured on a
powder of carbide VC 0.875 subjected to preliminary heating at
400 K to remove water. For comparison, measurements were taken
of the positron lifetime in a coarse-grained sintered specimen of
vanadium carbide VC 0.875 . The measured positron lifetime spectra
are shown in Fig. 2.20. The spectra show that the mean positron
lifetime in the nanopowder is considerably longer than that in the
polycrystal. The spectrum of the coarse-grained specimen of the
vanadium carbide contains only a short component 157 2 ps which
corresponds to the annihilation of positrons in the structural
vacancies of the carbon sublattice [156, 157]. The quantitative
analysis of the spectrum of the nanocrystalline specimen shows that
in addition to the short component, equal to 157 2 ps, it contains
a long component 500 ps with a relative intensity of I 2 = 7%.
According to [164], the long component is determined by the
annihilation of positrons in defects on the surface of the particles.
Trapping of the positron by a structural vacancy indicates the
absence of diffusion of the positron over large distances; in this
case, the intensities of the components are proportional to the
volume fractions of the phases containing defects of different type.
Thus, the relative intensity of the long component, I 2, coincides with
73

Nanocrystalline Materials





&RXQWV DUELWUDU\ XQLWV







QDQRFU\VWDOOLQH 9 &






FRDUVHJUDLQHG 9 &







7LPH QV

Fig. 2.20. Positron lifetime spectra of nanocrystalline and coarse-grained vanadium


carbide V 8C 7 (VC 0.875) [146, 155].

the volume fraction of the surface V surf = S/V in the vanadium


carbide nanopowder. Estimates show that the thickness of the
surface layer is = 0.50.7 nm and corresponds to 34 atomic
monolayers.
Thus, it follows from the measurements of the positron lifetime that
the internal part of the nanocrystallites contains only non-metal
structural vacancies, and the surface layer of the nanocrystallites of
the vanadium carbide contains defects of the type of vacancy
agglomerates.
Authors [150, 157159] proposed the following model of the
structure of vanadium carbide nanocrystallites. The nanocrystallites
represent strongly bent plates-sheets 400600 nm in diameter and 1520 nm thick. The internal part of the nanocrystallites consists of
ordered carbide V 8C 7 with a high degree of the long-range order and
a negligible small content of dissolved oxygen. The surface layer of
the nanocrystallites contains 3.1 wt.% of chemisorbed oxygen and a
large number of vacancy agglomerates indicating to a loose structure
of a layer. The thickness of the surface phase does not exceed 0.7
nm or 4 atomic monolayers.
This morphology of the nanocrystalline powder of the nonstoichiometric vanadium carbide may be a consequence of cracking of
74

Synthesis of Nanocrystalline Powders

the particles at the interfaces between the disordered and ordered


phases. In fact, high temperature X-ray measurements [149] show that
at a temperature of 141320 K the orderdisorder VC 0.875 V 8 C 7
phase transition results in a sudden large increase of the lattice
constant of basic B1 structure by 0.0004 nm (Fig. 2.21a); the size of
the domains of the ordered phase is ~20 nm. According to [151, 152],
the ordering VC 0.875 V 8C 7 takes place via the mechanism of phase
transition of the first kind at a temperature of 1368 12 K; at 300
K constant a B1 of the basic crystal lattice of the quenched disordered

,




,

QP

9&  9&
















 .

, QP






9&

9&













&9 DWRP UDWLR

Fig. 2.21. Variation of the basic lattice constant a B1 (B1 cubic structure, space
group Fm3m ) of non-stoichiometric vanadium carbide as a result of ordering: a)
sudden change in the basic lattice constant a B1 at the order disorder V 8C 7 VC0.875
transition at a temperature of T trans = 1413 10 K [149]; b) the dependence of the
basic lattice constant a B1 on the composition of vanadium monocarbide VC y in the
disordered (full circle) and ordered (open circle) states at a temperature of 300 K
[151, 152].
75

Nanocrystalline Materials

carbide VC 0.875 is 0.0002 nm lower than that of the ordered carbide


with the same carbon content (Fig. 2.21b). The difference in the
volumes of the disordered and ordered phases leads to the formation
of stresses and subsequent cracking at the phase boundaries.
Another mechanism of formation of nanostructures is also
possible. The first kind of the disorderorder phase transition results
in the formation of domains of the ordered phase between which
stresses form as a result of mismatch of the atomic structure of
antiphase domains. With time, the stresses lead to cracking of the
grains of the initial disordered phase at the boundaries of the
antiphase domains of the ordered phase.
Special investigations are required for reliable explanation of the
mechanism of refining of the grains due to order-disorder
transformation in vanadium carbide.
Thus, the formation of the nanostructure of the powder of nonstoichiometric vanadium carbide VC 0.875 is determined by the
disorderorder VC 0.875 V 8 C 7 phase transformation taking place
in this material. The nanocrystallites of the ordered nonstoichiometric vanadium carbide in the form of distorted discs with
a diameter not exceeding 600 nm and a thickness of 1520 nm form
as a result of cracking of particles up to 1 m in size. The surface
layer of the nanocrystallites has thickness 0.50.7 nm and contains
chemisorbed oxygen and a large number of vacancy agglomerates.
This indicates its loose structure.
It may be assumed that the disorderorder transformations,
taking place with spasmodic volume changes, may be used for the
formation of the nanostructured state of other materials, including
strongly non-stoichiometric compounds and substitutional solid
solution .
2.8 SYNTHESIS OF HIGH-DISPERSED OXIDES IN LIQUID
METALS
This new method of synthesis of highly dispersed oxides is proposed
by the authors of [168]. The working medium is represented by
melts of gallium at a temperature of 323423 K, melts of lead at
653873 K or of a leadbismuth alloy at 453873 K.
Synthesis takes place in two stages. In the first stage, metal M
is dissolved in the melt. The chemical affinity of this metal for
oxygen is higher than that of the metal forming the melt. The
solubility of metal M in the melt should not be lower than 0.1 wt.%.
In the second stage, the dissolved metal M is oxidised by bubbling
76

Synthesis of Nanocrystalline Powders

the melt with water vapour or a gas mixture (H 2 O + Ar). The


content of the water vapours in the oxidation gas mixture is 15
30 vol.%. Below, as an example, we present the reactions taking
place in the process of oxidation of metal M in a gallium melt:
2Ga (liquid) + 3H 2 O (gas) = Ga 2 O 3 (solid) + 3H 2,

(2.6)

Ga 2 O 3 (solid) Ga 2 O 3 (dissolved) ,

(2.7)

x M (dissolved) + y Ga 2 O 3 (dissolved) = M x O 3y (amorphous) + 2y Ga (liquid) . (2.8)


Selective oxidation leads to the formation of amorphous highly
dispersed metal oxides. In particular, the selective oxidation of
aluminium, dissolved in gallium in an amount of 1 wt.%, results in
formation of flakes of an amorphous high-porosity substance with
chemical composition Al 2O 3H 2O. Examination of the microstructure
shows that the substance consists of fibres with a diameter of 5100 nm, oriented in one direction. The distance between the residual
fibres is 5400 nm. The produced nanomaterial has a porosity of
96.599.5 vol.% and a specific surface from 30 to 800 m 2 g 1 may
be used for thermal insulation. The stability of the composition,
structure and properties of the nanomaterial is retained during longterm annealing in the temperature range up to 1000 K.
Selective oxidation of indium, dissolved in lead, has made it
possible to produce indium oxide In 2 O 3 in the flaky form with a
macroparticle size of up to 5 mm. Examination by scanning electron
microscopy shows that flakes of In 2 O 3 consists of needles with a
diameter of 50100 nm spaced at 80200 nm.
The number of metals, whose solubility in melts of gallium, lead
or leadbismuth alloys is 0.1 wt.% or higher, is relatively large. By
means of selective oxidation of the system Ga (liquid) M (dissolved) at
a temperature of up to 423 K it is possible to produce highdispersion oxides Na 2 O, Al 2 O 3 , MgO and CaO. Selective oxidation
in the systems Pb (liquid) M (dissolved) and Pb/Bi (liquid) M (dissolved) at
temperatures of up to 873 K makes it possible to synthesise
nanosized oxides SbO 2 , TeO, NiO, GeO 2 , SnO 2 , In 2 O 3 , K 2 O, ZnO,
Ga 2 O 3 , Na 2 O, MnO, Li 2 O, Al 2 O 3 , BaO, SrO, MgO and CaO.
According to [168], the method developed is suitable for producing
highly dispersed nitrides, sulfides and halides in melts. In this case,
the melt with a dissolved metal should be subjected to the treatment
by an inert gas with nitrogen N 2 , hydrogen sulfide H 2 S or gaseous
chlorides of gallium or lead.
77

Nanocrystalline Materials

2.9. SELF-PROPAGATING HIGH-TEMPERATURE SYNTHESIS


The self-propagating high-temperature synthesis (SHS) represents
a rapid process of solid combustion of reagents (a metal and carbon
for carbides or a metal in nitrogen for nitrides) at a temperature
from 2500 to 3000 K [169]. Usually carbides are synthesized in a
vacuum or an inert atmosphere (argon). The mean size of grains
in carbides produced by the SHS method is 1020 mm, while the
size of nitride grains usually is smaller and equals 510 mm. SHSsynthesized carbides and nitrides of group IV and V transition
metals have, as a rule, an inhomogeneous composition and require
additional grinding and annealing for homogenisation. To decrease
the grain size in synthesized carbides or nitrides, the starting
mixture is diluted with the final product (for example, up to
20 mass % TiC carbide is added to the Ti + C mixture). For the
same purpose, some carbon in the mixture is replaced by polymers
(polystyrene, polyvinylchloride) during synthesis of carbides. As a
result, carbides and nitrides with grains 15 mm in size on the mean
can be synthesized.
It is proposed [169] to use an inert dilutant in SHS of nanosized
powders of titanium carbide. The inert dilutant is NaCl, which is
chemically inert to starting powders of metallic titanium Ti and
carbon C, as well as to the final titanium carbide TiC. During
combustion of titanium and carbon, NaCl, whose melting point
T melt = 1074 K, forms a melt, which prevents the growth of the
synthesized carbide particles. Moreover, NaCl readily dissolves in
water and can be easily separated from the synthesized carbide.
Titanium carbide is synthesized using a Ti + C + mNaCl mixture,
where m was 0.2 to 0.7 mole of sodium chloride per Ti or C atom.
The size of particles in the starting powders of Ti, C and NaCl did
not exceed 50, 0.1 and 150 mm respectively. Samples 30 mm
across and 40 mm long are pressed from the powder mixture. The
synthesis is realized in a standard reactor under an argon
atmosphere at a pressure of 0.5 MPa. Combustion is initiated by
a current pulse (the voltage U = 1520 V and the pulse time of
1.01.5 s), which is fed to the sample via tungsten wire. When the
quantity of sodium chloride, m, increased from 0.2 to 0.7, the
combustion temperature dropped from 2500 to 1950 K.
The starting mixture is homogenised spontaneously during
combustion thanks to melting of Ti and NaCl. Microscopic
examination shows that the titanium carbide particles are distributed
in the NaCl melt and are isolated one from another by a thin layer
78

Synthesis of Nanocrystalline Powders

of the melt which prevents their growth. No intermediate phases


form. The size of the titanium carbide particles decreases with
growing concentration of NaCl in the starting mixture. After NaCl
is washed off, the synthesized powder of titanium carbide is
analyzed by X-ray diffraction and electron microscopic methods.
The synthesized titanium carbide has a B1 cubic structure with the
lattice spacing of 0.433 nm. The powder includes irregularly shaped
particles 20 to 300 nm in size with an mean size of about 100 nm.
According to [169], the optimal concentration of NaCl for the
synthesis of titanium carbide nanoparticles is m = 0.4 or about
30 mass %. The chemical composition of the synthesized titanium
carbide is not determined in [169]. If the synthesized titanium
carbide is free of impurities, the comparison of the lattice constant
measured in [169] and an exact experimental dependence a B1 (y)
for TiC y [55, 140, 141, 158, 170] suggests that the carbide has a
nearly stoichiometric composition TiC 1.0 .

References
1.
V. Kohlschutter, C. Ehlers. Versuche uber Kondensation von Metalldampfen.
Z. Electrochem. 18, 373-380 (1912).
2.
V. Kohlschutter, N. Noll. Uber feine Metallzerteilungen. Z. Electrochem.
18, 419-428 (1912).
3.
L. Harris, D. Jeffries, B. M. Siegel. An electron microscopy study of gold
smoke deposits. J. Appl. Phys. 19, 791-794 (1948).
4.
M. Ya. Gen, M. S. Ziskin, Yu. I. Petrov. A study of the dispersion of aluminium
aerosols as dependent on the conditions of their formation. Doklady AN
SSSR 127, 366-368 (1959) (in Russian).
5.
5. M. Ya. Gen, E. A. Velichenkova, I. V. Eremina, M. S. Ziskin. On formation conditions and properties of AgCu alloy in the high-dispersed state.
Fizika Tverd. Tela 6, 1622-1626 (1964) (in Russian).
6.
6. M. Ya. Gen, I. V. Eremina, E. A. Fedorova. Producing of Fe-Co alloys
in high-dispersed state and investigation of their crystal structure. Fiz. Metall.
Metalloved. 22, 721-724 (1966) (in Russian).
7.
V. Tikhomirov, A. Lidov. Notes on electrolysis. Zh. Russk. Fiziko-Khimichesk.
Obschestva 15, 421-423 (1883) (in Russian).
8.
G. Bredig. Einige Anwendnungen des elektrischen Lichtbogens. Z. Electrochem.
4, 514-515 (1898).
9.
T. Svedberg. Colloid Chemistry: Wisconsin Lectures (Chemical Catalog Comp.,
New York 1924) 265 pp.
10.
M. J. Gen, J. L. Zelmanoff, A. J. Schalnikoff. Uber Herstellung kolloider
Losungen der Alkalimetalle. Kolloid-Zeitschrift 63, 263-268 (1933).
11.
M. Ya. Gen, Yu. I. Petrov. Dispersed condensates of metallic vapor. Uspekhi
Khimii. 38, 2249-2278 (1969) (in Russian).
12.
B. M. Smirnov. Clusters with close packing. Uspekhi Fiz. Nauk 162, 119138 (1992) (in Russian).
13.
V. I. Vladimirov, A. E. Romanov. Disclinations in Crystal (Nauka, Lenin79

Nanocrystalline Materials

14.
15.
16.
17.
18.
19.
20.

21.

22.

23.

24.
25.

26.

27.

28.

29.

30.
31.

grad 1986) 223 pp. (in Russian)


Yu. I. Petrov. Clusters and Small Particles (Nauka, Moscow 1986) 368 pp.
(in Russian)
J. Muhlbuch, E. Recknagel, K. Sattler. Inert gas condensation of Sb, Bi and
Pb clusters. Surface Sci. 106, 188-194 (1981).
Yu. I. Petrov. Absorption of light by small particles of Ag, Cu, Al and
Se. Optika i Spektroskopiya 27, 665-673 (1969) (in Russian).
K. Kimoto, I. Nishida. A study of lithium clusters by means of a mass
analyzer. J. Phys. Soc. Japan. 42, 2071-2072 (1977).
B. Gunther, A. Kampmann. Ultrafine oxide powders prepared by inert gas
evaporation. Nanostruct. Mater. 1, 27-30 (1992).
H. Hahn, R. S. Averback. The production of nanocrystalline powders by
magnetron sputtering. J. Appl. Phys. 67, 1113-1115 (1990).
G. Skandan, H. Hahn, J. C. Parker. Nanostructured yttria: synthesis and
relation to microstructure and properties. Scripta Metal. Mater. 25, 23892393 (1991).
S. Iwama, K. Hayakawa, T. Arizumi. Ultrafine powders of titanium nitride
and aluminium nitride produced by a reactive gas evaporation technique with
electron beam heating. J. Cryst. Growth. 56, 265-269 (1982); Growth of
ultrafine paarticles of transition metal nitrides produced by the reactive
gas evaporation technique with electron beam heating. J. Cryst. Growth.
66, 189-194 (1984)
M. S. El-Shall, W. Slack, W. Vann, D. Kane, D. Hanley. Synthesis of nanoscale
metal oxide particles using laser vaporization/condensation in a diffusion
cloud chamber. J. Phys. Chem. 98, 3067-3070 (1994).
M. S. El-Shall, D. Gravier, U. Pernisz, M. I. Baraton. Synthesis and characterization of nanoscale zinc oxide particles: I. Laser vaporization/ condensation technique. Nanostruct. Mater. 6, 297-300 (1995).
B. H. Kear, P. R. Strutt. Chemical processing and applications for
nanostructured materials. Nanostruct. Mater. 6, 227-236 (1995).
V. N. Troitskii, S. V. Gurov, V. I. Berestenko. Peculiarities of production
of highly dispersed powders of nitrides of IV group metals by reduction
of chloride in low temperature plasma. Khimiya Vysokikh Energii 13, 267272 (1979) (in Russian).
T. N. Miller. Plasmochemical synthesis and properties of powders of highmelting compounds. Izv. AN SSSR. Neorgan. Materialy 15, 557-562 (1979)
(in Russian).
T. Ya. Kosolapova, G. N. Makarenko, D. P. Zyatkevich. Plasmochemical
synthesis of high-melting compounds. Zh. Vsesouzn. Khimichesk. Obschestva
24, 228-233 (1979) (in Russian).
T. N. Miller, Ya. P. Grabis. Plasma chemical synthesis of high-melting nitrides.
In: Methods of Preparation, Properties and Application of Nitrides (Zinantne,
Riga 1980) pp.5-6 (in Russian).
T. N. Miller. Some properties of highly dispersed powders of high-melting nitrides. In: Nitrides: Methods of Preparation, Properties and Application. V.1 (Zinantne, Riga 1984) pp.8-9 (in Russian).
R. W. Chorley, P. W. Lednor. Synthetic routes to high-surface area nonoxide
materials. Advanced Mater. 3, 474-485 (1991)
R. Uyeda. Studies of ultrafine particle in Japan: crystallography. Methods of preparation and technological applications. Progr. Mater. Sci. 35,
1-96 (1991).

80

Synthesis of Nanocrystalline Powders


32.

33.

34.

35.

36.
37.

38.
39.

40.

41.

42.

43.

44.

45.
46.
47.
48.
49.

50.

V. F. Petrunin, V. A. Pogonin, L. I. Trusov, A. S. Ivanov, V. N. Troitskii.


Structure of ultrafine grained particles of titanium nitride. Izv. AN SSSR.
Neorgan. Materialy 17, 59-63 (1981) (in Russian).
V. F. Petrunin, Yu. G. Andreev, T. N. Miller, Ya. P. Grabis. Neutron diffraction analysis of ultrafine grained powders of zirconium nitride. Poroshkovaya
Metallurgiya No 9, 90-97 (1987) (in Russian).
V. F. Petrunin, Yu. G. Andreev, V. N. Troitskii, O. M. Grebtsova. Neutron diffraction investigation of niobium nitrides in ultrafine grained state.
Poverkhnost No 11, 143-148 (1982) (in Russian).
I. V. Blinkov, A. V. Ivanov, I. E. Orekhov. Synthesis of ultradispersed carbide
powders in pulsed plasma. Fizika i Khimiya Obrabotki Mater. No 2, 7376 (1992).
K. Kijima, H. Nogushi, M. Konishi. Sintering of ultrafine silicon carbide
powders prepared by plasma CVD. J. Mater. Sci. 24, 2929-2933 (1989).
F. Thevenot. Laboratory methods for the preparation of boron carbides.
In: The Physics and Chemistry of Carbides, Nitrides and Borides. Ed. R.
Freer. (Kluwer Academic Press, Netherlands, Dordrecht 1990) pp.87-96;
Boron carbide a comprehensive review. J. Eur. Ceram. Soc. 6, 205-225
(1990).
A. I. Gusev. Physical Chemistry of Nonstoichiometric Refractory Compounds
(Nauka, Moscow 1991) 286 pp. (in Russian).
N. V. Alekseev, A. V. Samokhin, E. N. Kurkin, K. N. Agafonov, Yu. V. Tsvetkov.
Synthesis of the alumina nanoparticles in the process of metal oxidation
in thermal plasma flow. Fizika i Khimiya Obrabotki Mater. No 3, 33-38
(1997) (in Russian).
J. S. Haggerty, W. R. Cannon. Laser-induced chemical processes. Sinterable
powders from laser-driven reactions. In: Laser-Induced Chemical Processes.
Ed. J. I. Steinfeld (Plenum Press, New York 1981) pp.165-241.
J. D. Casey, J. S. Haggerty. Laser-induced vapor-phase synthesis of boron and titanium diboride powders. J. Mater. Sci. 22, 737-744 (1987); Laserinduced vapor-phase synthesis of titanium dioxide. J. Mater. Sci. 22, 43074312 (1987).
R. A. Bauer, J. G. M. Becht, F. E. Kruis, B. Scarlet, J. Schoonman. Laser synthesis of low-agglomerated submicrometer silicon nitride powders
from chlorinated silanes. J. Amer. Ceram. Soc. 74, 2759-2768 (1991).
M. Suzuki, Y. Maniette, Y. Nakata, T. Okutani. Synthesis of silicon carbide silicon nitride composite ultrafine particles using a carbon dioxide laser.
J. Amer. Ceram. Soc. 76, 1195-1200 (1993).
N. V. Karlov, M. A. Kirichenko, B. S. Lukyanchuk. Macroscopical kinetics
of thermochemical processes at laser heating: state of art and perspectives.
Uspekhi Khimii 62, 223-248 (1993).
E. A. Rohlfing, D. M. Cox, A. Kaldor. Production and characterization of
supersonic carbon cluster beams. J. Chem. Phys. 81, 3322-3330 (1984).
H. W. Kroto, J. R. Heath, S. C. OBrien, R. F. Curl, R. E. Smalley. C 60 :
buckminsterfullerene. Nature 318, 162-163 (1985).
R. F. Curl, R. E. Smalley. Probing C 60 . Science 241, 1017-1022 (1988)
E. Osawa. Superaromaticity. Kagaku (Kyoto) 25, 854-863 (1970) (in Japan).
D. A. Bochvar, E. G. Galpern. Carbododecahedron, s-icosahedron, and carbos-icosahedron hypothetical systems. Doklady Akad. Nauk SSSR 209, 610612 (1973) (in Russian).
T. L. Makarova, B. Sundqvist, R. Hohne, P. Esquinazi, Ya. Kopelevich, P.
81

Nanocrystalline Materials

51.

52.

53.

54.

55.

56.
57.
58.

59.
60.
61.

62.

63.
64.
65.
66.
67.
68.

69.
70.

Schraff, V. A. Davydov, L. S. Kashevarova, A. V. Rakhmanina. Magnetic


carbon. Nature 413, 716-718 (2001).
L. Hultman, S. Stafstrom, Z. Czigany, J. Neidhardt, N. Hellgren, I. F. Brunell,
K. Suenaga, C. Colliex. Cross-linked nano-onions of carbon nitride in the
solid phase: Existence of a novel C 48 N 12 aza-fullerene. Phys. Rev. Lett. 87,
225503-1 225503-4 (2001).
T. Guo, M. D. Diener, Y. Chai, M. J. Alford, R. E. Haufler, S. M. Mclure,
T. Ohno, J. H. Weaver, G. E. Scuseria, R. E. Smalley. Uranium stabilization
of C 28 : a tetravalent fullerene. Science 257, 1661-1664 (1992).
Yu. N. Makurin, V. V. Sofronov, A. I. Gusev, A. L. Ivanovsky. Electronic
structure and chemical stabilization of C 28 fullerene. Chem. Physics 270,
293-308 (2001).
T. Guo, R. E. Smalley, G. E. Scuseria. Ab initio theoretical predictions of
C 28 , C 28 H 4 , C 28 F 4 , (Ti@C 28 )H 4 and M@C 28 (M = Mg, Al, Si, S, Ca, Sc,
Ti, Ge, Zr, and Sn). J. Chem. Phys. 99, 352-359 (1993).
A. I. Gusev. Phase equilibria, phases and compounds in the TiC system.
Uspekhi Khimii 71, 507-532 (2002) (in Russian). (Engl. transl.: Russ. Chem.
Reviews 71, 439-465 (2002)).
H. W. Kroto, A. W. Allaf, S. P. Balm. C 60 : buckminsterfullerene. Chem.
Rev. 91, 1213-1235 (1991).
A. V. Eletskii, B. M. Smirnov. Fullerenes and carbon structures. Uspekhi
Fiz. Nauk 165, 977-1009 (1995) (in Russian).
Fullerenes. Synthesis, Properties, and Chemistry of Large Carbon Clusters.
Eds. G. S. Hammond and V. J. Kuck. / ACS Symposium Series. V.481. (Amer.
Chem. Soc. Publ., Washington 1992) 195 pp.
B. C. Guo, K. P. Kerns, A. W. Castleman. Ti 8 C +12 -metallo-carbohedrenes:
a new class of molecular clusters? Science 255, 1411-1413 (1992).
Z. Lin, M. B. Hall. Theoretical studies of stability of M 8 C 12 clusters. J.
Amer. Chem. Soc. 115, 11165-11168 (1993)
B. C. Guo, S. Wei, J. Purnell, S. Buzza, A. W. Castleman. Metallo-carbohedrenes
[M 8 C 12 (M = V, Zr, Hf, and Ti)]: a class of stable molecular cluster ions.
Science 256, 515-516 (1992).
S. Wei, B. C. Guo, J. Purnell, S. Buzza, A. W. Castleman. Metallocarbohedrenes
as a class of stuble neutral clusters: formation mechanism of M 8 C 12
(M = Ti and V). J. Phys. Chem. 96, 4166-4168 (1992).
J. S. Pilgrim, M. A. Duncan. Photodissociation of metallo-carbohedrene (MetCars) cluster cations. J. Amer. Chem. Soc. 115, 4395-4396 (1993).
B. C. Guo, A. W. Castleman. Metallo-carbohedrenes: a new class of molecular clusters. Adv. Met. Semicond. Clusters 2, 137-164 (1994).
A. W. Castleman. Cluster reactions. Ann. Rev. Phys. Chem. 45, 685-719
(1994).
M. Methfessel, M. van Schilfgaarde, S. Scheffler. Electronic structure and
bonding in the metallocarbohedrene Ti 8C 12. Phys. Rev. Lett. 70, 29-32 (1993).
B. V. Reddy, S. N. Khanna, P. Jena. Electronic, magnetic and geometric structure
of metallo-carbohedrenes. Science 258, 1640-1643 (1992).
A. A. Sofronov, Yu. N. Makurin, M. V. Ryzhkov, A. L. Ivanovsky. Investigation of electronic structure and chemical bonds in Ti 8 C 12 compound. Zh.
Koordin. Khimii 25, 597-603 (1999) (in Russian).
S. N. Khanna, B. V. Reddy. Geometry, stability and properties of metallocarbohedrenes. Comput. Mater. Sci. 2, 638-642 (1994).
J. H. Weaver, J. L. Martins, T. Komeda, Y. Chen, T. R. Ono, G. H. Kroll,

82

Synthesis of Nanocrystalline Powders

71.
72.
73.
74.

75.
76.
77.

78.

79.
80.

81.

82.

83.
84.

85.

86.
87.

88.
89.

N. Troullier, R. E. Haufler, R. E. Smalley. Electronic structure of solid sixtyatom


carbon (C 60 ): experiment and theory. Phys. Rev. Lett. 66, 1741-1744 (1991).
S. Saito, A. Oshiyama. Cohesive mechanism and energy bands of a solid
of carbon sixty-atom molecules. Phys. Rev. Lett. 66, 2637-2640 (1991).
B. C. Guo, K. P. Kerns, A. W. Castleman. Reactivities of Ti 8 C +12 at thermal energies. J. Amer. Chem. Soc. 115, 7415-7418 (1993).
J. S. Pilgrim, M. A. Duncan. Metallo-carbohedrenes: chromium, iron, and
molybdenum analogues. J. Amer. Chem. Soc. 115, 6958-6961 (1993).
S. Wei, B. C. Guo, H. T. Deng, K. Kerns, J. Purnell, S. Buzza, A. W. Castleman.
Formation of metcars and face-centered cubic structures: thermodynamically or kinetically controlled? J. Amer. Chem. Soc. 116, 4475-4476 (1994).
S. Wei, B. C. Guo, J. Purnell, S. Buzza, A.W. Castleman. Metallo-carbohedrenes:
formation of multicage structures. Science 256, 818-820 (1992).
M. Faraday. Experimental relations of gold (and other metals) to light.
Philosoph. Trans. Roy. Soc. (London) 147, 145-181 (1857).
R. Rossetti, J. L. Ellison, J. M. Gibson, L. E. Brus. Size effects in the
excited electronic states of small colloidal CdS crystallites. J. Chem. Phys.
80, 4464-4469 (1984).
N. Herron, J. C. Calabrese, W. E. Forneth, Y. Wang. Crystal structure and
optical properties of Cd 32 S 14 (SC 6 H 5 ) 36 DMF 4 , a cluster with a 15 angstrom
cadmium sulfide core. Science 259, 1426-1428 (1993).
J. Kuczynski, J. K. Thomas. Surface effects in the photochemistry of colloidal
calcium sulfide. J. Phys. Chem. 87, 5498-5503 (1983).
Y. Wang, A. Suna, J. Mchugh, E. F. Hilinski, P. A. Lucas, R. D. Johnson.
Optical transient bleaching of quantum-confined CdS clusters: the effects
of surface-trapped electronhole pairs. J. Chem. Phys. 92, 6927-6939 (1990).
N. Herron, Y. Wang, M. M. Eddy, G. D. Stucky, D. E. Cox, K. Moller,
T. Bein. Structure and optical properties of cadmium sulfide superclusters
in zeolite hosts. J. Amer. Chem. Soc. 111, 530-540 (1989).
Y. Wang, N. Herron. Optical properties of cadmium sulfide and lead (II)
sulfide clusters encapsulated in zeolites. J. Phys. Chem. 91, 257-260 (1987);
Photoluminescence and relaxation dynamics of cadmium sulfide superclusters
in zeolites. J. Phys. Chem. 92, 4988-4994 (1988).
V. N. Bogomolov. Liquids in ultra-fine channels. Uspekhi Fiz. Nauk 124,
171-182 (1978) (in Russian).
A. R. Kortan, R. Hull, R. L. Opila, M. G. Bawendi, M. L. Steigerwald,
P. J. Carrol, L. E. Brus. Nucleation and growth of cadmium selenide on
zinc sulfide quantum crystallite seeds, and vice versa, in inverse micelle
media. J. Amer. Chem. Soc. 112, 1327-1332 (1990).
A. Haesselbarth, A. Eychmuller, R. Eichberger, M. Giersig, A. Mews, H. Weller.
Chemistry and photophysics of mixed cadmium sulfide/mercury sulfide colloids. J. Phys. Chem. 97, 5333-5340 (1993).
P. V. Kamat, B. Patrick B. Photophysics and photochemistry of quantized
zinc oxide colloids. J. Phys. Chem. 96, 6829-6834 (1992).
I. Bedja, P. V. Kamat. Capped semiconductor colloids. Synthesis and
photoelectrochemical behavior of TiO 2 capped SnO 2 nanocrystallites. J. Phys.
Chem. 99, 9182-9188 (1995).
G. Schmid. Chemical synthesis of large metal clusters and their properties. Nanostruct. Mater. 6, 15-24 (1995).
A. Bleier, R. Cannon. Nucleation and growth of uniform monoclinic zirconium dioxide. In: Better Ceramics Through Chemistry (MRS Symp. Proc.

83

Nanocrystalline Materials

90.

91.

92.

93.
94.

95.

96.

97.

98.
99.

100.

101.
102.
103.

104.
105.
106.

73 (1986)) Eds. C. J. Brinker, D. E. Clark and D. R. Ulrich (MRS, Pittsburg,


1986) pp.71-78.
G. Franz, G. Schwier. Starting materials for advanced ceramics production and properties. In: Raw Materials for New Technologies. Ed. M. Kursten
(Nagele and Obermuller, Stuttgart 1990) pp.139-158.
V. M. Chertov, V. I. Litvin, I. F. Mironyuk, V. V. Tsyirina. Synthesis and
texture of xerogels based on ultradispersed powders of alumina and aluminium monohydroxide. Neorgan. Materialy 29, 1019-1020 (1993) (in Russian).
F. Hatakeyama, Sh. Kanzaki. Synthesis of monodispersed spherical b-silicon
carbide powder by a sol-gel process. J. Amer. Ceram. Soc. 73, 2107-2110
(1990).
L. E. McCandlish, B. H. Kear, B. K. Kim. Processing and properties of
nanostructured WC-Co. Nanostruct. Mater. 1, 119-124 (1992).
L. Wu, J. Lin, B. K. Kim, B. H. Kear, L. E. McCandlish. Grain growth
inhibition of nanostructured WCCo alloys. In: Proc. of the 13 th Intern.
Plansee Seminar. V.3. Eds. H. Bildstein and R. Eck. (Metallwerk Plansee,
Reutte 1993) pp.667-675.
Z. Fang, J. W. Eason. Study of nanostructured WC-Co composites. In: Proc.
of the 13th Intern. Plansee Seminar. V.3. Eds. H. Bildstein and R. Eck.
(Metallwerk Plansee, Reutte 1993) pp.625-638
P. Seegopaul, L. E. McCandlish, F. M. Shinneman. Production capability
and powder processing methods for nanostructured WC-Co powder. Intern.
J. Refr. Met. Hard Mater. 15, 133-138 (1997).
N. I. Borisenko, V. A. Moldaver, A. V. Lebedev, N. V. Kobzarev. Application of nanosize powders in technology of hard alloys. In: Physics and
Chemistry of Ultra-Disperse (Nano-) Systems / Proc. of VI th All-Russian
Conference,August 19-23, 2002, Tomsk, Russia. (Moscow Physico-Technical Institute, Moscow 2002) pp.339-340.
I. D. Morokhov, L. I. Trusov, S. P. Chizhik. Ultra-Dispersed Metallic Substances (Atomizdat, Moscow 1977) 264 pp. (in Russian).
K. E. Gonsalves, P. R. Strutt, T. D. Xiao, P. G. Klemens. Synthesis of
silicon (carbide, nitride) nanoparticles by rapid laser polycondensation/
crosslinking reactions of an organosilazane precursor. J. Mater. Sci. 27, 32313238 (1992)
K. E. Gonsalves, K. T. Kembaiyan. Synthesis of advanced ceramics and
intermetallics from organometallic/polymeric precursors. Solid State Ionics
32/33, 661-668 (1989).
M. Peuckert, T. Vaahs, M. Bruck. Ceramics from organometallic polymers.
Advanc. Mater. 2, 398-404 (1990).
M. G. Mirabelli, A. T. Lynch, L. G. Sheddon. Molecular polymeric precursors to boron-based ceramics. Solid State Ionics 32/33, 655-660 (1989).
T. Wade, J. Park, E. G. Garza, C. B. Ross, D. M. Smith, R. M. Crooks.
Electrochemical synthesis of ceramic materials. 2. Synthesis of aluminum
nitride (AlN) and an AlN polymer precursor: chemistry and materials characterization. J. Amer. Chem. Soc. 114, 9457-9464 (1992).
G. W. Rice, R. L. Woodin. Zirconium boronitride as a zirconium boride
precursor. J. Amer. Ceram. Soc. 71, C181-C183 (1988).
E. G. Avvakumov. Mechanical Methods of Activation of Chemical Processes
(Nauka, Novosibirsk 1988) 305 pp. (in Russian).
Mechanical Alloying in Inorganic Chemistry. Ed. E. G. Avvakumov. (Nauka,
Novosibirsk 1991) 264 pp. (in Russian).

84

Synthesis of Nanocrystalline Powders


107.

108.
109.
110.

111.

112.

113.

114.

115.
116.
117.
118.

119.
120.

121.

122.

123.

124.

125.

Mechanical Alloying (Proc. Intern. Symp. On Mechanical Alloying, Kyoto,


Japan, May 7-10, 1991) Ed. P. H. Shingu. Materials Science Forum 8990 (1991) (Trans Tech Publications, Switzerland 1992) 816 pp.
A. R. Yavari, P. J. Desre, T. Benameur. Mechanically driven alloying of
immiscible elements. Phys. Rev. Lett. 68, 2235-2238 (1992).
H.-J. Fecht. Nanostructure formation by mechanical attrition. Nanostruct.
Mater. 6, 33-42 (1995).
J. Balogh, L. Bujdoso, G. Faigel, L. Granasy, T. Kemeny, I. Vincze, S. Szabo,
H. Bakker. Nucleation controlled transformation in ball milled FeB. Nanostruct.
Mater. 2, 11-18 (1993).
A. A. Popovich, V. N. Vasilenko. Mechanical alloying of high-melting compounds. In: Mechanical Alloying in Inorganic Chemistry. Ed. E. G. Avvakumov.
(Nauka, Novosibirsk 1991) pp.168-176 (in Russian).
A. A. Popovich, V. P. Reva, V. N. Vasilenko. Mechanochemical synthesis
kinetics and structure formation of refractory compounds. Neorgan. Materialy
28, 1871-1876 (1992) (in Russian).
V. Yu. Davydkin, L. I. Trusov, P. Yu. Butyagib, V. V. Moskvin, I. V. Kolbanev,
V. I. Novikov, S. S. Plotkin. Structure of high-melting carbides synthesized
by mechanical alloying. In: Mechanical Alloying in Inorganic Chemistry. Ed.
E. G. Avvakumov. (Nauka, Novosibirsk 1991) pp.183-185 (in Russian).
N. Atsumi, K. Yoshioka, T. Yamasaki, Y. Ogino. Nitriding of transition metals
by mechanical alloying in nitrogen gas. Funtai oyobi Funmatsu Yakin (J.
Japan. Soc. Powd. and Powd. Metall.). 40, 261-264 (1993) (in Japanese).
A. Teresiak, H. Kubsch. X-ray investigations of high energy ball milled
transition metal carbides. Nanostruct. Mater. 6, 671-674 (1995).
D. Oleszak, H. Matyja. Nanocrystalline Fe-based alloys obtained by mechanical
alloying. Nanostruct. Mater. 6, 425-428 (1995.
M. A. Xueming, J. I. Gang. Nanostructured WC-Co alloy prepared by
mechanical alloying. J. Alloys and Comp. 245, L30-L32 (1996).
M. S. El-Eskandarany, M. Omori, T. Kamiyama, T. J. Konno, K. Sumiyama,
T. Hirai, K. Suzuki. Mechanically induced carbonization for formation of
nanocrystalline TiC alloy. Sci. Reports of Res. Inst. Tohoku Univ. (Sendai,
Japan) 43, 181-193 (1997).
P. S. de Carli, J. C. Jamieson. Formation of diamond by explosive shock.
Science 133, 1821-1822 (1961).
L. F. Trueb. Microstructural study of diamonds synthesized under conditions
of high temperature and moderate explosive shock pressure. J. Appl. Phys.
42, 503-510 (1971).
S. A. Gubin, V. V. Odintsov, V. I. Pepekin. Diagramm of carbon phase states
and its account in calculation of detonation parameters. Khimich. Fizika
5, 111-120 (1986) (in Russian).
S. A. Gubin, V. V. Odintsov, V. I. Pepekin. Thermodynamic calculation of
ideal and nonideal detonation. Fizika Goreniya i Vzryva 23, 75-84 (1987).
(in Russian)
A. M. Staver, N. V. Gubareva, A. I. Lyamkin, E. A. Petrov. Ultrafine grained
diamond powders prepared with using of explosion energy. Fizika Goreniya
i Vzryva 20, 100-103 (1984) (in Russian).
A. I. Lyamkin, E. A. Petrov, A. P. Ershov, G. V. Sakovich, A. M. Staver,
V. M. Titov. Production of diamonds from explosives. Doklady AN SSSR
302, 611-613 (1988) (in Russian).
I. Yu. Malkov, L. I. Filatov, V. M. Titov, B. V. Litvinov, A. L. Chuvilin,
T. S. Teslenko. Diamond formation from liquid carbon phase. Fizika Goreniya
85

Nanocrystalline Materials

126.

127.

128.

129.
130.
131.

132.

133.

134.

135.

136.

137.

138.

139.

140.

141.

i Vzryva 29, 131-134 (1993) (in Russian).


A. E. Aleksenskii, M. V. Baidakova, A. Ya. Vul, V. Yu. Davydov, and Yu.
A. Pevtsova. Diamond-graphite phase transition in ultradisperse-diamond
clusters. Fizika Tverd. Tela 39, 1125-1134 (1997) (in Russian). (Engl. Transl.:
Physics of the Solid State 39, 1007-1015 (1997)).
V. L. Kuznetsov, M. N. Aleksandrov, I. V. Zagoruiko, A. L. Chuvilin, E.
M. Moroz, V. N. Kolomiichuk, V. A. Likholobov, P. M. Brylyakov, G. V.
Sakovich. Study of ultradispersed diamond powders obtained using explosive
energy. Carbon 29, 665-668 (1991).
M. Yoshikawa, Y. Mori, H. Obata, M. Maegawa, G. Katagiri, H. Ishida,
Ishitani. Raman scattering from nanometer-sized diamond. Appl. Phys. Lett.
67, 694-696 (1995).
M. Gamarnik. Energetical preference of diamond nanoparticles. Phys. Rev.
B54, 2150-2156 (1996).
A. L. Vereshchagin. Detonation Nano-Diamonds (Biisk Technological Institute of Altai State Technical Unversity, Barnaul 2001) 177 pp. (in Russian).
A. G. Beloshapko, A. A. Bukaemskii, A. M. Staver. Formation of ultrafine
grained compounds at impact wave loading of porous aluminium. Investigation of obtained particles. Fizika Goreniya i Vzryva 26, 93-98 (1990)
(in Russian).
A. G. Beloshapko, A. A. Bukaemskii, I. G. Kuzmin, A. M. Staver. Ultrafine
grained powder of stabilized zirconium dioxide. Synthesis by dynamic method.
Fizika Goreniya i Vzryva 29, 111-112 (1993) (in Russian).
M. M. Martynyuk. Importance of evaporation and boiling of liquid metal
in process of electric explosion of conductors. Zh. Tekhnich. Fiziki 44, 12621276 (1974).
Yu. A. Kotov, N. A. Yavorski. Study of particles forming at electric explosion of conductors. Fizika i Khimiya Obrabotki Mater. No 4, 24-29 (1978)
(in Russian).
V. A. Burtsev, N. V. Kalinin, A. V. Luchinskii. Electric Explosion of Conductors and its Application in Electrophysical Devices (Energoatomizdat,
Moscow 1990) 289 pp. (in Russian).
A. P. Ilin, O. B. Nazarenko, D. V. Tikhonov. Physical Principles of Preparing Dispersed Metals by Impulses of Electric Current. In: Ultrafine Grained
Powders, Nanostructures, Materials. Proc. of 2 nd Interdistrict Conf.,
Krasnoyarsk, October 5-7, 1999. (Krasnoyarsk State Technical University,
Krasnoyarsk 1999) pp.31-33 (in Russian).
A. P. Ilin. On excess energy of ultrafine grained powders prepared by electric
explosion of wires. Fizika i Khimiya Obrab. Mater. No 3, 94-97 (1994)
(in Russian).
A. P. Ilin. Features of energy saturation of a structure of small metal particles
prepared in strongly non-equilibrium conditions. Fizika i Khimiya Obrab.
Mater. No 4, 93-97 (1997) (in Russian).
Properties, Production and Application of Refractory Compounds / Handbook. Ed. T. Ya. Kosolapova (Metallurgiya, Moscow 1986) 928 pp. (in
Russian).
A. I. Gusev, A. A. Rempel. Nonstoichiometry, Disorder and Order in Solids
(Ural Division of the Russ. Acad. Sci., Yekaterinburg 2001) 580 pp. (in
Russian).
A. I. Gusev, A. A. Rempel, A. J. Magerl. Disorder and Order in Strongly
Nonstoichiometric Compounds: Transition Metal Carbides, Nitrides and Oxides

86

Synthesis of Nanocrystalline Powders

142.

143.

144.
145.
146.

147.

148.

149.

150.

151.

152.

153.
154.
155.

156.
157.

158.

159.

(Springer, Berlin - Heidelberg - New-York, 2001) 607 pp.


A. I. Gusev, A. A. Rempel. Phase diagrams of metalcarbon and metal
nitrogen systems and ordering in strongly nonstoichiometric carbides and
nitrides. Phys. Stat. Sol. (a) 163, 273-304 (1997)
A. I. Gusev. Orderdisorder transformations and phase equilibria in strongly
nonstoichiometric compounds. Uspekhi Fiz. Nauk 170, 3-40 (2000) (in Russian).
(Engl. transl.: Physics - Uspekhi 43, 1-37 (2000)).
A. I. Gusev, A. A. Rempel. Structural Phase Transitions in Nonstoichiometric
Compounds (Nauka, Moscow 1988) 308 pp. (in Russian).
A. A. Rempel. Effects of Ordering in Nonstoichiometric Interstitial Compounds (Nauka, Yekaterinburg 1992) 232 pp. (in Russian).
A. A. Rempel. Atomic and vacancy ordering in nonstoichiometric carbides.
Uspekhi Fiz. Nauk 166, 33-62 (1996) (in Russian). (Engl. transl.: Physics Uspekhi 39, 31-56 (1996)).
A. I. Gusev, A. A. Rempel. Superstructures of non-stoichiometric interstitial compounds and the distribution functions of interstitial atoms. Phys.
Stat. Sol. (a) 135, 15-58 (1993).
V. N. Lipatnikov, A. I. Gusev. Ordering of Titanium and Vanadium Carbides (Ural Division of the Russ. Acad. Sci., Yekaterinburg 2000) 265 pp.
(in Russian).
T. Athanassiadis, N. Lorenzelli, C. H. de Novion. Diffraction studies of
the order-disorder transformation in V8C7. Ann. Chim. France 12, 129-142
(1987).
A. A. Rempel, A. I. Gusev. Nanostructure and atomic ordering in vanadium carbide. Pisma v ZhETF 69, 436-442 (1999) (in Russian). (Engl. transl.:
JETP Letters 69, 472-478 (1999)).
V. N. Lipatnikov, W. Lengauer, P. Ettmayer, E. Keil, G. Groboth, E. J. Kny.
Effects of vacancy ordering on structure and properties of vanadium carbide. J. Alloys Comp. 261, 192-197 (1997).
V. N. Lipatnikov, A. I. Gusev, P. Ettmayer, W. Lengauer. Phase transformations in non-stoichiometric vanadium carbide. J. Phys.: Condens. Matter
11, 163-184 (1999)
D. Rafaja, W. Lengauer, P. Ettmayer, V. N. Lipatnikov. Rietveld analysis
of the ordering in V8C7. J. Alloys Comp. 269, 60-62 (1998).
A. I. Gusev. Nanocrystalline Materials:Preparation and Properties (Ural
Division of the Russ. Acad. Sci., Yekaterinburg 1998) 200 pp. (in Russian)
A. I. Gusev. Effects of the nanocrystalline state in solids. Uspekhi Fiz.
Nauk 168, 55-83 (1998) (in Russian). (Engl. transl.: Physics - Uspekhi 41,
49-76 (1998)).
A. I. Gusev, A. A. Rempel. Nanocrystalline Materials (Nauka-Fizmatlit,
Moscow 2000) 224 pp. (in Russian)
A. A. Rempel, A. I. Gusev, O. V. Makarova, S. Z. Nazarova. Nanocrystalline nonstoichiometric vanadium carbide with high microhardness. In:
Ultrafine Grained Powders, Nanostructures, Materials. Proc. of 2nd Interdistrict
Conf., Krasnoyarsk, October 5-7, 1999. (Krasnoyarsk State Technical
University, Krasnoyarsk 1999) pp.168-174 (in Russian).
A. A. Rempel, A. I. Gusev, O. V. Makarova, S. Z. Nazarova. Physical and
chemical properties of nanostructured vanadium carbide. Perspektivnye
Materialy No 9, 9-15 (1999) (in Russian).
A. I. Gusev, A. A. Tulin, V. N. Lipatnikov, A. A. Rempel. Nanostructure
of dispersed and bulk nonstoichiometric vanadium carbide. Zh. Obsh. Khimii

87

Nanocrystalline Materials

160.

161.

162.
163.

164.

165.

166.

167.

168.

169.

170.

72, 1067-1076 (2002) (in Russian). (English transl.: Russ. J. General Chem.
72, 997-1006 (2002)).
A. A. Rempel, A. I. Gusev, R. R. Mulyukov, N. M. Amirkhanov. Microstructure, microhardness and magnetic susceptibility of submicrocrystalline
palladium. Nanostruct. Mater. 7, 667-674 (1996).
A. A. Rempel, A. I. Gusev, S.Z. Nazarova, R. R. Mulyukov. Imputity
superparamagnetism in plastically deformed copper. Doklady Akad. Nauk
347, 750-754 (1996) (in Russian). (Engl. transl.: Physics - Doklady 41,
152-156 (1996)).
R. Caudron, J. Castaing, P. Costa. Electronic structure of face centred cubic
titanium and vanadium carbide alloys. Solid State Commun. 8, 621-625 (1970).
A. S. Borukhovich, N. M. Volkova. On conduction band of nonstoichiometric
vanadium monocarbide. Izv. AN SSSR. Neorgan. Materrialy 7, 1529-1532
(1971) (in Russian).
R. Wurschum, H.-E. Schaefer. Interfacial free volumes and atomic diffusion in nanostructured solids. In: Nanomaterials: Synthesis, Properties and
Applications. Eds. A. S. Edelstein and R. C. Cammarata. (Institute of Physics
Publishing, Bristol 1996) pp.277-301.
A. A. Rempel, M. Forster, H.-E. Schaefer. Positron lifetime in carbides with
B1 structure. Doklady Akad. Nauk SSSR 326, 91-97 (1992) (in Russian).
(Engl. transl.: Sov. Physics Doklady 37, 484-487 (1992)).
A. A. Rempel, M. Forster, H.-E. Schaefer. Positron lifetime in non-stoichiometric carbides with a B1 (NaCl) structure. J. Phys.: Condens. Matter 5, 261-266 (1993).
A. A. Rempel, L. V. Zueva, V. N. Lipatnikov, H.-E. Schaefer. Positron lifetime
in the atomic vacancies of nonstoichiometric titanium and vanadium carbides. Phys. Stat. Sol. (a) 169, R9-R10 (1998).
R. Sh. Askhadullin, P. N. Martynov. Synthesis of ultrafine oxides in nonalkaline liquid metals (Ga, Pb, Pb-Bi): Properties and application possibilities of substances obtained. In: Physics and Chemistry of Ultra-Disperse
(Nano-) Systems / Proc. of VI th All-Russian Conference. (Moscow PhysicoTechnical Institute, Moscow 2003) pp.451-455 (in Russian)
H. H. Nersisyan, J. H. Lee, C. W. Won. Self-propagating high-temperature
synthesis of nano-sized titanium carbide powder. J. Mater. Res. 17, 28592864 (2002).
L. V. Zueva, A. I. Gusev. Effect of nonstoichiometry and ordering on the
period of the basis structure of cubic titanium carbide. Fizika Tverd. Tela
41, 1134-1141 (1999) (in Russian). (Engl. Transl.: Physics of the Solid
State 41, 1032-1038 (1999))

88

Preparation of Bulk Nanocrystalline Materials

+D=FJAH!
3. Preparation of Bulk Nanocrystalline
Materials
Regardless of the large variety and development of the methods of
production of nanocrystalline materials (this relates in particular to
the most likely known methods of gas-phase evaporation and
precipitation from colloidal solutions), the investigations of the
structure and properties of nanoparticles are very complicated. This
is associated with, in particular, the high reactivity of the particles
resulting from their highly developed surface. Therefore, bulk
nanocrystalline materials are the subject of fundamental and applied
science. In many cases, these materials are more suitable for
investigation and application. The main methods of production of
bulk nanomaterials were described in a review [1]. None of these
methods is universal, because each of them is suitable for a limited
number of substances.
The conventional methods of powder technology are used most
widely, i. e. different types of pressing and sintering (cold pressing
and sintering, hot axial and isostatic pressing, magnetic pulsed
pressing). Powder technology also includes the method of vacuum
compacting of nanoparticles, produced by condensation from the
gas phase, proposed by H. Gleiter [25]. The main difficulty when
using powder technologies for production of pore-free or with least
porosity ware from nanopowders is associated with intensive
recrystallisation and the residual porosity. By shortening the duration
of the high temperature treatment, it is possible to decrease the
degree of recrystallisation and grain growth during sintering. The
application of high static or dynamic pressure for the pressing of
nanopowders at room or high temperatures reduces the residual
porosity and increases the relative density of produced materials.
Powder technology is suitable for chemical elements, compounds and
alloys.
The deposition of films and coatings makes it possible to produce
89

Nanocrystalline Materials

pore-free material. Films with a thickness lower than 100150 nm are


considered as nanocrystalline materials. Films are universal in
composition, and the size of crystallites in them can changes in a wide
range. There are amorphous, crystalline and multilayered films. This
offers great possibilities for using films in instrument industry and
electronics. In fact, regardless of the small thickness of the coating,
they greatly increase the mechanical properties of ware. For example,
coatings of titanium nitride TiN or titanium carbonitride TiC xN y greatly
increase the wear resistance and cutting properties of metal-working
tools, the corrosion resistance of metals and alloys. Films of different
composition are used widely in electronic microcircuits. Films and
coatings are produced by chemical (CVD) and physical (PVD) vapour
deposition from the gas phase, electrodeposition and by means of sol
gel technology.
Pore-free nanostructured materials may also be produced by
crystallisation from the amorphous state, but this method is suitable
only for alloys, which can be quenched from the melt to the
amorphous state. Crystallisation of amorphous alloys is carried out
at normal and high pressure, and is combined with deformation
treatment.
Severe plastic deformation makes it possible to produce porefree metals and alloys with a grain size of approximately 100 nm
and is suitable mainly for plastically deforming materials.
The formation of the nanostructure in non-stoichiometric
compounds such as carbides, nitrides and oxides MX y of transition
metals (M = Ti, Zr, Hf, V, Nb, Ta; X = C, N, O) and in
substitutional solid solutions A x B 1x is possible by means of atomic
ordering. This method is applicable if the disorderorder
transformation is the phase transition of the first kind and is
accompanied by an abrupt change of volume.
3.1. COMPACTION OF NANOPOWDERS
The method for production of bulk nanocrystalline materials,
proposed by the authors of [26] in 19821986, is well known and
used widely. The technology described in [26] uses the method
of evaporation and condensation for the production of
nanocrystalline particles, which are deposited on the cold surface
of a rotating cylinder. Evaporation and condensation are carried out
in the atmosphere of a rarefied inert gas, mostly helium He. At the
same gas pressure, the transition from helium to xenon, i. e. from
a less dense inert gas to an inert gas with a higher density, is
90

Preparation of Bulk Nanocrystalline Materials

accompanied by a large increase of the size of the particles.


Usually, the particles of the surface condensate are faceted. In the
same evaporation and condensation conditions, the metals with
higher melting points form smaller particles. The deposited
condensate is removed from the cylinder surface using a special
scraper and is collected in a collector. After removing the inert gas,
the nanocrystalline powder is compacted in a vacuum first at a
pressure of ~1 GPa and finally at a pressure up to 10 GPa
(Fig. 3.1). The setup for producing bulk nanocrystalline substances
by Gleiters methods [2, 5] is shown in Fig. 3.2. This setup is used
to produce sheets with a diameter of 515 mm, a thickness of 0.2
3.0 mm and a density equal to 7090% of the theoretical density
of the appropriate material. For example, the density of
nanocrystalline metals reaches 97%, and that of nanoceramic
materials up to 85% [7]. The bulk nanomaterials, produced by this
5RWDWLQJ FROG ILQJHU
OLTXLG QLWURJHQ
6FUDSHU

8+9 YDFXXP
FKDPEHU
,QHUW JDV
H J +H
(YDSRUDWLRQ
VRXUFHV
9DOYH

)XQQHO

9DFXXP SXPSV
%HOORZV
$QYLO

)L[HG SLVWRQ
6OLGH

6OHHYH
3LVWRQ

/RZ SUHVVXUH
FRPSDFWLRQ XQLW
+LJK SUHVVXUH
FRPSDFWLRQ XQLW
3LVWRQ

Fig. 3.1. Schematic representation of a gas-condensation chamber for synthesis


of nanocrystalline substances [5]. The material, which is evaporated or sputtered
from one or several sources, condenses to crystallites in a noble-gas atmosphere
and is transported via convection to a finger cooled down by liquid-nitrogen. The
clusters are scraped off, collected and compacted in two compaction units in situ
after evacuation.
91

Nanocrystalline Materials

Fig. 3.2. External view of setup for producing nanocrystalline substances by the
method of evaporation, condensation and compacting (Institut fr Theoretische
und Angewandte Physik, Universitt Stuttgart).

method, consists of particles with a mean diameter <D> from


12 nm to 80100 nm, depending on the evaporation and
condensation conditions.
One of the initial investigations, concerned with the production
of bulk nanomaterials, was carried out in 1983 by a group of
Russian authors using a powder of nanocrystalline nickel [8]. The
nickel powder with a mean particle size of 60 nm was produced by
the evaporation and condensation method. To produce bulk
specimens, the powder was pressed for 30 seconds at temperatures
from 673 to 1173 K and at a pressure up to 5 GPa. As a result of
short term heating, it was possible to retain the nanostructure in the
produced bulk specimens. It was reported in [8] that the hardness
of the bulk nanostructured specimens of nickel is considerable
higher than that of the coarse-grained nickel specimen.
The exclusion of contact with the environment in the production of
the nanopowder and during its pressing makes it possible to prevent
contamination of the bulk nanocrystalline specimens. This is very
important when investigating the nanocrystalline state of metals and
alloys. The setup described in [26] may be used for producing bulk
nanocrystalline oxides and nitrides. In this case, the metal evaporates
into an oxygen- or nitride-containing atmosphere. As an example,
Fig. 3.3 shows a bulk specimen of nanocrystalline oxide ZrO 2 ,
92

Preparation of Bulk Nanocrystalline Materials

Fig. 3.3. Bulk specimen of nanocrystalline oxide ZrO 2 prepared by evaporation,


condensation and compacting, i. e. by the method proposed by H. Gleiter [2, 5].
Specimen diameter is 5 mm, thickness is approximately 1 mm. The mean grain
size of the specimen is 20 nm. The specimen was prepared by Gregor Knner,
Ralf Rwe and Alexej Rempel (Institut fr Theoretische und Angewandte Physik,
Universitt Stuttgart, 2001).

produced by this method. The mean grain size in the specimen is 20


nm. The diameter of the specimen is 5 mm, the thickness
approximately 1 mm.
The porosity of nanoceramics, produced by compacting of
powders, is associated with triple points of the crystallites. A
decrease of the particle size of the powders is accompanied by a
large decrease of their compacting capacity in pressing using the
same pressure [9]. The decrease and the more uniform distribution
of porosity are achieved by pressing at higher temperature, which
do not yet result in intensive recrystallisation. For example,
conventional sintering of the highly dispersed powder of zirconium
oxide with a particle size of 4060 nm at 1370 K for 10 seconds
makes it possible to produce a relative density of 72% at a mean
grain size in the sintered specimen of 120 nm. Hot pressing at the
same temperature and a pressure of 1.6 GPa results in a sintered
material with a relative density of 87% and a mean grain size of
130 nm [10]. A decrease in sintering temperature to 1320 K and
an increase in sintering time to 5 hours enable production of bulk
zirconium oxide ZrO 2 with a relative density higher than 99% and
a mean grain size of 85 nm [11]. The authors of [12] used hot
pressing of titanium nitride powder (D ~ 80 nm) at 1470 K and a
93

Nanocrystalline Materials

pressure of 4 GPa to produce bulk specimens with a density of


98 % of theoretical density but (according to diffraction data), the
mean grain size after hot pressing was not smaller than 0.3 m,
because of intensive recrystallisation. Investigation [13] showed that
the densest (relative density 98%) specimens of titanium nitride are
produced by sintering specimens pressed from the finest
nanopowders (D ~ 825 nm) with a narrow size distribution of the
grains.
On the whole, to produce bulk nanocrystalline materials, especially
ceramic ones, it is promising to use pressing of nanopowders with
subsequent high temperature sintering. When using this method, it
is important to avoid growth of grains when sintering the pressed
specimens. This is possible at a high density of pressed specimens
(not lower than 0.7 of theoretical density), when the sintering
processes are relatively rapid and take place at a relatively low
temperature T 0.5T melt (T melt is the melting point). The production
of dense pressed specimens is an important task because the
nanocrystalline powders are difficult to press and the conventional
methods of static pressing do not result in sufficiently high density.
The physical reasons for the low pressing capacity of the
nanopowders is the existence of interparticle adhesion forces whose
relative value rapidly increases with decreasing particle size.
The application of the dynamic methods of compaction of the
nanopowders makes it possible to overcome the adhesion forces of
the particles and obtain, at the same pressure, more dense bulk
specimens in comparison with the static pressing conditions.
The magnetic pulsed method, proposed by the authors of [14
17] is used efficiently for compacting of nanocrystalline powders.
This method represents dry high-intensity pressing of the powders.
The magnetic pulsed pressing makes it possible to generate pulsed
compression waves with an amplitude of up to 5 GPa and a duration
of up to several microseconds. The method is based on
concentration of the force effect of the magnetic field of powerful
pulsed currents and allows to control quite easily the parameters of
the compression wave. The magnetic pulsed pressing does not led
to pollution of environment and is far safer than the dynamic
methods using explosive substances.
The principal scheme of uniaxial magnetic pulsed pressing is
shown in Fig. 3.4 [14]. The inductor (1) generates a pulsed
magnetic field B. The mechanical pulse of force F, compressing the
powder is generated as a result of the interaction of the pulse
magnetic field with the conducting surface of the concentrator (2).
94

Preparation of Bulk Nanocrystalline Materials


,







Fig. 3.4. Diagram of uniaxial magnetic pulsed pressing [17]: (a) compression stage,
(b) stage of ejection of the finished specimen, (1) inductor, (2) concentrator, (3)
upper and lower plungers, (4) powder, (5) die, (6) device for removing the specimen

The concentrator activates the upper plunger (3), which


compresses the powder. The displacement of the concentrator is
based on using of diamagnetic effect of pushing out of the
conductor from the region of the pulsed magnetic field. The die and
the specimen are placed in the vacuum chamber and all operations
with the powder are carried out in vacuum.
In contrast to the stationary pressing methods, the pulsed
compression waves are accompanied by rapid heating of the powder
as a result of rapid generation of energy during friction of particles
in the course of packing. If the size of the particles is relatively
small (D 0.3 micrometer), the duration of heating of the particles
by diffusion of heat from the particle surface is considerably shorter
than the characteristic duration of the pulsed compression wave (1
10 s). Under specific conditions, by optimising the parameters of
the compression wave, it is possible to carry out dynamic hot
compacting of the ultrafine powder as a result of the high powder
surface energy. At the same pressing pressure, the magnetic pulsed
method makes it possible to produce more dense specimens than
static pressing (Fig. 3.5). As an example, Fig. 3.6 shows the
variation of the pressing pressure, shrinkage rate and density of
95

Nanocrystalline Materials


QDQR$O2 



J FP




















5 *3D

Fig. 3.5. Dependence of the density of the nanocrystalline oxide n-Al 2 O 3 on


the pressure in stationary and magnetic pulsed pressing [16]: (1), (2), (3) stationary
pressing at 300, 620 and 720 K, respectively; (4) magnetic pulsed pressing.






'





'



*3D

P V

J FP
















9



Fig. 3.6. Dynamic parameters of magnetic pulsed pressing of nanocrystalline oxide


n-Al 2O 3 [16]: variation of pressing pressure P, shrinkage rate V and density during
the passage of a pulsed compression wave.

nanocrystalline oxide nano-Al 2 O 3 during the passage of a pulsed


compression wave [16]. Aluminium nitride AlN powder, produced
by means of electric explosion, is pressed by the magnetic pulsed
method under a pressure of 2 GPa to a density of 95% of the
96

Preparation of Bulk Nanocrystalline Materials

theoretical value, and Al 2 O 3 to a relative density of 86%. The


magnetic pulsed pressing is used for producing ware of different
shape, and in the majority of cases these ware do not require any
additional mechanical treatment. In particular, in operation with
superconducting oxide ceramics [15] it was possible to produce
ware with a density higher than 95% of the theoretical value. In
a general case, the application of pulsed pressure results in a higher
density of the specimens in comparison with static pressing as a
result of the effective overcoming forces of interparticle interaction
during rapid movement of the particles in powder. The short heating
time of the nanopowder allows reducing the recrystallisation at high
temperature and retaining the small particle size.
The magnetic pulsed method has been used for pressing
nanocrystalline Al 2 O 3 [18, 19] and TiN [20]. The results published
in [20] show that an increase of pressing temperature to ~900 K
is more efficient than an increase of pressure in cold pressing. At
a pulsed pressure of 4.1 GPa and a temperature of 870 K it is
possible to produce bulk specimens of nanocrystalline titanium
nitride with a grain size of ~80 nm and a density of approximately
83% of the theoretical value. A decrease in pressing temperature
to 720 K is accompanied by a decrease in density to 81%.
To produce gasdense ceramic pipes with an external diameter of
up to 15 mm and up to 100 mm long from nanopowders, the authors
of [21] used radial magnetic pulsed pressing. The powder is placed
in a cylindrical gap between a hard metallic bar and an external
cylindrical copper shell. Pressing is carried out by the radial
compression of the outer shell by pulsed current. The developed
pulsed pressure may reach 2 GPa. The starting material is
represented by nanopowders of Al 2 O 3 and Y 2 O 3 ZrO 2 with a mean
particle size of 1030 nm. As a result of radial magnetic pulsed
pressing of these nanopowders, it was possible to produce pipes
with a relative density of ceramics of more than 95%.
Dry cold ultrasound pressing [22, 23] is a promising and efficient
method for compacting ceramic nanopowders without the application
of plasticising agents. The treatment of the powder by powerful
ultrasound in the pressing process decreases the friction amongst
the particles and the friction of the powder with the walls of the
pressing mould, breaks up the agglomerate and large particles,
increases the surface activity of the powder particles and the
uniformity of distribution of these particles in the volume. This
increases the density of the pressed ware, accelerates diffusion
processes, restricts the grain growth in subsequent sintering and
97

Nanocrystalline Materials

Fig. 3.7. Ceramic ware produced by ultrasound pressing of nanopowders [24].

results in the retention of the nanostructure. For example, ultrasound


pressing of the nanopowder of ZrO 2 , stabilised with Y 2 O 3 oxide
with subsequent sintering of the specimens in air at a temperature
of 1923 K was used to produce ceramics with a relative density
~90%. The mean size of the particles in the starting powder was
~50 nm. The mean grain size in sintered ceramics depends on the
power of ultrasound oscillations during pressing: Increase in the
power of ultrasound from 0 to 2 kW decreases the mean grain size
from 440 to 200 nm.
Ultrasound pressing of the nanopowders is especially efficient
for producing ware of complicated shapes: sleeves, conical gears,
spirals, etc. (Fig.3.7) [24]. The produced ceramic ware are
characterised by a uniform microstructure and density.
Thus, there are several methods for compacting nanocrystalline
powders. These methods can be used to produce ceramic
specimens with high relative density and homogeneity. Subsequent
sintering of these ceramic specimens allows to retain their higher
density and, to a lesser degree, the nanostructure. In fact,
recrystallisation in nanomaterials takes place at a relatively high rate
and the crystallites and grains grow even at room temperature [25
27] (see also Section 6.1). Taking this into account, it is clear that
the role of sintering in producing nanostructured ceramics is very
98

Preparation of Bulk Nanocrystalline Materials

important. Conventional methods of sintering and their effect on the


evolution of the microstructure of the powder materials have been
discussed in [28, 29].
A new method of sintering ceramic materials by means of
superhigh frequency radiation is quite interesting [3035]. This
method is based on superhigh frequency heating of sintered
specimens. Heating takes place by radiation in the millimeter range
(frequency range from 24 to 84 GHz). The volume absorption of
superhigh frequency energy leads to uniform simultaneous heating
of the entire specimen because the heating rate is not limited by
heat conductivity, as in conventional sintering methods.
Consequently, sintered ceramics with a homogeneous microstructure
can be produced.
A processing gyrotron system for high temperature superhigh
frequency treatment of materials [30], developed at the Institute of
Applied Physics of the Russian Academy of Sciences in Nizhny
Novgorod, is shown in Fig. 3.8. The microwave energy source is
a continuous wave gyrotron (1) with a power of 10 kW at a

Fig. 3.8. Gyrotron system for microwave processing of materials [30, 33]: (1)
continuous wave gyrotron with a power of 10 kW at a frequency of 30 GHz, (2)
a microwave power transmission line transforming the gyrotron operating mode
into a Gaussian wave beam, (3) superhigh frequency radiation furnace as a supermultimode
cylindrical cavity (Institute of Applied Physics of the Russian Academy of Sciences,
Nizhny Novgorod, Russia).
99

Nanocrystalline Materials

frequency of 30 GHz. Electromagnetic radiation from the gyrotron


is shaped as a Gaussian beam and transferred by means of an open
quasi-optical line (2) into a superhigh frequency furnace (3). The
furnace is a supermultimode cylindrical cavity with a diameter of
50 cm, height 60 cm. Inside the cavity the radiation is uniformly
distributed by means of a spherical scatterer. The unifotm volume
distribution of microwave energy ensures uniform heating of the
material. The sintering temperature is 13002300 K and is regulated
with an accuracy of 0.2 K. The application of the gyrotron system
for sintering ceramic materials greatly reduces the risk of
overheating.
Microwave sintering of bulk specimens with a relative density
7080% and compacted from TiO 2 nanopowders with a mean
particle size of 2030 nm, makes it possible to produce sintered
specimens with a relative density of 9799% [31, 33, 34]. The
mean grain size of the sintered specimen is 200220 nm.
The conventional sintering methods do not always make it
possible to ensure strong joining of different ceramic materials. For
example, the conventional methods cannot be used to produce
mechanically strong joining between ZrO 2 and Al 2 O 3 , which is
required for developing devices of the thermal barrier type. The
application of nanocrystalline materials and of microwave sintering
enables this problem to be solved. The joining of ZrO 2 and Al 2 O 3
is ensured by means of sintered interlayer of nanosized composite
ceramic 60 vol.% ZrO 2 + 40 vol.% Al 2 O 3 with a mean grain size
of 100 nm. The relative density of the interlayer is 9698% of
theoretical density. Short term microwave heating of the ZrO 2 /
interlayer/Al 2O 3 assembly to 1700 K results in strong joining ZrO 2
and Al 2 O 3 oxides.
On the whole, the currently available methods of compacting
nanocrystalline powders and sintering compacted nanomaterials
already make it possible to produce high-density ceramic ware of
complicated shapes. However, the small grain size characteristic of
the starting nanopowders cannot be retained in sintered
nanomaterials. In the majority of sintered nanomaterials, the grain
size is 200300 nm, i. e. approximately 510 times larger than in
the starting nanopowders. To retain the small grain size, it is
necessary to decrease the sintering temperature, to shorten sintering
time, and carry out sintering at high dynamic or static pressure.

100

Preparation of Bulk Nanocrystalline Materials

3.2. FILM AND COATING DEPOSITION


Deposition on a cold or preheated substrate is used to produce films
or coatings, i. e. continuous pore-free layers of a nanocrystalline
material. In this method, in contrast to gas-phase synthesis,
nanoparticles form directly on the surface of the substrate and not
in the volume of the inert gas in the vicinity of the cooled wall. As
a result of the formation of the dense layer of the nanocrystalline
materials, it is not necessary to carry out pressing.
Deposition takes place from vapours, plasma or a colloid solution.
In vapour deposition, the metal is evaporated in vacuum, in oxygenor nitrogen-containing atmosphere, and vapours of metal or
compound (oxide, nitride) condense on the substrate. Varying the
evaporation rate and substrate temperature regulates the size of the
crystallites in the film. A film of zirconium oxide, alloyed with
yttrium oxide, with a mean crystallite size of 1030 nm was
produced by means of pulsed laser evaporation of metal in a beam
of oxygen ions and subsequent deposition of oxides on a substrate
heated to 350700 K [38].
In deposition from plasma, the electric discharge is sustained
using an inert gas. Varying the gas pressure and discharge
parameters regulates the continuity and thickness of the film and
the size of crystallites in the film. In deposition from plasma, a
metallic cathode with a high degree of ionisation (from 30100%)
is a source of metallic ions. The kinetic energy of ions is 10200
eV, deposition rate up to 3 m min 1 .
The authors of [39, 40], acting on chromium with plasma,
produced by arc discharge in low-pressure argon, deposited a
chromium film with a mean crystallite size of ~20 nm on a copper
substrate; a film with a thickness less than 500 nm had an
amorphous structure, and more thick films were in the crystalline
state. High hardness (up to 20 GPa) of the film is determined by
the formation of supersaturated solid solutions of interstitial
impurities (C, N) in chromium.
Ion-plasma coatings of titanium nitride and carbonitride are used
widely. Heating the substrate to 500800 K makes it possible to
retain the nanocrystalline structure of the coating. The methods of
production and properties of the coatings and films of refractory
compounds are discussed in detail in a review [41].
Deposition from plasma is carried out using mainly reactive
working media (mixtures of argon with nitrogen or hydrocarbons at
a pressure of ~0.1 Pa) and metallic cathodes. The main
101

Nanocrystalline Materials

disadvantage of ion-plasma arc sputtering is the formation of fine


metal droplets due to partial melting of the cathode and a possibility
of penetration of metallic droplets into deposited films.
Deposition from plasma is used to produce not only simple films
of nanometer thickness but also films with a nanostructure.
Recently, Fujimori et al [42] reported that CoAlO granular thin
films exhibit very large magnetoresistance in spite of their high
electrical resistivity. This unique property is related to the metal
oxide granular microstructure, which contains metallic nanoparticles,
embedded into the matrix of a non-metallic insulating oxide. Giant
magnetoresistance forms in the presence of superparamagnetism
and, consequently, the magnetic particles in the film must be in the
nanoscale range. To explain this, Ohnuma et al [43] investigated
a microstructure of films by means of high resolution transmission
electron microscopy and small angle X-ray scattering. Co-AlO
thin granular films were prepared on a glass substrate by Ar + O 2
reactive sputtering using a Co 72 Al 28 alloy target. Oxygen
concentration in the films was varied from 0 to 47 at.% by
controlling the partial pressure of oxygen in the gas mixture for
reactive sputtering. The results show that giant magnetoresistance
in the film appears when Co particles are fully surrounded by the
amorphous aluminium oxide. The microstructure of granular films
Co 61Al 26 O 13 and Co 52Al 20 O 28 is shown in Fig. 3.9. Light regions are
the amorphous aluminium oxide, and dark regions correspond to
metallic particles with a size of 23 nm. In the Co 52 Al 20 O 28 films,
the metallic particles consist of pure cobalt with an hcp or fcc
structure. In the Co 61 Al 26 O 13 films, containing a larger amount of
aluminium, metallic particles represent the CoAl phase with a CsCl

Fig. 3.9. HREM image of Co 61 Al 26 O 13 (a) and Co 52 Al 30 O 28 (b) films [43].


102

Preparation of Bulk Nanocrystalline Materials

structure. The value of the giant magnetoresistance changes quite


markedly in relation to the oxygen content in the film and is
maximum when the mean distance between the metallic
nanoparticles is minimum. Thus, regulating the deposition conditions
and the oxygen content in the Ar + O 2 gas mixture it is possible to
change the microstructure and properties of the CoAlO films.
A variant of deposition from plasma is magnetron sputtering
which uses not only cathodes of metals and alloys but also cathodes
produced from different compounds. In this case, the substrate
temperature can be reduced to 100200 K or lower. This widens
the possibilities of producing amorphous and nanocrystalline films.
However, the degree of ionisation, the kinetic energy of ions and
the deposition rate in magnetron sputtering are lower than when
using electric arc discharge plasma.
In [44] the method of magnetron sputtering of a Ni 0.75Al 0.25 target
and of deposition of metallic vapours on an amorphous substrate was
used to produce Ni 3Al intermetallic films with a mean crystallite size
of ~20 nm.
Oxide semiconductor films are produced by deposition from
colloid solutions on a substrate. This method includes the preparation
of the solution, deposition on the substrate, drying and annealing.
Deposition of nanoparticles of oxides was used to produce
semiconductor films of ZnO, SnO 2 , TiO 2 , WO 3 [4549].
Nanostructured films, containing nanoparticles of different
semiconductors, are produced by the co-precipitation method. The
preparation of nanocrystalline ZrO 2 films is described in [50].
Chemical and physical vapour deposition from the gas phase
(CVD and PVD) are conventional methods of depositing films.
These methods have been used for a very long time for producing
films and coatings for different applications. Usually, the crystallites
in these films are quite large but in multilayered or multiphase CVD
films it is also possible to produce nanostructures [41, 51].
Deposition from the vapour phase is usually associated with hightemperature gas reactions of metal chlorides in the atmosphere of
hydrogen and nitrogen or hydrogen and hydrocarbons. The
temperature range of deposition of CVD films is 12001400 K, the
deposition rate is 0.030.2 m min 1 , whereas laser radiation makes
it possible to reduce to 600900 K the temperature developed in
deposition from the gas phase. This leads to the formation of
nanocrystalline films.
Recently, organometallic precursors of the type of
tetradimethyl(ethyl)amids M[N(CH 3) 2]4 and M[N(C 2H 5) 2] 4 which have
103

Nanocrystalline Materials

a high vapour pressure are used often in deposition from the gas
phase. In this case, the dissociation of the precursor and activation of
the reagent gas (N 2 , NH 3 ) are carried out using electron cyclotron
resonance.
Films of transition metal nitrides, produced by different
deposition methods, are characterised by a superstoichiometric (in
comparison with phases MN 1.0 with a B1 basic structure) content
of nitrogen, for example Zr 3 N 4 , Hf 3 N 4 , Nb 4 N 5 , Ta 3 N 5 , etc.
3.3 CRYSTALLISATION OF AMORPHOUS ALLOYS
In this method, the nanocrystalline structure is produced in an
amorphous alloy by its crystallisation. Melt spinning, i.e. production
of thin ribbons of amorphous metallic alloys by means of rapid
cooling (at a rate 10 6 K s 1 ) of the melt on the surface of a
spinning disc or drum has been developed quite sufficiently. The
amorphous ribbon is annealed at a controlled temperature for
crystallisation. To develop a nanocrystalline structure, annealing is
carried out in such manner as to ensure the formation of a large
number of crystallisation centres and a low growth rate of the
crystals. The first stage of crystallisation may be the precipitation
of fine crystals of intermediate metastable phases. For example, the
authors of [52] in investigating an amorphous alloy of the NiP
system found that the initial stage is characterised by the formation
of small crystals of a metastable highly supersaturated solid solution
of phosphorous in nickel Ni(P) and crystals of nickel phosphides
appear only then. It is assumed that the amorphous phase is a
barrier for the growth of crystals.
Nanocrystalline ribbons can also be produced directly during melt
spinning. In [53], the spinning method is used to produce a ribbon
of Ni 65 Al 35 alloy. The ribbon consists of crystals of NiAl
intermetallic with a mean grain size of ~2m; in turn, these crystals
are characterised by a highly uniform microtwin substructure with
characteristic size of several tens of nanometers. This substructure
prevents the propagation of microcracks and increases the ductility
and toughness of the brittle intermetallic NiAl.
Figure 3.10 shows changes in the X-ray diffraction pattern of a
hard magnetic Fe 90 Zr 7 B 3 alloy in transition from the amorphous to
nanocrystalline state. The amorphous alloy is produced by melt
spinning and additionally subjected to relaxation annealing at a
temperature of 673 K for two hours. The nanocrystalline state is
produced by annealing at 873 K in a vacuum of 10 5 Pa for 1 hour.
104

&RXQWV DUELWUDU\ XQLWV

Preparation of Bulk Nanocrystalline Materials

QDQRFU\VWDOOLQH )H=U% DOOR\


DPRUSKRXV )H=U% DOOR\














 GHJUHHV

Fig. 3.10. Changes in the X-ray diffraction pattern of amorphous Fe 90 Zr 7 B 3 alloy


after annealing and transforming to nanocrystalline state (the grain size of the precipitated
bcc phase -Fe(Zr) is approximately 10 nm).

The grain size of the crystalline bcc phase -Fe(Zr), formed in the
amorphous phase, was determined by high resolution transmission
electron microscopy (HRTEM) and is equal to ~10 nm.
Crystallisation of amorphous alloys has been studied actively
because of the possibility of producing nanocrystalline
ferromagnetic alloys of FeCuMSiB (M = Nb, Ta, W, Mo, Zr)
systems with a low coercive force and high magnetic permeability,
i. e. soft magnetic materials. Examination of thin films of a NiFe
alloy is shown that soft magnetic properties improve with a
decrease in effective magnetocrystalline anisotropy [54]. This may
be achieved by increasing the number of grains taking part in
exchange interaction in thin magnetic films. In other worse, a
decrease in the grain size increases the exchange interaction,
decreases magnetocrystalline anisotropy and improves the soft
magnetic properties. This concept was later realised by experiments
in directional crystallisation of multicomponent amorphous alloys.
Soft magnetic materials are Si-containing steels and,
consequently, initial attempts to improve the soft magnetic properties
by means of crystallisation of amorphous alloys were carried on
alloys of the FeSiB system with copper additions. However, it
was not possible to produce alloys with a nanocrystalline structure
in this system. Only the addition to the FeSiB amorphous alloys
of group IVVII transition metals and copper and subsequent
105

Nanocrystalline Materials

Fig. 3.11. Nanocrystalline microstructure evolution of Fe 73.5 Cu 1 Nb 3 Si 13.5 B 9 alloy


by primary crystallization [56].

crystallisation enabled a nanocrystalline structure to be produced


[55]. An alloy with a homogeneous nanocrystalline structure was
produced by crystallisation of FeCuNbSiB amorphous alloys at
700900 K. In this alloy, grains of the bcc phase -Fe(Si) with a
size of ~10 nm and copper clusters smaller than 1 nm are uniformly
distributed in the amorphous phase.
Crystallisation of Fe 73.5 Cu 1 Nb 3 Si1 3.5 B 9 amorphous alloys is
investigated by high-resolution electron microscopy in [56]. The
consecutive change of the microstructure of the alloys during
crystallisation, based on these results, is shown schematically in
Fig. 3.11.
Preliminary (prior to crystallisation annealing) deformation by
rolling of FeCuNbSiB amorphous alloys or preliminary lowtemperature annealing enable a further decrease in the grain size
to ~5 nm [57, 58]. For example, cold rolling of Fe 73.5 Cu 1 Nb 3 Si1 3.5B 9
amorphous alloy up to a strain of ~6% and subsequent annealing
in vacuum at 813 K for 1 hour leads to the precipitation of
nanocrystalline grains of the bcc phase -Fe(Si) in the amorphous
phase; the mean grain size is ~68 nm. The mean grain size of the
nanocrystalline alloy subjected to only annealing at 813 K for 1 h
is 810 nm. Low-temperature annealing of Fe 73.5 Cu 1 Nb 3 Si1 3.5 B 9
amorphous alloy at a temperature 723 K for 1 hour combined with
subsequent short-term (for 10 seconds) high-temperature annealing
at 923 K results in a mean grain size of the bcc phase of 45 nm.
The decrease in the grain size of FeCuNbSiB alloy after
stepped annealing approximates the structure of this alloy to the
structure of a pure bulk nanocrystalline metal with a grain size of
25 nm, produced by compacting [26].
106

Preparation of Bulk Nanocrystalline Materials

Additional deformation or heat treatment, decreasing the grain


size, does not result in any changes in the phase composition of the
alloy. According to [58], this means that the phase composition of
Fe 73.5 Cu 1 Nb 3 Si1 3.5 B 9 alloy is finally formed in the last hightemperature stage of treatment. A decrease in the grain size of the
nanocrystalline phase as a result of preliminary deformation or heat
treatment is caused by the formation of additional crystallisation
centres in the amorphous phase.
Crystallisation of rapidly solidified amorphous aluminium alloys
AlCrCeM (M = Fe, Co, Ni, Cu), which contained more than
92 at.% Al, led to the formation of a structure containing an
amorphous phase and Al-rich icosahedral nanoparticles (D ~ 512
nm) precipitated in the amorphous phase [59]. As an example,
Fig. 3.12 shows an HRTEM image of a rapidly solidified
Al 94.5 Cr 3Ce 1 Co 1.5 alloy with dispersed precipitates of the icosahedral
phase; electron diffraction patterns are taken from several regions
with a diameter of 1 nm. It is interesting to note that the type of
electron diffraction pattern of the dispersed phase depends on the
size of the region in which diffraction of the accurately focused
electron beam takes place. For example, the electron diffraction
pattern obtained from a region of a diameter 1 nm, belonging to
region B, has an fcc structure (Fig. 3.12b), whereas the electron
diffraction patterns, obtained from a region with a diameter of 3

Fig. 3.12. High resolution TEM image (a) of a rapidly solidified Al 94.5 Cr 3 Ce 1 Co 1.5
alloy [59, 60]: icosahedral nanoparticles B, D and others are distributed in amorphous
matrix C; (b),(c) and (d) are nanobeam diffraction patterns taken from the regions
with a diameter of 1 nm marked B,C and D, respectively.
107

Nanocrystalline Materials

nm, show reflections corresponding to the symmetry axis of the 5 th


order. This means that the precipitated nanoparticles at distances
approximately 1 nm have a disordered structure (without the
symmetry of the 5 th order), and at distance of approximately 2 nm
or greater they have an icosahedral structure with long-range order
[60]. Al 94.5Cr 3 Ce 1 Co 1.5 alloys have extremely high tensile strength
(up to 1340 MPa), close to or higher than the strength of special
steels. The main reason for the high tensile strength is the formation
of spherical nanoparticles of the icosahedral phase and the
presence of a thin aluminium layer around these particles.
At present, the production of nanocrystalline alloys by
crystallisation from the quenched amorphous state is developing and
the number of alloys with the nanocrystalline structure produced by
this method rapidly increases.
3.4. SEVERE PLASTIC DEFORMATION
Severe plastic deformation is a very attractive method of producing
submicrocrystalline (or, which is the same, superfine grain)
materials with a mean grain size of 100 nm [6165]. This method
is based on the formation of a highly fragmented and misoriented
structure retaining the residual features of the recrystallised
amorphous state. The formation of this structure in metals and
alloys is a result of multiple severe shear plastic deformation with
in a true logarithmic degree of deformation of e = 47. Various
methods are used to achieve high deformation of the material:
rotating under quasi-hydrostatic pressure, equal channel angular
pressing, rolling, uniform forging. In addition to increasing the mean
grain size, the application of severe plastic deformation makes it
possible to produce bulk specimens without damage and residual
porosity. This cannot be achieved by compacting highly dispersed
powders.
Plastic deformation is an efficient means of the formation of the
structure of metals, alloys and some other materials. During
deformation, the dislocation density increases, the grains are
refined and the concentration of point defects and static faults
increases. The combination of these changes leads to the formation
of a specific microstructure. The main relationships governing the
formation of the structure during plastic deformation are determined
by the combination of parameters of the starting structural state of
the material and specific deformation conditions, and also by the
mechanics of the deformation process. With other conditions being
108

Preparation of Bulk Nanocrystalline Materials

Fig. 3.13. Schematic representation of severe plastic deformation setup [62]: (a)
high pressure torsion, (b) equal-channel angular pressing

equal, the main role in the formation of the structure and properties
of the material is played by the mechanics of the deformation
process: If the process ensures the uniform stress and strain states
throughout the entire volume of the material, the deformation
process is most efficient.
The main methods for creating high strains, which lead to
appreciable grain refining without failure of the specimens, are high
pressure torsion and equal channel angular pressing (Fig. 3.13.)
In high-pressure torsion, shear deformation is developed in discshaped specimens with a radius R and thickness l. The geometrical
form of the specimens is such that the main volume of the material
is deformed in the conditions of quasi-hydrostatic compression and
the specimen does not fail inspite of the high strain. The true
logarithmic degree of deformation e, obtained by high-pressure
torsion, is calculated from the equation [61]:
e = ln ( R/l),

(3.1)

where is the rotation angle in radians. The following equation is


109

Nanocrystalline Materials

used to calculate the degree of shear strain s at some point x


situated at distance R x from the axis of the specimen
s = 2N(R x /l),

(3.2)

where N is the number of rotations. To compare the degree of


shear strain in torsion with the strain in other deformation methods,
the value of s is usually converted to equivalent strain equiv =
s /3. Equation (3.2) shows that the strain should change in a linear
manner from 0 in the centre of the specimen to the maximum value
at the perimeter of the specimen. However, the results of a large
number of investigations show that after several revolutions, the
structure in the central part of the specimen is refined and, on the
whole, the structure is uniform throughout the specimen bulk [62].
The conventional methods of plastic deformation based on shear
(rolling, drawing, pressing, forging, torsion, etc.) make it possible
to obtain relatively high degrees of strain as a result of multiple
treatment but do not provide the uniform distribution of the
parameters of the stress and strain state. The formation of a uniform
structure is most efficient when using the stationary deformation
process based on a simple shear. The process is based on pushing
a workpiece through two channels with an equal cross section
intersecting under an angle of 2 = 90150 (Fig. 3.14). On the
z
P

P0

Fig. 3.14. Diagram of plastic deformation by the equal-channel angular pressing


[68]: is half of the angle of intersection of the channel, P is pressing pressure,
and P 0 the counter pressure from the side of the outlet channel
110

Preparation of Bulk Nanocrystalline Materials

plane of intersection of the channels, the uniform localised simple


shear deformation is concentrated with an intensity of
= 2cot.

(3.3)

Multiple cyclic treatment of the material by this method results in


superhigh strain intensity
= N = 2Ncot,

(3.4)

where N is the number of cycles. The produced material is in a


uniform stressstrain state, but the size of the workpiece does not
change. The true logarithmic degree of deformation is determined from
the equation
e = sinh 1(/2) = ln{(/2) + [(/2) 2 + 1] /2 }.

(3.5)

It is most efficient to use angle 2 close to 90 which leads to the


highest levels of deformation intensity at a small increase of contact
friction. A lubricant is used to minimise contact friction. This
deformation method, proposed by V. M. Segal [66] and developed
in [67, 68], is referred to as equal-channel angular pressing. In
comparison with other methods of plastic deformation, equal-channel
angular pressing has produced the most uniform submicrocrystalline
structure of material. Materials with such uniform structure possess
stable and reproducible physical properties. In a general case, the
structure of the material, produced by equal-channel angular
pressing, depends not only on the nature of material and the
magnitude of applied strain, but also on a number of technical
parameters such as the size and shape of the cross section of the
channels (the diagonal of the square section or the diameter of
circular channels), and the direction of passage of the workpiece
through the channels. If a material is difficult to deform, equalchannel angular pressing is carried at elevated temperatures.
Analysis of the results of investigation of the structure and
properties of submicrocrystalline materials has been carried out in
reviews [62, 64, 69].
The main feature of the structure of submicrocrystalline
materials, produced by the deformation methods, is the presence of
non-equilibrium grain boundaries which act as sources of high elastic
stresses. Triple points of the grains are another source of stresses.
The diffusion contrast of the boundaries and intragranular flexural
111

Nanocrystalline Materials

extinction contours, which are observed on TEM images of


submicrocrystalline materials, indicate that the grain boundaries are
non-equilibrium boundaries. According to different estimates, the
width of the intergranular boundaries in submicrocrystalline
materials varies from 2 to 10 nm. The non-equilibrium grain
boundaries contain large numbers of dislocations. Additionally, noncompensated disclinations are found in the grains boundaries. The
dislocation density in the submicrocrystalline materials, produced by
severe plastic deformation, is ~310 15 m 2 , and the power of
disclinations is 12. It should be noted that the density of the
dislocations inside the grains is considerably lower in comparison
with that at the boundaries. The dislocations and disclinations are
the reason for the excess energy of the grain boundaries and
generate long-range stress fields, which concentrate in the vicinity
of the grain boundaries and triple points. For example, the excess
energy of intergranular boundaries of submicrocrystalline copper
with a mean grain size of ~20 nm is almost 0.5 J m 2 .
Annealing of submicrocrystalline materials leads to the evolution
of their microstructure. Evolution can be conventionally divided into
two stages. In the first stage, as a result of annealing at temperature
about 1/3 rd of the melting point, stress relaxation takes place, the
grain boundaries transfer from the non-equilibrium to more
equilibrium state, and grains slowly grow. A further increase in
annealing temperature or an increase in annealing time results in
collective recrystallisation, i. e. grain growth.
The method of severe plastic deformation is used for producing
submicrocrystalline structures of such metals as Cu [7072], Pd
[7376], Fe [7779], Ni [70, 72, 8082], Co [83], alloys based on
aluminium [63], magnesium [84] and titanium [85, 86].
The authors of [72] noted different microstructures of Ni and Cu
produced by severe plastic deformation with the same magnitude
of deformation: in submicrocrystalline nickel, the size of the majority
of grains was approximately 100 nm, whereas the grain size in
submicrocrystalline copper was 5100 nm, and the copper grains
contained a larger number of defects (dislocations and twins) than
the grains of submicrocrystalline Ni. This means that in
submicrocrystalline nickel, the redistribution of dislocations into
energy favourable configurations (for example, rows of dislocations)
already takes place in the process of severe plastic deformation,
whereas in submicrocrystalline copper such redistribution does not
even start. The results of [72] show that the microstructure of any
material, produced by severe plastic deformations, should greatly
112

Preparation of Bulk Nanocrystalline Materials

differ in different stages of deformation. In addition to this, the


microstructure depends greatly on the type of deformation
(pressure, shear or torsion) and deformation parameters
(temperature, magnitude, rate and duration of deformation).
In fact, in order to understand the structure and properties of
submicrocrystalline materials, it is very important to take into
account the phase and structural transformations which take place
in these materials during heating and cooling, i.e. such phenomena
as recrystallisation, dissolving, and precipitation of a second phase,
etc. The threshold of the temperature stability of the submicrocrystalline structure depends on the state of intergranular boundaries
which, in turn, depends on the conditions of production of this
structure. The structure and recrystallisation of submicrocrystalline
materials also depend on the composition of the alloy and the type
of crystal lattice, but these problems have not been discussed in
sufficient detail.
Severe plastic deformation has been used for producing the
submicrocrystalline structure of not only metals, alloys and
intermetallic compounds with relatively high plasticity, but also
certain compounds with high brittleness. It is interesting that after
plastic deformation of a comparable magnitude, the grain size of
brittle compounds is smaller than that of metals. For example, in
[87, 88] torsion under quasi-hydrostatic pressure was used to
produce bulk nanocrystalline titanium carbide specimens (Fig. 3.15)
with a grain size of ~3040 nm from the coarse-grained (D ~ 2
5 m) powder of non-stoichiometric titanium carbide TiC 0.62 .
The formation of the submicrocrystalline structure by
deformation methods is accompanied by significant changes in the

Fig. 3.15. Compacted specimen of nanocrystalline titanium carbide TiC 0.62 with a
grain size of ~3040 nm, produced by severe plastic deformation (torsion under
quasi-hydrostatic pressure) of coarse-grained titanium carbide [87].

113

Nanocrystalline Materials

physical properties of metals, alloys and compounds. Metals with


a submicrocrystalline structure are suitable objects for experimental
investigation of the intercrystallite boundaries because they can be
examined by certified methods of metal physics and solid state
physics [77, 78].
3.5. DISORDERORDER TRANSFORMATIONS
In the conditions of thermodynamic equilibrium, the strongly nonstoichiometric carbides MC y and nitrides MN y may be in the
disordered or ordered state. The disordered state is
thermodynamically equilibrium at a temperature above 13001500
K and at a temperature below 1000 K only the ordered state [89
96] is thermodynamically equilibrium. However, the disordered state
of non-stoichiometric carbides and nitrides is easily retained at low
temperatures by quenching from high temperature. In this case, the
disordered state is metastable. The disorderorder phase
transformations in strongly non-stoichiometric compounds are usually
transitions of the first kind and are accompanied by an abrupt
change of the lattice constant. Consequently, disorderorder
transformations, i.e. ordering, can be used to produce a
nanostructure in non-stoichiometric compounds. The application of
ordering for producing a nanocrystalline powder of nonstoichiometric vanadium carbide VC 0.875 is described in Section 2.7.
However, using ordering, the nanostructure can also be produced
in bulk specimens of non-stoichiometric compounds.
Let us consider producing a nanostructure in bulk nonstoichiometric vanadium carbide VC 0.875 . The disordered carbide
VC 0.875 has a cubic B1 basic crystal structure. Bulk specimens of
VC 0.875 carbide were produced by hot pressing a powder of
disordered vanadium carbide VC 0.875 at a temperature of 2000 K in
a high-purity argon flow [97].
Examination of the surface of sintered disordered carbide VC 0.875
by the EDX method (Fig. 3.16a) showed the presence of vanadium
V, carbon C and impurities of metallic Ti, Nb and Ta. The presence
of Ti, Nb and Ta is the result of surface segregation of small
impurities during high-temperature sintering of the specimen. Earlier,
similar phenomena of surface segregation of a small impurity of
zirconium carbide ZrC were detected on sintered NbC specimens
containing approximately 1 mol.% ZrC [98, 99]. After polishing
sintered specimens of VC 0.875 with the removal of a surface layer
approximately 50m thick, the impurity disappeared (Fig. 3.16b).
114

Preparation of Bulk Nanocrystalline Materials


9
:

& 9

$O

1E

7L

7D :

: :

&RXQWV DUELWUDU\ XQLWV

& 9

$O

6L

& 9

.




NH9

Fig. 3.16. EDX spectra of the surface of the sintered specimen of vanadium carbide
VC 0.875 : (a) immediately after sintering the disordered carbide there are impurities
Ti, Nb, Ta and W on the surface; (b) grinding with the removal of the surface
layer approximately 50 m thick leads to the disappearance of impurities Ti, Nb,
Ta and W, the formation of traces of Al and Si associated with the use of AlSicontaining abrasive for grinding, (c) there are no impurities on the surface of the
annealed ordered carbide V 8 C 7 (VC 0.875), EDX spectra of quenched specimens also
have the same form and do not contain lines of impurity elements.

Sintered bulk specimens were treated under three thermal routes:


1) annealing at a temperature of 1370 K for two hours followed by
slow (at a rate of 100 K h 1 ) cooling to 300 K, 2) quenching from
1420 to 300 K, and 3) quenching from 1500 K to 300 K [97]. For
quenching the bulk specimens are put in quartz ampoules evacuated
to 10 3 Pa and are annealed at a maximum temperature (1420 or
1500 K) for 15 min. The ampoules with specimens are then
dropped in water; quenching rate was 100 K s 1 .Figure 3.17 shows
the surface of a bulk carbide VC 0.875 specimen after quenching from
1500 K. The grain boundaries of the basic cubic phase are clearly
visible. The grain size is 1060 m. The EDX spectra of annealed
(Fig. 3.16c) and quenched bulk specimens of VC 0.875 contain only
lines of main elements: vanadium V and carbon C.
The structure of annealed and quenched bulk specimens of the
VC 0.875 carbide was examined by X-ray diffraction. Both after
annealing and quenching the X-ray diffraction patterns contain
115

Nanocrystalline Materials

Fig. 3.17. Microstructure of the bulk specimen of carbide VC 0.875 after quenching
from 1500 K [97]. Sharp grain boundaries are due to their high susceptibility to
oxidation. The image obtained in a ISI-DS 130 electron microscope at a magnification
of 100

additional weak reflections, in addition to structural reflections


(Fig. 3.18). Judging by their position, additional reflections are
superstructure reflections and correspond to an ordered cubic V 8 C 7
phase with the P4 332 space group. Superstructure reflections of the
annealed and quenched specimens have nearly equal integral
intensity, but greatly differ in the width. The widest superstructure
reflections are observed in the X-ray diffraction pattern of the
specimen quenched from 1500 K.
According to the equilibrium phase diagram of the VC systems
[92, 94, 100] the formation of an ordered phase V 8 C 7 takes place
as a result of the disorderorder transformation at a temperature
of T trans = 1380 K; the experimental temperature of phase
transformation is 1413 20 K [101]. These data show that rapid
cooling from 1420 or from 1500 K must lead to quenching of the
disordered non-stoichiometric vanadium carbide VC 0.875 and saving
of the disordered VC 0.875 carbide as a metastable phase. However,
even if a specimen is quenched from 1500 K, an ordered V 8 C 7
phase appears and the relative intensity of superstructure reflections
is approximately equal to that for the specimens after quenching
from 1420 K or after annealing at 1370 K (see Fig. 3.18).
As a result of ordering, every grain of the disordered basic phase
breaks down into domains of the ordered phase. The degree of
116

Preparation of Bulk Nanocrystalline Materials



, DQQHDOLQJ %

B
 8  8



 

 .

 

 8

**

 8

 8



&RXQWV DUELWUDU\ XQLWV



 

- TXHQFKLQJ %  .

**
B
 8  8  8  8



 8

 

 8






. TXHQFKLQJ %  .

 

 


B
 8  8  8



 8










 GHJUHHV

Fig. 3.18. X-ray diffraction patterns of bulk VC 0.875 carbide specimens produced
by hot pressing and subjected to additional heat treatment (radiation CuK1,2, structural
reflections of B1 phase and superstructure reflections of phase V 8 C 7 are shown)
[97]: (a) annealing at 1370 K for 2 hours followed by slow cooling for 300 K; (b)
quenching from 1420 to 300 K at a rate of 100 K s 1 ; (c) quenching for 1500 to
300 K at a rate of 100 K s 1 . All X-ray diffraction patterns show superstructure
reflections of an ordered cubic (space group P4 332) phase V 8C 7; the largest broadening
of the superstructure reflections is seen on the X-ray diffraction pattern (c) of
specimen VC 0.875 quenched from 1500 K

ordering in a domain is high, while the mutual spatial distribution of


the domains is chaotic and depends on the ratio between the
structures of the ordered phase and the disordered matrix. Optical
microscopic examination at a magnification of 200 times shows that
the formation of the ordered phase starts at the grain boundaries
of the disordered phase. In the disordered carbide, the grains of
basic phase in the disordered carbide have sharp straight boundaries
(Fig. 3.19a), and after annealing these boundaries became broken
117

Nanocrystalline Materials

Fig. 3.19. Changes in the microstructure of a bulk non-stoichiometric vanadium


carbide VC 0.875 as a result of ordering: (a) in disordered carbide VC 0.875 the grains
have sharp straight boundaries, (b) in the ordered V 8 C 7 (VC 0.875 ) carbide the grain
boundaries of the basic phase appear to be eaten as a result of the formation of
domains of the ordered phase (image is obtained by V.N. Lipatnikov, Vienna University
of Technology, 1995).

as a result of ordering (Fig. 3.19b). It means that the domains of


ordered phase grow through the grains of the disordered basic
phase in the direction from the boundaries to the centre of the grain.
The small size of the domains and the same cubic symmetry of the
disordered and ordered phases do not allow the domain boundaries
to be examined by optical microscopy.
X-ray examination shows that the width of structural reflections
is independent of the thermal treatment conditions of the bulk
VC 0.875 specimens. Therefore, it may be assumed that the size of
grains of the disordered basic phase remains unchanged during
ordering. On the contrary, the superstructure reflections greatly
broaden. This broadening may be due to the small size of the
domains of the ordered phase which forms under different thermal
treatment conditions. Does a nanostructure appear in bulk specimens
118

Preparation of Bulk Nanocrystalline Materials

of VC 0.875 carbide subjected to thermal treatment and containing an


ordered V 8 C 7 phase? How small are the domains of the ordered
phase? These questions may be answered if one measures the
width of experimental diffraction reflections and compares it with
the instrumental width determined from the resolution function of
the diffractometer.
The resolution function of a Siemens D-500 X-ray
diffractometer is determined in a special diffraction experiment
using an annealed specimen of a stoichiometric tungsten carbide
with a grain size of 1020 m. The tungsten carbide has no
homogeneity range and there is no reflection broadening caused by
inhomogeneity. The size broadening of the diffraction reflections
does not take place with these large grains either. As a result of
annealing of the specimen there is no strain broadening. Thus, the
width of some diffraction reflections of the tungsten carbide
completely coincides with the resolution function R of the
diffractometer for the given diffraction angle . The dependence of
the second moment R of the resolution function of the Siemens D500 X-ray diffractometer on diffraction angle is shown in
Fig. 3.20.

H[S GHJUHHV



BBB

> WJ    @ OQ



TXHQFKLQJ  .
TXHQFKLQJ  .



DQQHDOLQJ  .

















GHJUHHV







Fig. 3.20. Broadening of diffraction reflections of bulk vanadium carbide VC 0.875


specimens after heat treatment and formation of the ordered phase V8C7 [97] (comparison
of the second moments exp of the experimental diffraction reflections with the
angular dependence of second moment R() of the resolution function of diffractometer):
( l ) annealing at 1370 K followed by slow cooling, ( ) quenching from 1420 K,
( ) quenching from 1500 K. Angular dependence of the resolution function of
diffractometer, R (), is shown by the solid line.
119

Nanocrystalline Materials

Comparison of the width of superstructure diffraction reflections


of bulk vanadium carbide specimens with the resolution function R
shows that experimental reflections broaden (Fig. 3.20). Since the
width of these reflections depends on the domain size and the
instrumental width, the size of domains can be determined from the
2
2R . To a first approximation,
measured broadening = 2.235 exp

it will be assumed that the strain broadening is absent and the


observed broadening is caused only by a small size of domains,
therefore = s . In turn, the size broadening s (2) = 2 s (),
measured in radians, is associated with the mean domain size <D>
by the equation

D =

(3.8)

where K hkl 1 is Scherrers constant whose value depends on the


shape of the domain (crystallite, particle) and on Miller indices
(hkl) of diffraction reflection; is the radiation wavelength.
Figures 3.18 and 3.20 show that the broadening of the
superstructure reflections is the largest for the specimen quenched
from 1500 K. The smallest broadening of the superstructure
reflections is observed for the vanadium carbide specimen annealed
at 1370 K. This means that domains of the ordered phase are the
smallest in the specimen quenched from 1500 K. In the annealed
vanadium carbide specimens, the size of the domains is the largest
because annealing and subsequent slow cooling create favourable
conditions for domain growth. The determination of broadening s
and subsequent evaluation of the mean domain size <D> of the
ordered phase using equation (3.8) gave the following results. In the
annealed specimens (Fig. 3.18a), the size of domains is 127 10 nm
with a probability of 95%; in the specimens quenched from
1420 K (Fig. 3.18b) and 1500 K (Fig. 3.18c) the size of domains
is equal to 60 9 nm and 18 12 nm with a probability of 95%.
Thus, annealing of bulk specimens of the non-stoichiometric
vanadium carbide VC 0.875 and quenching of these specimens from
a temperature of T trans 100 K cause the appearence of a
nanostructure consisting of domains of the ordered phase. The size
of domains increases with a decrease in annealing or quenching
temperature and with a decrease in the cooling rate.
The authors of [97, 102] also produced a bulk specimen of
120

Preparation of Bulk Nanocrystalline Materials

vanadium carbide from a nanopowder of non-stoichiometric ordered


carbide V 8 C 7 (VC 0.875 ). The method of production and the microstructure of the vanadium carbide nanopowder were descrided in
Section 2.7. The nanostructured vanadium carbide VC 0.875 powder
was compacted by the cold method at room temperature and a
pressure of 10 MPa. The density of compacts was 68% of the
theoretical density of vanadium carbide; this value is much larger
than the density of the nanopowder which is equal to 36%. Stepped
sintering of compacted specimen was carried out in a vacuum of
110 3 Pa at temperature from 400 to 2000 K with a step of 100
K; the holding time at each temperature was 2 hours. No large
change was detected in the density of the sintered specimen in
comparison with the density of the compact.
The relative variation of the mass m/m of the sintered specimen
as a function of sintering temperature is shown in Fig. 3.21. There
are three main stages of mass loss. The first stage (from room
temperature to 500 K) is associated with the loss of water during
heating. The second stage corresponds to the temperature interval
from 900 to 1500 K and is caused by the dissociation and removal
of the surface oxide phase thanks to the presence of free carbon
in the specimen. After sintering in this temperature range, the colour
of the specimen changed from black, determined by the surface
oxide phase, to grey corresponding to the pure vanadium carbide.
The third stage of the mass loss starts at a temperature of 1900
K and is associated with vacuum congruent evaporation of
vanadium carbide VC 0.875 which takes place without any change in
the composition of the specimen [90].


QDQR9&



22













     

Fig. 3.21. Relative variation of mass, m/m, of the VC 0.875 specimen, produced
from a nanopowder, in relation to vacuum sintering temperature T [97, 102].
121

Nanocrystalline Materials

In spite of the large porosity (approximately 30%) of the sintered


specimen, sintering allowed Vickers microhardness to be measured.
The sintered specimen was ground and microhardness
measurements were taken in automatic regime at a load of 200 g
and 500 g and loading time 10 s. Microhardness does not exhibit
any dependence on load within the measurement error. Microhardness H V was equal to 6080 GPa. For comparison, the
measured microhardness of a coarse-grained carbide VC 0.875
specimen, produced by hot pressing, is 21 GPa at a load of 0.1 kg.
According to the literature data [103], the microhardness of the
coarse-grained vanadium carbide is 29 GPa at a load of 0.1 and
0.2 kg. Thus, the microhardness of the specimen produced by
sintering of vanadium carbide nanopowder is 2 to 3 times higher
than the microhardness of the coarse-grained disordered vanadium
carbide and approaches that of diamond. Usually, at 300 K the
microhardness of nanomaterials is 27 times higher than H V of the
conventional polycrystalline materials [1, 104]. The high microhardness of the carbide VC 0.875 specimen, produced by sintering of
the nanopowder, may be explained by the HallPetch law
H V H 0 + kD 1/2 , i.e. H V is proportional to D 1/2 . Analysis of the
experimental data on the microhardness of bulk nanocrystalline
materials, published in [104], shows that the HallPetch law is
fulfilled when the grain size varies in the range 500 to 20 nm. The
examined carbide corresponds to this range of the nanocrystallite
sizes.
It should be noted that no nanostructure is found in the sintered
specimen of VC 0.875 . The study of sintered specimens in an ISIDS130 scanning electron microscope shows that its microstructure
represents a set of well sintered agglomerates with free spaces
between them (Fig. 3.22). This is consistent with the large porosity
of the sintered specimen. In order to determine whether the sintered
agglomerates have a nanostructure, additional high-resolution
transmission electron microscopic studies are necessary.
Diffraction investigations showed no superstructure reflections
in the X-ray diffraction patterns of the sintered specimen. This is
obvious because the maximum sintering temperature of 2000 K is
considerably higher than the temperature of the orderdisorder
phase transformation T trans . To obtain the ordered state in the
specimen, it is necessary to carry out thermal treatment in the
vicinity of temperature T trans .
The formation of the nanostructure in the bulk non-stoichiometric
vanadium carbide VC 0.875 is determined by the disorderorder
122

Preparation of Bulk Nanocrystalline Materials

Fig. 3.22. Scanning electron micrograph of the microstructure of a sintered vanadium


carbide nanopowder. Highly dense sintered conglomerates and the free space between
them with the size of up to tens of micrometers are clearly visible [97].

VC 0.875 V 8 C 7 phase transformation and by the appearance of


domains of the ordered phase. The size of the domains decreases
with an increase in the temperature of the start of thermal
treatment (quenching or annealing) and with an increase in the
cooling rate.
The discussed results together with the data presented in Section
2.7 indicate that the ordering is a new efficient method of producing
a nanostructure in bulk and powdered non-stoichiometric
compounds. The disorderorder transformations, which are accompanied by an abrupt change of the crystal volume, may be used for
producing a nanostructured state not only of strongly nonstoichiometric compounds but also of substitutional solid solutions,
including some alloys. The formation of a nanostructure in solid
solutions also takes place as a result of solid phase decomposition.
For example, it has been established that the decomposition of
carbide solid solutions (ZrC) 1x (NbC) x [105] leads to the formation
of a nanostructure with a grain size of approximately 70 nm. The
formation of a nanostructure in these solid solutions will be
considered in detail in Section 4.2 when discussing the determination
of the size of small particles from the broadening of diffraction
reflections.
123

Nanocrystalline Materials

References
1.

2.

3.
4.
5.
6.

7.
8.

9.
10.
11.
12.

13.
14.

15.

16.

17.

18.

A. I. Gusev. Effects of the nanocrystalline state in solids. Uspekhi Fiz.


Nauk 168, 55-83 (1998) (in Russian). (Engl. transl.: Physics - Uspekhi 41,
49-76 (1998))
H. Gleiter. Materials with ultra-fine grain size. In: Deformation of Polycrystals:
Mechanisms and Microstructures. Eds. N. Hansen, A. Horsewell, T. Leffers
and H. Lilholt (Ris Nat. Laboratory, Roskilde, Denmark 1981) pp.1521
H. Gleiter, P. Marquardt. Nanocrystalline structures on approach to new
materials? Z. Metallkunde 75, 263-267 (1984)
R. Birringer, U. Herr, H. Gleiter. Nanocrystalline materials: a first report.
Trans. Japan. Inst. Met. Suppl. 27, 43-52 (1986)
H. Gleiter. Nanocrystalline materials. Progr. Mater. Sci. 33, 223-315 (1989)
R. W. Siegel, H. Hahn H. Nanophase materials. In: Current Trends in Physics
of Materials. Ed. M. Yussouff. (World Sci. Publ. Co, Singapore 1987) pp.403420
R. W. Siegel. What do we really know about the atomic-scale structures
of nanophase materials? J. Phys. Chem. Solids 55, 1097-1106 (1994)
E. N. Yakovlev, G. M. Gryaznov, V. I. Serbin, V. N. Lapovok, L. I. Trusov,
V. Ya. Ganelin, E. V. Kapitanov, N. B. Kukhar , V. B. Begoulev. Production of nickel polycrystals with high hardness by compaction of ultradispersed
powders. Poverkhnost No 4, 138-141 (1983) (in Russian)
R. A. Andrievski. Material Science of Powder (Metallurgiya, Moscow 1991)
205 pp. (in Russian)
M. D. Matthews, A. Pechenik. Rapid hot-pressing of ultrafine PSZ powders.
J. Amer. Ceram. Soc. 74, 1547-1553 (1991)
D.-J. Chen, M. J. Mayo. Densification and grain growth of ultrafine 3 mol
% Y 2 O 3 -ZrO 2 ceramics. Nanostruct. Mater. 2, 469-478 (1993)
R. A. Andrievski, G. V. Kalinnikov, A. F. Potafeev, V. S. Urbanovich. Synthesis,
structure and properties of nanocrystalline nitrides and borides. Nanostruct.
Mater. 6, 353-356 (1995)
T. Rabe, R. Wsche. Sintering behaviour of nanocrystalline titanium nitride
powders. Nanostruct. Mater. 6, 357-360 (1995)
G. V. Ivanov, N. A. Yavorovskii, Yu. A. Kotov, V. I. Davydovich, G. A.
Melnikova. Self-propagating process of sintering of ultradispersed metallic powders. Doklady AN SSSR 275, 873-875 (1984) (in Russian)
V. V. Ivanov, S. N. Paranin, E. A. Gavrilin, A. V. Petrichenko, Yu. A. Kotov,
S. A. Lebedev, S. M. Cheshnitskii. Production of high-current superconducting ceramics Bi 1.6 Pb 0.4 Sr 2 Ca 2 Cu 3 O 10 by magnetic pulsed pressing.
Sverkhprovodimost: Fizika, Khimiya, Tekhnologiya 5, 1112-1115 (1992)
(in Russian)
V. V. Ivanov, S. N. Paranin, A. N. Vikhrev, A. A. Nozdrin. Efficienct of
dynamic method for densification of nano-sized powders. Materialovedenie No 5, 49-55 (1997) (in Russian)
V. V. Ivanov, A. A. Nozdrin, S. N. Paranin, S. V. Zayats. Device for magnetic
pulsed pressing of powders. In: Physics and Chemistry of Ultra-Disperse
(Nano-) Systems / Proc. of V th All-Russian Conference. V.1. (Ural Division
of the Russ. Acad. Sci., Yekaterinburg 2001) pp.229-233 (in Russian)
V. V. Ivanov, Yu. A. Kotov, O. M. Samatov, R. Bhme, H. U. Karow and
G. Schumacher. Synthesis and dynamic compaction of ceramic nanopowders
by techniques based on electric pulsed power. Nanostruct. Mater. 6, 287-

124

Preparation of Bulk Nanocrystalline Materials

19.

20.

21.

22.

23.

24.

25.

26.
27.

28.
29.

30.

31.

32.

33.

290 (1995)
V. V. Ivanov, A. N. Vikhrev, A. A. Nozdrin. Pressing of nanosized powders Al 2 O 3 at magnetic-pulsed loadng. Fizika i Khimiya Obrabotki Mater.
No 3, 67-71 (1997) (in Russian)
R. A. Andrievski, A. N. Vikhrev, V. V. Ivanov, R. I. Kuznetsov, N. I. Noskova,
V. A. Sazonova. Magnetic-pulse and high-pressure shear-strain compaction
of nanocrystalline titanium nitride. Fiz. Metall. Metalloved. 81, 137-145
(1996) (in Russian). (Engl. Transl.: Phys. Met. Metallogr. 81, 92-97 (1996))
V. V. Ivanov, S. N. Paranin, A. V. Nikonov, V. R. Khrustov, S. V. Dobrov.
The principles of magnetic pulsed pressing of long-mesured unit from nanosized ceramic powders. In: Physics and Chemistry of Ultra-Disperse (Nano) Systems / Proc. of VI th All-Russian Conference,August 19-23, 2002, Tomsk,
Russia. (Moscow Physico-Technical Institute, Moscow 2002) pp.156-157
(in Russian)
O. L. Khasanov, E. S. Dvilis, Yu. P. Pokholkov, V. M. Sokolov. Mechanisms of ultrasonic pressing of ceramic nanopowders. Perspektivnye Materialy
No 3, 88-94 (1999) (in Russian). (Engl. transl.: J. Advanc. Mater. 5, 6975 (1999))
O. L. Khasanov, V. M. Sokolov, E. S. Dvilis, Yu. P. Pokholkov. Ultrasonic
technology for production of structural and functional ceramics. Perspektivnye
Materialy No 1, 76-83 (2002) (in Russian)
O. L. Khasanov, E. S. Dvilis, V. M. Sokolov, Yu. P. Pokholkov. Technique
of compaction of ceramics with complicated form from nanopowders. In:
Physics and Chemistry of Ultra-Disperse (Nano-) Systems / Proc. of VI th
All-Russian Conference,August 19-23, 2002, Tomsk, Russia. (Moscow PhysicoTechnical Institute, Moscow 2002) pp.194-195 (in Russian)
B. Gunther, A. Kumpmann, H.-D. Kunze. Secondary recrystallization effects in nanostructured elemental metals. Scripta Metall. Mater. 27, 833838 (1992)
V. Y. Gerstman, R. Birringer. On the room-temperature grain growth in
nanocrystalline Cu. Scripta Metall. Mater. 30, 577-581 (1994)
J. Weissmller, J. Lffler, M. Kleber. Atomic structure of nanocrystalline
metals studied by diffraction techniques and EXAFS. Nanostruct.Mater. 6,
105-114 (1995)
Ya. E. Geguzin. Physics of Sintering (Nauka, Moscow 1984) 312 pp. (in
Russian)
V. V. Skorokhod, Yu. M. Solonin, I. V. Uvarova. Chemical, Diffusion and
Rheology Processes in Technology of Powder Materials (Naukova Dumka,
Kiev 1990) 247 pp. (in Russian)
Yu. Bykov, A. Eremeev, V. Flyagin, V. Kaurov, A. Kuftin, A. Luchinin, O.
Malygin, I. Plotnikov, V. Zapevalov. The gyrotron system for ceramics sintering.
Ceramics Trans. 59, 133-140 (1995)
Yu. Bykov, A. Eremeev, S. Egorov, V. Ivanov, Yu. Kotov, V. Khrustov, A.
Sorokin. Sintering of nanostructural titanium oxide using millimeter-wave
radiation. Nanostruct. Mater. 12, 115-118 (1999)
Yu. V. Bykov, A. G. Eremeev, V. V. Holoptsev, C. Odemer, A. I. Rachkovskii,
H.-J. Ritzhaupt-Kleissl. Sintering of piezoceramics using millimeter-wave
radiation. Ceramics Trans. 80, 321-328 (1997)
Yu. V. Bykov, S. V. Egorov, A. G. Eremeev, K. I. Rybakov, V. E. Semenov,
A. A. Sorokin. Microwave processing of nanostructured and functional gradient
materials. In: Functional Gradient Materials and Surface Layers Prepared

125

Nanocrystalline Materials

34.

35.

36.
37.
38.

39.

40.

41.

42.
43.

44.

45.

46.

47.

by Fine Particles Technology. Eds. M.-I. Baraton, I. V. Uvarova. (Kluwer


Academic Publishers, Netherlands, Dordrecht 2000) pp.135-142
Yu. V. Bykov, S. V. Egorov, A. G. Eremeev, A. A. Sorokin, V. V. Holoptsev.
Microwave sintering of nanostructured ceramic materials. In: Physics and
Chemistry of Ultra-Disperse (Nano-) Systems / Proc. of V th All-Russian
Conference. V.2. (Ural Division of the Russ. Acad. Sci., Yekaterinburg 2001)
pp.14-19 (in Russian)
Yu. Bykov, V. Holoptsev, Y. Makino, S. Miayke, I. Plotnikov, T. Ueno.
Kinetics of densification and phase transformation at microwave sintering
of silicon nitride with alumina and yttria or ytterbia as additives. J. Japan. Soc. Powd. Powder Metallurgy 48, 558-564 (2001)
H. J. Hfler, H. Hahn, R. S. Averback. Diffusion in nanocrystalline materials. Defect and Diffusion Forum 75, 195-210 (1991)
S. Okada, F. Tany, H. Tanumoto, Y. Iwamoto. Anelasticity of ultrafine-grained
polycrystalline gold. J. Alloys Comp. 211-212, 494-497 (1994)
Yu. A. Bykovskii, V. P. Kozlenkov, Yu. B. Krasilnikov, I. N. Nikolaev.
Preparation of ZrO 2-Y 2 O 3 films by laser evaporation of metals in an oxygen
ion beam. Poverkhnost No 12, 69-73 (1992) (in Russian)
D. A. Dudko, V. G. Aleshin, A. E. Barg, N. V. Dubovitskaya, L. N. Larionov,
A. A. Smekhnov. On the nature of high hardness of vacuum-plated chromium. Doklady AN SSSR 285, 106-109 (1985) (in Russian)
A. E. Barg, V. N. Dubovitskaya, D. A. Dudko, L. N. Larikov. Formation
of amorphous phase based on chromium by ion-plasma deposition.
Metallofizika 9, 118-119 (1987) (in Russian)
R. A. Andrievski. The synthesis and properties of interstitial phase films.
Uspekhi Khimii 66, 57-77 (1996) (in Russian). (Engl. Transl.: Russ. Chem.
Reviews 66, 53-72 (1996))
H. Fujimori, S. Mitani, S. Ohnuma. Tunnel-type GMR in metal-nonmetal
granular alloy thin films. Mater. Sci. Eng. B 31, 219-223 (1995)
M. Ohnuma, K. Hono, H. Onodera, J. S. Pedersen, S. Mitani, H. Fujimori
H. Distribution of Co particles in Co-Al-O granular thin films. In: Advances
in Nanocrystallization. Proceedings of the Euroconference on Nanocrystallization
and Workshop on Bulk Metallic Glasses (Grenoble, France, April 21-24,
1998). Ed. A. R.Yavari. (Trans Tech Publications, Switzerland 1999). Materials
Science Forum 307, 171-176 (1999) / J. Metastable Nanocryst. Mater. 1,
171-176 (1999)
H. van Swygenhoven, P. Bni, F. Paschoud, M. Victoria, M. Knauss M.
Nanostructured Ni 3Al produced by magnetron sputtering. Nanostruct. Mater.
6, 739-742 (1995)
S. Hotchandani, P. V. Kamat. Charge-transfer processes in coupled semiconductor systems. Photochemistry and photoelectrochemistry of the colloidal
CdS - ZnO system. J. Phys. Chem. 92, 6834-6839 (1992)
I. Bedjia, S. Hotchandani, P. V. Kamat. Photoelectrochemistry of quantized
WO3 colloids. Electron Storage electrochromic and photoelectrochromic effects.
J. Phys. Chem. 97, 11064-11070 (1993); Preparation and photoelectrochemical
characterization of thin SnO 2 nanocrystalline semiconductor films and their
sensitization. J. Phys. Chem. 98, 4133-4140 (1994)
B. ORegan, M. Grtzel, D. Fitzmaurice. Optical electrochemistry. 1. Steadystate spectroscopy of conduction-band electrons in a metal oxide semiconductor
electrode. Chem. Phys. Letters 183, 89-93 (1991); Optical electrochemistry.
2. Real-time spectroscopy of conduction-band electrons in a metal oxide

126

Preparation of Bulk Nanocrystalline Materials

48.

49.

50.

51.

52.

53.
54.
55.

56.

57.

58.

59.

60.
61.

62.
63.
64.

semiconductor electrode. J. Phys. Chem. 95, 10525-10528 (1991)


H. Yoshiki, K. Hashimoto, A. Fujishima. Reaction mechanism of electroless
metal deposition using ZnO thin film (I): process of catalys formation. J.
Electrochem. Soc. 142, 428-432 (1995)
L. Kavan, T. Stoto, M. Grtzel, D. Fitzmaurice, V. Shklover. Quantum size
effects in nanocrystalline semiconducting TiO 2 layers prepared by anodic
oxidative hydrolysis of TiCl 3 . J. Phys. Chem. 97, 9493-9498 (1993)
K. Yamada, T. Y. Chow, T. Horihata, M. Nagata. A low-temperature synthesis
of zirconium oxide coating using chelating agents. J. Non-Cryst. Solids 100,
316-320 (1988)
R. A. Andrievski. The synthesis and properties of nanocrystalline refractory
compounds. Uspekhi Khimii 63, 431-448 (1994) (in Russan). (Engl. transl.:
Russ. Chem. Reviews 63, 411-428 (1994))
K. Lu, J. T. Wang, W. D. Wei. Thermal expansion and specific heat capacity of nanocrystalline Ni-P alloy. Scripta Metal. Mater. 25, 619-623
(1991)
T. Cheng. Nanometer substructures and its effects on ductility and toughness
at room temperature in nickel-rich NiAl. Nanostruct. Mater. 2, 19-28 (1993)
H. Hoffmann. Magnetic properties of thin ferromagnetic films in relation
to their structure. Thin Solid Films 58, 223-233 (1979)
Y. Yoshizawa, S. Oguma, K. Yamauchi. New iron-based soft-magnetic alloys composed of ultrafine grain structure. J. Appl. Phys. 64, 6044-6046
(1988)
K. Hono, D. H. Ping. APFIM studies of nanocomposite soft and hard magnetic
materials. In: Advances in Nanocrystallization. Proceedings of the Euroconference
on Nanocrystallization and Workshop on Bulk Metallic Glasses (Grenoble,
France, April 21-24, 1998). Ed. A. R.Yavari. (Trans Tech Publications,
Switzerland 1999). Materials Science Forum 307, 69-74 (1999) / J. Metastable
Nanocryst. Mater. 1, 69-74 (1999)
N. I. Noskova, N. F. Vildanova, A. P. Potapov, A. A. Glazer. Influence
of deformation and annealing on structure and properties of amorphous alloys.
Fiz. Metall. Metalloved. 73, No 2, 102-110 (1992) (in Russian)
N. I. Noskova, E. G. Ponomareva, A. A. Glazer, V. A. Lukshina, A. P. Potapov.
Influence of preliminary deformation and low temperature annealing on a
size of nanocrystallites of Fe 73.5 Cu 1 Nb 3 Si 13.5 B 9 alloy obtained at crystallization of amorphous ribbon. Fiz. Metall. Metalloved. 76, No 5, 171-173
(1993) (in Russian)
A. Inoue, H. M. Kimura, K. Sasamori, T. Masumoto. Ultrahigh strength
of rapidly solidified Al 68-x Cr 3 Ce 1 Co x (x = 1, 1.5 and 2 %) alloys containing an icosahedral phase as a main component. Mater. Trans. Japan. Inst.
Met. 35, 85-94 (1994)
A. Inoue. Preparation and novel properties of nanocrystalline and
nanoquasicrystalline alloys. Nanostruct. Mater. 6, 53-64 (1995)
N. A. Smirnova, V. I. Levit, V. P. Pilugin, R. I. Kuznetsov, M. V. Degtyarev.
Peculiarities of low temperature recrystallization of nickel and copper. Fiz.
Metall. Metalloved. 62, 566-570 (1986) (in Russian)
R. Z. Valiev, I. V. Aleksandrov. Nanostructured Materials Obtained by Severe
Plastic Deformation (Logos, Moscow 2000) 272 pp. (in Russian)
R. Z. Valiev, N. A. Krasilnikov, N. K. Tsenev. Plastic deformation of alloys with submicron-grain structure. Mater. Sci. Engineer. A137, 35-40 (1991)
R. Z. Valiev, A. V. Korznikov, R. R. Mulyukov. Structure and properties

127

Nanocrystalline Materials

65.
66.

67.

68.

69.

70.

71.

72.

73.

74.
75.

76.

77.

78.
79.

80.

81.

of metallic materials with submicrocrystalline structure. Fiz. Metall. Metalloved.


73, No 4, 70-86 (1992) (in Russian)
V. M. Segal. Materials processing by simple shear. Mater. Sci. Engineer.
A 197, 157-164 (1995)
V. M. Segal. Methods of investigation of strain-deformed state in processes
of plastic change of shape. Ph. D. Thesis (Physico-Technical Institute of
Acad. Sci. of Belorus. SSR, Minsk 1974) 30 pp. (in Russian)
V. M. Segal, V. I. Reznikov, F. E. Drobyshevskii, V. I. Kopylov. Plastic
treatment of metals by simple shear. Izv. AN SSSR. Metally No 1, 115123 (1981) (in Russian)
V. M. Segal, V. I. Reznikov, V. I. Kopylov, D. A. Pavlik, V. F. Malyshev.
Processes of Plastic Formation of Structure of Metals (Nauka i Tekhnika,
Minsk 1994) 232 pp. (in Russian)
R. Z. Valiev, A. V. Korznikov, R. R. Mulyukov. Structure and properties
of ultrafine-grained materials produced by severe plastic deformation. Mater.
Sci. Engineer. A168, 141-148 (1993)
N. A. Akhmadeev, R. Z. Valiev, V. I. Kopylov, R. R. Mulyukov. Formation of submicron-grained structure in copper and nickel with the use of
severe shearing strain. Metally No 5, 96-101 (1992) (in Russian)
A. A. Rempel, A. I. Gusev, S.Z. Nazarova, R. R. Mulyukov. Imputity
superparamagnetism in plastically deformed copper. Doklady Akad. Nauk
347, 750-754 (1996) (in Russian). (Engl. transl.: Physics - Doklady 41,
152-156 (1996))
A. Cziraki, I. Geracs, E. Toth-Kadar, I. Bakonyi. TEM and XRD study
of the microstructure of nanocrystalline Ni and Cu prepared by severe plastic
deformation and electrodeposition. Nanostruct. Mater. 6, 647-550 (1995)
A. A. Rempel, A. I. Gusev, R. R. Mulyukov, N. M. Amirkhanov. Microstructure and properties of palladium subjected to severe plastic deformation.
Metallofizika i Noveishie Tekhnologii 18, 14-22 (1996) (in Russian)
A. A. Rempel, A. I. Gusev. Magnetic susceptibility of palladium subjected
to severe plastic deformation. Phys. Stat. Sol. (b) 196, 251-260 (1996)
A. A. Rempel, A. I. Gusev, R. R. Mulyukov, N. M. Amirkhanov. Microstructure, microhardness and magnetic susceptibility of submicrocrystalline
palladium. Nanostruct. Mater. 7, 667-674 (1996)
A. A. Rempel, A. I. Gusev, S. Z. Nazarova, R. R. Mulyukov. Magnetic
susceptibility of palladium subjected to plastic deformation. Doklady Akad.
Nauk 345, 330-333 (1995) (in Russian). (Engl. transl.: Physics - Doklady
40, 570-573 (1995))
R. Z. Valiev, R. R. Mulyukov, V. V. Ovchinnikov, V. A. Shabashov, A. Yu.
Arkhipenko, I. M. Safarov. On physical width of intercrystallite boundaries.
Metallofizika 12, 124-126 (1990) (in Russian)
R. Z. Valiev, R. R. Mulyukov, V. V. Ovchinnikov. Direction of a grain-boundary
phase in submicrometer grained iron. Phil. Mag. Letters 62, 253-256 (1990)
R. Z. Valiev, R. R. Mulyukov, V. V. Ovchinnikov, V. A. Shabashov. Mossbauer
analysis of submicrometer grained iron. Scripta Metal. Mater. 25, 27172722 (1991)
R. Z. Valiev, R. R. Mulyukov, Kh. Ya. Mulyukov, L. I. Trusov, V. I. Novikov.
Curie temperature and saruration magnetization of nickel with submicrongrained structure. Pisma v ZhTF 15, 78-81 (1989) (in Russian)
R. Z. Valiev, Ya. D. Vishnyakov, R. R. Mulyukov, G. S. Fainstein. On the
decrease of Curie temperature in submicron-grained nickel. Phys. Stat. Sol.

128

Preparation of Bulk Nanocrystalline Materials

82.

83.

84.

85.

86.

87.

88.

89.
90.
91.

92.

93.

94.

95.
96.

97.

98.

(a) 117, 549-553 (1990)


Kh. Ya. Mulyukov, G. F. Korznikova, R. Z. Abdulov, R. Z. Valiev. Magnetic hysteresis properties of submicron grained nickel amd their variation
upon annealing. J. Magn. and Magn. Mater. 89, 207-213 (1990)
Kh. Ya. Mulyukov, G. F. Korznikova, R. Z. Valiev. Microstructure and magnetic
properties of submicron grained cobalt after large plastic deformation and
their variation during annealing. Phys. Stat. Sol. (a) 125, 609-614 (1991)
R. Z. Abdulov, R. Z. Valiev, N. A. Krasilnikov. Formation of submicrometergrained structure in magnesium alloy due to high plastic strains. Mater. Sci.
Letters 9, 1445-1447 (1990)
R. M. Galeyev, O. R. Valiakhmetov, G. A. Salishchev. Dynamic
recrystallization of coarse grained titanium-base VT30 alloy in (a + b) field.
Metally No 4, 97-103 (1990) (in Russian)
O. R. Valiakhmetov, R. M. Galeyev, G. A. Salishchev. Mechanical properties of titanium VT8 alloy with submicrocrystalline structure. Fiz. Metall.
Metalloved. 70, No 10, 204-206 (1990) (in Russian)
A. A. Rempel, A. I. Gusev, R. R. Mulyukov. Preparation of nanocrystalline
titanium carbide. In: Solid State Chemistry and New Materials. V.1. (Institute of Solid State Chemistry, Yekaterinburg 1996) pp.244-245 (in Russian)
A. A. Rempel, A. I. Gusev, R. R. Mulyukov. Preparation of nanocrystalline
titanium carbide by severe plastic deformation. In: Ultrafine grained powders,
materials and nanostructures. Proc. of Interdistrict Conf., Krasnoyarsk,
December 17-19, 1996. Ed. V. E. Redkin. (Krasnoyarsk State Technical
University, Krasnoyarsk 1996) pp.131-132 (in Russian)
A. I. Gusev, A. A. Rempel. Structural Phase Transitions in Nonstoichiometric
Compounds (Nauka, Moscow 1988) 308 pp. (in Russian)
A. I. Gusev. Physical Chemistry of Nonstoichiometric Refractory Compounds
(Nauka, Moscow 1991) 286 pp. (in Russian)
A. I. Gusev, A. A. Rempel. Nonstoichiometry, Disorder and Order in Solids
(Ural Division of the Russ. Acad. Sci., Yekaterinburg 2001) 580 pp. (in
Russian)
A. I. Gusev, A. A. Rempel, A. J. Magerl. Disorder and Order in Strongly
Nonstoichiometric Compounds: Transition Metal Carbides, Nitrides and Oxides
(Springer, Berlin - Heidelberg - New-York 2001) 607 pp.
A. I. Gusev, A. A. Rempel. Phase diagrams of metal-carbon and metal-nitrogen
systems and ordering in strongly nonstoichiometric carbides and nitrides.
Phys. Stat. Sol. (a) 163, 273-304 (1997)
A. I. Gusev. Order-disorder transformations and phase equilibria in strongly
nonstoichiometric compounds. Uspekhi Fiz. Nauk 170, 3-40 (2000) (in Russian).
(Engl. transl.: Physics - Uspekhi 43, 1-37 (2000))
A. A. Rempel. Effects of Ordering in Nonstoichiometric Interstitial Compounds (Nauka, Yekaterinburg 1992) 232 pp. (in Russian)
A. A. Rempel. Atomic and vacancy ordering in nonstoichiometric carbides.
Uspekhi Fiz. Nauk 166, 33-62 (1996) (in Russian). (Engl. transl.: Physics - Uspekhi 39, 31-56 (1996))
A. I. Gusev, A. A. Tulin, V. N. Lipatnikov, A. A. Rempel. Nanostructure
of dispersed and bulk nonstoichiometric vanadium carbide. Zh. Obsh. Khimii
72, 1067-1076 (2002) (in Russian). (English transl.: Russ. J. General Chem.
72, 985-993 (2002))
S. V. Rempel, A. I. Gusev. ZrC segregation to the surface of dilute solid
solutions of zirconium carbide in niobium carbide. Neorgan. Materialy 37,

129

Nanocrystalline Materials

99.

100.

101.

102.

103.

104.
105.

1205-1211 (2001) (in Russian). (Engl. transl.: Inorganic Materials 37, 10241029 (2001))
S. V. Rempel, A. I. Gusev. Surface segregation of ZrC from a carbide solid
solution. Fiz. Tverd. Tela 44, 66-71 (2002) (in Russian). (Engl. transl.: Physics
of the Solid State 44, 68-74 (2002))
A. I. Gusev. A phase diagram of the vanadiumcarbon system taking into
account ordering in nonstoichiometric vanadium carbide. Zh. Fiz. Khimii
74, 600-606 (2000) (in Russian). (Engl. transl.: Russ. J. Phys. Chem. 74,
510-516 (2000))
T. Athanassiadis, N. Lorenzelli, C. H. de Novion. Diffraction studies of
the order-disorder transformation in V 8 C 7 . Ann. Chim. France 12, 129-142
(1987)
A. A. Rempel, A. I. Gusev, O. V. Makarova, S. Z. Nazarova. Physical and
chemical properties of nanostructured vanadium carbide. Perspektivnye
Materialy No 9, 9-15 (1999) (in Russian)
L. Ramqvist. Variation of hardness, resistivity and lattice parameter with
carbon content of group 5b metal carbides. Jernkontors Annaler. 152, 467475 (1968)
A. I. Gusev. Nanocrystalline Materials:Preparation and Properties (Ural
Division of the Russ. Acad. Sci., Yekaterinburg 1998) 200 pp. (in Russian)
A. I. Gusev, S. V. Rempel. X-ray diffraction study of the nanostructure
resulting from decomposition of (ZrC) 1x (NbC) x solid solutions. Neorgan.
Materialy 39, 49-53 (2003) (in Russian). (Engl. trans.: Inorganic Materials 39, 43-46 (2003)).

130

Evaluation of the Size of Small Particles

+D=FJAH"
4. Evaluation of the Size of
Small Particles
Investigations of materials with superfine grains show that a
decrease in the crystallite size below some threshold value results
in large changes of the properties. Size effects appear if the mean
size of crystalline grains does not exceed 100 nm, and manifest
themselves most vividly when the grain size is smaller than 10 nm.
The particle size is one of the most important parameters
determining the peculiarities of the properties and the application
field of the nanomaterial. Particles, grains or crystallites of what
sizes can be called nanoparticles? Where is the limit behind which
we achieve the nanocrystalline state?
As a first step, one can accept the conventional division of any
substance on the basis of the grain size. On this level of
understanding, materials with a mean grain (particle) size greater
than 1 m are referred to as coarse-grained materials. Polycrystalline materials with a mean grain size of 100150 to 40 nm
are usually referred to as submicrocrystalline materials and those
with a mean grain size smaller than 40 nm are called nanocrystalline
materials. However, this conventional division is not scientifically
justified.
The physically justified determination of the nanocrystalline state
requires deep understanding of the problem. From the physical
viewpoint, the transition to the nanocrystalline state is associated
with the appearance of size effects on the properties. While there
are no size effects, there is no nanocrystalline state. Actually, if the
length of the solid in one, two or three dimensions is comparable
with or smaller than the characteristic correlation length of a
specific physical phenomenon or physical parameter, used in the
theoretical description of some property or process, size effects will
be detected on the appropriate properties. Such physical parameters
as the size of magnetic domains in ferromagnetics, the free path
131

Nanocrystalline Materials

of the electron, the de Broglie wavelength, the coherence length in


superconductors, the wavelength of elastic oscillations, and the size
of the exciton in semiconductors may be used for evaluating the
transition to the nanocrystalline state. Thus, the size effects refer
to a set of phenomena associated with the change in the properties
of a substance as a result of a decrease in the particle size and a
simultaneous increase in the fraction of the surface contribution to
the general properties of the system. The substance or material is
referred to as nanocrystalline substance when its particle size is
equal to or smaller than some characteristic physical parameters
having the dimension of length. Only in this case is it possible to
detect the real and not far-fetched changes of the properties, which
are determined by the size effects.
Now we understand what the nanocrystalline state is from the
physical viewpoint. Let us consider how the particle size in a
substance can be determined. It should be mentioned that in many
cases the size and shape of particles are determined to a large
extent by the method of preparing a nanomaterial.
4.1. ELECTRON MICROSCOPY
The most widely used method of determining the size of small
particles and objects is microscopy. In fact, the invention of the
optical microscope already resulted in a large number of inventions
in crystallography, medicine, biology and made microscopy very
popular and well known. The development of electron microscopy
made it possible to examine objects which are considerably smaller
than 1 m and resulted in its extensive application in solid state
physics, crystallography, solid state chemistry, materials science and
mineralogy. At present, high-resolution electron microscopy [13]
may be used to examine the distribution of individual atom columns,
the distribution of atom columns in crystalline and colloid particles,
various crystal lattice defects, the mutual location of the molecules
in biological objects, for example, in a DNA helix. Electron
microscopy is used for the direct determination of the size of
nanoparticles and nanocrystallites. It is used widely in investigations
of nanomaterials and in nanotechnologies for advanced electronics.
Electron microscopy is the only technique for determining the
constitution of the nanoparticles produced from colloid solutions.
Figure 4.1 shows a bimetallic colloid particle produced by reduction
of HAuCl 4 with tri-sodium citrate. Full-shell clusters are formed,
for instance, if palladium acetate is reduced in an acetic acid
132

Evaluation of the Size of Small Particles

Fig. 4.1. A bimetallic colloid, consisting of a gold core (dark area) and a shell of
palladium (light area) [4].

Fig. 4.2. HRTEM micrograph of a NiTiN nanocomposite particle [5].

solution by gaseous hydrogen in the presence of ligand molecules.


The dark core of the particle consists of gold and has a
characteristic hexagonal shape, the light shell is formed by palladium
atoms; parallel planes of Pd atoms are seen clearly. The particle
size is approximately 20 nm. In turn, the palladium shell is protected
by ligand molecules but on this image the shell is not visible.
Electron microscopy makes it possible to examine the interfaces
in nanocomposite materials. In particular, Fig. 4.2 shows a
composite nanoparticle NiTiN produced by active plasmametal
reaction methods [5], i.e. by evaporation and condensation of nickel
133

Nanocrystalline Materials

and titanium in nitrogen using DC arc plasma. The interface


between the particles of nickel and titanium nitride TiN is clearly
visible; the size of the Ni particle is 3050 nm.
Figures 4.3 and 4.4 were obtained at a magnification of
1 000 000 times in a JEOL electron microscope. They show clearly
bcc nanocrystallites of molybdenum produced by gas phase
condensation. The spherical particle of molybdenum is located
between eight smaller particles of molybdenum carbide, which form
the vertices of the cube (Fig. 4.3). The mean size of the crystallite
is 1215 nm, and the quality of high-resolution electron microscopic
image makes it possible to determine interplanar distances. For
example, the distance between (110) atomic planes is about
0.22 nm.
The application of electron microscopy makes it possible to
examine an unusual nanostructure produced as a result of ordering
of non-stoichiometric vanadium carbide [6, 7]. At a magnification
of 100 times the vanadium carbide powder is made up of large (up
to 20 m) separate irregular-shaped agglomerates which comprise
particles of about 1 m in size. However, at a larger magnification
it becomes clear that each 1 m particle has a complicated
structure and actually represents a set of a large number of

Fig. 4.3. BCC nanoparticles of molybdenum Mo (central part) with smaller particles
of molybdenum carbide MoC grown on it (image was obtained by W.M. Straub,
F. Phillipp and H.-E. Schaefer, Institut fr Theoretische und Angewandte Physik,
Universitt Stuttgart, 1994).
134

Evaluation of the Size of Small Particles

Fig. 4.4. BCC Mo nanoparticles produced by gas condensation method (image


was obtained by W.M. Straub, F. Phillipp and H.-E. Schaefer, Institut fr Theoretische
und Angewandte Physik, Universitt Stuttgart, 1994).

nanocrystallites. The SEM image of vanadium carbide nanopowder


(Fig. 4.5) was obtained at a magnification of up to 30000 times in
a DSM 982 Gemini high-resolution scanning electron microscope.
It is seen that the vanadium carbide nanocrystallites have the shape
of distorted lobes, which grow into each other and form a
nanostructure resembling corals.
Figure 4.6 shows a nanoparticle of titanium carbide Ti 44 C 56
powder produced by ball milling metallic titanium and carbon, which
was taken at a 44:56 ratio [8]. The size of the nanoparticle is 12
18 nm, the distance between the atomic planes is 0.25 nm and
corresponds to the (111) B1 interplanar spacing of the cubic titanium
carbides close to the stoichiometric composition TiC 0.91.0 .
With all advantages of electron microscopy as a method of
determination of the particle size, it must be taken into account that
it is a local method and provides information on the size of the
object only in the field of observation. However, the observed area
may not be representative, i. e. not characteristic of the entire
volume of the substance.
Let us consider a simple example. Electron microscopy is used
135

Nanocrystalline Materials

Fig. 4.5. Nanocrystallites of the powder of ordered vanadium carbide V 8C 7 (VC 0.875),
magnification up to 30 000 [6].

Fig. 4.6. HRTEM image of mechanically alloyed Ti 44 C 56 powder after 720 ks of


ball-milling [8].

to examine a section with a size of 1010 m (i.e. 10 10 m 2 ) of a


specimen with a surface area of 1 cm 2 . In this case, the examined
area contains 0.000001 part of the entire surface of the specimen.
Is this much or little? The surface area of our planet is equal to
approximately 510 million km 2, and a millionth part is 510 km 2. This
136

Evaluation of the Size of Small Particles

small area of the earth surface may be covered with water (ocean,
sea or large lakes) or ice (Antarctic, Greenland, and large mountain
glaciers), may be a desert or marsh, may be overgrown with
forests, etc. It is clear that examination of only one millionth part
of the earth surface is insufficient for making an accurate
conclusion on the structure of the entire planet surface. An
adequate representation may be obtained only by examining several
sections. Therefore, electron microscopy examination must be
carried out on several areas in order to obtain statistically averaged
out information for the entire substance.
Electron microscopy is the only direct method of determining the
size of very small particles. All other methods are indirect, because
the information on the mean particle size is extracted from the data
on the variation of some property of the substance or process
parameter. Indirect methods include diffraction, magnetic,
sedimentation and gas-adsorption methods.
4.2. DIFFRACTION
The diffraction method occupies the main position amongst indirect
methods for determining the particle size. At the same time, this
method is simplest and accessible because the X-ray examination
of the structure is used widely and efficient equipment is available.
In diffraction experiments on powder or polycrystalline specimens,
the defects of a structure are studied on the basis of the
broadening of diffraction reflections. However, in the application of
this method in practice, one often compares the width of the
diffraction reflections from a coarse-grained substance and from the
same substance in the nanocrystalline state. This determination of
the broadening and the subsequent evaluation of the mean particle
size sometimes are not accurate and could be characterised by a
very large error (up to several hundreds of percent). The point is
that the broadening should be determined in relation to diffraction
reflections from an infinitely large crystal. It means that the
measured width of the diffraction reflection should be compared
with the instrumental width, i.e. with the width of the resolution
function of the diffractometer. This function should be determined
in a special diffraction experiment. In addition to this, the accurate
determination of the width of diffraction reflections is possible only
by means of computational description of the shape of the
experimental reflection. This is very important because there may
also be other (in addition to the small crystallite size) physical
137

Nanocrystalline Materials

reasons for the broadening of diffraction reflections. Therefore, it


is important to determine not only the broadening but also to
separate the different contributions to the broadening, i. e. to find
the values of these contributions.
Since the diffraction method of determination of the particle size
is most widely used and is accessible, special features of this
application will be discussed in detail here.
The diffraction method is based on the effect of broadening of
diffraction reflections associated with the size of the particles
(crystallites). All types of defects cause displacement of the atoms
from the lattice sites. The author of [9] derived an equation for the
intensity of Bragg reflections from a crystal defect, which enabled
all the defects to be divided conventionally into two groups. The
defects in the first group only lower the intensity of diffraction
reflections but do not cause reflection broadening. The broadening
of reflections is caused by defects of the second group. These
defects are microdeformations, inhomogeneity (non-uniform
composition of the substance over their volume) and the small
particle size. The broadening, caused by microdeformations and
randomly distributed dislocations, depends on the order of reflection
and is proportional to tan , where is the scattering angle in the
diffraction experiment. In substitutional solid solutions A 1x B x and
in non-stoichiometric interstitial compounds MX y there is another
reason for broadening, which is called inhomogeneity. Inhomogeneity
causes a variation of the composition over the volume of the
specimen [10, 11]. The broadening, caused by inhomogeneity x, is
proportional to (sin 2 )/cos [1012]. In the case of nanocrystalline
substances, the broadening associated with a small size D of
crystallites (D < 200 nm) is most interesting, and in this case the
magnitude of broadening is proportional to sec . The broadening,
caused by the inhomogeneity or small particle size, is independent
of the order of reflection. Let us consider the derivation of the
equation, which takes into account the diffraction reflection
broadening, caused by the finite size of the particles of a
polycrystalline substance.
Let <l> V be the height of the column of planes of coherence
scattering averaged out with respect to volume, while <D> is the
particle diameter averaged out with respect to volume. For
spherical particles, integration leads to the equation:
<D> = 4<l> V /3 .

(4.1)

138

Evaluation of the Size of Small Particles

Let us consider the scattering vector s = 2sin / , where is the


radiation wavelength. Mathematically, its differential (or
indeterminacy from the physical viewpoint, because in a finite
crystal the wave vector becomes a poor quantum number) is equal
ds = 2cos d / = cos d(2 )/ .

(4.2)

In this equation, quantity d(2 ) is the integral width of the


diffraction reflection expressed in angle 2 and measured in
radians. The integral width is determined as the integral intensity
of the reflection divided by reflection height and does not depend
on the shape of the diffraction reflection. Therefore, it is possible
to use the integral width for analysis of X-ray, synchrotron or
neutron diffraction experiments, carried out by different techniques
with differing resolution functions of the diffractometer and in
different angle ranges.
The indeterminacy of the scattering vector, ds, is inversely
proportional to the <l> V . Therefore, the product of these quantities
is equal to unity <l> V ds = 1. It is clear from this relationship that
the indeterminacy ds is equal to zero at the infinite height of the
column (i.e. the infinitely large size of the crystals). If the height
of the column is small and tends to zero, indeterminacy ds of the
wave vector and, consequently, width d(2 ) of the diffraction
reflection become very large. We shall assume that all grains are
spherical. Considering equations <l> V = 1/ds, (4.1) and (4.2), the
mean grain size <D> for the diffraction reflection of an arbitrary
shape may be determined as
<D> =1.33 /(cos d(2 ))

(4.3)

where d(2 ) is the integral width of the diffraction reflection. This


width is often replaced by the full width at half maximum
(FWHM). The relationship between the integral width of reflection
and FWHM depends on the shape of the experimental diffraction
reflection and must be determined in every specific case. For the
reflections in the form of a rectangle and a triangle, the integral
width of the reflection is equal to FWHM exactly. For the Lorentz
and Gauss functions, this relationship has the form d(2 ) L
1.6FWHM L (2 ) and d(2 ) G 1.1FWHM G (2 ), respectively. For
the pseudo-Voigt function, which will be examined later, this
relationship is more complicated and depends on the ratio of the
Gauss and Lorentz contributions. For diffraction reflections in small
139

Nanocrystalline Materials

angles, the relationship between the integral broadening and FWHM


may be assumed to equal to d(2 ) 1.47FWHM(2 ). Substituting
this relationship into (4.3), we have the Debye formula [13]:
<D> = 0.9 /[(cos ) FWHM(2 )] .

(4.4)

In a general case, when the particles of the substance have an


arbitrary shape, the mean particle size can be determined by the
DebyeScherrer formula:
<D> = K hkl /[(cos ) FWHM(2 )] K hkl /[(2cos )FWHM( )]
(4.5)
where K hkl is the Scherrers constant whose value depends on the
shape of the particle (crystallite, domain) and on diffraction
reflection indices (hkl).
In a experiment as a result of the finite resolution of the
diffractometer, the reflection broadens and cannot be smaller than
the instrumental width of reflection. In other words, in equation
(4.5) the width FWHM(2 ) of the reflection should be replaced by
broadening of the reflection, , in relation to the instrumental width.
Therefore, in the diffraction experiment, the mean particle size is
determined using Warrens method [1416]:
<D> = K hkl /[(cos ) (2 )] K hkl /[(2cos ) ( )]

(4.6)

where = (FWHM exp )2 (FWHM R ) 2 is the broadening of the diffraction


reflection. It should be noted that (2 ) 2 ( ).
The full width at half maximum FWHM R or the instrument width
of reflection may be measured on an efficiently annealed and
completely homogeneous substance (powder) with a particle size of
about 110 m. In other words, the reference should be
represented by a reflection without any additional broadening, with
the exception of instrument broadening. If the resolution function
of the diffractometer is described by the Gauss function, and R is
its second momentum, then FWHM R = 2.235 R .
The diffraction reflections are described by the Gauss function
g( ) and Lorentz function l ( )

140

Evaluation of the Size of Small Particles

g ( ) = A exp[( 0 ) 2 /(2 G2 )] ,

(4.7)

l( ) = A[1 + ( 0 ) 2 / L2 ] 1

(4.8)

or by their superposition V( ) = cl( ) + (1 c)g( ), i.e. by the


pseudo-Voigt function

V ( ) = cA[1 + ( 0 ) 2 / L2 ] 1 + (1 c) A exp[( 0 ) 2 /(2 G2 )] ,


(4.9)
where c is the relative contribution of the Lorentz function to the
total reflection intensity; L and G are the parameters of the
Lorentz and Gauss distributions, respectively; A is the normalising
factor.
Let us consider special features of the Gauss and Lorentz
distributions required for further analysis. Parameter G is the
second momentum of the Gauss function. The second momentum
G , expressed in angles , is associated with the full width at half
maximum, measured in angles 2 , by the well-known relationship
G ( ) = [FWHM(2 )]/(22.355). This relationship can easily be
derived directly from the Gauss distribution. Figure 4.7a shows the
Gauss distribution described by the function

g ( ) = hG exp[( 0 ) 2 /(2 G2 )] ,

(4.10)

where G is the second momentum of the Gauss function, i.e. the


value of the argument corresponding to the inflection point of the
function where 2 g( )/ 2 = 0. If the value of function (4.10) is
equal to to half its height, i.e., h G exp [( 0 ) 2 /(2 G2 )] = h G /2,
then exp [( 0 ) 2 /(2 G2 )] = 1/2. From this it follows that
( 0 ) = G 2 ln 2 . As can be seen from Fig. 4.7a, the Gauss
function full width at half height is FWHM G = 2( 0 ) =
2G 2 ln 2 = 2.355 G .
Parameter L of the Lorentz distribution coincides with the half
width of this function at half height. Let it be that the Lorentz
function

141

Nanocrystalline Materials
a

g( ) = hG exp[ -( - 0)2/ 2 G2 ]


( ) = hL[1 + ( - 0)2/ L22] -1

FWHML = 2 L
L /3

()

g( )

___
FWHMG = 2 G2ln2

FWHM L

FWHMG

hL /2

hG /2

Fig. 4.7. Gauss g( ) and Lorentz l( ) distributions, used for model description
of the shape of the diffraction lines: (a) relationship between the second momentum
G of the Gauss function and the full width of the function g( ) at half of its
maximum, FWHM G ; (b) for the Lorentz function parameter L = FWHM L /2, and
the second momentum of the function l( ) is L / 3.

( 0 ) 2
l ( ) = hL 1 +

L2

(4.11)

has the value, which is equal to half of the height, i.e. l( ) =


h L /2 (Fig.4.7b). The value of the argument, which corresponds to
this value of the Lorentz function, is determined from the
( 0 )2

equation 1 +
L2

( 0 ) 2
1
= 1 and
. It then follows that
L2
2

= 0 + L . Thus, for the Lorentz function FWHM L = 2 L . The


second momentum of the Lorentz function, i. e. the value of the
argument, corresponding to the inflection point of the function, can
be determined from the condition 2l( )/ 2 = 0. Calculation shows
that the second momentum of the Lorentz function is equal to
L /3.
The pseudo-Voigt function (4.9) ensures the best description of
experimental diffraction reflection in comparison with the Gauss and
Lorentz functions.
Taking this into account, the resolution function of the
diffractometer R( ) is represented as a pseudo-Voigt function; it
will be assumed for simplicity that A = 1 in equation (4.9). Then we
have

142

Evaluation of the Size of Small Particles

R ( ) = c[1 + ( 0 ) 2 / L2 ] 1 + (1 c) exp[( 0 ) 2 /( 2 G2 )] . (4.12)


The resolution function is the superposition of the Lorentz and
Gauss functions. As the zero approximation, the width of the
resolution function may be approximated as
FWHM RpV = cFWHM L + (1 c)FWHM G = 2c L + 2.235(1 c) G . (4.13)

If L = G R , then FWHM RpV = 2(1.1775 0.1775c) R . Let


some effective Gauss function G eff ( ), the area under which
coincides with that of the pseudo-Voigt function, have the width
FWHM, equal to FWHM RpV . Then the second momentum of this
function is eff G = (1 0.15c) R . Thus, the pseudo-Voigt resolution
function R( ) and the effective Gauss function G eff ( ) are
equivalent with respect to half width. To a zero approximation, this
makes it possible to replace function (4.12) by the function
2
2
2
Reff ( ) = exp[( 0 ) 2 /( 2 eff
G )] exp[( 0 ) /( 2 R * )] , (4.14)

where

R* eff G = (1 0.15c) R

(4.15)

on the condition that L = G R .


The experimental function I( ), describing the shape of an
arbitrary diffraction reflection, is a convolution of the distribution
function g( ) and of the resolution function R eff ( ) (4.14), i. e.

I ( ) =

Reff ( ' ) g ( ' )d ' =

= I exp[( ' + 0 ) 2 /( 2 R2 * )] exp[( ' + 0 ) 2 /( 2 x2 )]d '=

2
2
= [I / 2 ( R*
+ x2 ) ] exp ( 0 ) 2 / [2( R*
+ x2 )] .

143

(4.16)

Nanocrystalline Materials

Equation (4.16) shows that the second momentum exp of the


2
= R2 * + x2 or, with allowance for
experimental function is exp
(4.15),
2
exp
= [(1 0.15c) R ] 2 + x2 .

(4.17)

Broadening of the diffraction reflection is expressed by means of


the full width of reflection at half height FWHM exp as
= (FWHM exp ) 2 (FWHM R* ) 2 . If the second momenta and the full

width are expressed in the same units (in angles or in angles 2 ),


then FWHM exp, R = 2.235 exp,R and the broadening of reflection (hkl)
is
2
2
2
= 2.235 exp
R*
= 2.235 exp
[(1 0.15c ) R ]2 .

(4.18)

As already mentioned, the values of the broadening, caused by


the small grain size, deformations and inhomogeneity are proportional
to sec , tan and (sin 2 )/cos , respectively. Three different types
of broadening can be separated because of the different angular
dependence. It is worth noting that the size of regions of coherent
scattering, determined from size broadening, may correspond to the
size of the individual particles (crystallites), but may also reflect the
subdomain structure and characterise the mean distance between
the stacking faults or effective size of mosaic blocks, etc. In
addition to this, it is important to point out that the shape of
diffraction reflection depends not only on the size but also on the
shape of nanoparticles [17, 18].
The separation of the broadening, caused by several different
factors, will be examined on the example of nanostructured carbide
solid solutions of the ZrCNbC systems. In X-ray investigation of
these solid solutions it was established that diffraction reflections
on the X-ray diffraction pattern of a (ZrC) 0.46 (NbC) 0.54 specimen are
greatly broadened. It is well known [19] that these solid solutions
have a tendency to decomposition in the solid state. However,
according to the X-ray data this specimen consists of a single
phase. To explain the reason for the broadening of reflections
(inhomogeneity, small grain size or microdeformations), the
quantitative analysis of the profile of diffraction reflections was
carried out using the pseudo-Voigt function (4.9). Analysis shows
144

Evaluation of the Size of Small Particles

that the width of all diffraction reflections greatly exceeds the width
of the resolution function of the diffractometer.
Broadening h , caused by the inhomogeneity x of the solid
solution A 1x B x , whose composition changes in the range x x, is
proportional to (sin 2 )/cos [10]:

sin 2
x [degree]
h 2 + k 2 + l 2 cos
360

h ( ) = 2.235 a' B1( x ) x = x

(4.19)
or, measured in radians, is equal to

h ( ) = 2.235 a' B1( x ) x = x

sin 2
x [rad],
h 2 + k 2 + l 2 cos
2

(4.20)

where a'B1 (x)da(x)/dx is the derivative of the concentration


dependence of the lattice constant of the solid solution with respect
to composition x.
The broadening s(2 ) 2 s ( ), caused by the small particle size
(crystallite, domain) is associated with the mean particle size
<D> ~ V 1/3 (V is the volume of the particle) by equation (4.6) which
can be written in the form [9, 20]:

s ( ) =

K hkl
2 < D > cos

[rad]

or

s ( ) =

90 K hkl
< D > cos

[degree].
(4.21)

In a cubic crystal lattice, the crystallites have the size of the same
order in three perpendicular directions. In this case, Scherrer s
constant K hkl for reflections with different crystallographic Miller
indices (hkl) of a cubic crystal lattice, may be calculated from the
equation [21]:

K hkl =

6h

(h 2 + k 2 + l 2 )1 / 2 (6h 2 2 hk + kl 2 hl )

(4.22)

Values of K hkl for reflections (hkl) of fcc lattice are presented in


Table 4.1.

145

Nanocrystalline Materials
Table 4.1 Scherrers constant K hkl for the diffraction reflections (hkl) from crystals
with fcc lattice
(hkl)

(111)

(200)

(220)

(311)

(222)

(400)

Khkl

1.1547

1.0000

1.0606

1.1359

1.1547

1.0000

(333)

(511)

(hkl)

(331)

(420)

(422)

Khkl

1.1262

1.0733

1.1527

1.1547
1.1088
for (333) and (511) reflections mean value is
Khkl = 1.2113

(440)
1.0607

Deformation distortions and induced non-uniform displacements of


the atoms from the lattice sites may form in the case of random
distribution of the dislocations in the volume of the specimen. In this
case, the atom displacements are determined by the superposition
of displacements from every dislocation. This may be regarded as
a local variation of interplanar distances. In other words, the
distance between the planes continuously changes from d 0 d to
d 0 + d (d 0 and d are the interplanar distance in the ideal crystal
and the mean variation of the distance between the planes (hkl) in
the volume V of the crystal, respectively). In this case, the quantity
= d/d 0 is the lattice microstrain which characterises the uniform
strain averaged out with respect to the crystal. The diffraction
maximum from the regions of the crystal with the changed
interplanar distance appears at angle , which slightly differs from
angle 0 for the ideal crystal. Therefore, the reflection broadens.
The equation for the reflection broadening, caused by the
microdeformations of the lattice, can be easily derived by
differentiation of the WullfBragg equation d = n /(2sin ); the result
is d/ =(n /2)(cos /sin 2 ) = d/tan . The reflection broadening to one side of the maximum of the reflection, corresponding to
the interplanar distance d, with the variation of the interplanar
distance by +d is = (d/d)tan , and in the case of variation
by d, it is = (d/d)tan . The total reflection broadening is
equal to the sum of these broadenings and its value with respect
to angle is:

d ( ) = (2d/d)tan = 2 tg [rad]
d ( ) = (360/ ) tan [degree].

or
(4.23)

Taking into account d (2 ) 2 d ( ), the strain broadening with


respect to angle 2 is [9]:

146

Evaluation of the Size of Small Particles

d (2 ) = 4 tan .

(4.24)

It should be mentioned that in (4.19)(4.21) and (4.23), the


broadening is determined in degrees , and not 2 .
The full width FWHM exp and the second momentum exp of every
experimental diffraction reflection of the solid solution
(ZrC) 0.46 (NbC) 0.54 are determined by approximation of the
reflections by functions (4.9), where A = 1. Broadening of the
(hkl) reflections is determined from equation (4.18). The angular
dependence of the second momentum R ( ) (Fig. 4.8a) of the
resolution function of the X-ray diffractometer was determined in
special experiments on annealed coarse-grained compounds with no
homogeneity interval: a single crystal of hexagonal silicon carbide
6HSiC, and on stoichiometric tungsten carbide WC. Large grain
size, absence of strain distortions and the homogeneity of the
composition of these specimens prevented broadening of the
reflections. Comparison of the obtained values of exp ( ) with the
resolution function R( ) made it possible to determine the angular
dependence exp ( ) of the experimental reflection broadening.
Numerical analysis shows that the dependences, which are
characteristic for the broadening associated with the inhomogeneity
and with the size of the crystallites, are closest to the linearity,
whereas there is no strain broadening (and randomly distributed
dislocations) in the investigated specimens.
Thus, it may be assumed that the observed broadening of the
diffraction reflections is a superposition of only two factors h and
s . Taking this into account, Fig.4.8 shows the dependences of the
experimental broadening exp ( ) of the diffraction reflections of
(ZrC) 0.46 (NbC) 0.54 specimen on (sin 2 ) / ( h 2 + k 2 + l 2 cos ) and
K hkl sec . Since the observed broadening is the result of two
independent mechanisms, it may be determined as a convolution of
two functions. The first function corresponds to the broadening
caused by inhomogeneity and the second function corresponds to
the broadening caused by the small grain size. In this case, the total
broadening can be determined from the equation

= h2 + s2 .

(4.25)

Calculation shows that the separated contributions of the


broadening h , caused by inhomogeneity, and size broadening s are

147

Nanocrystalline Materials

0.12

exp

___

( ) (degrees)

R (degrees)

0.14

R = [(0.0052tg2 + 0.0056)1/2] / (42ln2)

0.10
0.08
0.06
0.04

0.4

0.3
0.2
0.1

0.02
40

60

(degrees)

80

0.1

exp

c
0.4

0.2

s
1

0.3

0.3

0.1

0.2

(sin2 )/[(h2+k2+l 2)1/2cos ]

( ) (degrees)

( ) (degrees)

20

0.4
0.3
0.2

exp
= (h2 + s2)1/2
h
s

0.1

10

20

30

40

50

60

(degrees)

Khklsec

Fig. 4.8. The second momentum R( ) of the resolution function (a) of the diffractometer
and the separation of contributions to the broadening of diffraction reflections of
the solid solutions (ZrC) 0.46(NbC) 0.54 : (b) dependence of the experimental broadening
exp on (sin 2 )/[h 2 +k 2 +l 2 ) 1/2cos ] and the separated contribution of inhomogeneous
broadening h; (c) dependence of experimental broadening exp on Khklsec and separated
contributions of size broadening s ; (d) dependences of experimental exp , size s
and inhomogeneous h broadening on the diffraction angle and the approximation
of the experimental broadening by superposition of the broadenings h and s in
the form = [( h ) 2 +( s ) 2 ] 1/2 .
2
2
2
2
linear functions of (sin ) /( h + k + l cos ) and K hkl sec ,
respectively (Fig. 4.8a, 4.8b). This confirms the presence of
exactly these contributions. Figure 4.8c shows the angular
dependences of the determined broadening h ( ) and s ( ) and
approximation of the experimental broadening ( ) by their
superposition. The mean crystallite size <D> and the degree of
inhomogeneity x of the solid solution (ZrC) 0.46 (NbC) 0.54 were
calculated using equations (4.19) and (4.21) and the obtained
values of broadening. According to calculation results, <D> is equal
to 7010 nm and x = 0.041.
The calculation results show that the annealed solid solution

148

Evaluation of the Size of Small Particles

(ZrC) 0.46 (NbC) 0.54 is inhomogeneous and has a nanostructure. The


grains differing in the content of zirconium and niobium and
corresponding to the isostructural cubic phases of different
composition, have a size of approximately 70 nm. This
nanostructure is a decomposed solid solution with coherent
precipitates of two isostructural cubic phases (ZrC) 0.42 (NbC) 0.58 and
(ZrC) 0.50 (NbC) 0.50 with similar composition. Indeed, the diffraction
reflections of the solid solution (ZrC) 0.46 (NbC) 0.54 are efficiently
described by the superposition of two contributions, corresponding
to the phase (ZrC) 0.42 (NbC) 0.58 and (ZrC) 0.50 (NbC) 0.50 .
Thus, the diffraction study, in addition to obtaining standard
information on the crystal structure, makes it possible to determine
fine details of the microstructure.
On the whole, the sequence of the diffraction experiments in
determination of the grain (crystallite) size is the following:
(1) measuring of the diffraction pattern of the reference
substance, computational analysis of the profile of diffraction
reflections, determination of the width, construction of the
experimental dependence of the width of reflections on the
diffraction angle and calculation of the approximating resolution
function of diffractometer;
(2) measuring of the diffraction pattern of the investigated
substance, computational analysis of the profile of diffraction
reflections and determination of the width of reflections;
(3) determination of the reflection broadening for the investigated
substance as a function of the diffraction angle;
(4) separation of the contributions to the broadening, determined
by the small particle size, inhomogeneity of the investigated
substance, deformation distortions of the crystal lattice;
(5) determination of the mean grain (particle, crystallite) size.
Using the WullfBragg condition, it may easily be shown that
d hkl /d hkl = [cotan( hkl )] ; on the other hand, d hkl /d hkl = / ,
and therefore / = [cotan( )] . It then follows that the
broadening hkl (2 ) = 2( / )tan hkl . We shall assume that the
smallest reflection width is equal to the spectral width /
10 3 of the reflection. Then, from the expression obtained it is easy
to find the largest size of the crystallites which cause a measurable
reflection broadening. In this case, the broadening hkl =
210 3 tan hkl and <D> = /( hkl cos hkl )=10 3 /[(2tan hkl ) (cos hkl )]
= 10 3 l/(2sin hkl ) = 10 3 d hkl 200300 nm. Thus, the diffraction
method the particle size smaller than 300 nm to be determined.
For more precise measurement of the width of the reflection it
149

Nanocrystalline Materials

is much more efficient to use the filtered-off radiation with a


wavelength of 1 , because in this case the shape of the diffraction
reflection is symmetric and simple for analysis because the
reflection does not consist of a doublet.
Figure 4.9 shows diffraction reflections (111) from coarsegrained titanium carbide TiC 0.62 with a grain size of approximately
5 m and from the same titanium carbide subjected to severe plastic
deformation by pressure torsion [22, 23]. The broadening of
reflection indicates refining of the grain size. If the deformation
contribution to broadening is neglected, the grain size in the
deformed titanium carbide is equal to approximately 40 nm.
Figure 4.10 shows X-ray diffraction patterns of metallic nickel
with a grain size of approximately 210 m and compacted
nanocrystalline nickel with a grain size of less than 20 nm. The
width of the reflections for coarse-grained nickel coincides almost
completely to the smallest width that be measured in the
diffractometer; the low intensity of reflection (111) is determined
by the presence of a texture. It may be seen that the transition to
the nanocrystalline state leads to a large broadening of the
diffraction reflections.

 

7L& 

&RXQWV DUELWUDU\ XQLWV

QDQR7L& 

):+0 7L&



):+03,34
7L&













 GHJUHHV

Fig. 4.9. Broadening of the diffraction reflection (111) B1 of the nanocrystalline


titanium carbide TiC 0.62 produced by severe plastic deformation of the coarsegrained (with grain size ~5 m) carbide TiC 0.62 [22, 23]: reflections of nanocrystalline
and coarse-grained titanium carbide TiC 0.62 are indicated by the solid and broken
lines, respectively; FWHM is the full width of reflection at half maximum. The
grain size and the nanocrystalline carbide TiC 0.62 is 3050 nm (radiation CuK 1 ).
150

Evaluation of the Size of Small Particles






FRDUVHJUDLQHG 1L

&RXQWV DUELWUDU\ XQLWV

QDQRFU\VWDOOLQH 1L





















 GHJUHHV

Fig. 4.10. X-ray diffraction patterns of coarse-grained bulk Ni with a grain size
of about 210 m and compacted nanocrystalline Ni with a grain size of
D < 20 nm. Low intensity of reflection (111) of coarse-grained Ni is caused by
the presence of a texture. Diffraction reflections of nanocrystalline Ni are broadened
strongly

The method of evaluation of the mean grain size from the


broadening of diffraction reflections may be used for both bulk and
powdered nanocrystalline materials.
4.3. SUPERPARAMAGNETISM, SEDIMENTATION, PHOTON
CORRELATION SPECTROSCOPY AND GAS ADSORPTION
The size of nanoparticles can be measured by magnetic
measurements. In this case, the determination of the particle size
is associated with the superparamagnetism effect. Ferromagnetic
materials have a domain structure, which is energy-favourable
because of the closure of magnetic fluxes inside the ferromagnetic.
With a decrease of the size of the ferromagnetic particle, the
closure of the magnetic fluxes inside the ferromagnetic is less
advantageous from the energy viewpoint. When reaching some
critical size D c , every particle contains only one domain and a
further decrease in the size of the particles decreases the coercive
force to zero as a result of transition to the superparamagnetic
state. For typical ferromagnetics with a Curie temperature of 500
1000 K, the disappearance of ferromagnetism and transition to the
superparamagnetic state are possible when the particles become
smaller than 220 nm [24].
151

Nanocrystalline Materials

Superparamagnetism is interesting not only as a specific magnetic


phenomenon but also as a non-destructive method of determination
of the sizes, shape, concentration and composition of the particles
of the precipitated magnetic phase, and also the size distribution of
the particles. A relationship of superparamagnetism with the size
of ferromagnetic particles is discussed in more detail in Section 5.4
where the magnetic properties of isolated nanoparticles are
analysed.
A suitable example of the application of magnetic measurements
for determining the size of the nanoparticles are investigations [25
28] in which copper with iron nanoparticles was studied taking into
account the superparamagnetism effect. The starting copper was
diamagnetic and contained a very small quantity of dissolved iron.
The temperature and annealing dependences of the magnetic
susceptibility of starting copper and of copper subjected to severe
plastic deformation were measured. Analysis of these dependences
shows that severe plastic deformation leads to the precipitation of
iron particles previously dissolved in copper. The results show that
the mean volume and size of the precipitated superparamagnetic
particle of iron are ~1.810 20 cm 3 and ~3 nm, respectively. The
grain size of the produced submicrocrystalline copper smc-Cu is
130150 nm.
Determination of the particle size by sedimentation is based on
measuring the time during which a particle travels the fixed distance
S in a liquid medium with known viscosity . Viscosity determines
the force of resistance of the liquid (or gas) to the movement of
a body. At a low rate of movement V = S/t and a small size of the
body, the force of resistance to movement of the body with a mean
size R is described by the well-known Stokes equation
F = (4 / ) RV = (4 / ) RS/t

(4.26)

or (by a means of the volume v ~ R 3 )


F = (4 / ) v 1/3 S/t .

(4.27)

In a general case, coefficient depends on the shape of the body


and has different values. For spherical bodies = 2/3. For a steady
rate of movement, the falling body is affected, in addition to the
resistance force, by the force of gravity P = mg = v g, where g
is the acceleration of gravity, and the mass m of the body is equal
to the product of the volume v by density . To improve the
152

Evaluation of the Size of Small Particles

accuracy of calculations, a correction for the injecting Archimedes


force is introduced. Then, the force causing the body to fall in the
given medium (liquid or gas) is:
F d = P v m = vg( m ) ,

(4.28)

where m is the density of the medium. From the equality of the


force F d and the force of resistance to fall, F, it is easy to find the
volume and averaged out size of the falling body:
(4 / ) v 1/3 S/t = vg( m ) .

(4.29)

It then follows that

(4 / )S
v=

g ( m ) t

3/ 2

(4.30)

and
R~

(4 / )S
~ k /t .
g ( m ) t

(4.31)

For spherical particle F = 6 RS/t and F d = vg( m ) =


(4/3) R 3 g( m). Therefore, it follows from equality F = F d that

Rsph =

9S
2g ( m ) t .

(4.32)

Thus, the size of the falling particle in the sedimentation method


is inversely proportional to the time during which the particle travels
a fixed path. In currently available sedimentographs, the travel of
the particles and their number is recorded by means of laser
radiation. This technique is used to determine not only the size of
separate particles but also the size distribution of the particles, i.e.
the dispersion of the particle size. A disadvantage of the
sedimentation method is that it cannot be used to measure the size
of very small (less than 50 nm) particles. In addition to this, in order
to obtain high precision of measurements, the liquid must efficiently
153

Nanocrystalline Materials

wet the surface of the particles. In the case of poor wetting, the
surface tension force will maintain the small particles on the
surface, and a gas shell will surround the large particles. During
measurements it is also necessary to avoid coagulation of the
separate particles which may greatly distort the results. Detailed
description of sedimentation methods is given in monograph [29].
An efficient method of determination of the sizes of small
particles is based on using of Brownian movement [30], on the one
hand, and analysis of the spectral composition of the light scattered
by the suspension or a colloid solution, on the other hand.
A. Einstein invented this method of determination of the size of small
particles. In fact, his first article [31] from a serious of studies into
the theory of Brownian movement [3134] is referred to as Eine
neue Bestimmung der Molekldimensionen (New determination of
the size of molecules). It is interesting to note that this work was
in fact an Einsteins dissertation for the title of Doctor of
philosophy. The study was presented in 1905 at the Natural
Mathematics Section of the Higher Faculty of Philosophy of the
Zurich University. Almost independently and at the same time as
Einstein, the theory of Brownian motion was developed by Polish
scientist M. Smoluchowski [35].
The size of small particles could be measured, using the wellknown Einstein equation, describing the Brownian motion of
spherical particles in a liquid:
R = k B T/(6 D diff ),

(4.33)

where h is the liquid shear viscosity, D diff is the diffusion coefficient


of Brownian particles. At a known viscosity of the liquid, to
determine radius R of the particles, it is necessary to know the
diffusion coefficient D diff . The diffusion coefficient of Brownian
particles is determined measuring the full width of a non-displaced
(central) component in the spectrum of scattered light using an
optical displacement spectrometer [36]:
FWHM = 2D diff K 2 ,

(4.34)

where FWHM is the spectral line full width at half maximum, K is


the variation of wave vector during light scattering. This method of
determination of the size of small particles is known as photon
correlation spectroscopy (PCS) [3640].
Photon correlation spectroscopy is based on analysis of the
154

Evaluation of the Size of Small Particles

spectral composition of the light, scattered by the examined


specimen. Photon correlation spectroscopy is used widely for
measuring the sizes of submicrocrystalline and nanosized particles
in transparent suspensions and colloid systems. Measurements are
carried out using the radiation of a HeNe laser with a wavelength
of = 633 nm, laser power up to 10 mW. Light radiation is focused
in the centre of a cuvette containing a liquid with a specimen
(suspension). The cuvette is placed in a thermostat for temperature
stabilisation. The light is scattered under the given angle and is
received by a photoelectric multiplier. A multichannel correlator then
measures the autocorrelation function G( ) of scattered light.
Approximation of the autocorrelation function by the exponential
dependence makes it possible to determine the width of the spectral
line and then find the diffusion coefficient and the particle size. An
increase in radiation power results in an improvement of the
accuracy and accelerates the measurements.
Industrial analysers of the particle size, using the PCS method,
take measurements under several angles and make it possible to
determine the size of the particles in the range from 3 nm to 3m
and also the size distribution of the particles.
To determine the size of small particles in colloid solutions and
suspensions, it is also efficient to use the mass distribution of
particles using an ultracentrifuge. The method of separation of small
particles by a means of a centrifugal force in an ultracentrifuge was
developed the Swedish scientist T. Svedberg [41]. If the particle
density is known, measuring the mass of the particle it is easy to
find its volume and linear size.
The gas adsorption method is used to determine the specific
surface of the powder. The measured value of the specific surface
may be used to estimate the mean particle size. To measure the
specific surface, helium is passed through the prepared powder in
a special chamber. Helium atoms are adsorbed by the particle
surface. Helium-saturated powder is heated to remove the entire
amount of adsorbed helium and the amount of adsorbed helium is
determined from the variation of the mass. Assuming that helium
atoms form a monolayer on the surface of the particles, from the
volume of the adsorbed gas it is possible to determine the total
surface area of the particles and the specific surface of the
powder (in the units of area per units of mass, for example,
m 2 g 1 ). If it is assumed that the size and shape of all the particles
are the same, a simple relationship can be derived between the
specific surface S sp and the linear particle size R. For example, if
155

Nanocrystalline Materials

the mass of a spherical particle is m = (4 /3)R 3 and the surface


area s = 4 R 2 , then its specific surface is
S sp = s/m = 3/ R .

(4.35)

Thus, as the specific surface increases, the particle size decreases.


If a function describing the size distribution of the particles is
available, then from the measured specific surface it is possible to
determine the particle size, taking the size distribution function into
account.
The described methods of determination of the size of
nanoparticles are used most widely.

References
1.
2.
3.
4.
5.
6.

7.

8.

9.
10.

11.

12.

G. Tomas, M. J. Goringe. Transmission Electron Microscopy of Materials


(Wiley, New York 1990) 404 pp.
J. C. H. Spence. High-Resolution Electron Microscopy 3 rd edition. (Clarendon
Press, Oxford 2003) 400 pp.
D. Shindo, K. Hiraga. High-Resolution Electron Microscopy for Materials
Science (Springer, Tokyo 1998) 190 pp.
G. Schmid. Chemical synthesis of large metal clusters and their properties. Nanostruct. Mater. 6, 1524 (1995)
Y. Sakka, S. Ohno. Hydrogen sorptiondesorption characteristics of mixed
and composite NiTiN nanopartices. Nanostruct. Mater. 7, 341353 (1996)
A. I. Gusev, A. A. Tulin, V. N. Lipatnikov, A. A. Rempel. Nanostructure
of dispersed and bulk nonstoichiometric vanadium carbide. Zh. Obsh. Khimii
72, 10671076 (2002) (in Russian). (English transl.: Russ. J. General Chem.
72, 985993 (2002))
A. A. Rempel, A. I. Gusev. Nanostructure and atomic ordering in vanadium carbide. Pisma v ZhETP 69,436442 (1999) (in Russian). (Engl. Transl.:
JETP Letters 69, 472478 (1999))
M. S. El-Eskandarany, M. Omori, T. Kamiyama, T. J. Konno, K. Sumiyama,
T. Hirai, K. Suzuki. Mechanically induced carbonization for formation of
nanocrystalline TiC alloy. Sci. Reports of Res. Inst. Tohoku Univ. (Sendai,
Japan) 43, 181193 (1997)
M. A. Krivoglaz. Theory of X-ray and Thermal-Neutron Scattering by Real
Crystals (Plenum Press, New York 1969) 405 pp.
A. A. Rempel, A. I. Gusev. Preparation of disordered and ordered highly
nonstoichiometric carbides and evaluation of their homogeneity. Fiz. Tverd.
Tela 42, 12431249 (2000) (in Russian). (Engl. Transl.: Physics of the Solid
State 42, 12801286 (2000))
A. I. Gusev, A. A. Rempel. Nonstoichiometry, Disorder and Order in Solids
(Ural Division of the Russ. Acad. Sci., Yekaterinburg 2001) 580 pp. (in
Russian)
A. I. Gusev, A. A. Rempel, A. J. Magerl. Disorder and Order in Strongly
Nonstoichiometric Compounds: Transition Metal Carbides, Nitrides and Oxides

156

Evaluation of the Size of Small Particles

13.
14.
15.

16.
17.

18.
19.

20.

21.
22.

23.

24.
25.

26.

27.

28.
29.
30.

(Springer, Berlin Heidelberg New York, 2001) 607 pp.


B. D. Cullity. Elements of X-ray Diffraction (Addison-Wesley Publ., London
1978) 555 pp.
B. E. Warren, B. L. Averbach, B. W. Roberts. Atomic size effect in the X-ray
scattering by alloys. J. Appl. Phys. 22, 14931496 (1951)
H. P. Klug, L. E. Alexander. X-ray Diffraction Procedures for Polycrystalline
and Amorphous Materials (Wiley, New York 1954) 491 pp. (second edition:
X-ray Diffraction Procedures (Wiley, New York 1974))
B. E. Warren. X-Ray Diffraction (Dover Publications, New York 1990) 381
pp.
Defect and Microstructure Analysis by Diffraction (Intern. Union Crystallogr.,
Monographs on Crystallography. V.10) Eds. R. L. Snyder, J. Fiala, H. Bunge.
(Oxford University Press, New York 1999) 808 pp.
V. Ya. Shevchenko, A. E. Madison, Yu. I. Smolin. Peculiarities of diffraction on nanoparticles. Fizika i Khimiya Stekla 28, 465476 (2002) (in Russian)
S. V. Rempel, A. I. Gusev. Phase Equilibria in the ZrNbC Ternary System.
Zh. Fiz. Khimii 75, 15531559 (2001) (in Russian). (Engl. transl.: Russ.
J. Phys. Chem. 75, 14131419 (2001))
Ya. S. Umanskii, Yu. A. Skakov, A. N. Ivanov, L. N. Rastorguev. Crystallography,
X-Ray Diffraction and Electron Microscopy (Metallurgiya, Moscow 1982)
632 pp. (in Russian)
R. W. James. The Optical Principles of the Diffraction of X-rays (G. Bell
& Sons Ltd., London 1954) 624 pp.
A. A. Rempel, A. I. Gusev, R. R. Mulyukov. Preparation of nanocrystalline titanium carbide by severe plastic deformation. In: Ultrafine grained
powders, materials and nanostructures. Proc. of Interdistrict Conf., Krasnoyarsk,
December 1719, 1996. Ed. V. E. Redkin. (Krasnoyarsk State Technical
University, Krasnoyarsk 1996) pp.131132 (in Russian)
A. I. Gusev. Phase equilibria, phases and compounds in the Ti C system. Uspekhi Khimii 71, 507532 (2002) (in Russian). (Engl. transl.: Russ.
Chem. Reviews 71, 439465 (2002))
S. V. Vonsovskii. Magnetism (Nauka, Moscow 1971) 1032 pp. (in Russian)
A. A. Rempel, A. I. Gusev, S.Z. Nazarova, R. R. Mulyukov. Imputity
superparamagnetism in plastically deformed copper. Doklady Akad. Nauk
347, 750754 (1996) (in Russian). (Engl. transl.: Physics - Doklady 41,
152156 (1996))
A. A. Rempel, S. Z. Nazarova. Magnetic properties of iron nanoparticles
in submicrocrystalline copper. In: Advances in Nanocrystallization. Proceedings
of the Euroconference on Nanocrystallization and Workshop on Bulk Metallic
Glasses (Grenoble, France, April 2124, 1998). Ed. A. R. Yavari. Materials Science Forum 307, 217222 (1999). (Trans Tech Publications, Switzerland
1999) / J. Metastable Nanocryst. Mater. 1, 217222 (1999)
A. A. Rempel, S. Z. Nazarova, A. I. Gusev. Intrinsic and extrinsic defects
in palladium and copper after severe plastic deformation. In: Structure and
properties of nanocrystalline materials (Ural Division of the Russ. Acad.
Sci., Yekaterinburg, 1999) pp.265278
A. A. Rempel, S. Z. Nazarova, A. I. Gusev. Iron nanopatricles in severeplastic-deformed copper. J. Nanoparticle Researh 1, 485490 (1999)
C. Bernhardt. Particle Size Analysis. Classification and Sedimentation Methods
(Kluwer Academic Publishers, Netherlands, Dordrecht, 1994) 448 pp.
R. Brown. A brief account of microscopical observations made in the months
on June, July, and August, 1827, on the particles contained in the pollen
157

Nanocrystalline Materials

31.

32.

33.
34.
35.
36.
37.

38.
39.

40.

41.

of plants; and on the general existence of active molecules in organic and


inorganic bodies. Phil. Mag. 4, 161173 (1828)
A. Einstein. Eine neue Bestimmung der Molekldimensionen. Inaugural
Dissertation. Zrich Universitt (Buchdruck. K. J. Wyss, Bern 1905) 28
pp.; Ann. Physik 19, 289306 (1906)
A. Einstein. ber die von der molekularkinetischen Theorie der Wrme
geforderte Bewegung von in ruhenden Flssigkeiten suspendierten Teilchen.
Ann. Physik 17, 549560 (1905)
A. Einstein. Zur Theorie der Brownschen Bewegung. Ann. Physik 19, 371
381 (1906)
A. Einstein. Elementare Theorie der Brownschen Bewegung. Z. Elektrochem.
14, 235239 (1908)
M. von Smoluchowski. Zur kinetischen Theorie der Molekularbewegungen
und der Suspensionen. Z. Physik 21, 756780 (1906)
Photon Correlation and Light Beating Spectroscopy. Eds. H. Z. Cummins,
E. R. Pike. (Plenum Press, New York 1974) 584 pp.
M. A. Anisimov, Yu. F. Kiyachenko, G. L. Nikolaenko, I. K. Yudin. Measurement of liquid viscosity and sizes of suspended particles by means of
correlation spectroscopy of light beating. Inzhenerno-Fiz. Zh. 38, 651655
(1980) (in Russian)
Measurement of Suspended Particles by Quasi-Elastic Light Scattering. Ed.
B. E. Dahneke. (John Wiley, New York 1983) 584 pp.
Light Scattering and Photon Correlation Spectroscopy. Eds. E. R. Pike, J. B.
Abbiss. (Kluwer Academic Publishers, Netherlands, Dordrecht, 1997)
470 pp.
I. K. Yudin, G. L. Nikolaenko, V. I. Kosov, V. A. Agayan, M. A. Anisimov,
J. V. Sengers. A compact photon-correlation spectrometer for research and
education. Int. J. Thermophys. 18, 12371248 (1997)
T. Svedberg. Colloid Chemistry: Wisconsin Lectures (Chemical Catalog Comp.,
New York 1924) 265 pp.

158

Properties of Isolated Nanoparticles and Nanocrystalline Powders

+D=FJAH#
5. Properties of Isolated Nanoparticles and
Nanocrystalline Powders
The physical properties of small atomic aggregates, referred to as
clusters, small particles, isolated nanocrystals, have been determined
in a large number of original experimental studies whose results have
been generalised in a number of detailed reviews and monographs
[112]. Consequently, this section can be restricted to discussing
only the size effects associated with the structure of the
nanoparticles and also some new experimental results obtained in
recent years.
5.1 STRUCTURAL AND PHASE TRANSFORMATIONS
The developed surface of isolated nanoparticles provides a
significant contribution to their properties. The non-additive nature
of thermodynamic functions, associated with the contribution of the
interfaces and taken into account by introducing surface tension ,
leads to size effects of thermodynamic quantities. In the case of
nanoparticles it is also necessary to take into account the
dependence of surface tension on the particle size. The effect of
surface energy is manifested in the thermodynamic conditions of
phase transformations. Phases which do not exist in the given
substance in the large particles may form in nanoparticles. With a
decrease of the particle size, the contribution of the surface energy

FS =

(n)ds

(where (n) is the surface tension which depends

on the direction of unit vector n, normal to the surface) to the free


energy F = F V + F s (F V is the bulk contribution to the free energy)
increases. If the phase 1 is stable in a coarse-grained (bulk)
specimen at some temperature, i.e. F V(1) < F V(2) , then with a decrease
in the particle size and with an increase in surface energy F s it may
be shown that
159

Nanocrystalline Materials

F V ( 2 ) + F s ( 2 ) F V ( 1 ) +F s ( 1 )

(5.1)

and the phase 2 will be stable in small particles.


Since surface energy is a significant part of free energy and is
high enough in comparison with bulk energy, equation (5.1) shows
that deformation of the crystal which leads to decreasing of the
surface energy may be efficient in decreasing the total energy of
the system. A similar energy decrease may be realised by changing
the crystal structure of the nanoparticles in comparison with a
coarse-grained bulk specimen. The close-packed structures have
the lowest surface energy. Therefore, the face-centered cubic (fcc)
or hexagonal close-packed (hcp) structures are most suitable for
nanocrystalline particles [1, 2]. This has also been confirmed by
experiments. Electron diffraction examination [13] of niobium,
tantalum, molybdenum and tungsten nanocrystals with a size of 5
10 nm shows that they have an fcc or hcp structure, whereas in
coarse-grained powders and in the bulk state these metals have a
body-centered cubic (bcc) lattice. Nanoparticles of beryllium and
bismuth contain cubic phases, although these elements have a
hexagonal lattice in the bulk state [14]. The bulk gadolinium, terbium
and holmium have an hcp structure. The authors of [15,16]
investigated the structure of Gd, Tb and Ho particles with a size
of 110 to 24 nm and observed traces of the fcc phase. They showed
that the content of the fcc phase increases and the content of the
hcp phase decreases with a decrease in the particle size of these
metals. The hcp phase, which is typical of the bulk Gd specimens,
was not found at all in Gd nanocrystals with a size of 24 nm.
However, the author [4] has doubts on the accuracy of conclusions
[15, 16] on the hcp fcc transition, because the diffraction
reflections, detected on the X-ray diffraction patterns of the Gd, Tb
and Ho nanoparticles, could belong to low-temperature cubic
modifications of oxides of these metals. A decrease in the particle
size of certain elements (Fe, Cr, Cd, Se) resulted in a loss of
crystal structure and appearance of an amorphous structure [14,
17]. It was noted in a review [12] that a decrease in the surface
energy of particles may take place not only by means of a complete
change of the crystal structure but also in the case of deformation
of the structure. For example, small particles may have a multiply
twinned structure which in bulk specimens exists only as a
metastable structure.
In particular, attention should be given to the structure of clusters,
i.e. particles, which contain less than (12)10 3 atoms. The results
160

Properties of Isolated Nanoparticles and Nanocrystalline Powders

of a large number of theoretical calculations show that in addition


to the fcc structure, typical of the bulk crystal, clusters may have
a crystallographic symmetry characterised by the fifth fold
symmetry axis [1821]. Two main assumptions are used in
modelling the structure of clusters containing small numbers of
atoms: a cluster should have close packing and should be stable
from the energy viewpoint. It follows from the first assumption that
clusters are constructed from the simplest stable atomic
configurations and the fraction of tetrahedra among stable
configurations should be large because the tetrahedron is a stable
atom configuration with the smallest volume. Structural elements of
the clusters are the tetrahedron, the octahedron, the cube, the
cubeoctahedron, the pentagonal pyramid, the icosahedron, etc. The
smallest stable cluster with the symmetry axis of the 5 th order
contains seven atoms and has the form of a pentagonal bipyramid,
the next stable configuration with the symmetry axis of the fifth
order is a cluster in the form of an icosahedron consisting of
thirteen atoms.
The stable configurations (isomers) of the cluster consisting
n atoms are determined by those atomic coordinates which
correspond to the minima of the surface of potential energy in
(3n6) dimensional space. Clusters with n > 10 have tens or even
hundreds of isomers [21]. Examination of the relative stability of
different structural modifications shows that icosahedral forms are
more stable for clusters containing less than 150300 atoms. The
smallest icosahedron contains 13 atoms, 12 of which are located at
equal distances around the central atom. The icosahedron of 13
atoms may be regarded as a figure consisting of 20 identical
tetrahedra which have the common vertex and are connected
together with common faces which are twinning planes. In
icosahedral groupings, every k-th atomic shell (layer) contains
(10k 2 + 2) atoms, and the total number of atoms of the icosahedral
cluster is n = (2 N + 1) + 10

k2,

k =1

where k is the order number of

the atomic shell, N is the number of atomic shells. It should be


mentioned that the icosahedral particles are characterised by
a 6-angle profile on electron microscopic images.
Each icosahedral particle has a particle-twin with a structure of
a non-deformed fcc lattice. According to numerical calculations [19]
the energy of a 13-atom icosahedral cluster is 17 % lower than the
energy of an fcc cluster, and fcc clusters spontaneously transfer to
161

Nanocrystalline Materials

the icosahedral form.


An increase in the number of atoms in a cluster leads to a rapid
increase of the energy of elastic deformation, which is proportional
to the volume. As a result, the increase in the elastic energy in a
large cluster is larger than the decrease in the surface energy and
it results in the destabilisation of the icosahedral structure. Thus,
there is some critical size above which the icosahedral structure
becomes less stable than cubic or hexagonal ones. Cubic and
hexagonal structures are characteristic of nanoparticles larger than
10 nm.
The application of a uniform hydrostatic pressure to a
nanoparticle may increase the density of the structure. Under
hydrostatic pressure, a nanocrystal of CdSe, having the wurtzite
ZnS structure, acquires structure B1 [22]. The surface energy of
a nanoparticle increases with a decrease in the nanoparticle size.
Therefore, the pressure required for changing the crystal structure
of the nanoparticles should also increase. This dependence of
pressure on the size of the nanoparticles extracted from colloid
solutions (or prepared by colloidal methods) was detected for CdSe
[22] (Fig. 5.1), CdS, Si and InP [23,24].
The dependence of surface energy on the particle size
predetermines the relationship between the melting point and the size
of the nanoparticles. Let us consider a system which is a solid
spherical isotropic particle, located in its own melt. If the Gibbs





97,38

*3D

&G6H












7 QP

Fig. 5.1. Size dependence of hydrostatic pressure p trans required for the transition
of nanoparticles of CdSe from hexagonal (wurtzite type) structure to the B1 structure
[22].

162

Properties of Isolated Nanoparticles and Nanocrystalline Powders

tension surface is introduced as the surface separating the two


phases, then we shall have three subsystems: condensed phase 1,
melt (phase 2) and interface 3. In equilibrium conditions, the total
variation of the energy of these subsystems is equal to zero. This
equality is fulfilled if the temperature and chemical potentials are
the same in all subsystems, i.e. T 1 = T 2 = T 3 and 1 = 2 = 3 ,
and the pressure in phase 1 exceeds the pressure in phase 2 by the
value 2 /r (Laplace pressure), determined by the curvature of the
interface:
(p 1 p 2 ) = 2 /r.

(5.2)

Taking into account the equality of chemical potentials of the


phases 1 and 2 and relationship (5.2), one obtains the well-known
Thomson equation describing the dependence of the melting point
of the particle, T melt (r), on its size (radius r)
[T melt (r) T melt ]/ T melt = (v 1 /L)(2 /r),

(5.3)

where T melt and L are the temperature and heat of melting of a bulk
solid, v 1 is the volume of 1 g of substance, i.e. the quantity
reciprocal to density. Thomsons equation (5.3) predicts the
universal decrease of the melting point of the particles inversely
proportional to their radius. For the particle melt system,
equation (5.3) contradicts the initial assumption on the equilibrium
of the solid particle with the surrounding medium. According to
(5.3), when the system is heated the small particle should melt
earlier than a melt of the bulk solid forms. In other words, any
particle of a finite size should have a lower melting point than the
bulk solid. It is clear that in this case the observed equilibrium of
the crystal with a liquid is not possible. The invalidity of the
Thomsons equation is determined by the assumption made when
deriving the equation. According to this assumption, the volume of
the solidmelt system is constant and the changes of the volume
and mass of the phases are dependent in a simple way: the volume
of the melt is proportional to the mass of the melt.
Later, it was proposed to determine the melting point of small
crystals as the temperature at which the solid and liquid spherical
particles of the same mass are in equilibrium with their vapour [25].
Indeed, T melt (r) is the temperature at which the transfer of matter
through the vapour from the solid to the liquid and vice versa does
not take place in a mixture of solid and liquid particles with equal
163

Nanocrystalline Materials

masses. Using and developing the concept proposed in [25], the


authors of [2628] obtained the following equations for the
equilibrium melting point T melt(r) of solid particles

Tmelt (r ) = Tmelt {1

2
[ s l ( s / l ) 2 / 3 ]}
s Lr

(5.4)

and particles coated with a layer of melt with thickness ,

Tmelt (r ) = Tmelt{1

2 sl l
[
+
(1 s )]} ,
r
l
s L r

(5.5)

where s , l , sl is the surface tension of the solid and liquid


particles, and also at the interface of the solid and liquid phases;
s , l are the densities of the solid and liquid particles.
For the thermodynamic equilibrium conditions, the melting point
is determined as a temperature at which the total free energies of
the solid and liquid phases are equal to each other. Allowing for the
surface energy in the equation for the total free energy, the author
[29] proposed the following equation for the melting point of a
spherical particle:

Tmelt (r ) = Tmelt {1

3
[ s l ( s / l ) 2 / 3 ]}
s Lr

(5.6)

This formula gives the smallest possible value of melting point


T melt (r). There are other equations, which describe the decrease of
the melting point of small particles with a decrease of their size.
The equation (5.4)(5.6), obtained by different authors for
describing the size effect of the melting point of nanocrystalline
particles, may be presented in the form
Tmelt (r ) = Tmelt (1 / r )

(5.7)

where is a constant which depends on the density, heat of melting


and surface energy of the material. It may easily be seen that
dependence (5.7) is similar to Thomsons equation (5.3). Recently,
it was proposed [30] to use expansion into a power series

T melt (r) = T melt (1 + r 1 + r 2 + ...)

164

(5.8)

Properties of Isolated Nanoparticles and Nanocrystalline Powders



6Q

%PHOW












7 QP

Fig. 5.2. Decrease in melting point T melt of Sn nanoparticles in relation to their


reciprocal radius r 1 ; solid line corresponds to the melting point calculated from
equation (5.5) [27].

for describing the experimental data for T melt (r), where , are
empirical constants.
In experiments, a decrease in the melting point of small particles
was detected in a large number of studies: Sn [27], Pb, In [28], Ag,
Cu, Al [31], In [32], Au [33, 34], Pb, Sn [35], Pb, Sn, Bi, In, Ga
[3639], Ag, Au [40].
Electron diffraction examination of Sn nanoparticles with a
diameter of 880 nm [27] showed a large deviation of the
experimental dependence T melt (r) from the linear dependence
T melt ~1/r. The linear dependence comes from Thomsons equation
which is wrong. Approximation of the results of [27] by equation
(5.5) showed good agreement between experimental and calculated
melting points (Fig. 5.2). Calculation was performed at the following
values of the parameters of equation (5.5): s = 7.1810 3 kg m 3 ,
l = 6.9810 3 kg m 3 , l = 0.58 N m 1 , sl = 0.0622 N m 1 , L =
58.5 kJ kg 1 , = 3.210 9 m, and T melt = 505 K. For tin, equation
(5.5) at these values of the parameters has the form

3.74 1 ,
Tmelt (r ) = 505 40

r 3.2 r

(5.9)

where the size of r is in nanometers. The size dependences of the


relative melting point T melt (r)/T melt of Sn nanoparticles, determined
165

Nanocrystalline Materials



QDQR$X

%2









7 QP

Fig. 5.3. Dependence of melting point T melt on the radius r of Au nanoparticles:


the solid line is the melting point calculated from equation (5.5), the dashed line
is the melting point of a macroscopic bulk specimen of Au [34].

in [27, 35], are in complete agreement in the range r = 1040 nm.


Previously, the same dependence T melt (r)/T melt on r was detected in
[33] for nanoparticles of gold with a radius smaller than 40 nm.
Equation (5.5) was used to describe the size dependences T melt (r)/
T melt for Pb, Sn, In, and Bi nanoparticles with a radius greater than
2 nm [37].
The results of electron diffraction examination of the dependence
T melt (r) for gold particles with a radius greater than 1 nm [34] (Fig.
5.3) are described efficiently by both equation (5.4) and (5.5)
because the accuracy of experimental measurements was
insufficient for discovering a difference between the models (5.4)
and (5.5).
A large decrease of the melting point of Sn, Ga and Hg clusters
with a size of ~1 nm, obtained in cavities of zeolites is described
in [7, 41]. Specimens were produced by filling under pressure the
cavities in zeolite with liquid metals. The largest decrease of the
melting point of clusters of Sn, Ga and Hg was 152, 106 and 95 K,
respectively, whereas no melting of clusters In, Pb and Cd was
detected [41].
The very large (by several hundreds of degrees) decrease in the
melting point was determined in [42] for colloid CdS nanoparticles
with a radius of 14 nm (Fig. 5.4).
In [43], equation (5.6) was used to calculate the dependence of
the melting points of Al, Cu, Ni, and Ti nanoparticles on their
166

Properties of Isolated Nanoparticles and Nanocrystalline Powders

&G6



%PHOW








7 QP

Fig. 5.4. Dependence of melting point T melt on radius r of nanoparticles of cadmium


sulfide CdS [42].


%PHOW





7L
1L
&X
$O





7 QP

Fig. 5.5. Dependence of melting point T melt on reciprocal radius r 1 of nanoparticles


of Al, Cu, Ni and Ti calculated from equation (5.6) using the parameters in Table
5.1 [43].

inverse radius 1/r (Fig. 5.5). The parameters of equation (5.6), used
to calculate the dependences T melt (1/r), and coefficient a for
equation (5.7) are presented in Table 5.1.
The estimates obtained in [43] show that the melting point of a
nanoparticle tends to zero when the nanoparticle radius drops below
0.50.6 nm.
The majority of authors hold that melting of the nanoparticles
starts from the surface because of the spatial inhomogeneity of the
nanoparticle. In this case, the experimental size dependence T melt (r)
should be described more efficiently by equation (5.5) which takes
the presence of the liquid shell into account. However, it is shown
in [35] that the same data are described quite accurately by
167

Nanocrystalline Materials
Table 5.1 Parameters of melting point function (5.6) of nanoparticles [43]

(J m2)

l 1 0 5

1 0 10
fo r e q ua tio n (5 . 7 )
(m)

0.926

0.865

0.894

4.43

1.592

1.320

1.310

1.250

4.07

17470

2.104

1.400

1.750

1.350

3.82

14150

1.797

0.910

1.500

0.868

5.80

M e ta l

T melt
(K )

L
(J mo l1)

s 1 0 5

(J m2)

(mo l m3)

Al

934

10700

1.032

Cu

1358

13050

Ni

1728

Ti

1943

equation (5.4) which does not consider the liquid shell. A partial
confirmation of the formation of the liquid shell is made by
computer modelling of melting of gold particles [44]. According to
[44], the liquid shell forms on the particles containing at least 350
atoms or larger.
Surface melting was observed experimentally on films of Pb
where melting of the surface started at a temperature equal to 0.75
of melting point T melt of bulk lead. The thickness of the molten layer
increased with approach to T melt [45, 46]. Surface melting was also
observed in the case of Ar [47], O 2 [48], Ge [49], and Ne [50].
The authors of [51] have proposed a different physical model of
melting of the nanoparticles. According to [51], clusters with a
given number of atoms are characterized by two temperatures T f
and T melt . Temperature T f is the lower limit of the temperature of
stability of the cluster in the liquid state. Temperature T melt is the
upper temperature limit of stability of the cluster in the solid state.
A set of identical clusters behaves as a statistical ensemble, which
consists of solid and liquid clusters in a specific temperature and
pressure range. The ratio of the number of solid and liquid clusters
is equal to exp(F/T), where F is the difference of the free
energies in the solid and liquid states. The equilibrium between the
solid and liquid clusters is dynamic and each individual cluster
transfers from the solid to liquid state and vice versa. Since the
frequency of transition between the liquid and solid state of the
cluster is low, equilibrium properties are certain for each phase.
The results published in [51] were obtained using the theoretical
analysis of the density of states of the cluster. The limiting
temperatures correspond to reaching the maximum or minimum of
free energy, i.e. appearing or loss of local stability by the phase.
168

Properties of Isolated Nanoparticles and Nanocrystalline Powders

Subsequent computer modelling [52] confirmed these conclusions.


The behaviour of gold nanoparticles at T < T melt , observed by the
authors of [53], was very similar to the model of melting proposed
in [51]. Gold particles were deposited on a SiO 2 substrate. Au
particles with a size of ~2 nm, excited by a beam of an electron
microscope, transferred from the single crystalline structure to
multiply twinned structure and vice versa. The lifetime of every
structure was approximately 0.1 s. Interphase fluctuations were
absent in particles larger than 10 nm. Earlier, analogous phenomena
were observed by the authors of [54].
Analysis of the data obtained by various authors for the size
dependence of the melting point of small particles shows that the
melting points of bulk crystals and small particles with a size greater
than 10 nm are almost the same. A strong decrease of the melting
point, caused by the size effect, is detected when the nanoparticle
size is below 10 nm.
5.2 CRYSTAL LATTICE CONSTANT
The transition from bulk crystals or large particles to nanoparticles
is accompanied by a change of the interatomic distances and crystal
lattice constants [25, 12]. The main task is to determine whether
the lattice constants decrease or increase with a decrease in the
particle size, and also to determine that nanoparticle size behind
which this change is measurable. The experimental data on the size
dependence of the lattice constant reported in the literature are
often ambiguous and contradictory.
Analysing the variation of the lattice constant of the
nanoparticles, it is important to take into account the possibility of
transition from less dense bcc and hexagonal structures to more
dense fcc structures accompanying by a decrease in the particle
size, mentioned in Section 5.1. According to electron diffraction
data [55], the hexagonal close-packed structure and the lattice
constants, which are characteristic of bulk metals, remains
unchanged with a decrease of diameter D of the particles Gd, Tb,
Dy, Er, Eu and Yb from 8 to 5 nm. With a further decrease of the
particle size, the lattice constants decrease rapidly. However, these
changes are accompanied by a change of electron diffraction
patterns, indicating a structural transformation, i.e. transition from
the hcp to fcc structure, and not a decrease of the interplanar
distances of the hcp lattice. In fact, investigation of nanoparticles
of rare-earth metals by the X-ray diffraction technique showed a
169

Nanocrystalline Materials

, QP



QDQR$O










QP

Fig. 5.6. Dependence of lattice constant a on the diameter D of aluminium nanoparticles


[57].

structural transition from the hcp to fcc lattice [15, 16]. Thus, to
ensure reliable detection of the size effect on the lattice constant
of the nanoparticles, it is also necessary to take into account the
possibility of structural transformations. The effect of the size of
nanoparticles on the lattice constant could be determined most
reliably by investigating of substances with the fcc crystal lattice
because the probability of a structural transition for the fcc solids
is very low.
Electron diffraction is one of the methods of determining the
lattice constant of the nanoparticles. Analysis of the systematic
errors of this method shows that only certain diffraction reflections
could be used for accurate determination of the lattice constant of
the nanoparticles. For example, it is recommended to use (220)
reflection for cubic nanocrystals [56]. Consideration of the
broadening of this diffraction reflection show that in Ag particles
with a diameter of 3.1 nm and in Pt particles with a diameter of
3.8 nm the lattice constant decreases by 0.7% and 0.5% in
comparison with bulk silver and platinum [56].
Electron diffraction using moire patterns shows that the variation
of the diameter of Al particles from 20 to 6 nm decreases the lattice
constant by 1.5% (Fig. 5.6) [8, 57, 58]. Previously [59] no decrease
of the lattice constant of Al particles with a diameter of 3 nm
was found. A decrease in the lattice constant from
0.405 nm for the bulk Al specimen to 0.402 nm for an Al nanoparticle with a diameter of 40 nm was detected by neutron
diffraction [2].
170

Properties of Isolated Nanoparticles and Nanocrystalline Powders

QDQR$J

QDQR$X



,,

 














QP

Fig. 5.7. Relative variation of the lattice constant a/a in relation to the diameter
D of gold and silver nanoparticles (2) [65].

The absence of the size dependence of the lattice constant was


noted for particles of Pb and Bi with a diameter of 5 and
8 nm respectively [60], for Au particles with a diameter of 6
23 nm [61, 62], and for Cu clusters with a diameter of 5 nm
[63]. However, a decrease in the size of copper clusters to 0.7 nm
decreased the lattice constant by 2 % in comparison with bulk metal
[63]. Electron diffraction investigation [64] detected a small
(~0.3 %) decrease of the lattice constant of an Au nanoparticles
with a diameter of 2.514 nm. Compression of the lattice constant
by ~0.1 % was established [65, 66] for the nanoparticles of Ag and
Au with a diameter ranging from 40 to 10 nm (Fig. 5.7).
The effect of the nanoparticle size on the lattice constant was
noted not only for metals but also for compounds.
A decrease in the lattice constant of submicrocrystalline titanium,
zirconium and niobium nitrides with decrease in the particle size is
described in [6770]. Nitride powders were produced by the plasma
chemical method. The dependence of the lattice constant a on the
specific surface S sp of the titanium nitride submicrocrystalline
powder is given in [70]: a(nm) = 0.424130.38410 8 S sp where S sp
changes from 410 4 to 110 5 m 2 kg 1 . It should be mentioned that
the dependence of the lattice constant on the size of the titanium
nitride particles, established in [70], does not take into account the
fact that the composition of the powders with different particle sizes
differs. The nitrogen content in the nitride powder decreased with
a decrease of the particle size in the powder. Unfortunately, the
authors of [70] did not attempt to separate the effect of the
171

Nanocrystalline Materials

composition and the effect of particle size of titanium nitride on the


lattice constant.
The decrease of the lattice constant of the cubic zirconium
nitride, explained in [68] by a decrease in the particle size of the
powder, took place with a large change in the composition of the
nitride. A large decrease of the lattice constant from 0.4395 nm for
the bulk niobium nitride to 0.4382 nm for the NbN powder with a
particle size of ~40 nm was found in [69].
The submicrocrystalline nitrides, produced by the plasma chemical
method, usually contain a large (up to 7 at.%!) amount of impurity
oxygen. The introduction of oxygen into the carbides and nitrides
decreases their lattice constant [71]. The lattice constant of the
cubic nitrides of group IV and V transition metals rapidly decreases
with a decrease in the nitrogen content [72, 73]. Taking this into
account, the conclusions made in [6769] on the decrease of the
lattice constant of the cubic nitrides with a decrease of the particle
size cannot be regarded as reliable.
In some cases, the lattice constant of the nanoparticles increases
instead of decreasing [74, 75]. A decrease of the size of Si particles
from 10 to 3 nm leads to an increase of the lattice constant by
1.1 % [76]. An increase of the lattice constant of CeO 2 oxide with
a decrease of the particle size from 25 to 5 nm (Fig. 5.8) is found
in [77]. It is possible that the increase of the lattice constant of the
cerium oxide is caused by the adsorption of water, as observed in
the case of MgO [78].
Thus, the experimental data on the size effect of the lattice
constant of the nanoparticles are ambiguous. Above all, this
ambiguity may be associated with the adsorption of the impurities
by nanoparticles. In the case of compounds, which have a
homogeneity interval, the different chemical composition of the
coarse particles and nanoparticles could lead to ambiguity. Another
possible reason for the ambiguity of the results are the structural
transformations, caused by a decrease of the particle size. The next
reason may be inaccurate measurement of the lattice constant.
The most reliable results did not show any decrease of the
lattice constant with a decrease of the particle size to 10 nm,
whereas for smaller particles, the decrease of the interatomic
distances in comparison with the bulk material is relatively realistic.
This is confirmed by the experimental data on interatomic distances
in metallic dimers (clusters consisting of two metal atoms). For
dimers, these distances are smaller than those for the bulk metals.
For example, the interatomic distances for clusters of Cu 2, Ni 2, Fe 2
172

Properties of Isolated Nanoparticles and Nanocrystalline Powders

&H2



QP













QP

Fig. 5.8. Dependence of the lattice constant a on the diameter D of nanoparticles


of CeO 2 [77] .

are equal to 0.222, 0.2305 and 0.187 nm respectively, and the


interatomic distances in these bulk metals are 0.256, 0.249 and 0.248
nm [79, 80].
Many authors assume that a decrease in the lattice constants of
the nanoparticles is a consequence of the excess Laplace pressure
p = 2 /r generated by surface tension . According to the
elasticity theory, the relative variation of the volume of the particle
V/V is proportional to p, i.e. V/V = k T (2/r), where k T is
isothermal compressibility. Since V/V = 3(a/a), then a/a =
k/r, where k is the proportionality coefficient. However, the values
of k for the same substance obtained by different authors greatly
differ. In addition to this, in certain cases the expansion of the small
particles is observed instead of compression. If the Laplace
pressure compressed the nanoparticles, then only decreasing of
lattice constant in nanoparticles would be their universal property.
In [8185], the decrease of the lattice constant of metallic
particles was explained by the formation of thermal vacancies and
by an increase of their concentration with a decrease in the particle
size. The high concentration of vacancies was regarded as a
consequence of uniform compression under the pressure p =
2/r. The latter claim is doubtful. In fact, the increase of the
concentration of vacancies in metals with increasing temperature
is well known. The temperature and pressure in the free energy
equation have opposite sign and, therefore, in a general case an
increase of pressure should affect the vacancy concentration in the
same manner as a decrease of temperature, i.e. it should lead to
173

Nanocrystalline Materials

a decrease (not increase) of the number of vacancies. In turn, a


decrease of the concentration of vacancies, following the logic in
[8185] cannot result in a decrease of the lattice constant.
The physical meaning of the Laplace pressure was analysed in
[86]. According to [86] the Laplace pressure cannot cause uniform
compression of solids. The Laplace pressure tends to change the
form of a solid in such a manner as to ensure the minimum of its
surface energy E S . For a liquid droplet it is assumed that E S is
proportional to the surface area S, i.e. E S = S, where surface
tension is constant. The surface area of the droplet can be
reduced by two methods: making the droplet spherical without
decreasing the volume, or compress the droplet in order to decrease
the surface area of even the spherical droplet. However, the
phenomenological relationship E S =S is valid only if the variation
of the surface area takes place at a constant volume. This means
that surface tension determines the equilibrium form of the
surface of small particles but does not lead to their compression.
A phenomenological approach was used by authors of [86] for
the analysis of the Laplace pressure. According to [86], the
Laplace pressure is a purely mathematical concept which makes it
possible to treat the chemical potential of the atoms in a solid of
finite size at true pressure p as a chemical potential of a solid of
infinitely large size at pressure (p + p), where p is Laplace
pressure. At the thermodynamic equilibrium, the shape of a small
solid should ensure the minimum of surface energy. It is assumed
that a particle with size D, surface area S and atomic density n 1 ,
i.e. phase 1, is in equilibrium with phase 2 at pressure p and
temperature T. The number of atoms in the particle is N 1 and is
proportional to n 1 DS or n 1 rS. In this case, the free Gibbs energy
of the considered system is F = F 1 (p,T) + F 2 (p,T) + S = F 1 (p,T)
+ F 2 (p,T) + 2N 1 /n 1 r. Differentiating F with respect to N 1 , it is
possible to find the chemical potential of the particle
1 (S, p,T) 1 (D, p,T) = 1 (, p,T) + 2 /n 1 r, where 1 (, p,T)
is the chemical potential of the specimen of an infinite size.
Expanding 1 into a power series and limiting considerations to the
terms of expansion of the first order, the authors of [86] obtained
that 1 (S, p,T) = 1 (, p + p, T). Thus, Laplace pressure p
makes it possible to express the chemical potential of a small
particle by means of the chemical potential of a bulk specimen with
S . When the dependence of the chemical potential of the
small particle on the surface area S is taken into account explicitly,
it is not required to introduce the Laplace pressure.
174

Properties of Isolated Nanoparticles and Nanocrystalline Powders

Analysis [86] shows that the Laplace pressure does not cause
compression of the solids and, consequently, cannot be the reason
for a decrease in the lattice constants of the nanoparticles.
The most probable reason for the decrease of the lattice
constant of the small particles in comparison with large particles or
bulk material is the fact that interatomic bonds of the surface atoms
are not compensated in contrast to the atoms located inside the
particle. As a result of this, the distances between the atomic planes
in the vicinity of the surface of the particle decrease because
surface relaxation takes place. In fact, an atom in a surface layer
has a smaller number of neighbours than in the volume because
there are neighbours exclusively in a surface plane and under the
surface plane. This disrupts equilibrium and symmetry in the
distribution of the forces and masses and leads to a change of the
equilibrium interatomic distances, shear deformation, smoothing of
vertices and edges of a particle. Surface relaxation affects several
surface layers and causes corrections in the volume of the particle
of the order of D 1 (D is the particle size). According to [8688],
the largest surface relaxation takes place on the surface of
nanoparticle, decreases from the surface to the centre of the
particle and in certain conditions may be oscillating. Oscillating
surface relaxation decreases in the direction to the centre of the
particle and is associated with Friedel oscillations of the density of
a degenerated electronic gas. The Friedel oscillations are caused
by any defects which disrupt the translation symmetry of the crystal.
In this case, the surface is such a two-dimensional defect. The
Friedel oscillations are transferred to the lattice by means of
electron-phonon interaction and lead to a change in the interplanar
distances. According to [88], in the free electron model, the
amplitude of Friedel oscillations decreases with increase of the
distance from the surface. It should be mentioned that surface
relaxation may not only lead to a decrease but also to an increase
in the volume of the crystal depending on the lattice constant and
the size of the crystal.
Surface relaxation in nanoparticles is efficiently confirmed by
measurements of the lattice constant of separate Al particles grown
epitaxially on a substrate of single crystal MgO [8, 57].
Compression of the lattice was divided separated into two
contributions. The first of them is compression of the lattice volume
with a decrease of the size of Al nanoparticles. Second contribution
is surface relaxation, i.e. a decrease of the lattice constant in the
direction from the centre to the surface of the particle.
175

Nanocrystalline Materials

Unfortunately, the authors of [57] did not consider the interaction


of epitaxial particles with a substrate, and this may have affected
interpretation of the results obtained.
According to [5], the main reason for the change of the
interatomic distances and the lattice constants in the nanoparticles
with a diameter smaller than 5 nm is a decrease in the number of
atoms forming these particles. In fact, the restriction of the number
of interacting atoms leads to a difference of the radial distribution
of the atoms in the nanoparticles from that in bulk crystals [89].
The theoretical analysis of the dependence of the crystal lattice
constant of a nanocrystalline substance on the nanoparticle size was
carried out by Qin et al [90]. They considered a nanomaterial as
a two-component system formed by discontinuous crystallites and
a continuous matrix of grain boundaries containing highly disordered
substance. According to the results of electronpositron annihilation
[91, 92], the grain boundaries contain two types of free volume,
namely monovacancies and vacancy clusters. Therefore, the density
of substances at the grain boundaries is lower than the density of
perfect crystallites not containing atomic defects. The excess
volume of the grain boundaries is determined as V = (V V 0 )/V 0 ,
where V and V 0 are the molar volumes of defective and perfect
crystallites, respectively. The grain boundary defects form a stress
field. Under the effect of stresses, the atoms of the nanocrystallites
deflect from their normal position in the lattice sites and this leads
to a distortion of the lattice. If a is the mean lattice constant of
the distorted crystals, and a 0 is the lattice constant of the perfect
crystal, then the quantity a / a 0 characterises the deviation of the
lattice constant of the nanocrystallite from the bulk value.
According to [90], the value of a / a 0 is determined directly by
parameters of the nanostructures such as the excess volume of the
grain boundaries V, the mean width of the grain boundaries ,
and mean diameter D:

a
1 ( + 2r0 )
( 1 + V 1) ,
=
a 0 2 D + r0

(5.10)

where r 0 is the distance between the nearest adjacent atoms in a


perfect lattice.
In a general case, the excess volume of the grain boundaries V
is a function of grain size D. According to [93], the excess volume
increases almost linearly with increasing grain size. Taking into
176

Properties of Isolated Nanoparticles and Nanocrystalline Powders

account the data [93], it is assumed in [90] that V ~ D. In this


case, equation (5.10) shows that a decrease of the grain size in a
nanocrystalline material should result in an increase of the lattice
constant. In [90] it is noted that the expansion of the crystal lattice
takes place mainly in the vicinity of the grain boundaries, in a thin
layer with a thickness of approximately (3 2 1) / 2 . Thus, a
decrease of a grain size in a nanomaterial should be accompanied
by an increase of the lattice constant. This conclusion follows from
the assumption that the excess volume of the grain boundaries
increases with increasing grain size. However, the assumption is
based on the data of a single study [93] and is not essential. It is
well known that a decrease of the grain size accompanied by an
increase of the relative number of atoms distributed at the grain
boundaries and, consequently, it may be assumed that a decrease
of the grain size will result in an increase of the excess volume of
the grain boundaries, i.e. V ~ D 1 . In this case, a decrease of the
grain size should increase the lattice constant. However, if V ~
D 2 or V ~ D 3 , the lattice constant may be constant or decrease
with a decrease of the nanoparticle size. Thus, theoretical analysis
does not provide an unambiguous answer to the question of how
should the lattice constant of the nanocrystalline substance vary in
relation to the nanoparticle size. Evidently, the lattice constant may
both increase and decrease with a decrease of the nanoparticle size.
5.3 PHONON SPECTRUM AND HEAT CAPACITY
The main reason for the variation of the thermodynamic
characteristics of nanocrystals in comparison with bulk materials is
the variation of the phonon spectrum, i.e. the variation of the
distribution function of frequencies of atomic vibrations (the term
frequency distribution function will be used). This is confirmed by
the results of investigation [94] of single crystal Si and silicon
powder by the inelastic scattering of slow thermal neutrons. The
frequency distribution function, g( ), for a fine powder and bulk
silicon greatly differ. Neutron scattering was also used for obtaining
phonon spectra of MgO particles (D ~ 11, 16 and 23 nm), TiN
particles (D ~ 30 nm) and bulk specimens of MgO and TiN [95
97]. According to [19], the phonon spectrum of small particles does
not contain low frequency modes which are present in the spectra
of bulk crystals. The waves whose half-length exceeds the longest
size D of the particle may form in the nanoparticles. Therefore, the
phonon spectrum is restricted by some minimum frequency min ~
177

Nanocrystalline Materials

c/2D (where c is the velocity of sound) on the lower limit of


vibration frequency. There is no such restriction in bulk solids. The
numerical value of min depends on the properties of the substance
and the shape and size of the particles. It may be expected that
a decrease of the particle size should displace the phonon spectrum
to the range of high frequencies. Special features of the vibration
spectra of the nanoparticles primarily reflect in heat capacity.
The distribution of eigentones in the presence of restrictions on
the side of lower limit of frequencies was discussed by the authors
of [98, 99]. They proposed similar expressions for describing the
number of eigentones n( ) of a rectangular particle taking into
account its geometrical characteristics. The equation obtained in
[99] for n( ) in a slightly modified form was used in [100] to
describe the size effect on low-temperature heat capacity.
According to [98], the frequency distribution function g( ) of the
phonon spectrum of a small rectangular particle with edges L x , L y,
L z , has the form

g () = V 2 / 2 2 c3 + S / 8c2 + L /16c1

(5.11)

where V=L x L y L z , S=2(L x ,L y +L x L z +L y L z ), L=4(L x +L y +L z ) is the


volume, surface area and total length of the edges; cl , ct are the
velocities of propagation of the longitudinal and transverse elastic
j

oscillations; c j 1 = c l + 2c t is the effective velocity (here for


transition from function g(), presented in [100], to function g( )
1
g ( = / 2 ) ). It should be
2
mentioned that the elasticity theory determines accurately only the

we use the expression

g ( ) =

1
3
3
quantity c3 = c l + 2ct corresponding to the effective velocity of
sound in solid with a large solid V, where the boundary
conditions, determined by the presence of a surface, could be
1

neglected. The correct expression for c 2 , derived in [101103]


taking into account the boundary conditions in an accurate manner

( A B ) divu + B ( u
l

kl

+ k ul ) = 0 , (A + B) 1/2 = cl and B 1/2 =

c t , has the form

178

Properties of Isolated Nanoparticles and Nanocrystalline Powders

c2 =

2c t4 3c t2 c l2 + 3c l4
c t2 c l2 (c l2 c t2 )

(5.12)

Equation (5.12) takes into account the mode scrambling effect


determined by the finite size of the particles [104]. Up to now, the
correct expression for c11 is unknown.
A spectral distribution similar to (5.11) but more accurate can
be found in [105]. The total number of normal modes for a particle
containing N atoms is

3N =

max

g ( )d .

(5.13)

min

With allowance for (5.12) it follows from equation (5.13) that

max min

18 2 Nc 3

+
V

1/3

2 /3

18 2 c 3
S
2 /3
1

(
)
N

+
144c 2 N 1/3 V

(5.14)

where (N 2/3 ) are the correction term of the order N 2/3 . Since
= 2 , then at min = 0 equation (5.14) completely coincides with
the analogous equation for max derived in [100]. Taking into
account the boundaries of the phonon spectrum, the heat capacity
of the small particle can be determined from

Cv =

max

min

( , T )
g ( )d ,
T

(5.15)

where ( , T ) = (h / 2)cotan (h / 2k BT ) is the mean energy of the


oscillator. According to [100], the heat capacity of a small particle
at a temperature T 0 and in the approximation min = 0 may be
written in the form

( kB L max / 8 c1 )( kBT / h max )I 2


3
2
C v (r ) = ( 4k BV max
/ 2 c 3 )(k BT / h max ) 3 I 4 + (k B S max
/ 2c 2 )

(k BT / h max ) I 3 +
2

179

(5.16)

Nanocrystalline Materials

where I m = (4m!/2

m+1

N
N =1

(4m!/2 m+1 ) (m) and (m) is the

Riemann zeta-function (I 4 = 4 /30, I 3 = 1.8031, I 2 = 2 /6). One


may assume that max = (18 2 Nc 3 /V ) 1/3 , i.e. coincides with the
maximum frequency for the bulk crystal. Then the first term in
equation (5.16) represents Debyes contribution (12 4 Nk B /5)(T/ D ) 3
to heat capacity and D = h max / k B is the Debye temperature for
the bulk crystal. Expression (5.16) can be written in the form
C v (r) = a 3 VT 3 + a 2 ST 2 + a 1 LT ,

(5.17)

where a 1, 2, 3 are the positive constants. Equations (5.16) and (5.17)


show that in the case of small particles, the heat capacity contains
a contribution determined by the large surface area of these
particles. Therefore, one may expect that the low-temperature heat
capacity will increase and the Debye temperature will decrease
with decreasing particle size.
The analysis of the size effects of the phonon spectrum, carried
out in [100, 101], is based on a quasi-continuum approach. The
quantum approach [106108] for calculating the frequency
distribution function g( ) of a small particle with radius r,
containing N atoms, is based on the expression

g () = ( l , s ) ,

(5.18)

l,s

where is the energy interval between the adjacent permitted


states; l, s = k l,s ct = ct a l ,s / r ; kl ,s is the wave vector; c t is the
velocity of transverse elastic vibrations; a 'l , s is the s-th zero of the
derivative of the spherical Bessel function of order l ; the
degeneracy of kl ,s is (2 l + 1).
According to [106], in the limit of long wave vectors k the total
number of the modes for a spherical particle, which contains N
atoms and has radius r, is

N = (2 / 9)r 3 kn3 + (1/ 4 ) r 2 kn2 + ( 2 / 3 ) rkn ,

(5.19)

where k n is the boundary wave vector corresponding to the


maximum frequency of vibrations, max , of the small particle. The
terms in the right part of equation (5.19) take into account the
180

Properties of Isolated Nanoparticles and Nanocrystalline Powders

volume, surface and linear contributions.


The molar heat capacity of a finite crystal with radius r has the
form
3N

( h j /k B T ) 2 exp( h j /k B T )

j =1

[exp(h j /k B T )] 2

C v ( r ) = 3 NA k B

(5.20)

According to (5.19), the frequencies j depend on the particle size.


The asymptotic expansion of heat capacity (5.20) using Poissons
equation for the temperature range h min /k B << T < D = h max /k B
leads to the expressions [106, 107]
3

C v (r ) = C v + vm

9 (3)k BT

2 2

4h c r

+ vm

k BT
6hcr

(5.21)

where v m is the molar volume and (3) = 1.20206 is the Riemann


zeta-function. At r the second and third terms in (5.21)
convert to zero, and the first term is the Debye equation for heat
capacity
/T

C v = 9 R (T / D )

x 4 exp xdx
= 9 R (T / D ) 3 D( x) .
2
(exp x 1)

(5.22)

At high temperature T >> D , the upper integration limit in (5.22)


is 1 >> D/T 0, the integral D(x) 13 ( D /T) 3 and heat capacity
C v 3R, i.e. tends to a limiting value determined by the Dulong
Petit law.
According to the exact solution [106], the sum of the second and
third terms in (5.21), i.e. the enhancement of the vibrational part
of the heat capacity of the small particle with an allowance for the
quantum size effect has the form

C = C v (r ) C v = vm
l s

3(2l + 1)k B
4r

exp
,
(exp 1) 2

(5.23)

where = hcal ,s / k B rT ; the summation in (5.23) is performed over all

s and l up to l max determined from the condition


181

l max

(2l 1) N
l=0

Nanocrystalline Materials

From equations (5.21) and (5.23) and also from equations (5.16) or
(5.17) it follows that at h w min /k B << T < D the heat capacity of
the small particle C v (r) is higher than the heat capacity C v of the
bulk crystal. With an increase of the particle size r the
difference of the heat capacities is C = C v (r) C v 0.
In the range of low temperatures (T 0), heat capacity (5.20)
may be asymptotically represented [107] as

C v (r ) 3N A k B (h min / k BT ) 2 exp(h min / k BT ) .

(5.24)

+HDW FDSDFLW\ 

At T 0 the heat capacity C v (r) (5.24) decreases more rapidly


than the Debye heat capacity C v (5.22). Therefore, the difference
C = C v (r) C v < 0 is negative in the low-temperature range.
This means that there is some temperature T 0 below which
C < 0, and at T > T 0 this difference is positive, i.e. C > 0 (Fig.
5.9).
The size effect of the vibrational (lattice) part of heat capacity
was considered by the authors of [100, 105108]. The electron
subsystem of bulk metals in low- and high-temperature ranges
provides a significant electronic contribution C el = e T to heat
capacity. Estimation of the electronic heat capacity of the
nanoparticles is complicated by discrete electronic energy levels
which form as a result of a limited number of atoms. In the case
of small particles and low temperatures, when the mean distance

% 

% 

Fig. 5.9. Temperature dependence of the heat capacity C: solid line is the heat
capacity temperature dependence in according to the Debye theory, horizontal dashed
line is the DulongPetit limit of heat capacity, and dashed curve is the deviation
of heat capacity from the Debye theory, which is associated with the quantum
size effect [107].

182

Properties of Isolated Nanoparticles and Nanocrystalline Powders

between the level is = h p F /2m * D > k B T (p F is the Fermi


momentum, D is the particle size, and m* is the effective mass of
the conduction electron), electronic heat capacity C el may greatly
differ from that of bulk metal. The dependence C el (T) is determined
by the distribution of energy levels. The linear dependence of the
electronic heat capacity on temperature with the coefficient
2
e was obtained by the authors of [109, 110]. They used the
3
assumption on the random distribution of electronic levels. The
theoretical analysis of the heat capacity in two-dimensional systems
[111] showed that the electronic part of heat capacity is a linear
function of temperature and the vibrational part is a quadratic
function of temperature. This is in agreement with the conclusion
made in [100, 105108], according to which the temperature
dependence of the heat capacity of small particles contains the
quadratic term bT 2 determined by the surface contribution.
For a bulk crystal with boundary wave vector k the authors of
[112] transformed equation (5.19) and obtained the size dependence
of the Debye temperature. If k = (6 2 /v) 1/3 is the boundary wave
vector in a bulk solid and v = V/N is the atomic volume, then
N = (2/9p)k 3 r 3 . Taking this into account, equation (5.19) may be
written in the form
*e =

k 3r 2 = r 2 kn3 + (9 / 8) rkn2 + 3kn

(5.25)

or, with the accuracy of the terms of the first order, k n = k(1 +
k / k n)

k n D (r )
3r 2 k 2 + (9 / 8)rk
=
2 2
.
D
k
3r k + (9 / 4)rk + 3

(5.26)

If the last term rk n in equation (5.19) is ignored, this gives a simple


size dependence of the Debye temperature

k n D ( r ) 1 + (3 / 8rk )
=

k
1 + (3 / 4rk ) .
D

(5.27)

Equation (5.27) may be written with a small decrease in accuracy


in the form
183

Nanocrystalline Materials

D (r)/ D 1 3 /(8rk) .

(5.28)

To estimate D (r) of a small particle of an arbitrary shape,


having the volume V and the surface are S, equations (5.27) or
(5.28) may be used taking the approximation r 3V/S into account:

D (r ) 1 + (S / 8Vk )

1 + (S / 4Vk ) .
D

(5.29)

In [112], surface tension was additionally taken into account


and equation (5.28) was transformed to the form

D (r)/ D 1 + [(2K /r) (3 /8rk T )],

(5.30)

where is the Gruneisen constant and k T is isothermal compressibility.


The dependence of the Debye temperature D (r) on the
effective velocity of sound c may also be found using the expression
proposed by the authors of [106] for the maximum vibration
frequency max = a max c( r ) / r for the particle:

c( r ) / k B r .
D (r ) = h max / k B = ha max

(5.31)

As a rule, the Debye temperature D (r) for the nanoparticles is


lower than D for the bulk materials. Allowing for this, equation
(5.31) shows that the effective velocity of sound in the nanoparticles
decreases with a decrease of the particle size and is proportional
to r m , where m > 1.
The changes in the phonon spectrum of the small particles should
also affect the value of the root-mean-square dynamic atomic
displacements

1 max ( , T )
u =
g ( )d .
Nm
2
2

(5.32)

min

Heat capacity is one of the most extensively studied properties


of nanoparticles. Let us discuss the results of investigation of the
heat capacity of colloid nanoparticles of silver and gold.
184

Properties of Isolated Nanoparticles and Nanocrystalline Powders

Measurements were performed at very low temperatures from 0.05


to 10.0 K in the magnetic field B from 0 to 6 T [113]. At
T < 1 K, the heat capacity of the nanoparticles of Ag (D = 10 nm)
and Au (D = 4, 6 and 18 nm) is 310 times higher than the heat
capacity of bulk specimens of silver and gold. The heat capacity
of the largest Au particles (D = 18 nm) in the temperature range
from 0.2 to 1.0 K almost coincides with that of the bulk specimen.
With a decrease of the Au particle size from 18 to 6 nm an
additional positive contribution to heat capacity initially increases
and with a further decrease of the particle size to 4 nm it decreases
but does not disappear and remains positive even for clusters Au 55
with a size of 1.5 nm. Measurements of the heat capacity of Ag
nanoparticles with a size of 10 nm in the magnetic field with B =
6 T detected the quantum size effect: at T < 1 K, the heat capacity
of Ag nanoparticles was lower and at T > 1 K it was higher than
the heat capacity of bulk silver. In the absence of the magnetic
field, the heat capacity of colloid nanoparticles of silver in the entire
investigated temperature interval was higher than that of bulk Ag
(Fig. 5.10). This experimental result is in good agreement with
theoretical conclusions [107] on the quantum size effect of the heat

10-2
Ag nanoparticles

10-3

D = 10 nm
B=0

C (J g-1 K-1)

B=6T

10-4

bulk Ag

10-5

10-6

10-7
0.1

10

T (K)

Fig. 5.10. Specific heat C versus temperature T of colloid Ag with D = 10 nm


[113]: measurements were performed in magnetic fields B up to 6 T, dashed line
corresponds to the specific heat of bulk silver.

185

Nanocrystalline Materials

capacity of the nanoparticles. Such an effect was not detected in


the case of colloid particles of gold because their heat capacity
becomes immeasurably small with increasing density of the magnetic
flux.
Measurements of the heat capacity of Pb nanoparticles with a
diameter of 2.2, 3.7 and 6.6 nm and In nanoparticles with a diameter
of 2.2 nm [114, 115] show that at T < 10 K the heat capacity C v (r)
of nanoparticles is 2575 % higher than the heat capacity C v of
bulk metals. The maximum deviation C = C v (r) C v was
observed in the temperature range 35 K. The sharp decrease of
C v (r) at T 2 K was caused by the low-frequency truncation of
the phonon spectrum as a result of the size effect. Results of [114,
115] were explained [106] by means of equation (5.23). Theoretical
analysis [106] was performed for the case of a homogeneous
spherical nanoparticle with a diameter of 2.2 nm, consisting of 184
atoms and accounting for the first 183 vibrational modes. The
authors of [116] measure C of vanadium nanoparticles with a
diameter of 3.8 and 6.5 nm and palladium nanoparticles with a
diameter of 3.0 and 6.6 nm, produced by vapour condensation. The
heat capacity of vanadium particles at T < 10 K is determined in
the main by the electronic contribution and the value of C,
determined by the size effect of the lattice heat capacity, is
relatively low. An increase in the heat capacity of Pd nanoparticles
in comparison with bulk palladium at 1.4 < T < 30.0 K (Fig. 5.11)
is completely determined by the additional lattice contribution
because the electronic heat capacity, regardless of the particle size,


P- PRO


%




3G













Fig. 5.11. Temperature dependences of the heat capacity of Pd nanoparticles with


a diameter 3.0 nm (1) and 6.6 nm (2) and bulk palladium (3) [116].

186

Properties of Isolated Nanoparticles and Nanocrystalline Powders

is described by the normal linear law e T, and the coefficient e is


the same as for bulk palladium.
The similar size effect on the heat capacity of nanocrystalline
powder of palladium with a mean particle size of 8 nm was
observed in [117]. The temperature dependence of the heat capacity
of nanocrystalline palladium n-Pd at 1 K T < 20 K was described
by the function C(T) = aT + bT 2 + cT 3 , similar to equation (5.17)
at a fixed value of r. The C(T) dependence of bulk palladium did
not contain the quadratic term bT 2 . The coefficient of electronic
heat capacity, e = a, of n-Pd was slightly lower, and the
temperature coefficient of lattice heat capacity, c, was twice as
high as the coefficients a and c for bulk palladium (Table 5.2). The
results obtained in [117] are in good agreement with the data in
[115] for the heat capacity of n-Pd.
The heat capacity of bulk copper and nanocrystalline powders
of Cu and CuO with a particle size of ~50 nm was investigated in
temperature ranges 120 K and 300800 K [118]. At a temperature
below 20 K the heat capacity was described by the polynomial
C(T) = aT + bT 2 + cT 3 (the values of the coefficient of the
polynomial are presented in Table 5.2). The quadratic term bT 2 was
presented only in the temperature dependence of the heat capacity
of Cu nanoparticles. It should be mentioned that the coefficients of
the linear and cubic terms of heat capacity of n-Cu were higher
than the appropriate coefficients for the bulk copper (Table 5.2).
At all temperatures (from 1 to 20 K and from 300 to 800 K) the
highest heat capacity was recorded for the CuO nanopowder and
Table 5.2 Coefficients of heat capacity polynomial C(T) = aT + bT 2 + cT 3

S p e c ime n

a
(mJ mo l1K 2)

b
(mJ mo l1K 3)

c
(mJ mo l1K 4)

Re fe re nc e s

Pd
(b ulk )

9.70.2

0.100.03

11 7

na no - P d
(D~ 8 nm)

8.50.2

0.100.03

0.200.03

11 7

Cu
(b ulk )

0.68

0.01

0.051

11 8

na no C u
(D~ 5 0 nm)

1.03

0.32

0.066

11 8

C uO
(D~ 5 0 nm)

0.410

11 8

187

Nanocrystalline Materials

the lowest heat capacity was found for bulk copper. The heat
capacity of the Cu nanoparticles was from 1.2 to 2.0 times higher
than that for bulk copper at temperature up to 450 K. With a further
increase of temperature a growth of the Cu nanoparticles and a
decrease of heat capacity to the values corresponding to bulk
copper are observed.
According to [119], the heat capacity of the nickel nanoparticles
with a diameter of 22 nm is 2 times higher than the heat capacity
of bulk nickel at 300800 K. The C(T) dependence of n-Ni shows
a weak diffuse exothermic effect at 380480 K, associated with the
recrystallisation of nickel particles, and a high endothermic peak with
a maximum at 560 K determined by the magnetic phase transition.
In bulk nickel, a weak endothermic peak, corresponding to the
magnetic transformation, was recorded at 630 K.
The authors of [120] used inelastic neutron scattering at 100
300 K to study the phonon density of states of coarse-grained
polycrystalline nickel and nanocrystalline nickel with a particle size
of 10 nm in the form of a powder and a pressed compacted
specimen with a relative density of 80 %. The largest size effect
is an increase of the density of phonon states of n-Ni specimen in
comparison with coarse-grained Ni at energies below 15 meV (Fig.
5.12). According to [120], the variation of the phonon spectrum of
n-Ni is caused by the low density of substance in the grain
boundaries.
A decrease of the Debye temperature D , associated with a
decrease of the particle size, was observed by many researches
(Table 5.3). The relative values D (r)/ D were determined by
calorimetric and diffraction methods. However, examination of
small particles of Au and Fe (D = 57 nm) by the Mssbauer effect
shows that the Debye temperature for these particles is the same
as that for bulk crystals [126, 127]. Comparison of the lattice
constant of small particles of Au and Fe with the relative intensity
of scattered X-ray [128] also shows that the detected effects
cannot be explained only by a decrease of the Debye temperature.
According to [5], the contradiction of the experimental data for the
Debye temperature of small particles indicates that it is necessary
to take into account the oscillations of the clusters (metastable
atomic groupings with increased local stability), which form a
nanoparticle and have the symmetry different from that of the
crystal. It is also necessary to take into account anharmonic
effects, which are quite strong in the nanoparticles.

188

Properties of Isolated Nanoparticles and Nanocrystalline Powders


Table 5.3 Size dependence of Debye temperature D (r) for small metal particles
( D is Debye temperature of a bulk metal)
M e ta l

P a rtic le size (nm)

D( r ) / D

Re fe re nc e s

Ag

~ 20

0.75

121

Ag

1020

0.750.83

122

Ag

15

0.735

123

Al

1520

0.500.67

122

Au

2.0

0.69

124

Au

1.0

0.92

11 2

Au

10.0

0.995

11 2

In

0.80

11 5

Pb

2.2

0.87

11 5

Pb

3.7

0.90

11 5

Pb

6.0

0.92

11 5

Pb

20.0

0.84

125

Pd

3.0

0.640.83

11 6

Pd

6.6

0.670.89

11 6

3.8

0.83

11 6

6.5

0.86

11 6


,







PH9




F














K   PH9







Fig. 5.12. Phonon density of states, g( ),


of nanocrystalline compacted nickel Ni (a),
non-compacted powdered nanocrystalline Ni
(b) and coarse-grained bulk Ni (c) [120].

189

Nanocrystalline Materials

5.4 MAGNETIC PROPERTIES


The special features of the magnetic properties of the nanoparticles
are associated with the discrete electronic and phonon states. One
of these special features is the oscillation dependence of the
susceptibility of the nanoparticles of paramagnetic metals on the
strength of the magnetic field H. In addition to this, Curie
paramagnetism may greatly overlap Pauli paramagnetism of metals
due to a small size of the nanoparticles. The results of theoretical
and experimental investigations of the magnetic properties of
nanoparticles of paramagnetics are considered in reviews [10, 11].
The effect of the degeneration of the electronic states on the
magnetic susceptibility of the small particles of paramagnetic
metals taking into account the even or odd number of electrons in
them was discussed in [109, 110, 129]. In weak magnetic fields
p H << ( p is magnetic permeability, is the energy distance
between the adjacent electronic levels) the electronic spin
paramagnetism of the metallic particles with the even number of the
electrons at a low temperature k B T/ << 1 decreases to almost
zero, but does not disappear completely as a result of the existing
weak spin-orbital interaction. In particles with an odd number of
electrons, a decrease in temperature leads to increasing the
paramagnetic susceptibility in accordance with the Curie law [130].
At high temperature k BT/ >1 the paramagnetism of the metallic
particles with the even and odd numbers of the electrons
asymptotically tends to Pauli paramagnetism.
The variation of the magnetic susceptibility of the nanoparticles
of Li (D ~ 1 nm), Pt (D ~ 2 nm) and Al (D ~ 2 nm) in the lowtemperature range according to the Curie law was observed in [131,
132]. According to [133], the magnetic susceptibility of the lithium
nanoparticles with a diameter of 3.2 nm at high temperature
corresponds to Pauli paramagnetism, and in the range of low
temperatures it is governed by the Curie law.
The size dependence of the susceptibility was detected in
particles of selenium Se and tellurium Te with a particle size from
1 to 1000 nm [134]: a decrease in the size of the Se particles is
accompanied by an increase of the diamagnetism, whereas the
magnetic susceptibility of Te increases as a result of an increase
of the Van Fleck orbital contribution.
The measurements of the magnetic susceptibility of Hg 13 and
Ga 13 clusters in a magnetic field with a strength of up to 15 kOe
show that they are weak paramagnetics irrespective of temperature
190

Properties of Isolated Nanoparticles and Nanocrystalline Powders

[135, 136]. However, in a field with H > 20 kOe with a decrease


of temperature below 7080 K, the susceptibility of clusters Hg 13
increased in accordance with the Curie law to high paramagnetic
values ( = 1 cm 3 g 1 at H = 40 kOe), although bulk mercury is
a diamagnetic. According to [137, 138], the magnetic susceptibility
of Na clusters in zeolite is also governed by the Curie law, even
in strong magnetic fields. Variation of the magnetic susceptibility of
Ag clusters in zeolite in accordance with CurieWeiss law in the
temperature interval from 4 to 300 K was detected in [139]. The
increase of paramagnetic susceptibility of the Mg nanoparticles
(D ~ 3 nm) in comparison with bulk magnesium and a large
decrease of susceptibility of the nanoparticles at T 0 are noted
in [140]. According to the authors of [12], these experimental
results are explained by the fact that the very small clusters and
nanoparticles of these metals do not have metallic properties,
because their outer s-electrons are localised on atoms. Therefore,
a normal exchange interaction forms between the atoms in the
clusters. The clusters and nanoparticles of metals lose the metallic
properties with decreasing size. For example, investigation of
photoemission from clusters Pt 6 [141] and tunneling phenomena in
Fe 13 clusters with a volume of 0.15 nm 3 (D ~ 0.5 nm) [142] shows
that these clusters are not metals (although the Fe 35 cluster already
have metallic properties). According to [143], mercury clusters,
containing from 20 to 70 atoms, are characterised by a transition
from the Van-der-Waals crystal to the metal.
In [117] the magnetic susceptibility of nanocrystalline palladium
particles (D = 8 nm) and bulk palladium was measured in the
temperature range of 1.8300.0 K. In the entire temperature range,
nano-Pd and bulk Pd are paramagnetics, a decrease in temperature
increases susceptibility. The (T) dependence of bulk palladium at
T 80 K shows a weak diffuse maximum which was not found on
the similar dependence in the case of nano-Pd. At T > 20 K and
up to 300 K, the susceptibility of nano-Pd is 2025 % lower than
that of bulk palladium. According to [117], the absence of a
maximum on the (T) dependence of Pd nanoparticles indicates a
large difference of the electronic energy spectra of nano-Pd and
bulk Pd in the vicinity of the Fermi energy. The results of magnetic
measurements [117] cause certain doubts because the temperature
dependence of the susceptibility of bulk palladium greatly differs
from that recorded in reliable and accurate experiments [144, 145].
The anomalies of magnetic susceptibility of the nanoparticles
were manifested in investigations by the EPR method. According
191

Nanocrystalline Materials

to [146], a decrease in the size of the nanoparticles should reduce


the width of the EPR lines and this effect should be detected for
particles smaller than 10 nm. However, examination by the EPR of
small particles of Na with a size from 600 to 2 nm [147, 148]
showed a reverse dependence: with a decrease in the size of
sodium particles, the width of the EPR lines increased. A large
broadening of the EPR lines of the Gd nanoparticles (D ~ 10 nm)
in comparison with bulk Gd was detected by the authors of [149].
The nanostructured state affects the properties of ferromagnetics. Ferromagnetic materials have a domain structure which
forms as a result of minimization of the total energy of the
ferromagnetic in the magnetic field. According to [150], this energy
includes the exchange energy, which is minimum for the case of
parallel electron spins; the energy of the crystallographic magnetic
anisotropy, determined by the presence in the crystals of axes of
easy and hard magnetization; magnetostriction energy, associated
with the change of the equilibrium distances between the lattice
sites and the length of the domain; magnetostatic energy, linked with
the existence of magnetic poles both inside the crystal and on its
surface. Closure of the magnetic fluxes inside the domains,
distributed along the axes of easy magnetization, decreases the
magnetostatic energy, whereas any disruptions of the homogeneity
of the ferromagnetic (especially, interfaces) increase its internal
energy.
With a decrease in the size of the ferromagnetic, the closure of
the magnetic fluxes inside the ferromagnetic is less and less
advantageous from the viewpoint of energy. Whilst the ferromagnetic
particles have a multi-domain structure, its interaction with the
external magnetic field is reduced to the displacement of the
boundary layer (wall) between the domains. With approach of the
ferromagnetic particles to the single-domain state, a coherent
rotation of the majority of magnetic moments of the separate atoms
becomes the main mechanism of remagnetisation. This is inhibited
by the anisotropy of the shape of the particles, crystallographic and
magnetic anisotropy. When reaching a certain critical size D c, the
particles consist of single domain and this accompanied by an
increase in the coercive force H c to the maximum value (for
remagnetising the single-domain spherical particle by coherent
rotation it is necessary to apply a reverse magnetic field, i.e.
maximum coercive force H c = 2K/I s , where K is the anisotropy
constant and I s is saturation magnetisation). According to [151], the
largest size of the single-domain particles of Fe and Ni does not
192

Properties of Isolated Nanoparticles and Nanocrystalline Powders

exceed 20 and 60 nm respectively. A further decrease of the


particle size leads to a large decrease of the coercive force to zero
as a result of transition to the superparamagnetic state. The critical
linear size of the particles, D c , corresponds to a disappearance of
ferromagnetism at a temperature below the Curie point owing to
thermal fluctuations of the orientation of the magnetic moment.
Taking into account the Heisenbergs uncertainty principle it was
shown in [150] that the critical linear size of the particles is about
1 nm.
Indeed, if the size of a ferromagnetic particle is equal to some
value 0 , then the momentum p of the electron, which freely
propagates in the volume of the particle, has the uncertainty p.
Heisenbergs uncertainty principle shows that p h / 0 . Part of
the electron energy, determined by the limited size of the particles,
is equal to
0 = (p) 2 /2m e h 2 /2m e 0 2

(5.33)

or, taking into account the values of h and m e


0 6.110 39 / 0 2 ,

(5.34)

where energy 0 is measured in J, and size 0 is measured in


meters. Energy 0 has a disordering effect on the magnetic
moments, similar to the effect of thermal vibrations.
With disruption of the homogeneity of magnetisation, a correction
must be introduced to the energy of exchange interaction. This
correction is maximum when the magnetisation vector changes its
direction to reversed at distances of the order of the distance
between the adjacent metallic atoms, i.e. at distances at the order
of lattice constant a. The physical meaning of the correction is that
the exchange energy tends to maintain the isotropy of magnetisation
at any disruption. In other words, the exchange energy is the energy
of magnetic ordering. The maximum correction of the exchange
max
energy is exch
AV / a 3 , where A is the exchange energy, V is the
volume of the solid. Complete disruption of the isotropy of
magnetisation and disorientation of the magnetic moments take place
at Curie temperature T C when the spontaneous magnetisation of
max
should be
ferromagnetic decreases. Therefore, the correction exch
equal to or slightly lower than the thermal energy k BT CV/a 3 . From
this it follows that the exchange energy is

193

Nanocrystalline Materials

A k BT C .

(5.35)

Equating the disordering energy 0 (5.34) and the ordering energy


of exchange A (5.35), it is possible to evaluate the critical linear
size 0 of the ferromagnetic particle at which the ferromagnetism
disappears at all temperatures because of disordering of the
magnetic moment under the effect of energy 0 :

0 210 8 T C1/2 [m].

(5.36)

According to (5.36), for ferromagnetics with a Curie temperature


of 5001000 K the critical size of the particle at which
ferromagnetism disappears and transition to the superparamagnetic
state takes place, is ~1 nm. The exchange energy is slightly lower
than k B T C and, therefore, quantity d 0 might be slightly higher, as
indicated by the estimation using equation (5.36). For typical
ferromagnetics, the transition to the superparamagnetic state is
possible when the particle size becomes smaller than 110 nm.
Analysis of the literature data on the dependence of coercive
force H c on the mean size of the ferromagnetic particles [4]
confirms the increase of H c with a decrease in the particle to some
critical size. The maximum values of H c are obtained for particles
of Fe, Ni and Co with a mean diameter of 2025, 5070 and 20 nm,
respectively. These values are close to theoretical estimates of D c
of single-domain particles [151]. As regards a decrease of H c at
D < D c, it may be associated not only with the superparamagnetism
effect but also with other magnetic properties of the surface layer.
For example, if the anisotropy of the surface layer is reduced, the
layer will be remagnetised in weaker fields and facilitate the
magnetisation of the entire nanoparticle [151]. The dependence of
the relative residual magnetisation I r /I s (here I s is saturation
magnetisation of bulk metal) on the size of particles of Fe, Co and
Ni also pass through a maximum in the vicinity of values of D c for
these metals [4].
A decrease of saturation magnetisation with a decrease of the
size of Fe, Ni and Co nanoparticles in ferromagnetic alloys was
detected in many studies [152159]. The authors of [4, 152156]
treat the decrease of I s as the result of oxidation of the surface
layer of metallic nanoparticles, whereas in [157159] the decrease
of I s was explained directly by the size effect.
The authors of [159] have investigated the magnetic properties
of spherical particles of iron with a diameter of 40100 nm at
194

Properties of Isolated Nanoparticles and Nanocrystalline Powders

temperatures at 4.2300 K in fields with a strength of up to


25 kOe. Particles were suspended in paraffin and their volume
concentration was 0.01. Examination by nuclear gamma resonance
shows that the iron particles are not oxidised. Measurements of the
coercive force of particles of different sizes at 4.2, 77 and 300 K
showed a distinctive maximum of H c at D ~ 24 nm. According to
[159], this maximum is determined by the superposition of two
process: increase of H c in transition of the particles to the singledomain state and the appearance of superparamagnetism in singledomain particles when they reach the critical size. The saturation
magnetisation I s, even for the largest particles of iron (D ~ 98 nm),
was smaller than saturation magnetisation of bulk iron; with a
decrease of the particle size from ~ 40 to 35 nm, I s decreases and
with a further decrease of size the saturation magnetisation remains
constant. The maximum of the I r /I s ratio is observed for particles
with a size of 24 nm (I r is a residual magnetisation). According
to [159], transition of Fe particles from ferromagnetic to
superparamagnetic state is observed when particle size is 24 nm.
Investigations of the saturation magnetisation of bulk Ni and
nanocrystalline powder of Ni (D = 12, 22 and 100 nm) at 10300
K [119] show that with a decrease of the particle size to 12 nm,
I s is almost halved in comparison with bulk nickel. At temperatures
below 50 K for nanoparticles of Ni with D < 50 nm, the magnetic
hysteresis loop was asymmetric. According to [119] the
displacement of the hysteresis loop and a decrease in I s are
associated with a presence of the surface oxide shell and are
determined by the anisotropy of the exchange interaction of
ferromagnetic nickel with the antiferromagnetic oxide NiO, which
forms the shell of nanoparticles.
On the whole, at present there is no common view regarding the
reasons for the change in the saturation magnetisation of
ferromagnetic nanoparticles. It is important to mention the results
of investigation of compacted nanocrystalline nano-Ni with a mean
grain size of 70100 nm [160]. The saturation magnetisation of
compacted nano-Ni was approximately 10 % lower than that of
coarse-grained nickel; the same was also reported for
submicrocrystalline nickel produced by deformation and heat
treatment [161]. To explain this effect, the authors of [162]
assumed that the atoms distributed in the vicinity of grain
boundaries, being in the non-equilibrium state, are dynamically more
active than the atoms in the grains and form a grain-boundary
phase. The difference in the magnetic properties of the grain195

Nanocrystalline Materials

boundary phase in relation to the properties of the phase inside the


grain is the reason for a decrease of the saturation magnetisation
of submicrocrystalline nickel. It may be possible that the decrease
of the saturation magnetisation of nanoparticles may be associated
not only with oxidation of the surface or with the size effect, but
also with the special state of the surface layer of isolated
nanoparticles or powder nanoparticles. At the same time, the
presence of the developed surface of nanoparticles itself is a
consequence of their small size.
Recently, the dependence of the coercive force on the size of
nanoparticles of Fe, Ni and nanoparticles of Fe 0.91 Si 0.09 alloy was
studied by the authors of [163, 164]. The nanocrystalline powders
of Fe, Ni and Fe 0.91 Si 0.09 alloy with a minimum particle size of 8,
12 and 6 nm, respectively, were produced by ball milling for 380,
350 and 180 hours. The magnetic measurements showed that a
decrease in the size of iron nanoparticles from 80 to 810 nm is
accompanied by an increase in coercive force H c by almost a
factor of 3. The dependence of coercive force on the size of
particles nano-Ni showed a maximum corresponding to nanoparticles
with a diameter of 1535 nm; with a decrease in the particle size
from 1512 nm, the value of H c rapidly decreases almost 5 times.
Saturation magnetisation I s of nano-Ni particles (D = 10 nm) was
37 % higher than that of bulk nickel. However, this may be
associated with the appearance of an impurity of 15 at.% Fe as a
result of milling. A decrease in the size of nanoparticles of
Fe 0.91 Si 0.09 from 40 to 6 nm is accompanied by an increase of the
coercive force 5 times.
The size and temperature dependences of the coercive force of
cobalt nanoparticles with nitrided and oxidised surfaces were
investigated in [165]. The size of Co nanoparticles was equal to
1560 nm. The coercive force of the nitrided and oxidised particles
of Co increased with a decrease of temperature from 240 K and
200 K for particles with a size of ~ 10 nm and 3050 nm
respectively. The highest value H c 2 kOe was recorded at 5 K
for nanoparticles with a diameter of 34 nm. According to [165],
oxidation leads to a larger increase of H c of the cobalt nanoparticles
than nitriding. It should be mentioned that the increase of H c as a
result of oxidation of the Fe and Co nanoparticles was detected
previously in [166168].
A large (up to ~ 800 K) increase of the Neel temperature was
found in bcc Cr nanoparticles with a diameter of 3875 nm [169],
although bulk chromium is antiferromagnetic with a Neel
196

Properties of Isolated Nanoparticles and Nanocrystalline Powders

temperature of 311 K.
Interesting results were obtained in examining the magnetic
properties of nanoparticles of hematite -Fe 2 O 3 [170]. In the
normal state, hematite is an antiferromagnetic. Measurements show
that with a decrease in the particle size from 300 to 100 nm, the
magnetic susceptibility of nanocrystalline hematite does not change
and equal to the magnetic susceptibility of the bulk crystal. A
decrease in the particle diameter from 100 to 20 nm resulted in a
rapid increase of magnetic susceptibility.
Analysis of this spontaneous magnetisation of the nanoparticles
[171] carried out in the approximation of the molecular field, shows
the presence of the size dependence of Curie temperature.
According to [171], a decrease in the Curie temperature becomes
noticeable for particles with a size of D < 10 nm; for particles with
D = 2 nm, a decrease in T C in comparison with bulk metal does not
exceed 10 %. However, examination of the thermodynamics of
superparamagnetic particles by the Monte-Carlo method does not
detect any dependence of the Curie temperature on the particle size
[172]. In fact, the transition of the nanoparticles from the
superparamagnetic state to paramagnetic state is smooth, without
any sharp point of magnetic transformation. Measurements of the
Curie temperature of nanoparticles of nickel (D = 2.16.8 nm)
[173], saturation magnetisation and the Curie temperature of Fe films
with a thickness of 1.5 nm [174], saturation magnetisation of Fe
nanoparticles (D = 1.5 nm) [175] and Co nanoparticles (D =
0.8 nm) [176] show that these values coincide (within the
measurement error range) with those for bulk metals. According to
[4, 5], the Curie temperature of the ferromagnetic particles with a
decrease of their size to 2 nm does not differ from that of bulk
metals. However, the authors of [177] detected a decrease of T C
by 7 and 12 % for nickel nanoparticles with a diameter of 6.0 and
4.8 nm, respectively. It should be noted that the superparamagnetism
phenomenon greatly complicates the investigation of the size
dependences of the coercive force, saturation magnetisation and
Curie temperature of the ferromagnetic nanoparticles.
A decrease in the size of the single-domain particle leads to the
transition from ferromagnetic state to superparamagnetic state.
Thermal fluctuations may cause rotation of magnetic moments, if
the mean thermal energy k B T is equal to or higher than the
anisotropy energy of E = KV, where K is the constant of anisotropy
and V is the volume of the particle. The total magnetisation of the
particle, which is achieved in an external magnetic field with
197

Nanocrystalline Materials

sufficient strength for saturation, becomes equal to zero after


switching the field off during relaxation time r . In the model of
discrete orientations [178], the relaxation time is

r = 0 exp( KV / k BT ) .

(5.37)

If measurement time m is greatly shorter than the relaxation time


r , the particle retains the initial ferromagnetic state. Otherwise,
when the measurement time m is longer than r , the thermal
fluctuations completely disorientate magnetic moments and the
particle will behave a superparamagnetic one. The transition from
the ferromagnetic to the superparamagnetic state takes place at a
certain blocking temperature T = T B , for which r = m . Taking
equation (5.37) into account, the blocking temperature is

TB = KV /[ k B ln( m / 0 )] .

(5.38)

A nanoparticle of a ferromagnetic material which has the volume


V, at T < T B behaves as a ferromagnetic, and at T > T B is in the
superparamagnetic state.
For the given temperature, the condition r = m also determines
critical volume V B (blocking volumes): a nanoparticle with V < V B
is in the superparamagnetic state, a nanoparticle whose volume is
larger than critical (V > V B ) is a ferromagnetic. Estimates [150]
show that for typical ferromagnetics or ferrimagnetics at 100 K the
critical volume is 10 27 10 23 m 3 and corresponds to nanoparticles
with linear size smaller than 115 nm.
Superparamagnetism is found in nanoparticles (D 10 nm) of Ni
in matrixes of silicagel [179] and Pb [80]; Co in the Cu matrix [181]
and in Hg [182]; Fe in Hg [175, 182] and in -brass [183]. The
experimental data on superparamagnetism have been reviewed in
detail in [4, 5]. Therefore, only recent experimental investigations
will be discussed briefly in this section.
Detailed investigation of the magnetic properties of cobalt
nanoparticles with a diameter of 1.8 to 4.4 nm, produced by
deposition from a colloid solution, was carried out in [184]. The
magnetic properties were measured using a SQUID-magnetometer
in the temperature 2300 K in a field with a strength of up to
50 kOe. At 300 K the Co nanoparticles were superparamagnetic.
The variation of the blocking temperature T B from 22 to 50 K with
an increase of the particle size from 1.8 to 4.4 nm was described
198

Properties of Isolated Nanoparticles and Nanocrystalline Powders

by the dependence T B = KV/30k B , i. e. by the function (5.38). Using


the dependence T B (V) and experimental results for T B and the
particle size, the authors of [134] found the size dependence of
anisotropy constant K: the anisotropy constant increases with a
decrease in the particle size and in the entire examined range
1.8 D 4.4 nm it is larger than K of bulk fcc cobalt. The size
dependence of coercive force H c was measured at 10 K at which
the nanoparticles of all sizes were ferromagnetic. Increase in H c
with an increase of the size of nano-Co particles completely
corresponds to the behaviour of single-domain particles. The size
dependences of T B , K, H c of the cobalt nanoparticles are in good
agreement with similar data for nanoparticles of other ferromagnetic
metals. A different situation exists in the case of magnetisation.
Measurements show that at T = 2 K the nanoparticles of Co do not
reach magnetic saturation even in a field of 55 kOe. Therefore, the
values of saturation magnetisation I s were obtained by the
extrapolation of the dependence I(1/H) to an infinitely large field,
i.e. 1/H 0. The value of I s increases with a decrease in the size
D, and I s of particles with D < 3.3 nm is higher than I s of bulk
cobalt. The saturation magnetisation of the smallest Co particles
(D = 1.8 nm) is 20 % higher than I s of bulk cobalt. An increase
of the magnetic moment of the cobalt atom in the nanoparticles is
predicted theoretically by the authors of [185, 186] and is observed
in experiments [187] on cobalt clusters.
The nanocrystalline powder of -Fe 2 O 3 (D ~ 47 nm) was
synthesised by the plasma chemical method using a microwave
generator [188]. The magnetic measurements show that
nanoparticles of -Fe 2 O 3 are superparamagnetic with a blocking
temperature of T B 80 K. With a decrease T < T B the particles
of -Fe 2 O 3 behave as a ferrimagnetic, their residual magnetisation
increases, reaching the maximum at 20 K, and then decreases.
The size dependence of the blocking temperature of the -Fe 2 O 3
nanoparticles with a size from 3 to 10 nm, distributed in a polymer
matrix, is determined in [189]. The dependence T B (V) is close to
linear and is described by a function of type (5.38). The blocking
temperature of the particles with a volume of ~100 nm 3 (D ~
45 nm) is equal to ~75 K and is in good agreement with the
results [188].
The superparamagnetism of Fe nanoparticles, distributed in a
copper matrix, was investigated in [180]. Starting copper, containing
approximately 0.01 at.% of dissolved Fe, was diamagnetic in the
entire examined temperature range 3001225 K. The temperature
199

Nanocrystalline Materials

dependence (T) of starting copper is in good agreement with the


data [191]. As a result of severe plastic deformation of starting
copper by equal channel angular pressing (the true logarithmic
degree of deformation e was 3.5), the authors obtained submicrocrystalline copper smc-Cu with a grain size of 130150 nm and
observed precipitation of iron particles, previously dissolved in
copper. Magnetic measurements were performed in a vacuum of
1.310 3 Pa (10 5 mm Hg) on a high-sensitive magnetic balance in
a field with an induction of 8.8 kGs. To understand the results, it
is important to know the measurement procedure: heating from 300
K to annealing temperature T; holding the specimen for 1 h at this
temperature and measurement of the susceptibility at the end of
holding cycle; cooling from annealing temperature to 300 K and
measuring the susceptibility at 300 K; heating to the next annealing
temperature, and so on (Fig. 5.13). The annealing temperature was
varied from 300 to 1225 K with a step of 25 K. The measurements,
made directly at the annealing temperature, relate to the
temperature dependence of susceptibility and will be henceforth
denoted by (T). The susceptibility measurements, carried out after
annealing and subsequent cooling to 300 K, refer to the annealing
curve and will be henceforth designated as (300, T). The results
of measurements are shown in Fig. 5.14.
The measurements show that the susceptibility of
submicrocrystalline copper is higher than that of starting copper. In
addition to this, it was observed that the susceptibility of smc-Cu
is inversely proportional to the magnetic field strength H. This
indicates the presence of a ferromagnetic impurity of iron in the









9 KRXUV

Fig. 5.13. Sequence of magnetic susceptibility measurements in annealing experiments:


( ) measurements in situ (at annealing temperature T), ( l ) measurements after
cooling from annealing temperature to 300 K [192].
200

Properties of Isolated Nanoparticles and Nanocrystalline Powders






 FP J




























Fig. 5.14. Magnetic susceptibility of nanocrystalline copper (n-Cu) matrix with


iron impurity in a magnetic field of H = 8800 Gs [190]: ( ) temperature dependence
c(T); (l) annealing dependence (300, T); () reverse temperature trend of susceptibility,
corresponding to susceptibility of copper with 0.01 % of dissolved iron impurity

specimen. It is known that dissolved iron is precipitated from


copper during rolling [193]. In [190], the precipitation of iron
particles, previously dissolved in copper, was initiated by severe
plastic deformation.
The annealing dependence (300, T) in the vicinity of the
nanotransition temperature T n 425 K (transition of copper from
submicrocrystalline to coarse-grained state) exhibits a pronounced
abrupt increase of susceptibility (Fig. 5.14). Over the temperature
interval from 450 to 650 K the (300, T) curve does not virtually
change. As the annealing temperature is raised further, the
(300, T) curve increases gradually, passes through a maximum at
a temperature of 975 K and then decreases drastically to
diamagnetic values, corresponding to the susceptibility of copper
free from ferromagnetic impurities (Fig. 5.14). The susceptibility
(300, T) of smc-Cu ceases to depend on H after annealing at
T > 1200 K.
Measurements of the temperature dependence of susceptibility,
(T), revealed a large decrease of even at a temperature under
425 K. After a slight increase of the susceptibility in the range
425475 K, the temperature curve descends to diamagnetic values
and at a temperature of 850 K transforms to the temperature
dependence of the susceptibility of copper free from ferromagnetic
impurities. The dependence of on the field strength H disappears
at T > 850 K.
201

Nanocrystalline Materials

The reverse temperature trend of the susceptibility from 1225 to


300 K (Fig. 5.14) corresponds to the susceptibility of copper with
0.01 at.% of iron impurity dissolved in it [194]. There is no
magnetic field strength dependence of susceptibility in this case.
One of the most interesting experimental results discovered in
[190, 195197] is a jump on the annealing and temperature
dependences of magnetic susceptibility at a temperature of
T n 425 K, corresponding to the transition of submicrocrystalline
copper to coarse-grain copper. Is the detected jump associated with
the change in the susceptibility of copper proper? The main
contributions to the magnetic susceptibility of bulk crystalline copper
are the diamagnetism of atomic cores, the Pauli spin paramagnetism
and the Landau diamagnetism of conduction electrons. The sum of
these contributions for copper is negative and, therefore, copper is
a diamagnetic. The weak quadratic temperature dependence of
susceptibility [191] is determined by the Pauli contribution. In the
considered case, the sign of the susceptibility of smc-Cu is positive
as a result of the precipitation of iron particles. The lowest
susceptibility of submicrocrystalline copper (Fig. 5.14, dependence
(300, T)) could be a consequence of the lower density of states
on the Fermi level and the lower effective mass of conduction
electrons. However, this should not result in any large change of
the temperature dependence of the susceptibility. Therefore, the
difference in susceptibility values on the annealing and temperature
curves of smc-Cu after the nanotransition ( = (300, T) (T) at
T ~ 500 K), cannot be explained only by change in the state of
copper.
The analysis carried out in [190, 195197] showed that the jump
of on the annealing and temperature dependences of susceptibility
at 425450 K is probably not associated with the change in the
susceptibility of copper but with a variation of the magnetic
contribution from the iron impurity, precipitated in the form of
nanoparticles at joints of copper grains.
If one assumes that the jump on the dependences (T) and
(300, T) at the nanotransition temperature in smc-Cu is associated
with a change in the magnetic contribution by an impurity, then one
can subtract from (T) the susceptibility of copper, Cu (T), and
thereby can determine the contribution of the ferromagnetic phase,
Fe (T), to susceptibility (Fig. 5.15). This phase may be a surface
phase or bulk phase. If this is a two-dimensional surface phase, it
can lie on the boundary of two grains. If the phase is a threedimensional one, it is most likely to be located at the joints of three
202

Properties of Isolated Nanoparticles and Nanocrystalline Powders






%.0

0  FP J


%3



CHECK SYMBOLS IN CAPTIOM

















Fig. 5.15. Approximation of the susceptibility versus temperature curve for the
iron impurity superparamagnetic phase [190]: (1) and (2) are the variations of
susceptibility after the nanotransition with and without allowance for the dissolution
of Fe impurity; (3) variation of susceptibility before the nanotransition; ( o ) value
of susceptibility of the superparamagnetic impurity at 300 K after the nanotransition
in copper. The vertical dashed lines show the region of the nanotransition near
temperature T n , and also Curie temperature TCFe of crystalline iron.

or more grains. Let us consider in detail the interpretation of Fe (T)


dependence, proposed by the authors of [190, 197]. Interpretation
was carried out in the approximation of precipitation of iron
particles of the same size and the independence of Curie
temperature T C of iron on the nanoparticle size.
For conventional ferromagnetics, the temperature dependence of
the susceptibility in saturating magnetic fields at low temperature
is not so strong as that observed experimentally (Fig. 5.15) [190].
A strong dependence at low temperature is possible in the case of
the superparamagnetism of precipitated iron particles. Expressed in
dimensionless units, the superparamagnetic contribution sp , at
temperature T in magnetic field H may be represented in the form
[150]:

sp = nspVsp

M s (T ) Vsp M s (T ) H
,
L

k BT
H

(5.39)

with L = [coth(x) 1/x] being the Langevin function, n sp the


number of superparamagnetic particles in the unit volume, V sp the
volume of the superparamagnetic particle, M s (T) the saturation
203

Nanocrystalline Materials

magnetisation of the ferromagnetic phase at temperature T.


The temperature dependence of the saturation magnetisation for
the crystalline ferromagnetic phase, M s (T), is determined by solving
the equation

M (T )TC
M s (T )
= tanh s
M s (0)
M s (0)T

(5.40)

where M s (0) is saturation magnetisation at T = 0 K. For crystalline


iron M s (0) = 1740 Gs and T C = 1043 K [198]. To the quoted
saturation magnetisation M s (0) = 1740 Gs corresponds the iron
atomic magnetic moment 2.22 B ( B is the Bohr magneton).
The ferromagnetic contribution disappears at a temperature of
850 K, which is much lower than the Curie temperature of
conventional polycrystalline iron (Fig. 5.14, 5.15). This may be due
to complete equilibrium dissolution of the ferromagnetic impurity in
copper even at a temperature of 850 K. According to the phase
diagram [193], the undissolved iron concentration in copper
decreases and the dissolved iron concentration, c Fe (T), in copper
decreases with an increase of temperature in accordance with the
equation
c Fe (T) = c Fe (0) Cexp(E m/k B T) ,

(5.41)

where c Fe (T) is the relative atomic concentration of dissolved Fe


at 0 K or the concentration at 300 K, which is virtually equal to
this concentration; C is a constant; E m is the energy of mixing.
According to [193] C = 43, E m/k B = 9217 K or E m = 0.79 eV.
To go from the dimensionless bulk susceptibility sp and atomic
concentration c Fe over to the mass susceptibility of the ferromagnetic phase, whose concentration varies owing to dissolution,
one can use the relation

Fe (T ) = sp (T , nsp ,Vsp ) cFe (T )

AFe
,
ACu Fe

(5.42)

where A Cu = 63.55 and A Fe = 55.85 are the atomic weights of


copper and iron, Fe = 7.86 g cm 3 is the density of iron.
Approximating the temperature dependence of susceptibility Fe (T)
in the temperature range from 425 to 1043 K by equation (5.42),
204

Properties of Isolated Nanoparticles and Nanocrystalline Powders

with allowance for equations (5.39)(5.41), and assuming the Curie


temperature equal to 1043 K and the mixing energy equal to
E m = 0.79 eV, the authors of [190, 197] obtained a good agreement
with the experimental data (Fig. 5.15, curve 1). Curve 1 also passes
at 300 K through the point labelled which correspond to
(300, T) values in the annealing temperature range 450600 K
and would lie on the temperature curve if there was no
susceptibility jump at the nanotransition in copper. Curve 2 (Fig.
5.15) was plotted without allowing for the dissolution of iron, i.e.
with the ferromagnetic impurity concentration remaining unchanged
up to the Curie temperature. The fitting (Fig. 5.15, curve 1) gave
the values of the volume of the superparamagnetic particle, V sp =
1.810 20 cm 3 , and the number of particles, n sp = 5.710 14 cm 3 ,
after nanotransition, and also constant C = 0.4. This value of
constant C is approximately 100 times lower than that in [193] and
shows that the rate of dissolution of iron, obtained by authors [190,
197], is higher than that indicated in [193]. The high rate of
dissolution of iron is a consequence of the fact that the precipitation
of nanoparticles of Fe and their existence in the copper matrix are
thermodynamically non-equilibrium.
If the dissolution of iron in copper up to a temperature of
650 K is ignored, then the relative volume concentration of the
superparamagnetic impurity, n sp V sp , in copper before and after
nanotransition is the same. Consequently, in accordance with (5.39),
the superparamagnetic contributions at 0 K are the same too.
Assuming that the Curie temperature does not depend on the size
of superparamagnetic particles and that the susceptibility of copper
remains unchanged during the transition, the experimental data for
the susceptibility prior to the nanotransition may be approximated
(Fig. 5.15, curve 3). Approximation shows that prior to the
nanotransition, the volume of the superparamagnetic particles was
a factor of 1.62 smaller and the mean particle size a factor of
(1.62) 1/3 = 1.17 smaller than after the transition. The difference in
the temperature dependences of susceptibility before and after the
nanotransition comes from the increase in the mean size of the
superparamagnetic particles from 2.8 to 3.3 nm.
The volume of copper per 1 superparamagnetic impurity particle
is V = 1/n sp . It is possible to determine the linear size of the copper
particle. It turns out that for each superparamagnetic particle there
is one copper particle 128 nm in diameter prior to the nanotransition
and a particle 150 nm in diameter after the transition. These sizes
are similar to the grain size of copper in this specimen prior to and
205

Nanocrystalline Materials

after the transition. Therefore, one may assume that the impurity
particles are distributed in copper uniformly and there is one iron
particle per every copper grain. An iron particle may be located,
for example, in the space where several grains are contacted.
When the copper grains coarsen, the number of grain joining points
decreases and the impurity iron atoms have to diffuse over the
surface of copper grains to the points that remain. The impurity
nanoparticles already present coarsen and their number decreases.
Similar processes also take place at higher annealing temperature,
when the copper grain growth continues.
In the temperature 450600 K, the annealing dependence
(300, T) (Fig. 5.14) is almost constant. This means that the state
of the superparamagnetic impurity, i.e. the number and size of the
particles, does not change when the specimen is heated in this
temperature range and subsequently cooled. For the temperature
dependence (T) this is confirmed by the results of calculation (Fig.
5.15, curves 1 and 2) which shows that the dissolution of iron at
450600 K is negligible. The increase of susceptibility (300, T)
(Fig. 5.14) by the amount ~110 7 cm 3 g 1 , observed after
annealing at temperatures between 650 and 975 K, is partially
associated with an increase in the size of superparamagnetic
particles in copper cooled to 300 K and with an increase in the
contribution from the impurity at 300 K. However, this factor may
be responsible for an increase of (300, T) by an amount of ~2
10 8 cm 3 g 1 only. The remaining increase in the susceptibility
should be associated with other factors, for example, the smaller
saturation magnetisation M s of nanoparticles in comparison with that
of a bulk crystal, or with the precipitation of a larger quantity of
ferromagnetic phase when the specimen is cooled.
According to [190, 195197], the decrease of susceptibility in the
temperature range 10001225 K is detected only at a high cooling
rate of the specimen and is a result of quenching of the high
temperature state, in which the whole of the ferromagnetic impurity
is dissolved in copper. If cooling after annealing is slow, the iron
impurity has a chance to transfer to the ferromagnetic phase and
the decrease in susceptibility (300, T), observed in Figure 5.14, is
absent after the maximum.
The investigations [190, 195197] show that the measurement of
magnetic susceptibility is an information-carrying method of
examining the behaviour of ferromagnetic nanoparticles in a
diamagnetic matrix. The presence of the matrix prevents the rapid
growth of the nanoparticles at a temperature of structural relaxation
206

Properties of Isolated Nanoparticles and Nanocrystalline Powders

of the ferromagnetic polycrystal and, at the same time, greatly


increases the temperature range of existence of the nanocrystalline
state of the ferromagnetic.
5.5 OPTICAL PROPERTIES
The scattering and absorption of light by the nanoparticles has a
number of special features in comparison with macroscopic solids
[199]. Experimentally, these special features are most evident when
examining a large number of particles. For example, colloid solutions
and granular films may be extensively coloured as a result of the
specific optical properties of nanoparticles. A classical object for
investigating the optical properties of the dispersed media is gold.
Even Faraday paid attention to the similarity of the colour of a
colloid solution of gold and a gold film and assumed the dispersion
state of the gold film.
In absorption of light by fine-grained metal films the visible part
of the spectrum contains absorption peaks which are absent in the
spectra of bulk metals (in metals the optical absorption by
conduction electrons takes place in a wide wavelength range). For
example, the granular films of gold particles with a diameter of
4 nm have a distinctive maximum absorption in a range of =560
500 nm [200, 201]. The spectra of absorption of nanoparticles Ag,
Cu, Mg, In, Li, Na, K also had maxima in the optical range [4,
202].
Another special feature of the granular films is a decrease of
their absorption in transition from the visible to infrared part of the
spectrum in contrast to continuous metal films in which the
absorption of radiation increases with increasing wavelength [4,
201, 203207].
The size effect of the optical properties are important for the
nanoparticles whose size is considerably smaller than the
wavelength and does not exceed 1015 nm [9, 199].
The differences in the absorption spectra of nanoparticles and
bulk metals are caused by the differences of their dielectric
permittivity = 1 + i 2 . The dielectric permittivity of the
nanoparticles with a discrete energy spectrum depends on both the
particle size and on radiation frequency. In addition to this, the
value of dielectric permittivity does not depend monotonically on
frequency but oscillates as a result of transitions between electronic
states [208]. The minimum number of particles, which requires for
the experimental investigation of the optical properties, is at least
207

Nanocrystalline Materials

10 10 . It is not possible to produce 10 10 10 13 particles of the same


size and shape and, consequently, in experiments in practice these
oscillations are smoothed out. Nevertheless, even the value
averaged out for the ensemble of particles differs from the value
of dielectric permittivity of the bulk substance. According to [208,
209], the imaginary part of dielectric permittivity is inversely
proportional to particle radius r

2 ( ) = 2 ( ) + A( )/r ,

(5.43)

where e 2 ( ) is an imaginary part of the dielectric permittivity of


the macroscopic crystal and A( ) is some function of frequency.
Experimental results [210, 211], obtained on particles of gold with
r = 0.93.0 nm at a constant wavelength = 510 nm, confirm the
dependence 2 ~ 1/r. The particle size also determines the width
of the absorption band and the shape of the low-frequency edge of
the absorption band. The broadening of the absorption band of light
by the gold and silver nanoparticles with a decrease of the size is
observed by the authors [210, 212, 213].
The shift of the resonance peak of absorption of light is another
size effect. The free path length of the electron in metallic
particles, whose diameter is smaller than the free path length of
electrons l in the bulk metal, is equal to particle radius r [4, 5].
In this case, the effective relaxation time ef may be represented
in the form

ef1 = 1 + vF / r ,

(5.44)

where = l /v F is the relaxation time in bulk metal; v F is the


velocity of the electron with the Fermi energy. If interband
transitions are neglected and the movement of only free electrons
takes into account [214], then

1 = 1

2P
,
12 + 1/ ef

(5.45)

where p = 4 Ne 2 /m * is the plasma frequency; N, e, m* is the


concentration, charge and effective mass of free electron. In Mies
theory [215] the maximum of light absorption is reached on the
condition m = 1 ( 1 ); taking this into account, for very small
208

Properties of Isolated Nanoparticles and Nanocrystalline Powders

particles with ef1 ~ vF / r equation (5.45) gives the expression for the
resonance frequency

p2
vF2

1 =
2
1
+
2

r
m

1/2

(5.46)

According to (5.46), the resonance frequency decreases with a


decrease in the particle size, i.e. the absorption band should be
displaced to the low-frequency range. The red shift of the
resonance peak of absorption of light with a decrease in the particle
size is predicted by theory [216]. On the other hand, quantummechanical calculations [208, 217] predict an increase in the
frequency of the resonance peak, i.e. the blue shift of the
absorption band with a decrease in the nanoparticle size.
Experimental results obtained for the displacement of the
frequency of resonance absorption in relation to the nanoparticle
size are also contradicting. In [218220] with a decrease in the size
of Ag particles from 10 to 1 nm there was strong red shift of the
absorption peak. According to the data [221223] the position of
the absorption peak of silver and gold particles with a diameter of
2.510.0 nm does not depend on the particle size. The blue shift
of the absorption peak of Ag nanoparticles with a decrease of their
size to 12 nm was established in [208, 212, 213, 224].
In [225, 226] it is shown that both blue and red shifts can be
observed. It depends on the degree of smearing of the electron
cloud on the surface of the particle. For transition from one effect
to another it is sufficient to change slightly the size of the region
of diffusion smearing of the electrons. According to [224, 226],
the width of the absorption band of light is a complicated function
of the particle size and reaches a maximum in the vicinity of
D 1.1 nm.
Recently, special attention has been paid to the investigations of
the size effect on the optical and luminescent properties of
semiconductor substances, because optical absorption is one of the
main methods of studying the band structure of semiconductors.
In semiconductors, the energy of interatomic interactions is high.
Therefore, when describing the electronic properties, the
macroscopic semiconductor crystal might be regarded as a single
large molecule. The electronic excitation of semiconductor crystal
may leads to the formation of a weakly bonded electronhole pair,
209

Nanocrystalline Materials

i.e. MottVanie exciton. The region of delocalisation for such the


exciton is considerably greater than the lattice constant of the
semiconductor. A decrease of the size of the semiconductor crystal
to the value comparable with the size of the exciton affects the
semiconductor properties.
Thus, the specific properties of semiconductor nanoparticles are
determined by the fact that the size of the nanoparticle is
comparable with the Bohr radius of the excitons, r ex n 2 h 2 / ex 2 ,
in the macroscopic crystal (here ex = m e m h /(m e + m h ) is the
reduced mass of the exciton; m e and m h are the effective masses
of the electron and the hole; n = 1, 2, 3). For semiconductors,
the Bohr radius of the exciton changes in a wide range: from 0.7
nm for CuCl to 10 nm for GaAs. The energy of the electronic
excitation of a molecule is usually higher than the energy gap (width
of the forbidden band) in the macroscopic semiconductor. It follows
from this that when going from a crystal to a molecule, i.e. with
a decrease in the particle size, there should be a range of particle
size in which the energy of the electronic excitation changes
smoothly from a lower to a higher value. In other words, a decrease
in the size of the semiconductor nanoparticles should be
accompanied by displacement of the absorption band to the highfrequency range. A manifestation of this effect is the blue shift of
the exciton band of absorption of semiconductor nanoparticles with
a decrease in their size [227231]. In the most widely studied
semiconductor CdS, the blue shift of the absorption band was found
for nanoparticles with D 1012 nm. The effect of the size of the
nanoparticles on the optical spectra was found for many types of
semiconductor [228243].
For a macroscopic crystal, the exciton energy E includes two
contributions. First of them is the width of the forbidden band E g
(the energy difference between the lower energy of conduction
band and the upper energy of valence band), reduced by the
binding energy of the electron and the hole, i.e. the effective
Rydberg energy E Ry = ex e 4 /2n 2 h 2 . The second contribution is
kinetic energy of the centre of masses of the exciton. For a
semiconductor nanoparticle with radius r, the second contribution
is equal to n 2 p 2 h2 /2 ex r 2 [227] and is inversely proportional to the
square of the radius. A more rigorous analysis [229, 244] of the
effect of the nanoparticle size on the exciton energy and taking into
account the Coulomb interaction between the electron and the hole
give the following expression:

210

Properties of Isolated Nanoparticles and Nanocrystalline Powders

E = E g 0.248E Ry + (n 2 2 h 2 /2 ex r 2 ) (1.78e 2 / r ) .

(5.47)

The sum of the first and third terms in (5.47) is the effective width
of the forbidden band. Equation (5.47) shows that a decrease in the
particle size should be accompanied by an increase of the effective
width of the forbidden band. In particular, this broadening effect is
observed for CdTe nanoparticles: when going from a bulk crystal
to nanoparticles with a diameter of 4 and 2 nm, the effective width
of the forbidden band increased from 1.5 eV to 2.0 and 2.8 eV
respectively [245]. An increase in the width of the forbidden band
of the very fine dispersed powder Si 3 N 4 in comparison with a bulk
crystal was detected in examination of infrared and fluorescent
emission spectra [246].
The excitation energy of the exciton is E = h where is the
frequency of incident light. Therefore it follows from equation (5.46)
that with a decrease in the size of the nanoparticles, the lines of
the optical spectrum should be shifted to the high-frequency range.
This shift of the absorption bands by 0.1 eV in the spectra of
nanoparticles of CuCl (D = 31, 10 and 2 nm), dispersed in glass,
was observed in [228].
As an example, Fig. 5.16 shows the optical spectra of absorption
of CdSe [243]: with a decrease of the diameter of the CdSe
nanoparticles, the absorption bands are shifted to the range of
higher energies, i.e. blue shift is observed. To a first approximation,
%  .

2SWLFDO GHQVLW\ DUELWUDU\ XQLV

 QP
 QP

 QP
 QP
 QP





















 H9
Fig. 5.16. Low-temperature (T = 10 K) optical linear-absorption spectra of CdSe
nanocrystals [243]. Mean particle diameters <D> are shown in figure. The spectra
have been scaled for clarifying.
211

Nanocrystalline Materials

the energy of the maximum of the absorption band is inversely


proportional to the square of the radius, r 2 , of the CdSe particles.
The large width of the absorption bands (~0.15 eV or
1200 cm 1 ) is determined by the dispersion of the nanoparticle size,
i.e. deviation of the particle diameter from the mean value by
5 %. In fact, the broadened absorption spectra are observed even
when investigating the most monodispersed specimens. In other
words, the so-called inhomogeneous broadening is observed.
Therefore, in [243] the femtosecond photonecho technique is used
to probe the dynamics of quantum-confined excitons in
nanocrystalline CdSe. This made it possible to exclude the
inhomogeneous broadening and determine the homogeneous line
width accurately corresponding to the given particle size.
Consequently, it was shown that a decrease in the nanoparticle
diameter increases the width of absorption lines (Fig. 5.17,
curve 4).
The authors of [243] separated three contributions in the
homogeneous absorption line width. The most significant contribution
(Fig. 5.17, curve 1) to the line width is caused by the elastic
scattering of radiation from imputity and crystalline defect sites.
This contribution depends of the nanoparticle size (or more
accurately, on the effective surface area of scattering, proportional
7

 .










FP
















'!





QP

Fig. 5.17. Measured homogeneous line width, , of optical linear-absorption spectra


of CdSe nanocrystals as a function of diameter <D> at a temperature of 15 K [243].
The three most important contributing mechanisms to this line width are also displayed:
( ) elastic scattering contribution, ( ) phonon broadening due to coupling of lowfrequency vibrational modes, () lifetime broadening. The solid line shows the size
dependence of the total line width ( W ) and displays the sum of three contributions.
212

Properties of Isolated Nanoparticles and Nanocrystalline Powders

to the ratio S/V, where S is the surface area and V is the volume
of the particle) and does not depend on temperature. The second
contribution (Fig. 5.17, curve 2) is determined by the coupling to a
heat bath of low-frequency vibrational modes of the nanocrystal.
This contribution depends strongly on temperature and causes line
broadening, which increases linearly with increasing temperature.
The phonon broadening, due to low-frequency modes, provides a
significant contribution (up to 2035 %) to the homogeneous line
width not only at high but even at low temperatures. The third
contribution to the line width (Fig.5.17, curve 3) is the smallest. It
is associated with the lifetime to that corresponds the fast decay
of the initial state into some other electronic configuration which
is less strongly coupled to the grand state. The change in the state
strongly depends on the particle size as a result of exciton trapping
to a localized surface state. If the trapping is driven by a simple
overlap between an interior wave function and a localised surface
state, then the trapping rate should vary roughly as the ratio of the
surface area of the particle to its volume, i.e. S/V.
The recombination of light-generated charges leads to the
luminescence of nanoparticles. Examination of the luminescent
spectra of nanoparticles ZnO [247], ZnS [248, 249], CdS [250253],
and CdSe [254, 255] also shows a blue shift, i.e. the shift of the
spectra to the short wave range with a decrease in the particle size.
For the given nanoparticle size, the decay time of luminescence
depends on the wavelength and decreases with an increase in the
energy and a decrease of the wavelength of the emitted light
quantum. The dependence of the lifetime of the excited state on the
wavelength of luminescence is determined by contribution of the
Coulomb interaction between the electron and the hole to the energy
of the emitted light quantum h = 2 h c/ [256]:
2 h c/ = E min (D h D e ) + (e 2 / r eh ) ,

(5.48)

where E min is the minimum energy of excitation of luminescence of


the nanoparticle with a radius r, D h and D e are the depth of the
traps of the hole and the electron, r eh is the distance between the
electron and the hole. The electronhole pairs with small distance
r eh in tunneling recombination of the holes and electrons emit the
light with a faster rate and with a smaller wavelength than the
pairs with high r eh .

213

Nanocrystalline Materials

References
1.
2.

3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.

14.
15.

16.

17.

18.
19.

20.
21.

I. D. Morokhov, L. I. Trusov, S. P. Chizhik. Ultra-Dispersed Metallic Substances (Atomizdat, Moscow 1977) 264 pp. (in Russian)
I. D. Morokhov, V. I. Petinov, L. I. Trusov, V. F. Petrunin. Structure and
properties of small metallic particles. Uspekhi Fiz. Nauk 133, 653-692 (1981)
(in Russian)
I. D. Morokhov, L. I. Trusov, V. N. Lapovok. Physical Phenomena in UltraDispersed Substances (Energoatomizdat, Moscow 1984) 224 pp. (in Russian)
Yu. I. Petrov. Physics of Small Particles (Nauka, Moscow 1982) 360 pp.
(in Russian)
Yu. I. Petrov. Clusters and Small Particles (Nauka, Moscow 1986) 368 pp.
(in Russian)
M. Ya. Gen, Yu. I. Petrov. Dispersed condensates of metallic vapor. Uspekhi
Khimii. 38, 2249-2278 (1969) (in Russian)
V. N. Bogomolov. Liquids in ultra-fine channels. Uspekhi Fiz. Nauk 124,
171-182 (1978) (in Russian)
S. A. Nepijko. Physical Properties of Small Particles (Naukova Dumka, Kiev
1985) 248 pp. (in Russian)
C. F. Bohren, D. R. Huffman. Absorption and Scattering of Light by Small
Particles (Wiley-Interscience, New York 1998) 544 pp.
J. Perenboom, P. Wyder, F. Meier. Electronic properties of small metallic patricles. Phys. Rep. 78, 173-292 (1981)
W. Halperin. Quantum size effects in metal particles. Rev. Modern Phys.
58, 533-606 (1986)
E. L. Nagaev. Small metal particles. Uspekhi Fiz. Nauk 162, 49-124 (1992)
(in Russian)
N. T. Gladkikh, V. N. Khotkevich. Determination of solid surface energy
by melting temperature of disperse particles. Ukr. Fiz. Zh. 16, 1429-1436
(1971) (in Russian)
A. Hori. Properties and expected applications of ultrafine metal powders.
Chem. Econ. Eng. Rev. 7, 28-33 (1975)
Yu. G. Morozov, A. N. Kostygov, V. I. Petinov, P. E. Chizhov. Low-temperature
paramagnetism of fine gadolinium particles. Fiz. Nizkikh Temp. 1, 14071408 (1975) (in Russian)
Yu. G. Morozov, A. N. Kostygov, A. E. Petrov, P. E. Chizhov, V. I. Petinov.
Disappear of magnetic ordering in small terbium particles. Fiz. Tverd. Tela
18, 1394-1396 (1976) (in Russian)
S. Fujime. Electron diffraction at low temperature. IV. Amorphous films
of iron and chromium prepared by low temperature condensation. Japan.
J. Appl. Phys. 5, 1029-1035 (1966)
M. R. Hoare, P. Pal. Statistics and stability of small assemblies of atoms.
J. Cryst. Growth. 17, 77-96 (1972)
M. R. Hoare, P. Pal. Physical cluster mechanics. Statics and energy surfaces for monatomic systems. Adv. Phys. 20, 161-196 (1971); Physical cluster
mechanics. Statistical thermodynamics and nucleation theory for monatomic
systems. Adv. Phys. 24, 645-678 (1975)
M. R. Hoare, P. Pal. Geometry and stability of spherical fcc microcrystallites.
Nature - Phys. Sci. 236, 35-37 (1972)
M. R. Hoare, J. McInnes. Statistical mechanics and morphology of very
small atomic clusters. Faraday Discuss. Chem. Soc. 61, Precipitation, 12-

214

Properties of Isolated Nanoparticles and Nanocrystalline Powders

22.

23.
24.
25.

26.

27.
28.
29.
30.
31.

32.

33.
34.
35.

36.

37.
38.

39.

40.

41.

24 (1976)
S. H. Tolbert, A. P. Alivisatos. Size dependence of a first order solid-solid
transition: The wurtzite to rock salt transformation in CdSe nanocrystals.
Science 265, 373-376 (1994)
M. Haase, A. P.Alivisatos. Arrested solid-solid phase transition in 4-nmdiameter cadmium sulfide nanocrystals. J. Phys. Chem. 96, 6756-6762 (1992)
A. P. Alivisatos. Perspectives on the physical chemistry of semiconductor nanocrystals. J. Phys. Chem. 100, 13226-13229 (1996)
P. Pawlow. ber die Abhngigkeit des Schmelzpunktes von der
Oberflchenenergie eines festen Krpers. Z. physik. Chem. 65, 1-35, 545548 (1908); No 1; ber den Dampfdruck der Krner einer festen Substanz.
Z. physik. Chem. 68, 316-322 (1910)
K.-J. Hanszen. Theoretische Untersushungen ber den Schmelzpunkt kleiner
Kgelchen: Ein Beitrag zur Thermodynamik der Grenzflchen. Z. Physik
157, 523-553 (1960)
C. R. M. Wronski. The size dependence of the melting point of small patricles
of tin. Brit. J. Appl. Phys. 18, 1731-1737 (1967)
C. J. Coombes. The melting of small particles of lead and indium. J. Phys.
F: Metal. Phys. 2, 441-449 (1972)
T. L. Hill. Thermodynamics of Small Systems (Dover Publications, New York
1994) 416 pp.
B. M. Patterson, K. M. Unruh, S. I. Shah. Melting and freezing behavior
of ultrafine granular metal films. Nanostruct. Mater. 1, 65-70 (1992)
N. T. Gladkikh, R. Niedermayer, K. Spiegel. Nachweis groer Schmelzpunkt
serniedrigungen bei dnnen Metallschichten. Phys. Stat. Sol. 15, 181-192
(1966)
B. T. Boiko, A. T. Pugachev, V. M. Bratsykhin. On melting of condensated
indium films of undercritical thickness. Fiz. Tved. Tela 10, 3567-3570 (1968)
(in Russian)
M. Blackman, J. R. Sambles. Melting of very small particles during evaporation
at constant temperature. Nature 226, 938-939 (1970)
P. Buffat, J. Borel. Size effect on the melting temperature of gold particles. Phys. Rev. A 13, 2287-2298 (1976)
J. R. Sambles. An electron microscope study of evaporating gold particles:
The Kelvin equation for liquid gold and the lowering of the melting point
of solid gold particles. Proc. Roy. Soc. London A 324, 339-351 (1971)
V. P. Koverda, V. N. Skokov, V. P. Skripov. Influence of fluctuations and
unequilibrial facing on fusion of small metallic crystals. Fiz. Metall. Metalloved.
51, 1238-1244 (1981) (in Russian)
V. P. Skripov, V. P. Koverda, V. N. Skokov. Size effect on melting of small
particles. Phys. Stat. Sol. (a) 66, 109-118 (1981)
V. N. Skokov, V. P. Koverda, V. P. Skripov. Liquid-crystal phase transition in islet gallium films. Fiz. Tverd. Tela 24, 562-567 (1982) (in Russian)
V. P. Koverda, V. N. Skokov, V. P. Skripov. Crystallization of small particles in insular films of tin, lead and bismuth. Kristallografiya 27, 358362 (1982) (in Russian)
T. Castro, R. Reifenberger, E. Choi, R. P. Andres. Size-dependent melting temperature of individual nanometer-sized metallic clusters. Phys. Rev.
B 42, 8548-8556 (1990)
V. N. Bogomolov, A. I. Zadorozhnyi, A. A. Kapanadze, E. L. Lutsenko,

215

Nanocrystalline Materials

42.
43.

44.
45.
46.
47.
48.
49.
50.
51.
52.
53.
54.
55.
56.

57.
58.
59.
60.

61.

62.
63.

V. P. Petranovskii. Size effect on melting temperature of 9-Angstrom metallic particles. Fiz. Tverd. Tela 18, 3050-3053 (1976) (in Russian)
A. N. Goldstein, C. M. Echer, A. P. Alivisatos. Melting in semiconductor nanocrystals. Science 256, 1425-1427 (1992)
H. Y. Kai. Nanocrystalline materials: A study of their preparation and characterization. PhD Thesis. Netherlands, Amsterdam: Universiteit van Amsterdam,
1993. 113 pp.
F. Ercolessi, V. Andreoni, E. Tosatti. Melting of small gold particles: Mechanism
and size effects. Phys. Rev. Lett. 66, 911-914 (1991)
J. W. M. Frenken, J. F. van der Veen. Observation of surface melting. Phys.
Rev. Lett. 54, 134-137 (1985)
J. W. M. Frenken, P. M. Maree, J. F. van der Veen. Observation of surface-initiated melting. Phys. Rev. B 34, 7506-7516 (1986)
Da-Ming Zhu, J. G. Dash. Surface melting and roughening of adsorbed argon
films. Phys. Rev. Lett. 57, 2959-2962 (1986)
J. Krim, J. Coulomb, J. Bouzidi. Triple-point wetting and surface melting of oxygen films adsorbed on graphite. Phys. Rev. Lett. 58, 583-586 (1987)
E. G. McRae, R. A. Malic. A new phase transition at Ge(111) surface observed
by low-energy-electron diffraction. Phys. Rev. Lett. 58, 1437-1439 (1987)
DaMing Zhu, J. G. Dash. Surface melting of neon and argon films: Profile of the crystal-melt interface. Phys. Rev. Lett. 60, 432-435 (1988)
R. S. Berry, J. Jellinek, G. Natanson. Melting of clusters and melting. Phys.
Rev. A 30, 919-931 (1984)
R. S. Berry, D. J. Wales. Freezing, melting, spinodals, and clusters. Phys.
Rev. Lett. 63, 1156-1159 (1989)
S. Iijima, T. Ichihashi. Structural instability of ultrafine particles of metals. Phys. Rev. Lett. 56, 616-619 (1986)
J. Bovin, R. Wallenberg, D. Smith. Imaging of atomic clouds outside the
surface of gold crystals by electron microscopy. Nature 317, 47-49 (1985)
F. Vergand. Effect of grain size on the atomic distance in heavy rare earth
metals. Phil. Mag. 31, 537-550 (1975)
H. J. Wasserman, J. S. Vermaak. On the determination of a lattice contraction
in very small silver particles. Surface Sci. 22, 164-172 (1970); On the determination of the surface stress of copper and platinum. Surface Sci. 32,
168-174 (1972)
S. A. Nepijko, E. Pippel, J. Woltersdorf. Dependence of lattice parameter
on particle size. Phys. Stat. Sol. (a) 61, 469-475 (1980)
J. Woltersdorf, S. A. Nepijko, E. Pippel. Dependence of lattice parameters
of small particles on the size of the nuclei. Surface Sci. 106, 64-69 (1981)
G. Deutscher, M. Gershenson, E. Grnbaum, Y. Imry. Granular superconducting
films. J. Vac. Sci. Technol. 10, 697-701 (1973)
A. Yokozeki, G. D. Stein. A metal cluster generator for gas-phase electron
diffraction and its application to bismuth, lead, and indium: Variation in
microcrystal structure with size. J. Appl. Phys. 49, 2224-2232 (1978)
J. Harada, S. Yao, A. Ichimiya. X-ray diffraction stidy of fine gold particles prepared by gas evaporation technique. I. General features. J. Phys.
Soc. Japan. 48, 1625-1630 (1980)
J. Harada, K. Ohshima. X-ray diffraction stidy of fine gold particles prepared
by gas evaporation technique. Surface Sci. 106, 51-57 (1981)
P. A. Montano, G. K. Shenoy, E. E. Alp, W. Schulze, J. Urban. Structure
of copper microclusters isolated in solid argon. Phys. Rev. Lett. 56, 2076-

216

Properties of Isolated Nanoparticles and Nanocrystalline Powders

64.

65.

66.

67.

68.

69.

70.

71.
72.
73.
74.

75.
76.

77.
78.
79.

80.
81.

82.

2079 (1986)
C. Solliard, P. Buffat. Crystal size variation of small gold crystals due to
size effect. J. de Physique - Colloque C2 38, Suppl. No 7, C2-167 - C2170 (1977)
M. Ya. Gamarnik, Yu. Yu. Sidorin. Change of the unit cell parameters in
highly dispersed gold, silver, and copper powders. Phys. Stat. Sol. (b) 156,
K1-K4 (1989)
M. Ya. Gamarnik, Yu. Yu. Sidorin. Change of the unit cell parameters in
the fine disperse platinum powders. Poverkhnost No 4, 124-129 (1990)
(in Russian)
V. F. Petrunin, V. A. Pogonin, L. I. Trusov, A. S. Ivanov, V. N. Troitskii.
Structure of ultrafine grained particles of titanium nitride. Izv. AN SSSR.
Neorgan. Materialy 17, 59-63 (1981) (in Russian)
V. F. Petrunin, Yu. G. Andreev, T. N. Miller, Ya. P. Grabis. Neutron diffraction analysis of ultrafine grained powders of zirconium nitride. Poroshkovaya
Metallurgiya No 9, 90-97 (1987) (in Russian)
V. F. Petrunin, Yu. G. Andreev, V. N. Troitskii, O. M. Grebtsova. Neutron diffraction investigation of niobium nitrides in ultrafine grained state.
Poverkhnost No 11, 143-148 (1982) (in Russian)
V. F. Petrunin, Yu. G. Andreev, T. N. Miller, Ya. P. Grabis, A. G. Ermolaev,
F. M. Zelenuk. Structure distortions of ultrafine grained powders of titanium nitride. Poroshkovaya Metallurgiya No 8, 12-15 (1984)
S. I. Alyamovski, Yu. G. Zainulin, G. P. Shveikin. Oxycarbides and Oxyntrides
of Group IV and V Metals (Nauka, Moscow 1981) 144 pp. (in Russian)
A. I. Gusev. Physical Chemistry of Nonstoichiometric Refractory Compounds
(Nauka, Moscow 1991) 286 pp. (in Russian)
A. I. Gusev, A. A. Rempel. Structural Phase Transitions in Nonstoichiometric
Compounds (Nauka, Moscow 1988) 308 pp. (in Russian)
E. Moroz, S. Bogdanov, V. Ushakov. X-ray characteristics of iridium and
platinum blacks of various dispersity. React. Kinet. Catal. Lett. 9, 109112 (1978)
M. Ya. Gamarnik. Size effect in quartz. Doklady AN Ukr. SSR (seriya B)
No 4, 6-8 (1982) (in Russian)
S. Veprek, Z. Iqbal, H. R. Oswald, F.-A. Sarott, J. J. Wagner, A. P. Webb.
Lattice dilatation of small silicon crystallites - implications for amorphous
silicon. Solid State Commun. 39, 509-512 (1981)
M. Ya. Gamarnik. Size effect in CeO 2. Fiz. Tverd. Tela 30, 1399-1404 (1988)
(in Russian)
A. Cimino, P. Porta, M. Valigi. Dependence of the lattice parameter of
magnesium oxide on crystallite size. J. Amer. Ceram. Soc. 49, 152-156 (1966)
G. Apai, J. F. Hamilton, J. Stohr, A. Thompson. Extended X-ray-absorption
fine structure of small Cu and Ni clusters: Binding-energy and bond-length
changes with cluster size. Phys. Rev. Lett. 43, 165-169 (1979)
P. A. Montano, G. K. Shenoy. EXAFS study of iron monomers and dimers
isolated in solid argon. Solid State Commun. 35, 53-56 (1980)
N. S. Lidorenko, S. P. Chizhik, N. T. Gladkikh, L. K. Grigorieva, R. N.
Kuklin, R. G. Melkadze. Shift of electron chemical potentital in high-disperse
systems. Izv. AN SSSR. Metally No 6, 91-95 (1981) (in Russian)
N. S. Lidorenko, S. P. Chizhik, N. T. Gladkikh, L. K. Grigorieva, R. N.
Kuklin. On an investigation of the nature of size vacancy effect. Doklady
AN SSSR 258, 858-861 (1981) (in Russian)

217

Nanocrystalline Materials
83.

84.

85.

86.

87.

88.

89.

90.
91.

92.

93.

94.

95.
96.
97.
98.
99.
100.
101.

S. P. Chizhik, N. S. Lidorenko, N. T. Gladkikh, L. K. Grigorieva, R. N.


Kuklin. Shift of phase equilibria in high-disperse systems. Izv. AN SSSR.
Metally No 1, 80-81 (1982) (in Russian)
N. S. Lidorenko, S. P. Chizhik, N. T. Gladkikh, L. K. Grigorieva, R. N.
Kuklin. Size vacancy effect in the theory of homogeneous nucleation. Izv.
AN SSSR. Metally No 1, 82-86 (1982) (in Russian)
S. P. Chizhik, N. T. Gladkikh, L. K. Grigorieva, R. N. Kuklin, R. G. Melkadze.
Sze shifts in phonon spectrum. Izv. AN SSSR. Metally No 3, 165-167 (1982)
(in Russian)
V. I. Gorchakov, E. L. Nagaev, S. P. Chizhik. Does the Laplace pressure
compress solid bodies? Fiz. Tverd. Tela 30, 1068-1075 (1988) (in Russian)
V. I. Gorchakov, L. K. Grigor eva, E. L. Nagaev, S. P. Chizhik. Oscillating relaxation of the surface with a great penetration depth. Zh. Eksper.
Teor. Fiz. 93, 2090-2101 (1987) (in Russian)
V. I. Gorchakov, E. L. Nagaev. Oscillating relaxation of the surface with
large penetration depth and the effect of the surface state. Poverkhnost
No 11, 28-35 (1988) (in Russian)
J. Lffler, J. Weissmller, H. Gleiter. Characterization of nanocrystalline
palladium by x-ray atomic density distribution functions. Nanostruct. Mater.
6, 567-570 (1995)
W. Qin, Z. H. Chen, P. Y. Huang, Y. H. Zhuang. Crystal lattice expansion
of nanocrystalline materials. J. Alloys Comp. 292, 230-232 (1999)
H. Kronmller, M. Fhnle, M. Domann, H. Grimm, R. Grimm, B. Grger.
Magnetic properties of amorphous ferromagnetic alloys. J. Magn. Magn.
Mater. 13, 53-70 (1979)
H.-E. Schaefer. Interfaces and physical properties of nanostructurd solids.
In: Mechanical Properties and Deformation Behavior of Materials Having
Ultrafine Microstructure. Ed. M. A. Nastasi, D. M. Parkin, H. Gleiter. (Kluwer
Academic Press, Netherlands, Dordrecht 1993) pp.81-106
K. Lu, R. Lck, B. Predel. Variation of the interfacial energy with grain
size in nanocrystalline materials. Mater. Sci. Engineer. A 179-180, 536-540
(1994)
B. A. Nesterenko, B. I. Gorbachev, V. A. Zrazhevskii, P. G. Ivanitskii, V.
T. Krotenko, M. V. Pasechnik, O. V. Snitko. Phonon spectra of silicon lattice.
Fiz. Tverd. Tela 16, 3513-3515 (1974) (in Russian)
K. H. Rieder, E. M. Hrl. Search for surface modes of lattice vibrations
in magnesium oxide. Phys. Rev. Lett. 20, 209-211 (1968)
K. H. Rieder. Vibrational surface thermodynamic functions magnesium oxide.
Surface Sci. 26, 637-648 (1971)
K. H. Rieder, W. Drexler. Observation of vibrational surface modes in the
acousto-optical bulk gap of TiN. Phys. Rev. Lett. 34, 148-151 (1975)
R. H. Bolt. Frequency distribution of eigentones in a three-dimensional
continuum. J. Acoust. Soc. Amer. 10, 228-234 (1939)
D.-Y. Maa. Distribution of eigentones in a rectangular chamber at low frequency
range. J. Acoust. Soc. Amer. 10, 235-238 (1939)
E. W. Montrol. Size effect in low temperature heat capacities. J. Chem.
Phys. 18, 183-185 (1950)
A. A. Maradudin, R. F. Wallis. Lattice-dynamical calculation of the surface specific heat of a crystal at low temperatures. Phys. Rev. 148, 945961 (1966)

218

Properties of Isolated Nanoparticles and Nanocrystalline Powders


102.

103.
104.
105.
106.
107.
108.
109.
110.
111.
112.
113.
114.

115.
116.

117.

118.
119.

120.
121.
122.

123.

Ya. A. Iosilevskii. On the dynamics of surface atoms. Phys. Stat. Sol. (b)
46, 125-135 (1971); The Debye frequency tensor of single crystals. Phys.
Stat. Sol. (b) 53, 405-418 (1972); The surface density of the acoustical
phonon states. Phys. Stat. Sol. (b) 60, 39-50 (1973)
M. G. Burt. A method for calculating the low temperature surface specific
heat of a crystal lattice. J. Phys. C: Solid State Phys. 6, 855-867 (1973)
H. P. Baltes. Phonon in small particles. J. de Physique - Colloque C2 38,
Suppl. No 7, C2-151-C2-156 (1977)
H. Baltes, B. Steinle, M. Pabst. Poincar cycles and coherence of bounded
thermal radiation fields. Phys. Rev. A 13, 1866-1873 (1976)
H. P. Baltes, E. R. Hilf. Specific heat of lead grains. Solid State Commun.
12, 369-373 (1973)
Th. F. Nonnenmacher. Quantum size effect on the specific heat of small
particles. Phys. Lett. A 51, 213-214 (1975)
R. Lautenschlger. Improved theory of the vibrational specific heat of lead
grains. Solid State Commun. 16, 1331-1334 (1975)
R. Kubo. Electronic properties of metallic fine particles. J. Phys. Soc. Japan.
17, 975-986 (1962)
R. Kubo. Discreteness of energy levels in small metallic particles. J. de
Physique - Colloque C2 38, Suppl. No 7, C2-69 - C2-75 (1977)
H. T. Chu. Two-dimension motion of charge of carriers in ultrafine films.
J. Phys. Chem. Solids 49, 1191-1196 (1988)
P. R. Couchman, F. E. Karasz. The effect of particle size on Debye temperature. Phys. Lett. A 62, 59-61 (1977)
G. Goll, H. Lhneyen. Specific heat of nanocrystalline and colloidal noble metals at low temperatures. Nanostruct. Mater. 6, 559-562 (1995)
V. Novotny, P. P. M. Meincke, J. H. P. Watson. Effect of size and surface on the specific heat of small lead particles. Phys. Rev. Lett. 28, 901903 (1972)
V. Novotny, P. P. M. Meincke. Thermodynamic lattice and electronic properties
of small particles. Phys. Rev. B 8, 4186-4199 (1973)
G. H. Comsa, D. Heitkamp, H. S. Rde. Specific heat of ultrafine vanadium particles in the temperature range 1.3-10 K. Solid State Commun. 20,
877-880 (1976); Effect of size on the vibrational specific heat of ultrafine
palladium particles. Solid State Commun. 24, 547-550 (1977)
Y. Y. Chen, Y. D. Yao, S. U. Jen, B. T. Lin, H. M. Lin, C. Y. Tung, S. S.
Hsiao. Magnetic susceptibility and low temperature specific heat of palladium nanocrystals. Nanostruct. Mater. 6, 605-608 (1995)
Y. Y. Chen, Y. D. Yao, B. T. Lin, C. T. Suo, S. G. Shyu, H. M. Lin. Specific heat of fine copper particles. Nanostruct. Mater. 6, 597-600 (1995)
Y. D. Yao, Y. Y. Chen, C. M. Hsu, H. M. Lin, C. Y. Tung, M. F. Tai, D.
H. Wang, K. T. Wu, C. T. Suo. Thermal and magnetic studies of nanocrystalline Ni. Nanostruct. Mater. 6, 933-936 (1995)
J. Trampenau, K. Bauszus, W. Petry, U. Herr. Vibrational behaviour of
nanocrystalline Ni. Nanostruct. Mater. 6, 551-554 (1995)
T. Fujita, K. Ohshima, T. Kuroishi. Temperature dependence of electrical
conductivity in films of fine particles. J. Phys. Soc. Japan. 40, 90-92 (1976)
K. Ohshima, T. Fujita, T. Kuroishi. The phonon softening in metallic fine
particles. J. de Physique - Colloque C2 38, Suppl. No 7, C2-163 - C2165 (1977)
Y. Kashiwase, I. Nishida, I. Kainuma, K. Kimoto. X-ray diffraction study

219

Nanocrystalline Materials

124.
125.

126.
127.

128.
129.
130.
131.

132.
133.

134.

135.

136.

137.

138.

139.
140.
141.

142.

on lattice vibration of fine particles. J. de Physique - Colloque C2 38, Suppl.


No 7, C2-157 - C2-160 (1977)
P. Buffat. Size effect modifications of the Debye-Waller factor in small gold
particles. Solid State Commun. 23, 547-550 (1977)
Yu. I. Petrov. Influence of high-temperature thermal defects in metals on
the intensity of scattering X-ray radiation. Fiz. Metall. Metalloved. 19,
667-674 (1965) (in Russian)
M. P. A. Viegers, J. M. Trooster. Mssbauer spectroscopy of small gold
particles. Phys. Rev. B 15, 72-83 (1977)
G. von Eynatten, H. E. Bmmel. Size and temperature dependence of the
Mssbauer Debye-Waller factor of iron microcrystals. Appl. Physics 14,
415-421 (1977)
Yu. I. Petrov, V. A. Kotelnikov. On the Debye-Waller factor and lattice
parameters for small metal particles. Fiz. Tverd. Tela 13, 313-315 (1971)
L. P. Gorkov, G. M. Eliashberg. Small metallic particles in an electromagnetic
field. ZhETF 48, 1407-1418 (1965) (in Russian)
A. Kawabata. Electronic properties of fine metallic particles. III. ESR absorption line shape. J. Phys. Soc. Japan. 29, 902-911 (1970)
R. Monot, C. Narbel, J. P. Borel. Electron spin resonance in small particles of silver. Quantum size effect. Nuovo Cimento Soc. Ital. Fis. B 16,
253-260 (1974)
R. F. Marzke, W. S. Glaunsinger, M. Bayard. Magnetic susceptibility of
uniform microcrystals of platinum. Solid State Commun. 18, 1025-1030 (1976)
J. Borel, J. Millet. Conduction electron spin resonance in small particles
suspended in solid matrix. J. de Physique - Colloque C2 38, Suppl. No
7, C2-115 - C2-119 (1977)
A. I. Zadorozhnyi, L. K. Panina, V. F. Sakash, L. P. Strakhov, S. V.
Kholodkevich. Change of selenium and tellurium magnetic properties in fine
dispersion state. Fiz. Tverd. Tela 22, 2934-2938 (1980) (in Russian)
V. N. Bogomolov, A. I. Zadorozhney, L. K. Panina. Low-temperature phase
transition of mercuty microclusters in strong magnetic fields. Physica B+C
107, 89-90 (1981)
V. N. Bogomolov, A. I. Zadorozhney, L. K. Panina, V. I. Petranovskii. Transition
of 13-atom mercury clusters into a highly paramagnetic state in the presence
of magnetic field. Pisma v ZhETF 31, 371-374 (1980) (in Russian)
N. R. Harrison, P. P. Edwards, J. Klinowski, J. M. Thomas. Ionic and metallic
clusters of the alkali metals in zeolite Y. J. Solid State Chem. 54, 330-341
(1984)
K. W. Blazey, K. A. Mller, F. Blatter, E. Schumacher. Conduction electron spin resonance of cesium metallic clusters in zeolite X. Europhys. Lett.
4, 857-861 (1987)
F. Blatter, K. W. Blazey. Conduction electron spin resonance of silver clusters
in zeolite Ag Y. Z. Physik D 18, 427-429 (1991)
K. Kimura, S. Bandow. Paramagnetic enhancement in the magnetic susceptibility
of ultrafine magnesium particles. Phys. Rev. Lett. 58, 1359-1362 (1987)
W. Eberhardt, P. Fayet, D. M. Cox, Z. Fu, A. Kaldor, R. Sherwood, D.
Sondericker. Photoemission from mass-selected monodispersed Pt clusters.
Phys. Rev. Lett. 64, 780-783 (1990)
P. N. First, J. A. Stroscio, R. A. Dragose, D. T. Pierce, R. J. Celotta. Metallicity
and gap states in tunneling to Fe clusters in GaAs(110). Phys. Rev. Lett.
63, 1416-1419 (1989)

220

Properties of Isolated Nanoparticles and Nanocrystalline Powders


143.

144.
145.
146.
147.

148.
149.

150.
151.

152.

153.
154.

155.

156.
157.

158.

159.

160.

161.

162.

R. Rademann, B. Kaiser, U. Even, F. Hensel. Size dependence of the gradual


transition to metallic properties in isolated mercury clusters. Phys. Rev.
Lett. 59, 2319-2321 (1987)
C. J. Kriesmann, H. B. Callen. The magnetic susceptibility of the transition elements. Phys. Rev. 94, 837-844 (1954)
J. E. van Dam, O. K. Andersen. Temperature dependence of the palladium
susceptibility. Solid State Commun. 14, 645-647 (1974)
A. Kawabata, R. Kubo. Electronic properties of fine metallic particles. II.
Plasma resonance absorption. J. Phys. Soc. Japan. 21, 1765-1772 (1966)
M. A. Smithard. Size effect on the conduction electron spin resonance of
small sodium particles in sodium azide NaN 3 . Solid State Commun. 14, 411415 (1974)
D. Gordon. Conduction-electron spin resonance in small particles of sodium. Phys. Rev. B 13, 3738-3747 (1976)
M. Rappaz, A. Chatelain, L.A. Boatner. EPR investigation of size effects
in the crystal field of small dielectric particles. J. de Physique - Colloque
C2 38, Suppl. No 7, C2-105 - C2-108 (1977)
S. V. Vonsovskii. Magnetism (Nauka, Moscow 1971) 1032 pp. (in Russian)
E. I. Kondorskii. Micromagnetism and remagnetization of quasi-single-domain
particles. Izv. AN SSSR. Seriya Fizicheskaya 42, 1638-1645 (1978) (in
Russian)
A. Tasaki, S. Tomiyama, S. Iida, N. Wada, R. Uyeda. Magnetic properties of ferromagnetic metal fine particles prepared by evaporation in argon gas. Japan. J. Appl. Phys. 4, 707-711 (1965)
Yu. I. Petrov, Yu. I. Fedorov. Electromagnetic properties of colloidal suspension
of nickel in paraffin. Zh. Tekhnich. Fiz. 37, 726-728 (1967) (in Russian)
A. E. Ermakov, O. A. Ivanov, Ya. S. Shur, R. M. Grechishkin, G. M. Ivanova.
Magnetic properties of monocrystalline nickel powders. Fiz. Metall. Metalloved.
33, 558-563 (1972) (in Russian)
I. N. Shabanova, A. E. Ermakov, V. A. Trapeznikov, Ya. S. Shur. Dependence
of saturation magnetization of nickel aerosols on particle surface state studied
by electron spectroscopy method. Fiz. Metall. Metalloved. 38, 314-322
(1974) (in Russian)
W. D. Corner, P. A. Mundell. Properties of ferromagnetic micropowders.
J. Magn. Magn. Mater. 20, 148-157 (1980)
A. E. Petrov, V. I. Petinov, I. V. Plate, E. A. Fedorova, M. Ya. Gen. Magnetic
properties of small aerosole particles of cobalt. Fiz. Tverd. Tela 13, 15731577 (1971) (in Russian)
A. E. Petrov, V. I. Petinov, V. V. Shevchenko. Magnetic properties of small
aerosole Ni particles at 4.2-300 K. Fiz. Tverd. Tela 14, 3031-3036 (1972)
(in Russian)
A. E. Petrov, A. N. Kostygov, V. I. Petinov. Magnetic properties of small
spherical particles of iron at a temperature 4.2-300 K. Fiz. Tverd. Tela
15, 2927-2931 (1973) (in Russian)
R. Z. Valiev, R. R. Mulyukov, Kh. Ya. Mulyukov, L. I. Trusov, V. I. Novikov.
Curie temperature and saruration magnetization of nickel with submicrongrained structure. Pisma v ZhTF 15, 78-81 (1989) (in Russian)
A. V. Korolev, A. I. Deryagin, V. A. Zavalishin, R. I. Kuznetsov. Peculiarities of magnetic state of highly deformed polycrystalline ultrafine-grained
nickel. Fiz. Metall. Metalloved. 68, 672-678 (1989) (in Russian)
R. Z. Valiev, Ya. D. Vishnyakov, R. R. Mulyukov, G. S. Fainstein. On the

221

Nanocrystalline Materials

163.
164.

165.

166.

167.

168.
169.

170.
171.

172.

173.

174.
175.

176.

177.
178.
179.

180.

decrease of Curie temperature in submicron-grained nickel. Phys. Stat. Sol.


(a) 117, 549-553 (1990)
L. Daroczi, D. L. Beke, G. Posgay, M. Kis-Varga. Magnetic properties of
ball milled nanocrystalline Ni and Fe. Nanostruct. Mater. 6, 981-984 (1995)
S. Szabo, I. Brovko, M. Kis-Varga, D. L. Beke, G. Posgay. Mssbauereffect and magnetic properties of nanocrystalline Fe and Fe(Si) alloys.
Nanostruct. Mater. 6, 973-976 (1995)
Hong-Ming Lin, C. M. Hsu, Y. D. Yao, Y. Y. Chen, T. T. Kuan, F. A. Yang,
C. Y. Tung. Magnetic study of both nitrided and oxidized Co particles.
Nanostruct. Mater. 6, 977-980 (1995)
S. Gangopadhyay, G. C. Hadjipanayis, O. Dale, C. M. Sorensen, K. J.
Klabunde, V. Papaefthymiou, A. Kostikas. Magnetic properties of ultrafine
iron particles. Phys. Rev. B 45, 9778-9787 (1992)
S. Gangopadhyay, G. S. Hadjipanayis, S. I. Shah, C. M. Sorensen, K. J.
Klabunde, V. Papaefthymiou, A. Kostikas. Effect of oxide layer on the hysteresis
behavior of fine Fe particles. J. Appl. Phys. 70, 5888-5890 (1991)
S. Gangopadhyay, G. C. Hadjipanayis, B. Dale, C. M. Sorensen, K. J. Klabunde.
Magnetism of ultrafine particles. Nanostruct. Mater. 1, 77-82 (1992)
S. Matsuo, I. Nishida. Magnetic properties of a new magnetic phase in fine
bcc chromium particles prepared by gas evaporation method. J. Phys. Soc.
Japan. 49, 1005-1012 (1980)
G. J. Muench, S. Arais, E. Matijevic E. Magnetic properties of monodispersed
submicronic hematite -Fe 2O 3 particles. J. Appl. Phys. 52, 2493-2495 (1981)
W. Hellenthal. Theorie der spontanen Magnetisierung kleiner ferromagnetischer
Teilchen. Z. Physik 170, 303-319 (1962); ber die spontane Magnetisierung
kleiner Teilchen, berechnet mittels einer erweiferten Molekularfeldtheorie.
Z. angew. Physik 14, 194-195 (1962)
K. Binder, H. Rauch, V. Wildpaner. Monte-Carlo calculation of the magnetization of superparamagnetic particles. J. Phys. Chem. Solids 31, 391397 (1970)
A. Knappwost, E. Burkard. Untersuchungen zur Teilchengrenfunktion der
Curietemperatur an kollektiv paramagnetischem Katalysatornickel. Ber.
Bunsenges. phys. Chem. 68, 163-169 (1964)
E. L. Lee, P. E. Bolduc, C. E. Violet. Magnetic ordering and critical thickness
of ultrathin iron films. Phys. Rev. Lett. 13, 800-802 (1964)
F. E. Luborski, P.E. Lawrence. Saturation magnetization and size of iron
particles less than 100 in diameter. J. Appl. Phys. - Suppl. 32, 231S232S (1961)
A. Knappwost, A. Illenberger. Temperaturunabhngige spontane Magnetisierung
von Kohrenzbereichen des Kobalts im Wirtsgitter des Kupfers.
Naturwissenschaften 45, 238-238 (1958)
E. Sffge, W. von Hrsten. a.c.-Susceptibility measurements in small fields
on the fine superparamagnetic nickel particles. Z. Physik B 42, 47-55 (1981)
C. P. Bean, J. D. Livingston. Superparamagnetism. J. Appl. Phys. - Suppl.
30, 120S-129S (1959)
A. Hahn. Untersuchungen an kleien superparamagnetischen Nickel krnern
zur Frage der Temperaturabhngigkeit der spontanen Magnetisierung. Ann.
Physik 11, 277-309 (1963)
R. Mller. Einflu der Teilchengrenverteilung aud die Magnetisierungskurve
eines superparamagnetischen Teilchenkollektivs. Z. angew. Physik 30, 5660 (1970)

222

Properties of Isolated Nanoparticles and Nanocrystalline Powders


181.

182.
183.
184.

185.

186.
187.
188.

189.

190.

191.

192.
193.
194.

195.

196.

197.
198.

A. Knappwost, A. Illenberger, L. J. Nunez. Temperature function of the


spontaneous magnetization of collectively paramagnetic spherical and laminar
ranges in the angstrom range. Z. phys. Chem. (Frankfurt, BRD) 23, 145163 (1960)
W. von Henning, E. Vogt. Zur magnetischen Korngrenbestimmung von
hochdispersem Eisen und Kobalt. Ztschr. Naturforsch. A 12, 754-755 (1957)
A. E. Berkowitz, P. J. Flanders. Precipitation in a -brass-iron alloys. J.
Appl. Phys. - Suppl. 30, 111S-112S (1959)
J. P. Chen, C. M. Sorensen, K. J. Klabunde, G. C. Hadjipanayis. Magnetic properties of nanophase cobalt particles synthesized in inversed micelles.
J. Appl. Phys. 76, 6316-6318 (1994)
C. Y. Yang, K. H. Johnson, D. R. Salahub, J. Kaspar, R. P. Messmer. Iron
clusters: Electronic structure and magnetism. Phys. Rev. B 24, 5673-5692
(1981)
Zhi-qiang Li and Bing-lin Gu. Electronic-structure calculations of cobalt
clusters. Phys. Rev. B 47, 13611-13614 (1993)
J. P. Bucher, D. C. Douglass, L. A. Bloomfield. Magnetic properties of
free cobalt clusters. Phys. Rev. Lett. 66, 3052-3055 (1991)
D. Vollath, D. V. Szabo, R. D. Taylor, J. O. Willis, K. E. Sickafus. Synthesis and properties of nanocrystalline superparamagnetic -Fe2O 3. Nanostruct.
Mater. 6, 941-944 (1995)
E. Tronc, P. Prene, J. P. Jolivet, D. Fiorani, A. M. Testa, R. Cherkaoui,
M. Nogus, J. L. Dormann. Magnetic dynamics of -Fe 2 O 3 nanoparticles.
Nanostruct. Mater. 6, 945-948 (1995)
A. A. Rempel, A. I. Gusev, S.Z. Nazarova, R. R. Mulyukov. Imputity
superparamagnetism in plastically deformed copper. Doklady Akad. Nauk
347, 750-754 (1996) (in Russian). (Engl. transl.: Physics - Doklady 41,
152-156 (1996))
M. Garber, W. G. Henry, H. G. Hoeve. A magnetic susceptibility balance
and the temperature dependence of the magnetic susceptibility of copper,
silver, and gold, 295-975 K. Canadian J. Phys. 38, 1595-1613 (1960)
A. A. Rempel, A. I. Gusev. Magnetic susceptibility of palladium subjected
to severe plastic deformation. Phys. Stat. Sol. (b) 196, 251-260 (1996)
G. Tammann, W. Oelsen. Die Abhngigkeit der Konzentration gesttigter
Mischkristalle von der Temperatur. Z. anorg. Chemie 186, 257-288 (1930)
F. Bitter, A. R. Kaufmann, C. Starr, S. T. Pan. Magnetic studies of solid
solutions II. The properties of quenched copper-iron alloys. Phys. Rev.
60, 134-138 (1941)
A. A. Rempel, S. Z. Nazarova. Magnetic properties of iron nanoparticles
in submicrocrystalline copper. In: Advances in Nanocrystallization. Proceedings
of the Euroconference on Nanocrystallization and Workshop on Bulk Metallic
Glasses (Grenoble, France, April 21-24, 1998). Ed. A. R. Yavari. Materials Science Forum 307, 217-222 (1999). (Trans Tech Publications, Switzerland
1999) / J. Metastable Nanocryst. Mater. 1, 217-222 (1999)
A. A. Rempel, S. Z. Nazarova, A. I. Gusev. Intrinsic and extrinsic defects
in palladium and copper after severe plastic deformation. In: Structure and
properties of nanocrystalline materials (Ural Division of the Russ. Acad.
Sci., Yekaterinburg, 1999) pp.265-278
A. A. Rempel, S. Z. Nazarova, A. I. Gusev. Iron nanopatricles in severeplastic-deformed copper. J. Nanoparticle Researh 1,485-490 (1999)
C. Kittel. Introduction to Solid State Physics. Seventh edition. (Wiley, New

223

Nanocrystalline Materials

199.
200.

201.
202.
203.
204.
205.

206.

207.
208.
209.
210.

211.

212.

213.

214.
215.

216.
217.
218.
219.

York 1996) 673 pp.


A. E. Hughes, S. C. Jain. Metal colloids in ionic crystals. Adv. Phys. 28,
717-828 (1979)
C. Granqvist, O. Hunderi. Optical properties of ultrafine Au particles prepared
by gas evaporation. J. de Physique - Colloque C2 38, Suppl. No 7, C2143 C2-146
C. G. Granqvist, O. Hunderi. Optical properties of ultrafine gold particles. Phys. Rev. B 16, 3513-3534 (1977)
H. Abe, K.-P. Charle, B. Tesche, W. Schulze. Surface plasmon absorption
of various colloidal metal particles. Chem. Phys. 68, 137-141 (1982)
J. P. Marton, M. Schlesinger. Optical constants of thin discontinuous nickel
films. J. Appl. Phys. 40, 4529-4533 (1969)
G. Rasigni, J. P. Palmari, M. Rasigni. Plasma resonance in granular deposits
and rough surfaces of magnesium. Phys. Rev. B 12, 1121-1131 (1975)
M. Rasigni, G. Rasigni, J. P. Palmari. Interband absorption, intraband absotption,
and collective oscillations in very thin lithium deposits studied under static
ultrahigh vacuum. Phil. Mag. 31, 1307-132575)
G. Rasigni, J. P. Petrakian, M. Rasigni, J. P. Palmari. Plasma resonance
and interband absorption in granular deposits of indium. J. Phys. C: Solid
State Phys. 9, L325-L328 (1976)
M. Rasigni, G. Rasigni. Optical properties of aggregated lithium deposits.
J. Opt. Soc. Amer. 67, 510-519 (1977)
L. Genzel, T. P. Martin, U. Kreibig. Dielectric function and plasma resonances
of small metal particles. Z. Physik B 21, 339-346 (1975)
P. Ruppin, H. Yatom. Size and shape effects on the broadening of the plasma
resonance absorption in metals. Phys. Stat. Sol. (b) 74, 647-654 (1976)
U. Kreibig. Electronic properties of small silver particles: The optical constants
and their temperature dependence. J. Phys. F: Metal Physics 4, 999-1014
(1974)
U. Kreibig. Anomalous frequency and temperature dependence of the optical absorption of small gold particles. J. de Physique - Colloque C2 38,
Suppl. No 7, C2-97 C2-103 (1977)
W. Schulze, H. Abe. Optical and vibrational studies of sikver molecules
and microcrystallites prepared by matrix aggregation and gas aggregation
technique. Faraday Symp. Chem. Soc. 14, 87-93 (1980)
H. Abe, W. Schulze, B. Tesche. Optical properties of silver microcrystals
prepared by means of the gas evaporation technique. Chem. Phys. 47, 95104 (1980)
J. C. Slater. Insulators, Semiconductors and Metals (McGraw-Hill Book Co.,
New York 1967) 720 pp.
G. Mie. Die optischen Eigenschaften kolloidaler Goldlsungen. Physikal.
Z. 8, 769-773 (1907); Beitrage zur Optik trber Medien, speziele kolloidaler
Metall-lsungen. Ann. Physik 25, 377-385 (1908)
D. M. Wood, N. W. Ashcroft. Quantum size effects in the optical properties of small metallic particles. Phys. Rev. B 25, 6255-6274 (1982)
A. A. Lushnikov, A. J. Simonov. Surface plasmons in small metal particles. Z. Physik 270, 17-24 (1974)
M. A. Smithard, R. Dupree. The preparation and optical properties of small
silver particles in glass. Phys. Stat. Sol. (a) 11, 695-703 (1972)
M. A. Smithard. Size effect on the optical and paramagnetic absorption
of silver particles in glass matrix. Solid State Commun. 13, 153-156 (1973)

224

Properties of Isolated Nanoparticles and Nanocrystalline Powders


220.

221.
222.
223.
224.
225.
226.
227.
228.

229.

230.
231.

232.

233.

234.
235.

236.
237.

238.

239.

240.

J. D. Ganiere, R. Rechsteiner, M. A. Smithard. On the size dependence


of the optical absorption due to small metal particles. Solid State Commun.
16, 113-115 (1975)
R. H. Doremus. Optical properties of small gold particles. J. Chem. Phys.
40, 2389-2396 (1964)
R. H. Doremus. Optical properties of small silver particles. J. Chem. Phys.
40, 414-417 (1965)
U. Kreibig, C. von Fregstein. The limitation of electron mean free path in
small silver particles. Z. Physik 224, 307-323 (1969)
E. Anno, R. Hoshino. Size effect on the width of plasma resonance absorption of silver island films. J. Phys. Soc. Japan. 51, 1185-1192 (1982)
P. Ascarelli, M. Cini. Red shift on the surface plasmon resonance absorption by fine metal particles. Solid State Commun. 18, 385-388 (1976)
M. Cini. Classical and quantum aspects of size effects. J. Opt. Soc. Amer.
71, 386-392 (1981)
Al. A. Efros, A. A. Efros. Light absorption in a simiconductor sphere. Fiz.
Tekhn. Poluprovodn. 16, 1209-1214 (1982) (in Russian)
A. I. Ekimov, A. A. Onushchenko. Quantum size effect in optical spectra of semiconductor microcrystals. Fiz. Tekhn. Poluprovodn. 16, 12151219 (1982) (in Russian)
L. E. Brus. Electron-electron and electron-hole interactions in small semiconductor crystallites: The size dependence of the lowest excited electronic
state. J. Chem. Phys. 80, 4403-4409 (1984)
A. I. Ekimov, Al. L. Efros, A. A. Onushchenko. Quantum size effect in
semiconductor microcrystals. Solid State Commun. 56, 921-924 (1985)
T. Itoh, Y. Iwabuchi, M. Kataoka. Study of the size and shape of CuCl
microcrystals embedded in alkali-chloride matrices and their correlation with
exciton confinement. Phys. Stat. Sol. (b) 145, 567-577 (1988)
A. I. Ekimov, A. A. Onushchenko. Size quantization of energy spectrum
of electrons for semiconductor microcrystals. Pisma v ZhETF 40, 337340 (1984) (in Russian)
R. Rosetti, R. Hull, J. M. Gibson, L. E. Brus. Hybrid electronic properties between the molecular and solid state limits: Lead sulfide and silver
halide crystallites. J. Chem. Phys. 83, 1406-1410 (1985)
L. E. Brus. Electronic wave functions in semiconductor clusters: Experiment and theory. J. Phys. Chem. 90, 2555-2560 (1986)
A. Henglein. Fluorescence, photochemistry, and size quantization effects
of colloidal semiconductor particles. J. Chim. Phys. Phys.-Chim. Biol. 84,
1043-1047 (1987)
A. Henglein. Q-particles: Size quantization effects in colloidal semiconductors.
Prog. Colloid. Polym. Sci. 73, 1-3 (1987)
L. Spanhel, M. A. Anderson. Synthesis of porous quantum-size CdS membranes: Photoluminescence phase shift and demodulation measurements. J.
Amer. Chem. Soc. 112, 2278-2284 (1990)
V. Swayambunathan, D. Hayes, K. H. Schmidt, Y. X. Liao, D. Meisel.Thiol
surface complexation on growing CdS clusters. J. Amer. Chem. Soc. 112,
3831-3837 (1990)
Y. Nosaka, N. Ohta, H. Miyama. Photochemical kinetics of ultrasmall semiconductor particles in solution: Effect of size on the quantum yield of electron
transfer. J. Phys. Chem. 94, 3752-3755 (1990)
D. Hayes, D. Meisel, O. I. Misic. Size control and properties of thiol capped

225

Nanocrystalline Materials

241.
242.
243.

244.

245.

246.
247.

248.

249.

250.

251.

252.

253.

254.

255.

256.

cadmium sulfide particles. Colloids Surf. 55, 121-136 (1991)


M. V. Rama Krishna, R. A. Friesner. Exciton spectra of semiconductor clusters.
Phys. Rev. Lett. 67, 629-632 (1991)
H. Weller. Quantized semiconductor particles: A novel state of matter for
materials science. Advanced Mater. 5, 88-95 (1993)
D. M. Mittleman, R. W. Schoenlein, J. J. Shiang, V. L. Colvin, A. P. Alivisatos,
C. V. Shank. Quantum size dependence of femtosecond electronic dephasing
and vibrational dynamics in CdSe nanocrystals. Phys. Rev. B 49, 1443514447 (1994)
Y. Kayanuma. Quantum-size effects of interacting electrons and holes in
semiconductor microcrystals with spherical shape. Phys. Rev. B 38, 97979805 (1988)
T. Raih, O. I. Misic, A. J. Nozik. Synthesis and characterization of surface-modified colloidal cadmium telluride quantum dots. J. Phys. Chem. 97,
11999-12003 (1993)
Chi-mei Mo, L. Zhang, C. Xie, T. Wang. Luminescence of nanometer-sized
amorphous silicon nitride solids. J. Appl. Phys. 73, 5185-5188 (1993)
S. Hotchandani, P. V. Kamat. Charge-transfer processes in coupled semiconductor systems. Photochemistry and photoelectrochemistry of the colloidal
CdS - ZnO system. J. Phys. Chem. 92, 6834-6839 (1992)
W. G. Becker, A. J. Bard. Photoluminescence and photoinduced oxygen
adsorption of colloidal zinc sulfide dispersions. J. Phys. Chem. 87, 48884893 (1983)
Hyeong Chan Youn, S. Baral, J. H. Fendler. Dihexadecyl phosphate, vesiclestabilized and in situ generated mixed cadmium sulfide and zinc sulfide
semiconductor particles: Preparation and utilization for photosensitized charge
separation and hydrogen generation. J. Phys. Chem. 92, 6320-6327 (1988)
P. V. Kamat, N. M. Dimitrijevic, R. W. Fessenden. Photoelectrochemistry
in particulate systems. 6. Electron-transfer reactions of small cadmium sulfide
colloids in acetonitrile. J. Phys. Chem. 91, 396-401 (1987)
K. R. Gopidas, P. V. Kamat. Photoinduced charge transfer processes in
ultrasmall semiconductor clusters. Photophysical properties of CdS clusters in Nafion membrane. Proc. - Indian. Acad. Sci. (Chem. Sci.) 105, 505512 (1993)
T. Raih, O. I. Misic, D. Lawless, N. Serpone. Semiconductor photophysics.
7. Photoluminescence and picosecond charge carrier dynamics in cadmium
sulfide quantum dots confined in a silicate glass. J. Phys. Chem. 96, 46334641 (1992)
Y. Nosaka, K. Tanaka, N. Fujii. Laser-irradiation effect on poly(vinyl alcohol) films doped with nanometer-sized CdS particles: Ablation and thied
harmonic generation. J. Appl. Polym. Sci. 47, 1773-1779 (1993)
N. M. Dimitrijevic. Electron-transfer reactions on cadmium selenide colloids as studied by pulse radiolysis. J. Chem. Soc. Faraday Trans.1 83,
1193-1201 (1987)
K. R. Gopidas, P. V. Kamat. Photophysical behavior of ultrasmall cadmium
selenide semiconductor particles in a perfluorosulfonate membrane. Mater.
Lett. 9, 372-378 (1990)
N. Chestnoy, T. D. Harris, R. Hull, L. E. Brus. Luminescence and photophysics
of cadmium sulfide semiconductor clusters: The nature of the emitting electronic
state. J. Phys. Chem. 90, 3393-3399 (1986)

226

Microstructure of Compacted and Bulk Nanocrystalline Materials

+D=FJAH$

6. Microstructure of Compacted and Bulk


Nanocrystalline Materials
The difference in the properties of nanocrystalline and coarsegrained polycrystalline substances is associated with the small size
of crystallites and extremely developed interfaces of the
nanocrystalline substances. The interfaces can include up to 50 %
of atoms of the nanocrystal. At present, many researchers of
nanocrystalline compacted materials assume that the physical
properties of these materials (in particular, mechanical properties)
are determined primarily by the large area and special structure of
the interfaces [1,2]. For this reason, the examination of
nanocrystalline compacted substances is concentrated mainly on the
analysis and describing of the special features of the structure of
grain boundaries.
Before we proceed to discuss the microstructure of
nanocrystalline materials, i.e. polycrystalline substance, let us define
more exactly some notions, which will be used below. Three types
of the grain contact are possible in a polycrystalline substance. They
include contact surfaces, contact lines, and contact points. Surfaces
of two grains, which contact one another, are called interfaces. A
contact line may represent a common line for three or more
adjacent grains. The contact line of three grains is called triple
junction. The triple junctions are most often appearing contact lines
among different contact lines, which exist in polycrystals. Usually
about six grains meet at a contact point. The boundary of a grain
is its surface. Grain boundaries, which are seen in metallographic
slides, represent the section of interfaces by the slide plane. A triple
point is the section of the triple junction (contact line of three
grains) by a plane.
227

Nanocrystalline Materials

6.1 INTERFACES IN COMPACTED MATERIALS


The density of nanocrystalline materials, produced by different
methods of compacting nanopowders [313] can reach from 7080
% to 9597 % of theoretical density. In the simplest case, a
nanocrystalline material, consisting of the atoms of the same type,
contains two phases (components) differing in the structure [14]:
ordered grains (crystallites) with a size of 520 nm, and
intergranular boundaries up to 1.0 nm wide (Fig. 6.1). All
crystallites have the same structure and differ only in the
crystallographic orientation and sizes. The structure of the
intergranular boundaries is determined by the type of atomic
interaction (metallic, covalent, ionic) and by mutual orientation of
the adjacent crystallites. The different orientation of adjacent
crystallites results in a small decrease of the density of the
substances at the intergranular boundaries. In addition to this, the
atoms at the intergranular boundaries have a different short-range
order in comparison with the atoms in the crystallites. In fact, Xray and neutron diffraction investigations of nanocrystalline
compacted nc-Pd [15, 16] show that the density of the substance
at the intergranular boundaries is 2040 % lower than that of
coarse-grained Pd, and the coordination number of the atoms,
belonging to the intergranular boundary, is smaller than the

Fig. 6.1. Computed two-dimensional model of atomic structure of a nanostructured


material [14]: ( ) atoms of crystallites; ( l ) atoms of interfaces are displaced by
more than 10 % of interatomic distance from the corresponding lattice sites; all
atoms are chemically identical, atomic structure of all crystallites is identical too.
The computations are performed by modeling the interatomic forces by the Morse
potential.

228

Microstructure of Compacted and Bulk Nanocrystalline Materials

coordination number of the atoms in the normal crystal. The width


of the intergranular boundaries, determined by different methods for
different nanocrystalline compacted materials, varies from 0.4 to
1.0 nm [1720].
According to initial modeling considerations [7, 21, 22], the
structure of intergranular substance is characterised by a random
distribution of the atoms and by the absence of not only the longrange but also short-range order. According to the authors of [7,
21, 22], this structure is a gas-like structure, taking into account
not the mobility of the atoms but only their distribution (Fig. 6.1).
Experimental confirmation of some disordering of the intergranular
substance in nanomaterials, produced by compacting of
nanopowders, has been provided by the results of diffraction
investigations [21, 22].
At the same time, the latest investigations show [2327] that the
structure of the interfaces in nanomaterials is similar to that in
conventional coarse-grained polycrystals and the degree of order in
the distribution of the atoms at the interfaces is considerably higher
in comparison with previous assumptions [7, 1416, 21, 22]. The
application of high-resolution electron microscopy shows [28] that
in nanomaterials, like in coarse-grained polycrystals, the atoms of
the interfaces are subjected to the effect of only two adjacent
crystallites. Pores were detected only in triple points and not along
the entire length of the interfaces; the density of the atoms in the
intergranular boundaries was almost the same as in the crystallites.
The data on the relatively high degree of order in the distribution
of atoms at the grain boundaries in specimens of nanocrystalline
compacted nc-Pd are reported by the authors of [17, 18, 22, 24].
In [25, 26], analysis of experimental data on X-ray diffraction and
X-ray absorption spectroscopy (EXAFS) of nanocrystalline
substances was carried out using the function of the radial
distribution of atomic density

(r) = x l l (r) + (1 x l )

gb

(6.1)

where r is the interatomic distance; x l is the fraction of atoms


occupying the sites of the crystal lattice; l (r) is the atomic density
distribution function of the nanocrystal, in which the atoms of the
external layers of the crystallite are distributed in the sites of the
lattice; <> gb is the atomic density of the grain boundaries in which
all atoms are distributed in random positions, not coinciding with the
229

Nanocrystalline Materials

sites of the crystal lattice. The value of coefficient x l can be


determined if one knows the experimental atomic density distribution
function (r). For this purpose, using the function (r), it is
necessary to construct the dependence of the relative coordination
number Z i / Z iideal (Z i is the experimental number of the atoms in the
i-th coordination sphere; Z iideal is the coordination number of the ith coordination sphere for the perfect crystal) on the interatomic
distance r. The limiting value of Z i / Z iideal at r = 0 corresponds to
the value x l , i.e. the fraction of the atoms occupying the sites of
the crystal lattice.
Figure 6.2 shows the function of the radial distribution of the
atomic density in a specimen of nanocrystalline compacted nc-Pd,
aged at room temperature for 4 months. The distribution function
(r) was calculated [25] on the basis of the experimental data
obtained by X-ray diffraction. The analysis carried out in [25] using
the function (r) shows that in the nc-Pd specimens, aged at room
temperature for several months or on the same specimens,
subjected to additional annealing at 973 K after aging, the ratio
Z i / Z iideal at r = 0, i.e. the value of coefficient x l is equal to unity
(Fig. 6.3a, 6.3b). In the nc-Pd specimens, investigated not later
than ten days after compacting, 814 % of the atoms were not
located in the lattice sites (Fig. 6c) and the degree of short-range





   

QP





























7 QP

Fig. 6.2. Atomic density distribution function (r) for an aged nanocrystalline
compacted nc-Pd specimen [25]: highest peaks of eight coordination spheres are
numbered.
230

Microstructure of Compacted and Bulk Nanocrystalline Materials


1.10

Zi /Ziideal

1.05

1.10

1.05

1.00

1.00

0.95

0.95

0.90

0.90

0.85

0.85

0.80

0.80

0.75

Zi /Ziideal = -0.17ri + 1.009

Zi /Ziideal = -0.06ri + 0.996 Zi /Ziideal = -0.07ri + 0.861

0.70

0.75
0.70

0.2 0.4 0.6 0.8 1.0 1.2 1.4


ri (nm)

0.2 0.4 0.6 0.8 1.0 1.2 1.4


ri (nm)

0.2 0.4 0.6 0.8 1.0 1.2 1.4


ri (nm)

Fig. 6.3. Relative coordination number Z i / Z iideal versus interatomic distance r i for
nanocrystalline compacted nc-Pd [25]: (a) specimen aged without annealing, (b)
same aged specimen after several annealing runs, the last at 973 K, (c) as-prepared
specimen nc-Pd.

order was very low. The results show that immediately after
production of the compacted specimens, the grain boundaries in ncPd are in the non-equilibrium state with a low short-range order.
This state is unstable even at room temperature and during 120
150 days changes to a more ordered state with an increase of the
crystallite size from 12 to 2580 nm (Fig. 6.4) [26]. The calculations
carried out in [2527] show that the coordination number of the
atoms, distributed at the interfaces in aged nc-Pd, is close to that
in coarse-grained crystalline Pd.
Examination of the short-range order in nanocrystalline
compacted nc-Pd and crystalline coarse-grained Pd by the EXAFS
method [29] showed that the functions of the radial distribution of
atomic density (r) are identical. The coordination number for the
first coordination sphere in the as-prepared and annealed (at
370 K) specimens of nc-Pd was 56 % lower than that for the
coarse-grained Pd. This is in agreement with the data [25] (Fig.
6.3 a,c). According to [29], the reduced coordination number of the
first coordination sphere in nc-Pd is a consequence of the
thermodynamically non-equilibrium state of the substance and of the
presence of lattice vacancies in the specimen.
Examination of compacted specimens of nanocrystalline iron nc-Fe
(produced in high vacuum) with a mean size of crystallites of 10 nm
[30] shows that 955 % of all atoms are located in the sites of the
bcc lattice. In an earlier study [22], the authors did not find any
significant short-range order in the distribution of the atoms at the grain
231

Nanocrystalline Materials



 QP

















7LPH GD\V
Fig. 6.4. Variation of grain size D with specimen age at room temperature for
several nc-Pd specimens [26].

boundaries of nc-Fe. In [30] it is shown that unusual results [22] are


associated with the oxidation of the surface of crystallites: in the
specimens of nc-Fe, produced in high vacuum with subsequent
exposure in air, only 725 % of the atoms occupied the sites of the
bcc lattice of iron, and the majority of the rest iron atoms (~23 %)
forms an amorphous oxide phase, and only a small part of the iron
atoms (~5 %) are located in the lattice sites of crystalline oxide phase.
Investigation of the short-range order in nanocrystalline
compacted cobalt nc-Co [31] with a mean crystallite size of 7 nm
shows that the specimens contained ~70 % of the disordered
amorphous phase and ~30 % of the ordered crystalline phase. The
authors of [31] noted that the disordered phase, located at the grain
boundaries, has no special features typical of the disordered gaslike phase. The relative content of the disordered phase in nc-Co
overestimates evidently because the specimens of nc-Co were
partially oxidised (this has been reported by the authors themselves);
in addition, in processing of the experimental EXAFS spectra, the
presence of lattice defects and free volumes were not taken into
account.
In fact, the interfaces of nanocrystalline compacted material may
232

Microstructure of Compacted and Bulk Nanocrystalline Materials

Fig. 6.5. Two-dimensional schematic model of a nanocrystalline material with


submicroscopic free volumes as detected by positron lifetime spectroscopy [32]:
vacancy-like free volumes (with the positron lifetime 1 ) in the interfaces, nanovoids
(agglomerates of about 10 vacancies) as triple junctions ( 2 ), and large voids ( 3 )
of the size of missing crystallites.

contain three types of defects [32]: single vacancies; vacancy


agglomerates or nanovoids, located at triple points of the
crystallites; large voids in the place of absent crystallites (Fig. 6.5).
These defects represent structural elements of the interfaces with
reduced density.
Ignoring the presence of free volumes leads to large errors in the
determination of the volume fraction of the interfaces in the
nanocrystalline materials. For example, examination of nc-Pd by the
method of small-angle neutron scattering [33] and subsequent
processing of the experimental data, disregarding porosity, resulted in
an incorrect conclusion according to which the volume fraction of the
crystallites and the interfaces in nc-Pd are equal to 0.3 and 0.7,
respectively. According to the estimations [33], the relative density of
intergranular substance was only 50 %. In a later examination of ncPd subjected to high degree of compacting [30] it is found that the
density of the second phase reaches to zero. This means that the voids
but not the interfaces with reduced density are scattering objects.
Comparison of the data [33, 34] and own results enabled the authors
of [26] to conclude that small-angle scattering may provide information
233

Nanocrystalline Materials

on the atomic structure of interfaces, but may not provide information


of the free volumes in the interfaces.
6.2. STUDY OF NANOCRYSTALLINE MATERIALS BY MEANS OF
POSITRON ANNIHILATION TECHNIQUE
Positron annihilation [32, 35] is the most efficient, sensitive and
reliable method of studying free volumes in nanocrystalline
compacted and nanocrystalline bulk materials. This method is most
promising for examining the electronic structure of solids with point
(zero-dimensional), linear (one-dimensional) and volume (threedimensional) defects. The method is sensitive to a very small
content of defects in the solid, from 10 6 to 10 3 defects per atom.
Because of the trapping by defects, the positrons are efficiently
used for analysis of the interfaces in nanostructured substances. The
free path (diffusion) of a positron in a defect-free ideal crystalline
solid is approximately 100 nm, i.e. it is longer than the size or
particles or grains of the nanocrystalline material. Therefore, after
thermalisation, i.e. slowing down, the positron is usually trapped at
the interfaces. In addition, some of the positrons can be trapped in
triple junctions of adjacent crystallites, voids and in the volumes of
missing nanocrystallites (Fig. 6.5) [32]. This circumstance providing
a unique possibility of solving one of the most complicated and
interesting problems of nanomaterials, i.e. understanding the
structure of the interface. In fact, the structure of the interface
together with the small grain size determine most the properties of
the nanomaterial. Other direct methods, including high-resolution
transmission electron microscopy and diffusion of atoms, are far
less suitable for investigation of the interfaces.
Although electronpositron annihilation is a modern method of
examining the defects of solids, including nanocrystalline materials,
this method is not known widely to many material scientists.
Therefore, its specific features, possibilities and applications will be
discussed briefly here.
Positron annihilation makes it possible to determine the
characteristics of the electronic system of perfect crystals and, at
the same time, is sensitive to imperfections of especially small sizes
in the solid, such as vacancies, vacancy clusters and free volumes
up to 1 nm 3 . There are three methods of electronpositron
annihilation: positron lifetime, angular correlation of annihilation
radiation, and coincident Doppler broadening of positronelectron
annihilation radiation. If the measurement of positron lifetime gives
234

Microstructure of Compacted and Bulk Nanocrystalline Materials

information on the electron density in the site of positron


annihilation, the two other methods provide the data on the
distribution of momenta of electrons, with which the positrons are
annihilated. Further, all three methods may be subdivided into two
groups. The first group of methods uses slow positrons, which make
it possible to investigate of the substance at a small depth below
the surface. Methods of second group use fast positrons, which
penetrate to a larger depth (up to 50 m) and give the data on the
type, concentration and distribution of defects in the entire volume
of the solid.
In most cases, nanomaterials are studied using fast positrons
because they are efficiently in providing information on the
structure of the interfaces. It is common practice to use radioactive
isotope 22 Na, 44 Ti or 58 Co having the maximum energy of emitted
positrons equal to 0.54, 1.50 and 0.47 MeV [36], respectively.
According to [37], the linear coefficient of absorption of the
positron in the solid is:

(16 1)
1.43
E max

[cm 1]

(6.2)

Substituting the maximum energy of the positron E max (in MeV) and
the density (in g cm 3 ) into equation (6.2), it is possible to
determine x = 1 , i.e. the penetration depth of a positron in a
nanosubstance at a given intensity of the beam. The isotopes used
in most cases emit positrons with a maximum energy of
approximately E max = 0.5 MeV, and the density of nanomaterials
is from 5 to 10 g cm 3 . Taking this into account, equation (6.2)
shows that the positron penetrates in a nanosubstance to a
maximum depth from 200 to 300 m (in this case, the intensity of
the positron beams weakens e 6 times), and the main information is
received from a depth of 2550 m (intensity weakens of e times).
The positron emitted from a radioactive source (emitter) is
thermalised after penetrating in a solid, i.e. rapidly loses its velocity
and energy, which decreases to the value k B T corresponding to the
temperature T of the crystal. The thermalisation time is about 1
5 ps and is negligibly small in comparison with the positron lifetime
in a solid. Because the grain size of nanomaterials is not larger
than 40 nm, after thermalisation the positrons are uniformly
distributed over the volume of many grains located at a large depth
from the crystal surface. Thanks to the small size of the grains, the
235

Nanocrystalline Materials

fraction of the positrons thermalised near the surface of the grains,


coincides in the order of magnitude with the fraction of the
positrons thermalised inside the grains.
A thermalised positron starts to diffuse through the nanomaterial
in the so-called free (delocalised) state and annihilates from this
state in the characteristic time f of approximately 100 ps. This
value of the positron lifetime in the free state (or free lifetime)
is characteristic of metals [38] and many compounds, for example,
carbides [39]. A positron can move in defect-free solids to a
distance of about 100 nm during the time f . This estimate follows
from the experimental data concerning the diffusion coefficient of
positron D 0+ = 12 cm 2 s 1 in such metals as aluminium, copper,
molybdenum, silver and in silicon carbide SiC [40, 41] where
positrons move to a distance of 100180 nm and 7010 nm
respectively during their free lifetime. If a substance is not defectfree then positron could be trapped by defect. If the distance
between defects is close to the diffusion length of positron, some
of the positrons will annihilate in free state and some of positron
will annihilate in trapped (localised) state. In this case the positron
lifetime spectrum consists of two components. If the distance
between defects is very short in comparison with the diffusion
length of positron, all the positrons will be trapped by defects. This
case is called saturation and leads to a one component positron
lifetime spectrum.
Since the grain size in nanomaterials is smaller than the diffusion
length of the positron in a defect-free grain, almost all positrons may
reach the grain surface and, consequently, the interface. In this
case the data mostly on defects in the interfaces and in triple
junctions could be obtained. If the interior of the grains contains
defects which trap positrons, only some of the positrons reach the
grain boundaries and, consequently, the data on intragranular
defects in nanomaterials may be obtained.
According to the results of a large number of investigations, the
positron lifetime in a substance is determined by the number of
valence electrons and not by the total number of electrons. This
takes place owing to the fact that a positively charged positron
escapes the positively charged atom nuclei. Therefore, almost all
positrons (approximately 9598 %) annihilate with valence electrons
in the internuclear space. The authors of [38, 42, 43] established
a correlation between the normalised rate of positron annihilation

= ( ) /

(6.3)
236

Microstructure of Compacted and Bulk Nanocrystalline Materials

and the normalised volume density of valence electrons

= zVB / V .

(6.4)

In equations (6.3) and (6.4), = 2.0010 9 s 1 ; = 1/ ; z is the


number of valence electrons, participating in annihilation and
corresponding to one formulae unit of the substance occupying a
volume V; V B = 6.20810 31 m 3 is the volume of a sphere with the
Bohr radius. For metals, the correlation between the normalised rate
of positron annihilation, *, and the normalised volume density of
valence electrons, , is nonlinear [42]:

= (24.8 ) 0.81 .

(6.5)

For the transition metal oxides [43], a deviation from this


dependence, associated with the participation in annihilation of
d electrons, is found. In carbides, in which the carbon vacancies
are surrounded by the atoms of transition metals, the relationship
between the normalised rate of annihilation (6.3) and the volume
density of d electrons, calculated from equation (6.4), is close to
the dependence (6.5) but is linear [38]
val
* = 17.9 M
.

(6.6)

The essence of the correlations (6.5) and (6.6) is that as the


volume density of the valence electrons increases, the positron
lifetime becomes shorter. In addition to this, these correlations make
it possible to estimate quantitatively the free positron lifetime and
the lifetime of the positrons in vacancies, if the atomic and crystal
structure and chemical composition of the substance are known.
From the fundamental studies of positron annihilation in metals
[38], semiconductors [44], nanocrystals [45, 46], amorphous
materials [47, 48] and ceramics [38, 43] it is known that the
positrons are trapped by vacancies on the metallic and nonmetallic
sublattices, vacancy clusters, grain boundaries, dislocations, etc.
The interstitial atoms, which also represent defects in a solid,
usually do not trap positrons because of the large positive charge
of the atomic nucleus. After trapping by a defect, a positron
annihilates from the localised state in a time exceeding f . The
237

Nanocrystalline Materials

electron density in the defect is smaller than the electron density


in the interatomic space of the defect-free crystal, and, consequently,
the positron lifetime in the defect is longer than the free lifetime.
The positron lifetime depends strongly on the size of the free
volume. As the free volume increases, the positron lifetime also
increases. Interesting investigations were carried out in [49] where
a specific relationship was found between the positron lifetime
and the number of vacancies, N V , in the free volume of silicon Si

= C + AN V/(B + N V) ,

(6.7)

where A = 266.57 ps, C = f = 218 ps, and B = 4.60. By


replacing the number of vacancies N V by the free volume V,
equation (6.7) may be simplified and presented in the following
approximate form:
Si

[ns] = 0.22 + 1.4V [nm 3] .

(6.8)

Although this relationship is obtained for the pure element substance


silicon, it may be used to evaluate the size of the free volume V
at the interfaces of nanocrystalline materials. For this purpose, the
equation should be presented in the following form:

[ns] = f [ns] + 1.4V [nm3] ,

(6.9)

where f is the positron lifetime in a coarse-grained defect-free


substance or a defect-free single crystal.
In addition to the possibility of investigating the size of the free
volume, the positron annihilation makes it possible to carry out the
chemical analysis of the atomic environment of the free volume
[5052]. This possibility has appeared recently as a result of the
rapid development and reduction in the cost of high-resolution
equipment for detecting -quanta. Relatively cheap and easy to
operate -quanta detectors with a high energy resolution of about
12 keV for the -quanta energies close to 511 keV have made it
possible to develop a two-detector method of Doppler broadening
of the electronpositron annihilation radiation. To obtain the
distribution of electron momenta up to very high values of electron
momentum, it is necessary to take special coincident measurements
using two detectors. This coincident method makes it possible to
increase the signalnoise ratio and decrease the effect of

238

Microstructure of Compacted and Bulk Nanocrystalline Materials

background on the spectrum.


The Doppler experiments provide data on the distribution of the
momentum of electrons annihilating with the positron. The thermalised
positron has momentum approaching zero and, consequently, the shift
of the energy of annihilation -quanta is determined exclusively by
electron momenta. High electron momenta are determined by the core
electrons of the atoms and, consequently, the profile of the distribution
at high momenta is a unique fingerprint of the orbitals of the core
electrons of the element and makes it possible to identify chemical
elements, which are located in the nearest surrounding of the site of
positron annihilation, i.e. in the surrounding of the free volume.
Calculations or numerical processing of the profile of the
experimental Doppler spectrum are possible if we know the threedimensional electron momentum distribution. The momentum distribution
may be determined using the Schrdinger equation and first principal
quantum calculations, but this is a very time-consuming and difficult
operation which does not always lead to success in the case of
complicated defective substances and nanocrystalline materials.
Analysis of the chemical environment of the defects may be carried
out using a semi-empirical model proposed in [53]. This model is based
on the possibility of measuring the mean kinetic energy of core
electrons, E k, and on comparison of this energy with the tabulated
values of binding energy E B, which are available at the present time
for the electronic orbitals of almost all elements of the Periodic Table
of Elements. These measurements are possible in experiments with the
angular correlation of annihilation radiation or in Doppler experiments.
The mean kinetic energy of the electrons of a specific orbital and the
binding energy of this electron in the atom are compared using the virial
theorem whose consequence is the equality E B = E k.
For calculating the Doppler spectrum, all electrons may be divided
into valence and core electrons. In the valence electrons, i.e. those
which take part in chemical bonding, it is necessary to distinguish
delocalised and localised electrons. The delocalised electrons are quasifree electrons of metallic bond, and localised ones are the electrons
of the covalent and ion bonds, and also the electrons of the core shells.
The positively charged positrons are pushed away from the nuclei of
the atoms because they also have a positive charge. Consequently, the
probability of the presence of a positron in the internuclear space is
maximum. However, the probability of a positron penetrating into the
region of core electrons is not equal to zero and amounts to several
percent. The basic information on the wave function of the positron
is obtained from the experimental spectra in which the contribution of
239

Nanocrystalline Materials

every component is proportional to the overlapping integral of the wave


function of the electron of the appropriate orbital and the positron wave
function.
In the description of the valence electrons in the model of free
(delocalised) electrons, the wave function of the gas of free
electrons is a Bloch wave with a continuous spectrum of momenta
within the limits of the Fermi sphere. The square of the positron
wave function in the region of the valence electrons is close to a
constant and, therefore, the Doppler broadening spectrum [53] is
a truncated inverted parabola

p 2
I p ( p z ) = hp 1 z f ( p z p F ) ,
p F

(6.10)

where p z is the z-th component of the electron momentum, h p is


the height of the parabola, p F is the truncation momentum of the

1, if | p z | p F
0, if | p z | > p F

parabola or the Fermi momentum, f (| p z | p F ) =

is the Heaviside function. This approach makes it possible to


determine by experiments the Fermi energy E F = (p F ) 2 /2m 0 .
Already in studies of non-stoichiometric carbides of transition
metals [53] it was established empirically that the high-energy part
of the spectrum of the angular correlation of annihilation radiation
is efficiently described by a Gaussian. The spectra of the angular
correlation and of Doppler broadening give the same physical data
on the momentum distribution and, consequently, the distribution of
electron momenta in the Doppler broadening spectra may be
represented in the form of a Gaussian:

p2
I g ( p z ) = hg exp z 2
2

(6.11)

where h g is the height of the Gaussian, is the second momentum


of the Gaussian. In accordance with the results of [54], from the
parameter it is possible to transfer directly to the binding energy
using equation E B = (2 ) 2 /2m 0 (m 0 is the electron rest mass).
To determine the type of chemical environment of the defect, the
240

Microstructure of Compacted and Bulk Nanocrystalline Materials

Doppler spectrum is expanded into the truncated inverted parabola


and several Gaussians

I ( p z ) = I p ( p z ) + I gi ( p z ) ,

(6.12)

where i is the number of the Gaussians, which is equal to the


number of different orbitals of the electronic shells, participating in
annihilation with the positrons.
The parabola corresponds to the contribution from positron
annihilation with free electrons, and each Gaussian corresponds to
the contribution from annihilation of positrons with a specific orbital
of the core electrons. For transition metals and compounds of these
metals, the application of this model is complicated by hybridization
of s, p and d bands. Consequently, only part of the valence
electrons can be described by the truncated inverted parabola; the
remaining valence electrons are usually described by a Gaussian.
Because of Coulomb repulsion of the positron from the atomic
nuclei, the contribution of the core electrons to the spectra of the
electron momentum distribution does not exceed several percent.
Therefore, the core components can be separated only in the range
of high electron momenta in which the contribution from the valence
electrons is negligibly small.
The experimental spectrum N(p z ) widens as a result of the finite
resolution of the detectors. Taking into account the instrumental
resolution function of the spectrometer, R(p z ), the experimental
spectrum N(p z ) can be presented as a convolution of the true
spectrum I(p z ) and resolution function R(p z ):

N ( pz ) =

R( p p z ) I ( p z )dp .

(6.13)

The resolution function is a Gaussian. Taking this into account and


integrating equation (6.13) with the use of functions (6.10)(6.12),
the following dependence is obtained in [55]:

N( pz ) =

hp ( p F2 R2 ) ( p z p 0 ) 2 p F ( p z p 0 )
p F + ( p z p 0 )
erf
1
+ erf
+
2
2
2
R 2
R 2
2 pF
( p F R )

241

Nanocrystalline Materials

hp R
( pF p z + p0 ) 2
p
p
p
+

[
(
)]
exp

z
F
0
p F2 2
2 R2

( p p )2
hgi gi
( p + p z p 0 ) 2
0
z
+ [ p F ( p z p 0 )] exp F
+
exp

+ BG,

2
2
2 R2
2( R + gi )

i R2 + g2i

(6.14)
where erf ( x) =

2 x
2
exp(t )dt is the error function integral, p 0 is
0

the momentum corresponding to maximum of the Doppler spectrum,


R is the second moment of the instrumental resolution function, BG
is the random coincidence background. The application of this model
requires fitting of the parameters by the weighted least squares
method, i.e. assumes the minimisation of the function:

=
2

[ N ( p z j ) N exper ( p z j )]2
N exper ( p z j )

(6.15)

It is now possible to transfer directly to the analysis of


investigations of nanocrystalline materials, carried out using the
previously described methods of positron annihilation. The positron
annihilation was used for the first time for examining the special
features of the structure of nanocrystalline materials by the authors
of [56] who investigated the vacancies in the nickel nanoparticles
(D~15 nm), measuring the position lifetime. Examination by the
positron annihilation revealed the existence of vacancies and
nanovoids in nanocrystalline metals Al, Cu, Mo, Pd, Fe and Ni, in
nanocrystalline silicon Si and zirconium oxide ZrO 2 [32, 35, 38, 57
60].
The results of these investigations show that the position lifetime
spectra usually contain three components with the intensities I 1 , I 2 ,
and I 3 , = 1 I 1 I 2 , to which the lifetimes 1 , 2 and 3
correspond (Table 6.1, see also Fig. 6.5).
The first two lifetimes 1 and 2 with the relative intensities I 1
and I 2 dominate the position lifetime spectrum and the long-lived
component 3 appears with only a small intensity I 3 . In the
nanocrystalline metals, lifetime 1 is close in value to the positron
lifetime 1V in lattice monovacancies of coarse-grained metals and,
consequently, 1 is regarded as the positron lifetime in vacancies
242

243

4
4
4

1911
1581
1451

(ps)
(na no c rysta lline )

1871
1581
1511

(p s )
(a mo rp ho us)

1481
1451
11 4 (F e 3S i)

c rysta lline )

(ps) (coarse- grained

* In a d d itio n, the ' fre e ' life time s f in d e fe c t- fre e c rysta ls, a s we ll a s p o sitro n life time in mo no va c a nc ie s iV a nd in a gglo me ra te s o f i va c a nc ie s ( iV) a re give n; D is the c rysta llite
(gra in) size

C o 33Zr67
F e 90Zr10
F e 73.5C u1N b 3S i13.5B9

Allo y

0.21

0.79
0.31

14500600

4156
3781

1892
1992

20
10

2
2

N iZr
ZrO 2

3 2 1 ( i= 9 )

2 7 2 ( i= 8 ) [ 6 6 ]

1 8 0 [3 8 ]

2 7 2 [4 4 ]

1 0 8 [6 5 ]
1 0 8 [6 5 ]
1 0 3 [3 8 ]

2 1 9 [4 4 ]
1 0 8 (P d )
[6 5 ]
1 4 2 [3 8 ]
1 7 5 [4 3 ]

0.10
0.04
0.84
0.17
0.12
0.79
0.84
0.90
0.96
0.16
0.83
0.88
0.20
0.16

3330300

359
398
3452
2979
3309
4 2 2 11
3513

202
172
2049
1712
1611
31435
1708

86
76
10
100
200
10
9

1
3
2
3
3
1
2

1 0 8 [6 5 ]

0.60

161

29

0.40

4 2 2 ( i= 1 3 )
3 3 4 ( i= 1 0 )
3 7 6 ( i= 1 3 )

11

* iV (p s)
[6 3 ]

10

1 V( p s )

316

4127
3376
3632
29940
3 11 1 3
3475
465
365

* f (p s )

2 5 1 [3 8 ]
1 7 5 [3 8 ]
1 8 0 [6 4 ]
1 7 9 [3 8 ]

11 5 [6 0 ]

I2

1 6 8 [3 8 ]
1 0 6 [3 8 ]
9 4 [3 8 ]
11 2 [3 8 ]

1 0 8 [3 8 ]
1 0 8 [3 8 ]
1 0 8 [3 8 ]

I1

0.41
0.75
0.72
0.44
0.59
0.66
0.41
0.53

3 (p s )

0.58
0.20
0.28
0.43
0.40
0.33
0.59
0.47

2 (p s )

1970160
90040
53070
47050
600200
1080170

1 (p s )

2534
1618
741
17510
1828
1825
2 11
186

40
10
12
20
14
12
20
20

1
1
1,2
1
2
1
1
1

Al
Fe
Ni
Cu
Cu
Pd
P d (is me a sure d in va c uum)
P d (is me a sure d a fte r
e xp o sure to a ir)
P d (c o mp a c tio n a t a p re ssure
o f 2 GP a a nd a t a te mp ra ture
of 373 K )
P d (b a ll milling)
Pd
Mo
Cu
Ni
Si
P d 3F e

D
(nm)

P re p a ra tio n
me tho d

Ma te ria l

Table 6.1
Positron lifetimes 1 , 2 , 3 and relative intensities I 1 , I 2 in nanocrystalline solids after compaction of crystallites prepared
by evaporation (1) or sputtering (2), in nanostructured metals prepared by severe plastic deformation (3) and moderate annealing, as
well as the mean positron lifetime in nanostructured alloys after crystallization of the amorphous alloys (4) [32, 35, 6062]

Microstructure of Compacted and Bulk Nanocrystalline Materials

Nanocrystalline Materials

in the interfaces. The size of these vacancies (vacancy-like free


volumes) corresponds to 12 missing atoms. The affiliation of these
free volumes to the interfaces, and not to the crystallites, is
confirmed by the fact that the lifetime 1 is observed even after
annealing nanocrystalline metals at a temperature higher than the
annealing temperature of lattice monovacancies. The positron
lifetime 2 characterises the positron annihilation in threedimensional lattice vacancy agglomerates (nanovoids) whose size
approximately corresponds to 10 missing atoms.
The positron lifetimes 1 and 2 in submicrocrystalline metals
Cu, Pd, Ni (Table 6.1) are similar to the values 1 and 2 for the
appropriate nanocrystalline metals. However, in the case of submicrocrystalline metals, the contribution of the second component
I 2 to the positron lifetime spectrum is considerably lower because
the size of the crystallites in the submicrocrystalline metals is larger
than that in the nanocrystalline metals.
The long positron lifetime 3 corresponds to the positron
annihilation in nanopores, i.e. large free volumes whose size is
close to the size of missing crystallite (see Fig. 6.5).
The annealing of nano-Pd at 700 > T > 400 K increases the mean
positron lifetime as a result of joining of the separate free
volumes and of increasing their size. This process of structural
relaxation of the interfaces is accompanied by an increase of the
density of substance of the interfaces. At higher annealing
temperatures the crystallites grow and at T > 1200 K, the mean
crystallite size is already larger than the free path length of the free
positron. Consequently, the contribution of the free volumes of the
interfaces to the positron annihilation process decreases and the
mean lifetime is shortened to the value corresponding to the free
positron lifetime f in coarse-grained metals.
Let us consider in detail the investigations of FeSiNb
nanocrystalline alloys [67] which threw light on the possibilities of
the positron annihilation method in the examination of
nanomaterials. This work indicates how the positron annihilation can
be used for detailed precision investigations of the complicated and
interesting processes, which take place on the microscopic atomic
level in the nanomaterials at elevated temperatures.
In [67], using the combination of the positron annihilation and Xray diffraction analysis, the authors investigated the formation of
thermal vacancies and of the long-range order of the D0 3 type in
the disordered (Fe 3Si) 95 Nb 5, produced by mechanical alloying [68].
Using radioactive 58 Co isotope as a source of positrons, it was
244

Microstructure of Compacted and Bulk Nanocrystalline Materials

possible to measure the positron lifetime in situ at high


temperatures and observe the formation of vacancies in the
volumes of the grains and at the interfaces. The two-detector
Doppler spectroscopy has detected Nb at the interfaces and triple
points and made it possible to clarify the mechanism of stabilisation
of the nanocrystalline state. The FeSiNb system was chosen
because of the unique combination of the high thermal stability of
the nanocrystalline state and a low vacancy formation enthalpy in
D0 3 -ordered Fe 3 Si crystallites. Previous studies of FINEMET
industrial alloy, produced by crystallisation from the amorphous
state, showed that a high diffusivity is the result of a high thermal
vacancy concentration in Fe 80 Si 20 nanocrystallites [48]. In [67], it
is shown that a special feature on the temperature dependence of
the mean positron lifetime, detected at 800 K, appears as a result
of competition of two processes: (1) positron trapping at nanovoids,
and (2) positron trapping in thermal vacancies.
Analysis of the behaviour of the positron lifetime using a simple
model, considered previously, made it possible to calculate the
vacancy formation enthalpy in nanocrystalline Fe 3Si silicide. In [67],
attention was also given to the correlation between the ordering
process and the variation of diffusivity as a result of the
modification of the local atomic distribution, i.e. the short-range
order.
The (Fe 3 Si) 95 Nb 5 nanocrystalline powder was prepared by
mechanical attrition [68] with a Spex 8000 laboratory mixer/mill. A
mixture of elemental powders with composition 71.25 at.% iron,
23.75 at.% silicon and 5 at.% niobium was milled in an inert argon
atmosphere at room temperature for 48 hours. The mill and the
milling balls were made of a hard alloy based on tungsten carbide.
The ball-to-powder weight ratio was 4:1. Specimens with a diameter
5 mm and 1 mm thick for the investigations by electronpositron
annihilation were compacted from the produced powder at room
temperature under uniaxial pressure of 1.5 GPa.
For investigations, radioactive isotope 58 Co was used as the
positron source. This isotope in the form of liquid chloride 58 CoCl 2
was deposited on the surface of one specimen. Because the
specimen was porous, it was partially impregnated by cobalt
chloride. After drying the specimen, cobalt was reduced from the
dichloride in a H 2 atmosphere at a temperature of 483 K for 2
hours. The specimen with the 58 Co positron source was stacked
between two other identical specimens. The prepared sandwich was
placed in a thin-walled Fe container and then container with
245

Nanocrystalline Materials

specimens sealed off in a quartz ampoule under high vacuum. The


iron container, characterised by a short positron lifetime
(approximately 100 ps), was used to ensure that in the course of
experiments it would be possible to control the absence of
contamination of the outer surface of sandwich by the positron
source. In the case of such contamination part of positrons should
annihilate with an iron. The component, responsible for positron
annihilation with iron (if such a component appears), would be
easily separated because it is considerably shorter than the other
components of the spectrum of the investigated substance. In the
absence of the iron container and in case of contamination of the
outer surface of sandwich, annihilation will take place with quartz,
which has a long positron lifetime. The component of the lifetime
spectrum, responsible for annihilation of positrons with quartz,
cannot be separated from the components of the spectrum of the
investigated material, because they are very similar in value. The
positron lifetime was measured by means of a fast-slow
-spectrometer with a time resolution (full width at half maximum
FWHM) of 215 ps. The minimum number of coincidence counts in
each spectrum was 0.510 6 .
A series of 2 hour annealing cycles with simultaneous
measurement of the positron lifetime was carried out in situ in the
spectrometer at temperatures T an = 623, 713, 833 and 1023 K. After
each annealing, the positron lifetime spectra were recorded at room
temperature.
The evolution of the structure of the specimens upon annealing
was studied by X-ray diffraction method. Attention was given to the
determination of the parameters of the structure such as the volumeweighted mean grain size D V , the root mean square (rms)
microstrains {2hkl } 1 / 2 , the long-range order parameters S D 03 , S B2
and lattice constant a. X-ray investigations were carried out at room
temperature in a Siemens D500 X-ray diffractometer, using CuK
radiation and a secondary graphite monochromator. A Rietveld-like
analysis of the X-ray spectra was performed, fitting each Bragg
reflection profile with two Voigt functions for the K 1 and K 2 lines
with the same width. The intensities of these doublets were
calculated from structure factors using the kinematic scattering
theory and the appropriate intensity factors [69]. The long-range
order parameters were determined for a compound with the
composition Fe 3 Si. The vacancies were not taken into account
because the concentration of vacancies was considerably smaller
than the concentration of antistructural defects (antisite) for both
246

Microstructure of Compacted and Bulk Nanocrystalline Materials

partially and completely ordered states. It was assumed that the


long-range order parameters depend linearly on the site occupancies
and change from the maximum values S D 03 = 1 and S B2 = 1/2 to
the minimum values S D 0 = S B2 = 0 for the completely disordered
3
state. The long-range order parameters were determined from the
ratio of the intensities of superstructure reflections {111} and {200}
to the intensity of the structural reflection {220}. The intensity of
2
the {111} reflection is proportional to S D 03 and is determined only
by the phase D0 3 , whereas the intensity of the {200} reflection
depends on the long-range order parameters of both phases and is
2
proportional to ( S D 03 + 2S B 2 ) . Consequently, it was possible to
determine the long-range order parameters for both phases D0 3 and
B2. From the line widths, the grain size and rms microstrains
{2hkl } 1 / 2 were determined with taking into account the anisotropy
of elastic distortions and adding the broadening of superstructure
reflections {111}, {200}, etc. by boundaries of ordered domains
[70].
To obtain information on the chemical environment of the atomic
free volumes where positrons are annihilated, the electron
momentum distribution was measured. These measurements were
performed by the coincident Doppler broadening method at room
temperature. The Doppler broadening spectra were analysed using
the previously considered model of description of the electron
momentum distribution (see the equations (6.13), (6.14) and (6.15)).
The separation of the spectra into components showed that in the
range of the electron momentum from 1810 3 m 0 c to 2410 3 m 0 c
(or in the energies of -quanta from 4.5 to 6 keV) the spectra are
determined mainly by the core electrons, and in the range from 0
to 1010 3 m 0 c the spectra are determined by the valence electrons.
In the intermediate range, the contributions from both valence and
core electrons are important.
In order to observe the special features of spectra with the
naked eye, i.e. without preliminary numerical analysis, the Doppler
broadening spectra could be normalised to the same area and after
that divided by the reference spectrum. The reference spectrum is
usually measured with high statistics on pure silicon Si specimen
(see Fig. 2 in [67]).
The temperature dependences of the structural parameters and
the mean positron lifetime , measured at ambient temperature, are
shown in Fig. 6.6. The as-milled material is characterised by a
small grain size D V = 15 nm. This is in good agreement with the
results of examination by transmission electron microscopy [68], a

247

Nanocrystalline Materials

high microstrain {2220} 1 / 2 = 1.1 % and a very low parameter of


D0 3 long-range order. The lattice constant a is higher than that for
ordered Fe 75 Si 25 (a 0 = 0.565 nm [71]). The reduced lattice constant
of the ordered phase is often found in a disorderorder phase
transformations in metallic alloys [72]. The decrease is usually
related to different sizes of the atoms, which may result in denser
packing in the case of ordered state. However, lattice constant
a = 0.5710 nm, measured for ball-milled and most disordered
(Fe 3 Si) 95 Nb 5, is even greater than that for the milled and disordered














$



$ 





QP



$   $  











!;

















!;

QP









SV












%,3 .
Fig. 6.6. Long-range order parameters S D 03 and S B2 , lattice constant a, volume2
1/ 2
weighted mean grain size D V , root mean square microstrain {220} , and mean
positron lifetime for n-(Fe 3 Si) 95 Nb 5 at ambient temperature after 2 h annealing
at T an [67].
248

Microstructure of Compacted and Bulk Nanocrystalline Materials

Fe 75 Si 25 (0.5663 nm [73] or 0.5695 nm [74]). Extrapolation of the


lattice constant of the solid solution with the Si content smaller than
10 % [75] by the lattice constant of the solution with a Si content
of 25 % also gives a low value of 0.5699 nm. Thus, the enhanced
lattice constant indicates that in addition to disordering, partial
dissolution of niobium atoms with a larger atomic radius also takes
place in the compound.
After annealing the specimen at 483 K for 2 hours, a small
decrease is observed in the lattice constant and in the microstrain
{2220} 1 / 2 , while the grain size D V is unchanged. The mean
positron lifetime = 2212 ps results from the combination of two
components 1 = 171 and 2 = 375 ps with relative intensities of
I 1 = 75 % and I 2 = 25 % (Table 6.2).
The value 1 coincides with the positron lifetime for monovacancies in ferromagnetic -Fe ( 1V = 175 ps [76]) and in iron
silicide Fe 3 Si with a structure D0 3 ( 1V = 175 ps) [77]) and
therefore characterises the positron lifetime in free volume of the
size of one missing atom. Component 2 corresponds to the positron
lifetime in nanovoids of the size of 1015 missing atoms [63] (void
diameter 0.50.8 nm). Both positron lifetime components reflect
general features of spectra of nanocrystalline metals [57]. The
value 1 relates to the positron trapping in structural vacancies in
interfaces or in non-equilibrium thermal vacancies in the crystallites,
and 2 is associated with nanovoids located at the intersections of
interfaces of close-packed nanocrystallites. The absence of a
lifetime component smaller than the free lifetime f indicates
saturation trapping of positrons by defects. Saturation takes place
because the mean positron diffusion length in metal crystals
(L + 100 nm) is considerably larger than the crystallite size;
therefore, the positrons during diffusion reach the interfacial traps
with a high probability.
Table 6.2 Mean positron lifetime and component analysis in starting specimen
(Fe 3 Si) 95Nb 5 and the same specimen after annealing at the temperature T an = 1023 K
for 2 h. I 1 and I 2 represent the relative intensities of the two components with
lifetimes 1 and 2 , respectively [67]

T an(K )

(p s)

1(p s )

2(p s )

I1(%)

I2(%)

483
1023

2222
2252

1711
1741

3752
3703

75.00.5
741

25.00.5
261

249

Nanocrystalline Materials

The electron momentum distribution shows that positrons


annihilate in the vicinity of niobium atoms [67]. This is indicated by
the profile of the electron momentum distribution, which resembles
the profile of pure Nb to a greater extent than the electron
momentum distribution profile of Fe or iron silicide Fe 3Si (see Fig.
2 in [67]). This shows that either Nb atoms are segregated to the
interfaces or the vacancies in the crystallites are surrounded by
niobium atoms.
A temperature of T an = 623 K is characterised by a large
decrease of the mean positron lifetime from 221 to 215 ps. This
decrease takes place as a result of structural relaxation at
interfaces associated with annealing of strains in crystallites and
with the start of segregation of niobium atoms at interfaces. It leads
to a large decrease of the lattice constant a. Upon annealing at
T an = 623 K, a small degree of the long-range order, similar to that
in the as-milled state, is retained. Nearly complete ordering of the
type D0 3 ( S D 03 0.9) is obtained after annealing at T an = 713 K.
Indeed, the shift of the superfine magnetic field measured by
Mssbauer spectroscopy [68, 78], shows that dissolved niobium
leave the crystallite body after annealing for 1 hour at 723 K.
Consequently, both processes (ordering and Nb segregation) take
place at this temperature. It leads to a further decrease of the
lattice constant a (by about 0.9 %) and of the microstrain {2220} 1 / 2
(see Fig. 6.6).
After annealing at T an = 833 and 1023 K, the complete long-range
order is established and, according to the decrease in the width of
the superstructure reflection {111}, all antiphase domains, in
particular those, which determined by the displacement vector
100 /2, are annealed out. In this stage, the rms displacement
decreases and the lattice constant a reaches a value of 0.5658 nm.
Assuming that all Nb atoms are segregated at the interfaces, this
value corresponds to a composition Fe 76 Si 24 [73], which is in
agreement with the energy-dispersive X-ray analysis (EDX) [68].
At a high temperature of 8001000 K, the most evident
mechanism of structural relaxation is grain growth. On the basis of
the lower thermal stability of Fe 3 Si without Nb additions it may be
concluded that the segregation of niobium at interfaces inhibits
grain growth. After annealing at 1023 K, when the mean grain size
D V reaches ~110 nm, there are no the component corresponding
to the free positron lifetime. Since the width of the instrumental
resolution function, determined on a polycrystalline silicon nitride
Si 3 N 4 , corresponds to a crystalline size of 100200 nm, the value
250

Microstructure of Compacted and Bulk Nanocrystalline Materials

D V 110 nm is highly indeterminate. The saturation trapping of


the positrons means that L + 100 nm is still lower than the grain
radius D V /2, and confirms the results of X-ray diffraction
analysis. It should be mentioned that the mean positron lifetime
and the ratio of the components in the annealed and starting state
are practically identical (see Table 6.2), although the crystallite size
differs greatly. In accordance with the conventional interpretation
according to which the nanovoids are distributed mainly at lines of
intersection of interfaces, the ratio I 2 /I 1 should greatly decrease
with increasing crystallite size. Evidently, a decrease in the value
of I 2 /I 1 is compensated by annealing of quenched thermal or
interfacial vacancies and, consequently, the intensity of I 1
decreases. Interfacial vacancies may be associated with a presence
of niobium atoms at interfaces, because there are no interfacial
vacancies in pure Fe 3Si. In addition to this, additional nanovoids may
appear during crystallite growth as a result of agglomeration of the
vacancies. This increases the intensity of I 2 . This is fulfilled for
pure iron, processed by ball milling. Two-component analysis of the
Doppler spectra confirms the hypotheses according to which the
observed increase in the mean positron lifetime after annealing
at T an = 1023 K takes place due to a decrease of I 1 or increase of
I 2 , and not because of the change in the size of the vacancy
clusters. It is important to note that after annealing at 823 K only
small changes were found in the electron momentum distribution.
This is in agreement with the previously described positron
annihilation with electrons of Nb atoms segregated in the interfaces
or in structural vacancies, surrounded by niobium atoms.
The most important result of [67] is the unusual behaviour of the
mean positron lifetime at high temperatures (T > 800 K). As shown
in Fig. 6.7, the mean lifetime has a maximum in the vicinity of
800 K and then rapidly decreases with increasing temperature. The
measured dependence (T ) is equilibrium and is fully reversible
with a subsequent decrease or increase of temperature.
Analysis of the positron lifetime spectra shows that no new
component with a short lifetime, which could be used to explain the
decrease in , forms. A decrease in correlates with an increase
in intensity I 1 as a result of a decrease in intensity of I 2 . With
increasing temperature, the probability of positron trapping by the
vacancies, formed in thermodynamic equilibrium inside the
crystallites, increases. Since the positron lifetime in equilibrium
thermal vacancies, V, T , is approximately equal to that in interfacial
vacancies, V, S , then the intensity of the shorter lifetime component
1 increases and that of component 2 decreases. This ensures the
251

Nanocrystalline Materials



%2


B
(SV

























%,3 .

Fig. 6.7. Mean positron lifetime in n-(Fe 3 Si) 95 Nb 5 [67] measured in isothermal
conditions with the following sequence: () 2 h annealing at T an = 1023 K was
performed; ( ) the specimen was cooled to 293 K and a series of data was measured
at temperature increasing up to 1003 K; ( ) further data were taken at decreasing
temperature. This sequence proves the full reversibility of the observed temperature
dependence. The solid line is a fit to the whole set of data according to the threestate model of positron trapping (Eq. 6.16). The inset shows the same data together
with an extension of the fitting curve to higher temperatures.

reversibility of equilibrium dependence (T ) .


The chemical environment of interfacial vacancies is enriched by
niobium, while the chemical environment of thermal vacancies
formed inside the crystallites is niobium poor. Therefore, the
electron momentum distribution at high temperatures should change
and the profile of the spectrum should approach that of pure coarsegrained Fe 3 Si.
According to the above interpretation, the temperature dependence of can be described by the combination of the temperature
behaviour of three different traps:
(1) Interfacial vacancies (vacancy-size free volumes at the
interfaces) with a positron trapping rate V,S C V,S , which is
independent of temperature and is described [79] in the form:
V, SC V, S = 6 / D , where C V, S is the concentration of interfacial
vacancies, D is the grain size, and is the trapping coefficient
of grain boundaries. If = 4.210 2 m s 1 , as determined in [80],
and D =110 nm from X-ray diffraction analysis, then the trapping
rate V,S C V,S = 2.310 10 s 1 is obtained.
(2) Nanovoids with a trapping rate void C void = (1 + (T
293))( void C void ) T=293 K , which increases linearly with temperature.
252

Microstructure of Compacted and Bulk Nanocrystalline Materials

The value of ( void C void )T = 293 K can be determined from the intensity
I 2 void C void
ratio I = C
by substituting the values I 1 and I 2, measured
1
V, S V, S

T = 293 K, and the value of V, SC V, S calculated above. As a result,


a trapping rate ( void C void )T = 293 K = 0.810 10 s 1 is obtained.
(3) Thermodynamically equilibrium thermal vacancies with a
V, T
specific
trapping
rate
and
a
concentration
F
F
F
F

C V, T = exp( S V,
/
k
)
exp(
H
/
k
T
)
,
where
H
and
S V,
T
B
V, T
B
V, T
T are the
effective vacancy formation enthalpy and entropy, respectively, and
k B is the Boltzmann constant.
In a three-state transition-limited model without detrapping, the
mean positron lifetime may be represented as

V, T V, T C V, T + V, S V, SC V, S + void void C void


V, T C V, T + V, SC V, S + void C void

(6.16)

Dependence (6.16) with the above values of V, SC V, S and


( void C void )T =293 K was used to fit the temperature dependence of the
mean positron lifetime after annealing a specimen at 1023 K. In
approximation, the positron lifetimes V, T = V, S = 1 = 174 ps and
void = 2 = 370 ps, determined by the two-component analysis of
the spectrum obtained at room temperature (Table 6.2), were kept
constant. The free parameters of the fit are the temperature
F
coefficient tr , combined preexponential factor V, T exp( S V,
T / kB ) ,
F
and vacancy formation enthalpy H V, T . Good agreement between
the modeling curve and the experimental data (Fig. 6.7) is ensured
by the temperature coefficient tr of the positron trapping rate by
nanovoids, which describes the low temperature part of the curve.
This coefficient tr is equal to (1.0 0.1)10 3 . This value must be
compared with tr 810 3 , determined for neutron irradiated
molybdenum with nanovoids of the size of 2.6 nm [81]. Taking into
account the fact that in the present case the nanovoid diameter is
0.50.8 nm. Since the coefficient tr depends linearly with the
square of the void diameter [82], the two results are in good
agreement and confirm that the temperature dependence of a
trapping rate, void , is the principal reason for the observed linear
increase of in the temperature range from 300 to 600 K.
Positron trapping in the nanovoids restricts the positron trapping in
253

Nanocrystalline Materials

thermal vacancies, and in (Fe 3Si) 95 Nb 5 this takes place up to higher


temperatures than in neutron-irradiated molybdenum, as a result of
a higher concentration of the nanovoids in (Fe 3 Si) 95 Nb 5 alloy. In
other words, the positron trapping in nanovoids remains transitionlimited.
Table 6.3 compares these vacancy formation parameters,
obtained by approximation, with those of coarse-grained Fe 79 Si 21
with D0 3 superstructure, Fe 75 Si 25 [77] and Fe 76 Si 24 [38, 68]. The
F
vacancy formation enthalpy H V,
T = 1.1 eV for nanocrystalline
(Fe 3Si) 95 Nb 5 coincides with that of other silicides, enriched in iron.
This coincidence confirms the validity of the used model of
describing temperature dependence (T ) . This model predicts that
the mean positron lifetime initially increases and then decreases
with increasing temperature towards the limiting value V, T when
positron trapping at thermal vacancies becomes the completely
dominant process. The inset in Fig. 6.7 shows the modeling curve
(T ) extrapolated to higher temperatures: the dependence (T )
should reach the limiting value = V, T for T 1400 K. This
temperature is close to the melting point T melt = 1500 K of iron
silicide. However, in this temperature range the grain growth should
be so fast that it is not possible to verify this assumption for the
nanocrystalline state.
In order to determine the main reason for ordering of
(Fe 3 Si) 95 Nb 5 , let us consider the process of atom diffusion.
Complete ordering of the absolutely disordered state requires mutual
rearrangement of at least half of the atoms of silicon and iron in
the crystallite. The jump of a single vacancy takes place more
often than the jump of a single atom. The maximum distance to sink
and source of atoms or vacancies in nanocrystals is very short and,
F
Vacancy formation enthalpy H V,
T and preexponential factor
/ k B ) in nanocrystalline n-(Fe 3 Si) 95 Nb 5 [67] and in coarse-grained or
monocrystalline Fe 75 Si 25 , Fe 76 Si 24 and Fe 79 Si 21

Table 6.3

F
V, T exp(S V,
T

Ma te ria l
n- (F e 3S i)95N b 5
F e 76S i24[8 3 , 8 4 ]
F e 75S i25[7 7 ]
F e 79S i21[7 7 ]

HFV.T
( e V)

V.T e xp (S FV.T/k B)

1.10.2
1.060.04
0.770.08
1.10.06

0.5
2.2*
0.5
4.4

(1 0 16 s 1)

* V.T= 6 . 6 1 0 14 s1 a nd S FV.T= 3 . 5 k B ta k e n fro m Re fs. [8 3 , 8 4 ] re sp e c tive ly


254

Microstructure of Compacted and Bulk Nanocrystalline Materials

consequently, only a small part of the atoms should move in order


to reach the equilibrium concentration of thermal vacancies (for
example, interfaces are a sink of vacancies in nanocrystals).
Therefore, the equilibrium diffusion properties must be extrapolated
to a very short diffusion length. It may be expected that ordering
in the nanocrystallites is completed when all atoms make a few
jumps over a diffusion length of about 1 nm. The evaluated
diffusion lengths of either Fe or Si at 650 K (the mean temperature
of the ordering process) in the absolutely disordered alloy with the
structure A2 is slightly longer than the ordering length equal to 1
nm (Table 6.4).
Activation enthalpy Q for D0 3 short-range ordering in nanocrystalline (Fe 3 Si) 95 Nb 5 is equal to 2.7 eV [78]. This value is higher
than the activation enthalpy estimated for the disordered state with
a structure A2 (see Table 6.4). In strongly deformed Fe 3 Si, the
activation enthalpy of short range ordering is 2.0 eV [68]. Partial
ordering of B2 type (see Fig. 6.6) or the presence of niobium in
grains may be the reason for the fact that the ordering rate in nano(Fe 3 Si) 95 Nb 5 is lower than expected. The diffusion length of Fe
atoms in the D0 3 -ordered state is considerably longer than that of
Si atoms, which is less than 1 nm (see Table 6.4). This indicates
that the ordering of partially ordered superstructure D0 3 is
controlled by the diffusion of Si rather than by the diffusion of Fe.
In other words, the movement of Si atoms is a limiting factor of
ordering. Assuming that the frequency factor D 0 of Si diffusion in
D0 3 -Fe 3 Si is equal to 0.19 m 2 s 1 (see Table 6.4) and the activation
enthalpy, measured for D0 3 short-range ordering, is equal 2.7 eV,
Table 6.4 Diffusion activation enthalpies Q, frequency factors D 0 and diffusion
length 4Dt an calculated for an annealing time t an = 2 h in different FeSi materials
[67]
F e S i ma te ria ls
59

F e in D0 3- F e 76S i24 [8 4 ]
Ge in D0 3- F e 76S i24 * [8 5 ]
59
F e in A 2 - F e 76S i24 * *
S i in A 2 - F e 76S i24* * *
71

Q
(eV)

D0
(m2 s1)

4Dt an at 650 K

1.64 0.04
3.23 0.04
2.2
2.2

11..33++000.5..540.410
1044
1011
1.9++000...9960.610
8104
3103

850
0.02
14
28

(nm)

* Diffusio n le ngth o f S i is c lo se to tha t o f Ge


* * Va lue s Q a nd D0 a re o b ta ine d b y e xtra p o la tio n o f d iffusio n o f iro n a to ms in F e 1- xS ix with A 2 struc ture a t lo w
c o nte nt o f S i up to x = 0 . 1 Ra tio Q/T melt = 0 . 0 0 1 5 e V K 1 is d e te rmine d fro m F ig. 1 4 in Re f. [8 5 ] using the
me lting te mp e ra ture T melt = 1 5 0 0 K o f F e 76S i24. At a te mp e ra ture lo we r tha n C urie te mp e ra ture the a c tiva tio n
e ntha lp y Q a nd fre q ue nc y fa c to r D0 sho uld b e to so me e xte nt la rge r.
* * * Ap p a re ntly, d iffusio n c o e ffic ie nt o f S i a to ms in a d iso rd e re d F e 76S i24 is 2 5 time s la rge r tha n tha t fo r F e
a to ms

255

Nanocrystalline Materials

it can be calculated that the diffusion length of Si atoms is 2.5 nm


at 650 K. This value is very close to the ordering length. In
conclusion it should be noted that the model, in which ordering is
controlled by the movement of Si atoms and by the formation of
thermal vacancies, is in good agreement with the diffusion data for
coarse-grained iron silicide Fe 3 Si.
Thus, in [67] the method of electronpositron annihilation was
used to study the ordering and formation of vacancies in
nanostructured (Fe 3 Si) 95 Nb 5 . The states from the as-milled
disordered to completely D0 3-ordered with a crystallite size of about
100 nm were investigated. The superstructure D0 3 forms after
annealing for 2 hours at 713 K. The kinetics of the ordering is
controlled by the diffusion mobility of both elements Si and Fe in
the disordered phase, but the diffusion mobility of silicon is smaller
than that of iron. The presence of niobium slows down the ordering
process in comparison with the kinetics of ordering in pure Fe 3 Si.
The activation enthalpy, which is equal to 2.7 eV [78] and is higher
than the activation energy for disordered Fe 3 Si, makes it possible
to explain the experimental results of [67]. Structural ordering leads
to thermal stability of the nanomaterial in which rapid self-diffusion
of iron takes place as a result of a small vacancy formation
enthalpy. In particular, by this relationship Fe 3 Si differs from
compound FeAl with B2 structure in which a small vacancy
formation enthalpy is not accompanied by the jumps of Fe atoms
to the closest sites of the crystal lattice.
The formation of thermodynamically equilibrium vacancies in
nanocrystallites of nanostructured (Fe 3Si) 95 Nb 5 compound has been
confirmed by the reproducible equilibrium temperature dependence
of the mean positron lifetime (T ) . The competing positron trapping
at nanovoids and thermal vacancies results in a maximum of the
dependence (T ) in the vicinity of 800 K. This behaviour is caused
by the unique combination of the high thermal stability of
nanocrystalline state in respect to the grain growth, which ensures
by the segregation of Nb atoms at interfaces, and by a low vacancy
formation enthalpy. The data for the positron lifetime are described
efficiently by the three-state trapping model with a low vacancy
F
F
formation enthalpy H V,
T = (1.10.2) eV; this value of H V, T
completely corresponds to the results obtained for coarse-grained
Fe 3 Si.
Thus, the increased thermal stability of the nanocrystalline state
in relation to grain growth is associated with the segregation of
niobium atoms at interfaces of the nanograins and with atomic
256

Microstructure of Compacted and Bulk Nanocrystalline Materials

ordering. Because of low diffusion mobility, stabilisers for other


nanostructured materials should include either transition metals of
group IV (Ti, Zr) or group V (V, Ti), or their carbides, nitrides and
oxides. It should be mentioned that such a phenomenon of structural
modification as atomic ordering, controlled by diffusion, was found
in nanocrystalline vanadium carbide. In future, the understanding of
atomic mechanisms, determining the processes of grain growth and
atomic ordering will lead to goal-oriented synthesis of nanomaterials
with the required microstructure.
Recently, a new experimental method, the time-differential length
change, has been proposed for studying point thermal defects [86].
The method is based on contact-free measurement of the change
in the length of the investigated specimen as a function of time t
at a sudden change of temperature T. The cylindrical specimen with
a ring-shaped ledge in the centre is spark cut from the ingot. One
end of the specimen is hardly fixed in a cooled specimen holder.
The ring-shaped ledge in the form of a step is located at a distance
l = 20 mm from the outer end of the specimen (Fig. 6.8).
The surface of the ledge is parallel to the surface of the outer
end of the specimen, and both surfaces are polished. The change
in specimen length l as a function of time t is measured in the
experiment. The change of length is recorded by a two-beam
Michelson laser interferometer, which radiation is reflected from
parallel polished surfaces of the outer end and the ledge.

20 mm

40 mm

specimen
laserinterferometer

cooled
specimen holder

Fig. 6.8. Sketch of the experimental setup. The length of the specimen, which is
suspended together with the furnace in a vacuum recipient, is measured by the
two laser beams on the polished parallel planes on the specimen front and ledge
[86].

257

Nanocrystalline Materials

Determination by this method of the vacancy formation enthalpy, the


activation enthalpy and the migration enthalpy of vacancies in
Fe 55 Al 45 showed that the results are in good agreement with the
values obtained previously by the positron lifetime spectroscopy
[87]. In [88] it was reported that the method described in [86] is
a methodological innovation in studying thermal vacancies.
On the whole, examination of positron annihilation in
nanocrystalline compacted metals and alloys gave the following
results:
1. The positron lifetime in nanocrystalline metals is longer than
the lifetime f of free positrons.
2. The fraction of positrons, trapped by the vacancies, increases
with increasing compaction pressure applied during preparation of
the nanocrystalline solids; this means that an increase of the
compaction pressure increases the interfacial area.
3. The positrons are trapped by monovacancies, nanovoids
(vacancy complexes) and also nanopores whose size is close to that
of the missing crystallite.
4. The free vacancy-like volumes, trapping the positrons at low
temperatures, belong to the interfaces and not the crystallites.
5. The positron trapping by dislocations of the crystallites is
unlikely to take place because the plastic deformation of metals
leads to a smaller change in the positron lifetime than the
preparation of nanocrystalline metals by compaction of nanopowders.
6.3 STRUCTURAL FEATURES OF SUBMICROCRYSTALLINE
METALS PREPARED BY A SEVERE PLASTIC DEFORMATION
At present it is clear that the model of the gas-like structure does
not describe to the actual structure of the interfaces in
nanocrystalline materials. An alternative of gas-like structure model
is the assumption on the non-equilibrium interfaces having high
energy because of the presence of dislocations directly at the
interfaces and non-compensated disclinations at triple points. The
long-range stress field of non-equilibrium interfaces is characterised
by a strain tensor whose components inside the grain are
proportional to r 1/2 (r is the distance to the grain boundary).
Consequently, the stress field leads to the formation of elastic
distortions of the crystal lattice, with the maximum distortions near
the interfaces. This model is proposed in [8994] when studying the

258

Microstructure of Compacted and Bulk Nanocrystalline Materials

submicrocrystalline materials produced by different methods of


severe plastic deformation.
The results obtained by electron microscopy show that the main
special feature of the structure of submicrocrystalline materials is
the presence of randomly misoriented grain boundaries. These grain
boundaries are in non-equilibrium state. The as-prepared submicrocrystalline metals and alloys before annealing are characterised by
the presence of extinction contours along grain boundaries.
Extinction contours indicates high elastic stresses [89, 90, 9599].
Since the dislocation density inside the grains is considerably lower
than at the interfaces, the non-equilibrium interfaces become the
main sources of elastic stresses. After annealing, many grains
become completely free from dislocations, the extinction contours
disappear and a banded contrast, typical of the equilibrium state,
appears at the grain boundaries. The latter indicates that the
relaxation of these boundaries have taken place.
As an example, Fig. 6.9 shows the microstructure of submicrocrystalline palladium, produced by severe plastic deformation with
using torsion under quasi-hydrostatic pressure, and its variation
after annealing for 1 hour at different temperatures [98]. Plastic
deformation of the starting coarse-grained palladium was affected
at room temperature in air. As a result of severe rotation of Pd,
the true logarithmic degree of deformation e was 7.0. After severe

Fig. 6.9. Microstructure of submicrocrystalline palladium Pd after annealing for


one hour at different temperatures [98]: (a) 475 K, (b) 505 K, (c) 535 K, (d) 575
K, (e) 795 K, (f) 855 K.

259

Nanocrystalline Materials

plastic deformation, the specimens acquired a finely divided


dislocation-tangle-laden structure with misoriented fragments of
mean size of about 150 nm. The volume density of dislocations at
the boundaries of individual grains was as high as V b =
410 11 cm 2 .
After annealing at 475 K, a submicrocrystalline structure arises
in the specimen, which has grains of a mean size of about 200 nm
(Fig. 6.9a). From approximately 4060 % of the grains lattice
dislocations were removed, while the dislocation density within the
rest of grains reached 2.510 10 cm 2 . The intragranular bending
extinction contours and the boundary diffusion contrast indicate that
the grain boundaries are out of equilibrium; the volume density of
grain boundary dislocations V b is equal to 1.1010 11 cm 2 . After
annealing at 505 K (Fig. 6.9b) the mean grain size increased to
300 nm, and the fraction of grains free from lattice dislocations
increases by a large amount and ranges up to about 60 or 70 %.
Annealing at 535 K (Fig. 6.9c) leads to non-uniform grain growth.
The grain size distribution becomes bimodal with the mean grain size
of 850 nm. About 30 to 40 % of grains have non-equilibrium
boundaries and a mean size of 350 nm. The rest of the grains are
approximately 1100 nm in size and nearly half of them are
characterised by a banded contrast common for the equilibrium
state. The presence of the contrast indicates relaxation of these
boundaries. The dislocation density V b in the non-equilibrium
boundaries is 610 8 cm 2 . After annealing at 575 K (Fig. 6.9d), the
grain size increases to 2.2 m. As previously, dislocations are
observed at the grain boundaries. In individual grains (their fraction
is about 1020 % of all grains) the lattice dislocation density
amounts to (0.31.5)10 9 cm 2 , while in the remaining grains V b
is low and does not exceed 210 7 cm 2 . This suggests that a large
number of grains have already been cleaned to remove dislocations.
Annealing at 795 K (Fig. 6.9e) results in an increase of the grain
size to 4.5 m. The interior of the grains is virtually free from
dislocations. Approximately half of the grain boundaries remain in
non-equilibrium state with a dislocation density of V b
410 9 cm 2 . After annealing at 855 K (Fig. 6.9f ) the grain size
amounts to as much as 9 m and half of the grains do not contain
any dislocations. However, the situation reverses dramatically for
the remaining grains: they contain tangles of lattice dislocations with
a density of up to 310 9 cm 2 . These dislocations form a cellular
structure with a cell size from 0.5 to 2.0 m. As a result of further
annealing, grain growth continues and defects disappear completely
260

Microstructure of Compacted and Bulk Nanocrystalline Materials

from the interior of the grains. After annealing at 1075 K the mean
grain size is about 20 m.
The results of investigation of the microstructure of
submicrocrystalline Pd [98] show that annealing is accompanied by
a two-stage increase in the grain size. The first substantial increase
in the grain size occurs at a temperature slightly above 475 K. The
second stage of grain growth starts at temperature above 795 K,
i.e. at a temperature higher than the secondary (collective)
recrystallisation temperature.
The experimentally observed local distortions of the crystallite
lattice near their boundaries confirm the presence of elastic
stresses at the interfaces [17]. Examination of submicrocrystalline
Fe by Mssbauer spectroscopy showed [10] that experimental
spectrum is a superposition of two spectra corresponding to two
different states of the iron atoms. One of them (the state of the
Fe atoms in crystallites) coincides with the state of the iron atoms
in coarse-grained -Fe. The second component of the experimental
spectrum reflects the special state of the iron atoms at the
interfaces, although the crystalline structure of the grains and their
boundaries was identical. According to [89, 90], the difference in
the parameters of the superfine structure of the Mssbauer spectra
of submicrocrystalline iron is caused by higher dynamic mobility of
the interfacial atoms.
The results obtained by electronpositron annihilation [62] (see
Table 6.1) point to some similarity of the microstructure of
nanocrystalline and submicrocrystalline materials; in particular, these
materials contain free volumes of the same type.
In addition to relaxation of the interfaces, the annealing of
submicrocrystalline materials is accompanied by grain growth, and
an abrupt jump-like change in the properties of submicrocrystalline
metals is observed after annealing at the same temperature when
grain growth starts. This is confirmed by the results of investigation
of the evolution of the microstructure and properties of
submicrocrystalline Ni and Ni 3 Al alloy [101]. Submicrocrystalline
nickel with a mean grain size of approximately 100 nm was
produced by means of severe plastic deformation carried out by
torsion under high pressure. An increase of annealing temperature
of submicrocrystalline Ni to 450 K was accompanied by a slow
decrease of residual electrical resistance 4.2 and microhardness H V
at an almost unchanged grain size (Fig. 6.10).
Annealing at 500525 K led to a rapid decrease of electrical
resistance and microhardness due to the beginning of rapid grain
261










 






+'

 

QP



'!

 

1L

*UDLQ VL]H 



  FP

+DUGQHVV

+'

*3D

Nanocrystalline Materials

'!













Fig. 6.10.Evolution of residual electrical resistance 4.2, microhardness H V and grain


size D in submicrocrystalline nickel Ni as a function of annealing temperature
T [101].

growth. With a further increase of annealing temperature the slow


growth of the grains of submicrocrystalline Ni continues and there
are small changes of electrical resistance and microhardness.
According to [101], the changes of electrical resistance and
microhardness in relation to annealing temperature are associated
directly with an increase in the grain size and are correlated only
slightly with internal stresses.
A model of the grain boundaries [9194, 102104], which takes
into account the presence of dislocations and disclinations, makes
it possible to evaluate quantitatively the magnitude of developed
stresses and the excess energy of the interfaces, and also crystal
volume change caused by the excess elastic energy. According to
[92], an interface represents a disordered dislocation networks, and
the rms strains disl , formed at an interface and related to the unit
area of the interface, is equal:

disl 0.23b[( /D)log(R/2b)] 1/2 0.13b[ V log(R/2b)] 1/2 .

(6.17)

Excess energy ex, disl , caused by external dislocations and related


to the unit area of the interface, is

ex, disl = Gb 2 log(R/2b)/[4 (1 )] .


262

(6.18)

Microstructure of Compacted and Bulk Nanocrystalline Materials

In equations (6.17) and (6.18) b is the Burgers vector of


dislocations, D is the grain size in the nanocrystal, and
V 3 /D are the linear and volume density of dislocations, and
R is the grain size in a coarse-grained polycrystal, G is the shear
modulus, and is Poisson coefficient.
High internal stresses result in changes in the volume of the
submicrocrystalline material. The change of volume due to
disordered dislocation networks is
V/V 0.13b 2 V log(R/2b) .

(6.19)

Similar equations for a system of disclinations, formed at the


junction of several grains are derived in a two-dimensional model
of the polycrystal with square grains [102]:

discl 0.1 2 1/2 ,

(6.20)

ex,discl =[G 2 Dlog2]/[16 (1 )] ,

(6.21)

where 2 is the rms power of the disclinations.


According to [105], the power of junction disclinations is 12,
and 2 1/2 0.03. The volume dislocation density V for
submicrocrystalline materials is 310 15 m 2 [43]. Taking these
values into account, for submicrocrystalline aluminium Al with a
grain size of 10 nm the excess energies are ex,disl 0.3 J m 2 and
ex,discl 0.06 J m 2 . An increase in the volume, caused by the
presence of the dislocations, is V/V 410 4 . Since the change
in the volume is proportional to elastic energy and ex,discl is up to
five times smaller than ex,disl , then the increase of the volume
caused by the presence of disclinations, will be ~0.810 4. The total
increase in the volume of submicrocrystalline Al is V/V 4.8
10 4 . This is half of the experimental value V/V 910 4 [106].
Evidently, the additional increase in the volume is associated with
the formation of vacancies during deformation. For submicrocrystalline copper with a grain size D = 200 nm, V 310 15 m 2
and 2 1/2 0.03, estimates obtained using equations (6.17)(6.21)
give: disl 7.510 3 , discl 310 3 , total internal elastic strain
= ( 2disl + 2discl ) 1/2 810 3 , ex,disl 0.41 J m 2 , ex,discl
0.09 J m 2 , and total excess energy ex = ex,disl + ex,discl
0.5 J m 2 . The resultant value of internal elastic strain is in
satisfactory agreement with the magnitude of internal strains in
submicrocrystalline copper, which was determined by X-ray
263

Nanocrystalline Materials

diffraction [93].
The model [94, 102, 103] assumes an important role of
disclinations formed at triple junctions of the grains. Special
interests are splitting of grain boundary disclinations [103, 104]. The
splitting of disclinations at triple junctions decreases the elastic
energy of the system. This splitting of a disclination into smallpower disclinations may be of the volume type (Fig. 6.11b); in this
case, the local amorphisation of the triple junction region takes
place. Another variant is the splitting of a disclination into three
rows of small-power disclinations located in the boundaries of
adjacent grains (Fig. 6.11c). Line splitting of grain boundary
disclination to a number of small-power disclinations, located along
the grain boundary (Fig. 6.11e), is also possible. According to [103,
104], a decrease of the elastic energy of the initial grain boundary
disclination becomes greater with an increase of the number of new
disclinations formed. In the case of line splitting, the largest
decrease of the elastic energy is reached as a result of the
formation of two disclinations, located at the maximum permissible
distance from each other, i.e. at the distance equal to the grain
boundary length. The splitting of the disclinations in the interfaces
of the nanocrystalline materials is an effective channel of relaxation
of the elastic energy. Disclination splitting is accompanied by a
change in the structure of interfaces (appearance of stacking faults),

Fig. 6.11. Splitting of a triple-junction disclination (a) and of a grain boundary


disclination (d) [103]: (b) splitting of initial triple-junction disclination into the
circle ensemble of small-power disclinations and formation of local amorphous
region; (c) the initial triple-junction disclination splits into three rows of smallpower disclinations located in grain boundaries; (e) line splitting of initial grain
boundary disclination into the row of small-power disclinations

264

Microstructure of Compacted and Bulk Nanocrystalline Materials

decreases the probability of nucleation of microcracks in the vicinity


of the interface and stimulates grain boundary diffusion.
Thus, in addition to the small grain size and large area of the
interfaces, the presence of a long-range field of elastic stresses is
one of the main special features of nanocrystalline materials.
Main concepts of the microstructure of nanocrystalline materials
are based to a large extent on the results of X-ray investigation of
the lattice parameters, internal stresses and atomic displacements.
In comparison with the coarse-grained materials, X-ray diffraction
patterns of nanocrystalline materials are characterised by a larger
width of diffraction reflections, some changes of diffraction
reflection shape and also shift of diffraction reflections. Broadening
of the diffraction reflections is caused by the small grain size,
microdeformations, stacking faults of the crystal lattice, and
inhomogeneity (non-uniform composition) of substance (see Section
4.2) The shape and intensity of reflections depend on the atomic
displacements. Shift of the reflections indicates changes in the
lattice constant.
In the general case, scattering intensity of a crystal can be
presented [107110] as the sum
I(q) = I 0 (q)exp(2M) + I DAD (q)[1 exp(2M)] + I D (q),

(6.22)

where I 0 (q) is the intensity of structural reflections in the absence


of atomic displacements; q denotes the diffraction vector
(|q| q = (2sin)/); exp(2M) exp(2 iqu j )] is the DebyeWaller
factor allowing for attenuation of structural reflections as a result
of static and dynamic (thermal) atomic displacements. Second term
I DAD (q)[1 exp(2M)] in (6.22) denotes the intensity of diffuse
scattering due to displacement u j of atoms from crystal lattice sites.
I D(q) is the diffuse scattering intensity resulting from the difference
between atomic scattering factors and correlations in the mutual
arrangement of atoms, i.e. short-range order. In other words, I D (q)
is the intensity of diffusion scattering by the disordered crystalline
solid solution. According to [109, 110], the intensity I D(q) can be
presented as the sum of the intensity of Laue white noise and the
diffusion scattering intensity, which is due to short-range order.
Important information on special features of the structure is
provided by the diffuse scattering which is the sum of three
contributions: the diffusion scattering caused by atomic displacements (second term in equation (6.22)); diffusion scattering by the
disordered crystalline solid solution, I D(q); and the diffuse scattering
265

Nanocrystalline Materials

due to the disordered distribution of atoms in amorphous substance.


As a result of amorphisation, some of the atoms leave the crystal
lattice sites; it leads to decrease of the intensity of structural
reflections from I 0 (q) to [I 0 (q) ], and appearance of additional
contribution, which equals , into diffuse scattering. Atomic
displacements cause a monotonic increase in the intensity of the
diffuse scattering with increase in reflection angle , and the
absence of the order in the atom distribution leads to monotonic
decrease in the intensity of diffuse scattering. The diffuse
scattering, which is associated with a short-range of the type of
ordering or phase separation, modulates the Laue white noise, i.e.
results in periodic changes in diffuse scattering intensity [109, 110].
According to the experimental data, a decrease in the grain size
of nanocrystalline substances may result both in a decrease [111
114], and in an increase [114116] of lattice constants a, b, c and
the volume V of the unit cell of the crystal lattice. For example, for
cubic metal such as Cr and Pd, the unit cell volume increases with
decreasing of the mean grain size [115]. In nanocrystalline nc-Se,
a decrease in the grain size is accompanied by an increase of the
lattice constant a of the unit cell whereas the lattice constant c
remains unchanged; in accordance with this, the volume of the unit
cell of nc-Se increases from 0.0819 to 0.0823 nm 3 with a decrease
of the grain size from 70 to 12.5 nm [116]. The largest increase
of the lattice constant a and the unit cell volume V of nanocrystalline selenium is observed when the grain size D becomes
smaller than 15 nm. However, a decrease of the volume and lattice
constant (or constants) is more likely. This can be observed as a
result of compression (reduction) of crystallite when the crystallite
size is smaller than 10 nm. The observed increase of the unit cell
volume and the lattice constant is in all likelihood a consequence
of the adsorption and dissolution of impurities by the surface of
crystallites, as in the case of isolated nanoparticles (see Section
5.2).
The broadening of diffraction reflections is found for all
nanocrystalline materials. In Section 4.2 it is shown that broadening
s , associated with a small crystallite size D, is described by the
equation

s (2 ) 2 s ( ) = K hkl /(Dcos ) ,

(6.23)

where K hkl is the Scherrers constant whose value depends on the


shape of the particle (crystalline, domain) and on the Miller indices
266

Microstructure of Compacted and Bulk Nanocrystalline Materials

(hkl) of diffraction reflection. Deformation broadening d , caused


by stacking faults, is calculated from the equation:

d(2 ) = 4A 2 1/2 tg ,

(6.24)

where A is a constant approximately equal to unity in the case of


uniform distribution of dislocations in the crystallite. In equations
(6.23) and (6.24), broadening is expressed in radians.
Thanks to a different dependence of the size and deformation
broadening on the order of diffraction reflection it is possible to
separate these contributions using the pairs of reflections (hkl)
differing only in the order of reflection. In this case the total
diffraction reflection broadening is

0.5[ s + ( s2 + d2 ) 1/2 ] .

(6.25)

It is assumed that size broadening s does not depend on the


index l. The separation of the size and deformation broadening of
the diffraction reflections shows that the rms deformation in
nanocrystalline Al, Ru, Pd, Cu, AlRu [23, 112, 117, 118) is equal to
13 % and is considerably greater than that in coarse-grained
metals.
Analysis of diffraction measurements in order to determine the
size of crystallites and the level of deformation is described in detail
in Section 4.2. The results of X-ray diffraction investigations of
nanocrystalline materials are examined to some extent in [119].
Modeling of the X-ray diffraction pattern of nanocrystalline
materials [120123], which takes into account the grain size,
distortions of the crystal lattice, the thickness and structure of
interfaces, is of interest. Modeling was carried out using the
kinematic theory of X-ray scattering. A polycrystal containing 361
cubic crystallites was investigated; the length of the crystallite edge
was equal to Na (a is the constant of the unit cell). The size of
the crystallites varied by varying N. When calculating atomic
displacements it is assumed that all atoms of the external layer of
the crystallites are displaced from the positions of the ideal lattice
in a random manner and the displacements varies from 0 to 0.5b
where b = a / 2 is the Burgers vector.
Calculations carried out in [120, 121] show that atomic
displacements on the surface of crystallites lead only to a decrease
of the diffraction reflection intensity but have no effect on the
267

Nanocrystalline Materials

shape, width and position of diffraction reflections. A decrease in


the crystallite size results in a large broadening of diffraction
reflections. The effect of the long-range field of elastic stresses on
the parameters of diffraction reflection was modeled by changing
the linear dislocation density at the interfaces. An increase of
dislocation density from 0 to 0.1 and 1.0 nm 1 resulted in
broadening of the reflections and their displacements to the range
of high angles . At = 0.1 nm 1 the size broadening is dominant
when the crystallite size is smaller than 30 lattice constants of the
unit cell (D < 30a). For larger crystallites, the main contribution to
the diffraction reflection broadening comes from the elastic
distortions of the crystallite lattice, caused by dislocations at the
interfaces. From this, it follows that the effect of interfacial elastic
stresses on the microstructure of nanocrystalline materials
decreases with an decrease of the crystallite size.
6.4 NANOSTRUCTURE OF DISORDERED SYSTEMS
Investigations of glasses and amorphous metallic alloys, carried out
since 1985, show that the disordered materials possess a peculiar
nanostructure. The results of diffraction and electron microscopic
investigations [124134], discussed in previous sections, confirm the
nano-heterogeneous structure of amorphous alloys.
Similar conclusions on the nano-heterogeneous structure of
glasses and amorphous substances were made independently on the
basis of investigations of low-energy vibrational spectra and the
properties determined by the spectral distribution of elastic
vibrations. The results of these investigations will be examined
briefly.
The vibrational spectra of various disordered systems such as
glasses and amorphous solids greatly differ from those of normal
crystals. The density of vibrational states of crystals in the lowenergy range is described by the Debye law

2
g D ( ) = 9 N 3 ,
D

(6.26)

where D is the maximum frequency for the Debye function of the


frequency distribution.
In contrast to the crystals, at energies lower than 1 K the
spectra of glasses and amorphous substances are characterised by
268

Microstructure of Compacted and Bulk Nanocrystalline Materials

a constant density of vibrational states, and in the energy range


from 2 to 10 meV (from 1525 to 120125 K) there is an excess
density of vibrational states in comparison with Debye density
g D( ). This excess density of state is found in all glasses and is
manifested in low-energy spectra of inelastic neutron scattering,
low-frequency spectra of Raman scattering, in infra-red adsorption
spectra, and in low-temperature heat capacity and heat conductivity.
According to modeling assumptions [135139], vibrational
excitations, responsible for the excess density of states in
disordered substances, are localised in a region containing from
several tens to hundreds of atoms and having the size from 1 to
several nanometers. Thus, the low-energy special features of
phonon spectra of disordered materials indicate that the structure
of amorphous substances and glasses are characterised by a
presence of a regions with characteristic spatial scale of order of
several nanometers. In other words, amorphous substances and
glasses have a nanostructure.
Indeed, comparison and analysis of the vibrational spectra of
nanoparticles and the experimental results on the phonon spectra
and low-temperature heat capacity of nanoparticles and nanomaterials (see Section 5.3) with the observed special features of
low-energy spectra of disordered glassy substances allow us to note
the following. In the energy range lower than 15 meV the density
of vibrational states and the heat capacity of nanocrystalline and
disordered glassy substances are higher than these for the coarsegrained crystals. In the low-frequency region the vibrational spectra
of nanocrystalline and glassy materials differ mainly by the form of
the excess density of states. In the glasses the excess density of
vibrational states has the form of a peak (Fig. 6.12), whereas in
nanocrystalline materials with a particle size of about 10 nm it has
the form of a kink (see, for example, Fig. 5.12 for the density of
phonon states g( ) of n-Ni).
Low-energy quasilocal excitations in glassy materials and their
crystalline analogues As 2S 3 , SiO 2 and Mg 70 Zn 30 were studied by the
methods of inelastic incoherent neutrons scattering and Raman
scattering [139]. The inelastic neutron scattering provide direct
information on the density of vibrational states (see Fig. 6.12). The
excess density of vibrational states g( ) = g( ) g D ( ), where
g D ( ) is the Debye density of vibrational states, was determined
from the experimental values of the velocity of sound. For all three
disordered materials the excess density of states, g( ), has the
form of a peak whose maximum corresponds to some frequency
269

Nanocrystalline Materials






  DUELWUDU\ XQLWV












 PH9
Fig. 6.12. The density of vibrational states in (1) crystalline and (2) glassy materials
[139]: (a) As 2 S 3 , (b) SiO 2 , (c) Mg 70 Zn 30 .

max and energy E max and exceeds g D ( max ) by a factor of 26. For
the glasses of different chemical composition, with different type
of the short-range order and with different type of chemical bond
(planar structure in As 2 S 3 , covalent bonds in SiO 2 , dense packing
in Mg 70 Zn 30 metallic glasses) the dependences of the reduced
excess density of states g = g( )/g( max ) on the reduced
energy E/E max have the same shape. The same spectral shape of the
reduced excess density of states, g, indicates the universal nature
of the structural features responsible for the appearance of lowfrequency anomalies in vibrational spectra of different amorphous
materials.
270

Microstructure of Compacted and Bulk Nanocrystalline Materials

In the low-frequency range, the Raman scattering spectra of


glasses are characterised by the presence of a wide peak, which
is not found in the spectra of appropriate crystals. Analysis of the
spectra shows that the low-frequency peak is associated with the
first order light scattering at vibrational excitations, which governed
by Bose statistics [139]. The characteristic frequency of the
maximum of the boson peak is in the range of acoustic vibrations
and for different materials covers the range from 1/3 to 1/7 of the
Debye frequency D. This means that the characteristic length of
localisation of vibrational excitations, which scatter light with the
appearance of a boson peak, is equal to several interatomic
distances. The boson peak is a reflection of the excess density of
vibrational states in the Raman spectrum. In the reduced
coordinates, the spectral form of the boson peak for different
glasses, produced by melt cooling, is the same [140].
The best function, which describes the experimental spectra of
the excess density of vibrational states and the boson peak in
glasses and contains the least number of fitting parameters, is the
logarithmiconormal function,
ln 2 ( / max )
g ' ( ) = exp

2 2

(6.27)

In the reduced coordinates, function (6.27) is determined by only


one dimensionless parameter: the dispersion , which is equal to
the same number, 0.480.05 [141], for all low-molecular glasses.
Thus, a new universal parameter appears in the physics of glasses,
i.e. the distribution dispersion, , associated with the basic
characteristics of the nanostructure (note that the same
logarithmiconormal function (2.1) is also used to describe the size
distribution of nanocrystalline particles, produced by evaporation and
condensation).
This enabled the authors of [139] to link the excess density of
vibrational states with the presence in glasses of the regions having
the characteristic length (radius) with the nanometer scale. It is
assumed that low-energy vibrational excitations, responsible for the
excess density of states, are localised on nano-heterogeneities of
the structure. This is confirmed, in particular, by the results of lowfrequency Raman scattering study of glasses whose matrix contains
grown clusters of a different chemical composition with a size of
several nanometers [142, 143]. The authors of [142] have been
271

Nanocrystalline Materials

investigated photochromic glasses with SiO 2B 2O 3 matrix containing


clusters of silver halide. The cluster size depends on the annealing
time of specimens and varies in the range from 4 to 8 nm. The
Raman scattering spectra of the starting glass contain boson peak
typical of the spectra of all glassy materials (Fig. 6.13). Annealing
of specimen led to an appearance of a new peak only in the lowfrequency part of the spectrum: namely, if the glass matrix contains
clusters in the amount not smaller than 2 %, then after annealing
the spectrum of inelastic scattering contain an additional band (see
Fig. 6.13).
The difference of the spectra of the annealed and starting
specimens corresponds to the spectrum of surface vibrational modes
of clusters; it should be noted that the experimental Raman
scattering spectrum contained only main modes with the frequency

T = 0.8v t /D ,

(6.28)

,__ *,* > 3  @ DUELWUDU\ XQLWV

where v t is the transverse velocity of sound, D is the diameter of


a spherical cluster, T is the frequency of the main torsion mode.
According to the results of [143], the identical spectrum of glasses





FP



Fig. 6.13. Low-frequency Raman scattering spectra for photochromic glasses with
SiO 2 B 2 O 3 matrix containing clusters of silver halide [139]: (1) starting specimen,
(2) annealed specimen with clusters of silver halide, (3) calculated spectrum of
contribution of acoustic vibrational excitations.

272

Microstructure of Compacted and Bulk Nanocrystalline Materials

Al 2 O 3 MgO with MgCr 2 O 3 MgAl 2 O 3 clusters contained only


spherical vibrational modes with the frequency
S = 0.8v l /D ,

(6.29)

where v l is the longitudinal velocity of sound. The difference


between the first and second case is associated with the different
ratio of the elastic constants of the cluster and matrix: in
photochromic glasses the matrix, which is harder than the cluster,
suppresses spherical vibrations.
Analysis carried out in [144] shows that the low temperature
heat conductivity of glasses with clusters has a plateau whose
position correlates with the cluster size. According to [144], the
IoffeRegel criterion for the phonon localisation is fulfilled in the
region of the plateau, i.e. ~ l, where l is the free path length (in
the case of strong scattering it is determined by the size of the
heterogeneity of the structure), and is the phonon wavelength.
Comparing these data with the results of measurements of heat
conductivity in glasses, where localisation scale corresponds to the
correlation length of the structure, the authors of [144] found that
for glasses the correlation length is 13 nm.
Direct calculations of the excess low-energy density of
vibrational states in elastic medium with fluctuating elastic constants
was carried out by the authors of [145] in the framework of the
theory of perturbation with respect to small fluctuations. Calculation
showed that the elastic constant fluctuations, having a correlation
radius R c 12 nm, leads to the appearance of an excess density
of states in the low frequency ( ~ v/R c ) range. It can be shown
that any rational elastic constant correlation function, decreasing
with distance, results in the displacement of part of the highfrequency vibrational modes to the low-frequency part of the
spectrum, thus forming an excess density of the vibrational state.
As already mentioned, the spectrum of excess density of vibrational
states is efficiently approximated by the logarithmiconormal
function (6.27) with the dispersion = 0.48. If the excess density
of states is caused by vibrational excitations, localised on nanometer
heterogeneities of the structure, the frequency of quasilocal
vibrations, , is linked with the heterogeneity size, D, by the
relationship = Kv/D, where K is the constant of the order of
unity. This means that the size distribution of nano-heterogeneities
can also be described by the logarithmiconormal function similar
to (6.27), with the dispersion :
273

Nanocrystalline Materials

ln 2 ( D / D0 )
F ( D) = exp
.
4 2

(6.30)

In (6.30) the value D 0 is the most probable size of the nanoheterogeneity.


The low-energy features of vibrational spectra of glasses,
associated with the presence of nano-heterogeneities, may have a
strong effect on the properties of glasses not only at low but also
high temperatures, up to the solidification temperature of glass.
These properties include those for which the effect of the lowenergy density of vibrational states is enhanced in comparison with
the region of the spectrum near the Debye frequency. For example,
the contribution of low-energy phonons to the magnitude of rms
thermal vibrations of the atoms is increased in proportion to the
inverse square of the frequency of vibrations. Consequently, as
shown in [146], the presence in glasses of an excess density of
vibrational states, which is equal to ~10 %, increases the amplitude
of thermal vibrations of the atoms by 3040 % in comparison with
the crystalline material having the same temperature.
The proposed [139] universal form of the low-energy vibrational
spectrum of glasses, glassy and amorphous substances indicates that
structure of these substances contains heterogeneities of the
nanometer size. In this chaos and disorder with which the structure
of amorphous materials and glasses is usually associated, there is
a universal spatial scale typical of glasses of different nature
(dielectric, semiconductor, metallic). According to [139], the
presence of nanoregions in the disordered materials may have the
same important role for the theory of glassy and liquid states as
the presence of the unit cell for the theory of the crystal structure.

274

Microstructure of Compacted and Bulk Nanocrystalline Materials

References
1.

2.
3.

4.
5.
6.

7.
8.
9.
10.

11.

12.

13.

14.
15.
16.

17.
18.

A. I. Gusev. Effects of the nanocrystalline state in solids. Uspekhi Fiz.


Nauk 168, 55-83 (1998) (in Russian). (Engl. transl.: Physics - Uspekhi 41,
49-76 (1998))
A. I. Gusev. Nanocrystalline Materials:Preparation and Properties (Ural
Division of the Russ. Acad. Sci., Yekaterinburg 1998) 200 pp. (in Russian)
H. Gleiter. Materials with ultra-fine grain size. In: Deformation of Polycrystals:
Mechanisms and Microstructures. Eds. N. Hansen, A. Horsewell, T. Leffers
and H. Lilholt (Ris Nat. Laboratory, Roskilde, Denmark 1981) pp.15-21
H. Gleiter, P. Marquardt. Nanocrystalline structures on approach to new
materials? Z. Metallkunde 75, 263-267 (1984)
R. Birringer, U. Herr, H. Gleiter. Nanocrystalline materials: a first report.
Trans. Japan. Inst. Met. Suppl. 27, 43-52 (1986)
R. W. Siegel, H. Hahn. Nanophase materials. In: Current Trends in Physics of Materials. Ed. M. Yussouff. (World Scientific Publ. Comp., Singapore 1987) pp.403-420
H. Gleiter. Nanocrystalline materials. Progr. Mater. Sci. 33, 223-315 (1989)
M. D. Matthews, A. Pechenik. Rapid hot-pressing of ultrafine PSZ powders.
J. Amer. Ceram. Soc. 74, 1547-1553 (1991)
D.-J. Chen, M. J. Mayo. Densification and grain growth of ultrafine 3 mol %
Y 2 O 3 -ZrO 2 ceramics. Nanostruct. Mater. 2, 469-478 (1993)
G. V. Ivanov, N. A. Yavorovskii, Yu. A. Kotov, V. I. Davydovich, G. A.
Melnikova. Self-propagating process of sintering of ultradispersed metallic powders. Doklady AN SSSR 275, 873-875 (1984) (in Russian)
V. V. Ivanov, S. N. Paranin, E. A. Gavrilin, A. V. Petrichenko, Yu. A. Kotov,
S. A. Lebedev, S. M. Cheshnitskii. Production of high-current superconducting
ceramics Bi 1.6 Pb 0.4 Sr 2 Ca 2 Cu 3 O 10 by magnetic pulsed pressing.
Sverkhprovodimost: Fizika, Khimiya, Tekhnologiya 5, 1112-1115 (1992)
(in Russian)
V. V. Ivanov, Yu. A. Kotov, O. M. Samatov, R. Bhme, H. U. Karow, G.
Schumacher. Synthesis and dynamic compaction of ceramic nanopowders
by techniques based on electric pulsed power. Nanostruct. Mater. 6, 287290 (1995)
V. V. Ivanov, A. N. Vikhrev, A. A. Nozdrin. Pressing of nanosized powders
Al 2 O 3 at magnetic-pulsed loadng. Fizika i Khimiya Obrabotki Mater. No
3, 67-71 (1997) (in Russian)
H. Gleiter. Nanostructured materials: state of art and perspectives. Nanostruct.
Mater. 6, 3-14 (1995)
H. Gleiter. Materials with ultrafine microstructure: retrospectives and perspectives. Nanostruct. Mater. 1, 1-19 (1992)
H. Gleiter. Nanostructured materials. In: Mechanical Properties and Deformation Behavior of Materials Having Ultrafine Microstructure. Eds. M. A.
Nastasi, D. M. Parkin, H. Gleiter. (Kluwer Academic Press, Netherlands,
Dordrecht 1993) pp.3-35
W. Wunderlich, Y. Ishida, R. Maurer. HREM-studies of the microstructure
of nanocrystalline palladium. Scripta Metall. Mater. 24, 403-408 (1990)
G. J. Thomas, R. W. Siegel, J. A. Eastmen. Grain boundaries in nanocrystalline
Pd: high resolution electron microscopy and image simulation. Scripta Metall.
Mater. 24, 201-206 (1990)

275

Nanocrystalline Materials
19.
20.
21.

22.

23.

24.

25.

26.

27.
28.

29.

30.

31.
32.

33.

34.

35.

36.

T. Mtschele, R. Kirchheim. Segregation and diffusion of hydrogen in grain


boundaries of palladium. Scripta Met. 21, 135-140 (1987)
H.-E. Schaefer, H. Kisker, H. Kronmller, R. Wrschum. Magnetic properties of nanocrystalline nickel. Nanostruct. Mater. 1, 523-529 (1992)
R. Birringer, H. Gleiter. Nanocrystalline materials. In: Encyclopedia of Material
Science and Engineering. Suppl. Vol.1. Ed. R. W. Cahn. (Pergamon Press,
Oxford 1988) pp.339-349
X. Zhu, R. Birringer, U. Herr, H. Gleiter. X-ray diffraction studies of the
structure of nanometer-sized crystalline materials. Phys. Rev. B 35, 90859090 (1987)
M. Fitzsimmons, J. A. Eastman, M. Mller-Stach, G. Wallner. Structural
characterization of nanometer-sized crystalline Pd by x-ray-diffraction techniques. Phys. Rev. B 44, 2452-2460 (1991)
J. A. Eastman, M. R. Fitzsimmons, M. Mller-Stach, G. Wallner, W. T. Elam.
Characterization of nanocrystalline Pd by x-ray diffraction and EXAFS.
Nanostruct. Mater. 1, 47-52 (1992)
J. Lffler, J. Weissmller, H. Gleiter. Characterization of nanocrystalline
palladium by x-ray atomic density distribution functions. Nanostruct. Mater.
6, 567-570 (1995)
J. Weissmller, J. Lffler, M. Kleber. Atomic structure of nanocrystalline
metals studied by diffraction techniques and EXAFS. Nanostruct. Mater.
6, 105-114 (1995)
T. Haubold, R. Birringer, B. Lengeler, H. Gleiter. X-ray studies of nanocrystalline
Pd. Phys. Lett. A. 135, 461-466 (1989)
Y. Ishida, H. Ichinose, T. Kizuka, K. Suenaga. High-resolution electron
microscopy of interfaces in nanocrystalline materials. Nanostruct. Mater.
6, 115-124 (1995)
Yu. A. Babanov, L. A. Blaginina, I. V. Golovshchikova, T. Haubold, F. Boscherini,
S. Mobilio. Defects in nanocrystalline palladium. Fiz. Metal. Metalloved.
83, No 4, 167-176 (1997) (in Russian). (Engl. transl.: Phys. Metal. Metallogr.
83, 444-451 (1997))
N. Schlorke, J. Weissmller, W. Dickenscheid, H. Gleiter. In vacuo x-ray
diffraction study of atomic short-range order in inert-gas-condensed Fe.
Nanostruct. Mater. 6, 593-596 (1995)
Yu. A. Babanov, I. V. Golovshchikova, F. Boscherini, S. Mobilio. Short range
order in nanocrystalline cobalt. Nanostruct. Mater. 6, 601-604 (1995)
H.-E. Schaefer. Interfaces and physical properties of nanostructured solids. In: Mechanical Properties and Deformation Behavior of Materials Having
Ultrafine Microstructure. Eds. M. A. Nastasi, D. M. Parkin, H. Gleiter. (Kluwer
Academic Press, Netherlands, Dordrecht 1993) pp.81-106
E. Jorra, H. Franz, J. Peisl, W. Petry, G. Wallner, R. Birringer, H. Gleiter,
T. Haubold. Small-angle neutron scattering from nanocrystalline Pd. Philosoph.
Mag. B 60, 159-168 (1989)
P. Sanders, J. R. Weertman, J. G. Barker, R. W. Siegel. Small angle neutron
scattering from nanocrystalline palladium as a function of annealing. Scripta
Met. 29, 91-96 (1993)
R. Wrschum, W. Greiner, H.-E. Schaefer. Preparation and positron lifetime spectroscopy of nanocrystalline metals. Nanostruct. Mater. 2, 55-62
(1993)
E. Browne, J. M. Dairiki, R. E. Doebler. Table of isotopes / 7 th edition. Eds.
C. M. Lederer, V. S. Shirley. (John Wiley & Sons, New York 1978) 1523

276

Microstructure of Compacted and Bulk Nanocrystalline Materials

37.

38.
39.

40.

41.

42.
43.

44.

45.

46.

47.

48.

49.

50.

51.

52.

pp.
R. M. Nieminen, M. J. Manninen. Positrons in imperfect solids: theory.
In: Positron in Solids. Ed. P. Hautojrvi. (Springer, Berlin 1979) pp.145195
H.-E. Schaefer. Investigation of thermal equilibrium vacancies in metals by
positron annihilation. Phys. Stat. Sol. (a) 102, 47-65 (1987)
A. A. Rempel, M. Forster, H.-E. Schaefer. Positron lifetime in non-stoichiometric carbides with a B1 (NaCl) structure. J. Phys.: Condens. Matter 5, 261-266 (1993)
J. Strmer, A. Goodyear, W. Anwand, G. Brauer, P. G. Coleman, W.
Tr i f t s h u s e r. Silicon carbide: a new positron moderator. J.Phys.: Condens.
Matter 8, L89-L94 (1996)
G. Brauer, W. Anwand, E.-M. Nicht, J. Kuriplach, M. ob, N. Wagner, P. G.
Coleman, M. J. Puska, T. Korhonen. Evaluation of some basic positronrelated characteristics of SiC. Phys. Rev. B 54, 2512-2517 (1996)
A. Seeger, F. Banhart. On the systematics of positron lifetimes in metals.
Phys. Stat. Sol. (a) 102, 171-179 (1987)
H.-E. Schaefer, M. Forster. As-grown metal oxides and electron-irradiated
Al 2 O 3 studied by positron lifetime measurements. Mater. Sci. Engineer. A
109, 161-167 (1989)
R. Wrschum, W. Bauer, K. Maier, A. Seeger, H.-E. Schaefer. Defects in
semiconductors after electron-irradiation or in high-temperature thermal
equilibrium as studied by positron annihilation. J. Phys.: Condens. Matter 1, Suppl.A, SA33-SA48 (1989)
R. Wrschum, H.-E. Schaefer. Interfacial free volumes and atomic diffusion in nano-structured solids. In: Nanomaterials: Synthesis, Properties, and
Applications. Eds. A. S. Edelstein and R. C. Cammarata. (Institute of Physics,
Bristol 1996) pp.277-301
R. Wrschum, K. Reimann, S. Gru, A. Kbler, P. Scharwaechter, W. Frank,
O. Kruse, H. D. Carstanjen, H.-E. Schaefer. Structure and diffusional properties
of nanocrystalline Pd. Philosoph. Mag. B 17, 407-417 (1997)
H.-E. Schaefer, R. Wrschum, R. Schwarz, D. Slobodin, S. Wagner. Amorphous
hydrogenated Si studied by positron lifetime spectroscopy. Appl. Phys.
A 40, 145-149 (1986)
R. Wrschum, P. Farber, R. Dittmar, P. Scharwaechter, W. Frank, H.-E. Schaefer.
Thermal vacancy formation and self-diffusion in intermetallic Fe 3 Si
nanocrystallites of nanocomposite alloys. Phys. Rev. Lett. 79, 4918-4921
(1997)
G. Amarendra, R. Rajaraman, G. Venugopal Rao, K. G. M. Nair, B. Viswanathan,
T. Suzuki, R. Ohdaira, T. Mikado. Identification of open-volume defects
in disordered and amorphized Si: a depth-resolved positron annihilation study.
Phys Rev. B 63, 224112-1 - 224112-6 (2001)
U. Myler, R. D. Goldberg, A. P. Knights, D. W. Lawther, P. J. Simpson.
Chemical information in positron annihilation spectra. Appl. Phys. Lett.
69, 3333-3335 (1996)
P. Asoka-Kumar, M. Alatalo, V. J. Ghosh, A. C. Kruseman, B. Nielsen, K. G.
Lynn. Increased elemental specificity of positron annihilation spectra. Phys.
Rev. Lett. 77, 2097-2100 (1996)
A. A. Rempel, K. Blaurock, K. J. Reichle, W. Sprengel, H.-E. Schaefer. Chemical
environment of atomic vacancies in electron irradiated silicon carbide measured
by a two-detector Doppler broadening technique. Materials Science Forum

277

Nanocrystalline Materials

53.

54.

55.
56.

57.

58.

59.

60.

61.

62.

63.
64.

65.

66.
67.

68.
69.

389-393, 485-488 (2002)


A. A. Rempel. Atomic and vacancy ordering in nonstoichiometric carbides.
Uspekhi Fiz. Nauk 166, 33-62 (1996) (in Russian). (Engl. transl.: Physics - Uspekhi 39, 31-56 (1996))
M. A. Mller, A. A. Rempel, K. Reichle, W. Sprengel, J. Major, H.-E. Schaefer.
Identification of vacancies on each sublattice of SiC by coincident Doppler broadening of the positron annihilation photons after electron irradiation.
Materials Science Forum 363-365, 70-72 (2001)
A. A. Rempel. Effects of Ordering in Nonstoichiometric Interstitial Compounds (Nauka, Yekaterinburg 1992) 232 pp. (in Russian)
V. N. Lapovok, V. I. Novikov, S. V. Svirida, A. N. Semenikhin, L. I. Trusov.
Nonequilibrium vacancy formation during ultrafine nickel powder
recrystallization. Fiz. Tverd. Tela 25, 1846-1848 (1983) (in Russian)
H.-E. Schaefer, R. Wrschum, R. Birringer, H. Gleiter. Structure of nanocrystalline iron investigated by positron lifetime spectroscopy. Phys. Rev.
B 38, 9545-9554 (1988)
H.-E. Schaefer, W. Eckert, O. Stritzke, R. Wrschum, W. Templ. The microscopic structure of nanocrystalline materials. In: Positron Annihilation.
Eds. L. Dorikens-Vanpraet, M. Dorikens, and D. Segers. (World Scientific
Publ.Comp., Singapore 1989) pp.79-85
R. Wrschum, M. Scheytt, H.-E. Schaefer. Nanocrystalline metals and semiconductors studied by positron lifetime spectroscopy. Phys. Stat. Sol. (a)
102, 119-126 (1987)
H.-E. Schaefer, R. Wrschum, T. Gessmann, G. Stckl, P. Scharwaechter,
W. Frank, R. Z. Valiev, H. -J. Fecht, C. Moelle. Diffusion and free volumes in nanocrystalline Pd. Nanostruct. Mater. 6, 869-872 (1995)
R. Wrschum, W. Greiner, G. Soyez, H.-E. Schaefer. Size and compressibility
of free volumes in nanocrystalline materials. Materials Science Forum 105110, 1337-1340 (1992)
R. Wrschum, W. Greiner, R. Z. Valiev, M. Rapp, W. Sigle, O. Schneeweiss,
H.-E. Schaefer. Interfacial free volumes in ultra-fine grained metals prepared
by severe plastic deformation, by spark erosion, or by crystallization of
amorphous alloys. Scripta Metall. Mater. 25, 2451-2456 (1991)
M. J. Puska, R. M. Nieminen. Defect spectroscopy with positrons: a general
calculational method. J. Phys. F: Metal Phys. 13, 333-346 (1983)
G. Dlubek, O. Brmmer, N. Meyendorf, P. Hautojarvi, A. Vehanen, J. YliKauppila. Impurity-induced vacancy clustering in cold-worked nickel. J.
Phys. F: Metal Phys. 9, 1961-1974 (1979)
O. K. Alekseeva, V. A. Onishckuk, V. P. Shantarovich, I. Ya. Dekhtyar, V. I.
Shevchenko. Influence of adsorded hydrogen on the positron lifetime spectra
in Pd, Pd-Ag, Pd-Cu. Phys. Stat. Sol. (b) 95, K135-K139 (1979)
M. J. Puska, C. Corbell. Positron states in Si and GaAs. Phys. Rev. B 38,
9874-9880 (1988)
L. Pasquini, A. A. Rempel, R. Wrschum, R. Reimann, M. A. Mller, B.
Fultz, H.-E. Schaefer. Thermal vacancy formation and D0 3 -ordering in nanocrystalline intermetallic (Fe 3Si) 95Nb 5. Phys. Rev. B 63, 134114-1 1341147 (2001)
Z. Gao, B. Fultz. The thermal stability of nanocrystalline Fe-Si-Nb prepared by mechanical alloying. Nanostruct. Mater. 2, 231-240 (1993)
N. E. Maslen, A. G. Fox, M. A. OKeefe, P. J. Brown, B. T. M. Willis.
Intensity of diffracted radiation. Trigonometric intensity factors. In: In-

278

Microstructure of Compacted and Bulk Nanocrystalline Materials

70.
71.

72.
73.

74.
75.

76.

77.
78.

79.
80.

81.

82.

83.

84.

85.
86.

87.

ternational Tables for X-ray Crystallography. V. C. Ed. A. J. C. Wilson (Kluwer


Academic Publishers, Dordrecht-Boston-London 1992) pp.476-519
B. E. Warren. X-Ray Diffraction (Dover Publications, New York 1990) 381
pp.
Pearsons Handbook of Crystallographic Data for Intermetallic Phases. V.3.
2 nd edition. Eds. P. Villars, L. D. Calvart. (ASM International, Metals Park,
Ohio 1991). P.2699-4034.
R. W. Cahn. Lattice parameter changes on disordering intermetallics.
Intermetallics 7, 1089-1094 (1999)
E. P. Yelsukov, V. A. Barinov, G. N. Konygin. Influence of order-disorder
transition on magnetic properties of bcc iron-silicon alloy. Fiz. Metal.
Metalloved. 62, 719-723 (1986) (in Russian)
M. Abdellaoui, C. Djega-Mariadassou, E. Gaffet. Structural study of FeSi nanostructured materials. J. Alloy Comp. 259, 241-248 (1997)
M. Polcarov, K. Godwod, J. Bak-Misiuk, S. Kadecov, J. Brdler. Lattice parameters of Fe-Si alloy single crystals. Phys. Stat. Sol. (a) 106, 1723 (1988)
A. Vehanen, P. Hautojrvi, P. Johannson, J. Yli-Kauppila, P. Moser. Vacancies and carbon impurities in a-iron: electron irradiation. Phys. Rev. B
25, 762-780 (1982)
E. A. Kmmerle, K. Badura, B. Sepiol, H. Mehrer, H.-E. Schaefer. Thermal formation of vacancies in Fe 3 Si. Phys. Rev. B 52, R6947-R6950 (1995)
Z. Q. Gao, B. Fultz. Inter-dependence of grain growth, Nb segregation, and
chemical ordering in Fe-Si-Nb nanocrystals. Nanostruct. Mater. 4, 939-948
(1994)
R. Wrschum, A. Seeger. Diffusion-reaction model for the trapping of positrons
in grain boundaries. Philosoph. Mag. A 73, 1489-1502 (1996)
E. Shapiro, R. Wrschum, H.-E. Schaefer, H. Ehrhardt, C. E. Krill, R. Birringer.
Structural stability and high-temperature positron lifetime study of mechanically
alloyed nanocrystalline Pd-Zr. Materials Science Forum 343-346, 726-731
(2000)
M. D. Bentzon, J. H. Evans. The influence of positron diffusion on positron
trapping at voids in molybdenum. J. Phys.: Condens. Matter 2, 10165-10172
(1990)
R. M. Nieminen, J. Laakkonen, P. Hautojrvi, A. Vehanen. Temperature
dependence of positron trapping at voids in metals. Phys. Rev. B 19, 13971402 (1979)
A. Broska, J. Wolff, M. Franz, Th. Hehenkamp. Defect analysis in FeAl
and FeSi with positron lifetime spectroscopy and Doppler broadening.
Intermetallics 7, 259-267 (1999)
R. Kerl, J. Wolff, Th. Hehenkamp. Equilibrium vacancy concentrations in
FeAl and FeSi investigated with an absolute technique. Intermetallics 7,
301-308 (1999)
A. Gude, H. Mehrer. Diffusion in the D0 3 type intermetallic phase Fe 3 Si.
Philosoph. Mag. A 76, 1-29 (1997)
H.-E. Schaefer, K. Frenner, R. Wrschum. Time-differential length change
measurements for thermal defect investigations: Intermetallic B2-FeAl and
B2-NiAl compounds, a case study. Phys. Rev. Lett. 82, 948-951 (1999)
R. Wrschum, C. Grupp, H.-E. Schaefer. A simultaneous study of vacancy
formation and migration at high-temperatures in B2-type Fe aluminides. Phys.
Rev. Lett. 75, 97-100 (1995)

279

Nanocrystalline Materials
88.
89.

90.

91.

92.
93.

94.
95.

96.

97.

98.
99.

100.

101.

102.

103.

104.

105.
106.

R. W. Cahn. Measures of crystal vacancies. Nature 397, 656-657 (1999)


R. Z. Valiev, A. V. Korznikov, R. R. Mulyukov. Structure and properties
of metallic materials with submicrocrystalline structure. Fiz. Metall. Metalloved.
73, No 4, 70-86 (1992) (in Russian)
R. Z. Valiev, A. V. Korznikov, R. R. Mulyukov. Structure and properties
of ultrafine-grained materials by severe plastic deformation. Mater. Sci. Engineer.
A 168, 141-148 (1993)
A. A. Nazarov, A. E. Romanov, R. Z. Valiev. On the structure, stress fields
and energy of non-equilibrium grain boundaries. Acta Metall. Mater. 41,
1033-1040 (1993)
A. A. Nazarov, A. E. Romanov, R. Z. Valiev. On the nature of high internal stresses in ultrafine grained materials. Nanostruct. Mater. 4, 93-102 (1994)
V. Yu. Gertsman, R. Birringer, R. Z. Valiev, H. Gleiter. On the structure
and strength of ultrafine-grained copper produced by severe plastic deformation.
Scripta Metall. Mater. 30, 229-234 (1994)
R. Z. Valiev. Approach to nanostructured solids through the studies of
submicron grained polycrystals. Nanostruct. Mater. 6, 73-82 (1995)
R. Z. Valiev, N. A. Krasilnikov, N. K. Tsenev. Plastic deformation of alloys with submicron-grain structure. Mater. Sci. Engineer. A 137, 35-40
(1991)
N. A. Akhmadeev, R. Z. Valiev, V. I. Kopylov, R. R. Mulyukov. Formation
of submicron-grained structure in copper and nickel with the use of severe shearing strain. Metally No 5, 96-101 (1992) (in Russian)
A. A. Rempel, A. I. Gusev, R. R. Mulyukov, N. M. Amirkhanov. Microstructure and properties of palladium subjected to severe plastic deformation.
Metallofizika i Noveishie Tekhnologii 18, 14-22 (1996) (in Russian)
A. A. Rempel, A. I. Gusev. Magnetic susceptibility of palladium subjected
to severe plastic deformation. Phys. Stat. Sol. (b) 196, 251-260 (1996)
A. A. Rempel, A. I. Gusev, R. R. Mulyukov, N. M. Amirkhanov. Microstructure, microhardness and magnetic susceptibility of submicrocrystalline
palladium. Nanostruct. Mater. 7, 667-674 (1996)
R. Z. Valiev, R. R. Mulyukov, V. V. Ovchinnikov, V. A. Shabashov, A. Yu.
Arkhipenko, I. M. Safarov. On physical width of intercrystallite boundaries. Metallofizika 12, 124-126 (1990) (in Russian)
A. Korznikov, O. Dimitrov, G. Korznikova. Thermal evolution of the structure
of ultrafine grained materials produced by severe plastic deformation. Ann.
Chim. France 21, 443-460 (1996)
A. A. Nazarov, A. E. Romanov, R. Z. Valiev. Models of the defect structure and analysis of the mechanical behavior of nanocrystals. Nanostruct.
Mater. 6, 775-778 (1995)
M. Yu. Gutkin, I. A. Ovidko, K. H. Mikaelyan. On the role of disclinations
in relaxation and deformation processes in nanostructured materials. Nanostruct.
Mater. 6, 779-782 (1995)
M. Yu. Gutkin, K. N. Mikaelyan, I. A. Ovidko. Line splitting of disclinationsin
polycrystals and nanocrystals. Fiz. Tverd. Tela 37, 552-554 (1995) (in
Russian)
V. V. Rybin, A. A. Zisman, N. Yu. Zolotarevsky. Junction disclinations in
plastically deformed crystals. Acta Metal. Mater. 41, 2211-2217 (1993)
R. Sh. Musalimov, R. Z. Valiev. Dilatometric analysis of aluminium alloy
with submicrometer grained structure. Scripta Metall. Mater. 27, 1685-1690
(1992)

280

Microstructure of Compacted and Bulk Nanocrystalline Materials


107.
108.

109.
110.

111.

112.

113.
114.
115.

116.

117.

118.
119.
120.
121.
122.

123.

124.

125.
126.

M. A. Krivoglaz. Theory of X-ray and Thermal-Neutron Scattering by Real


Crystals (Plenum Press, New York 1969) 405 pp.
H. P. Klug, L. E. Alexander. X-ray Diffraction Procedures for Polycrystalline
and Amorphous Materials (Wiley, New York 1954) 491 pp. (second edition: X-ray Diffraction Procedures (Wiley, New York 1974))
A. I. Gusev, A. A. Rempel. Structural Phase Transitions in Nonstoichiometric
Compounds (Nauka, Moscow 1988) 308 pp. (in Russian)
A. I. Gusev, A. A. Rempel, A. J. Magerl. Disorder and Order in Strongly
Nonstoichiometric Compounds: Transition Metal Carbides, Nitrides and Oxides
(Springer, Berlin - Heidelberg New York, 2001) 607 pp.
J. Balogh, L. Bujdos, G. Faigel, L. Grnsy, T. Kemny, I. Vincze, S. Szab,
H. Bakker. Nucleation controlled transformation in ball milled FeB. Nanostruct.
Mater. 2, 11-18 (1993)
J. A. Eastman, M. Fitzsimmons, L. J. Thompson. The thermal properties
of nanocrystalline palladium from 16 to 300 K. Philosoph. Mag. B 66, 667696 (1992)
X. Y. Qin, X. J. Wu, L. F. Cheng. Exothermal and endothermal phenomena
in nanocrystalline aluminium. Nanostruct. Mater. 2, 99-108 (1993)
M. L. Sui, K. Lu. Variation in lattice parameters with grain size of nanophase
Ni 3 P. Mater. Sci. Engineer. A 179-180, 541-544 (1994)
J. A. Eastman, M. Fitzsimmons, L. J. Thompson, A. C. Lawson. Diffraction studies of the thermal properties of nanocrystalline Pd and Cr. Nanostructr.
Mater. 1, 465-471 (1992)
H. Y. Zhang, K. Lu, Z. Q. Hu. Formation and lattice distortion of nanocrystalline selenium. Nanostruct. Mater. 6, 489-492 (1995); Y. H. Zhao, K.
Zhang, K. Lu. Structure characteristics of nanocrystalline element selenium
with different grain sizes. Phys. Rev. B 56, 14322-14329 (1997)
E. Hellstern, H. J. Fecht, F. Zhu, W. L. Johnson. Structural and thermodynamic properties of heavily mechanically deformed Ru and AlRu. J. Appl.
Phys. 65, 305-310 (1989)
G. W. Neiman, J. R. Weertman, R. W. Siegel. Mechanical behavior of nanocrystalline Cu and Pd. J. Mater. Res. 6, 1012-1027 (1991)
I. V. Alexandrov, R. Z. Valiev. Study of nanocrystalline materials by X-ray
analysis. Fiz. Metal. Metalloved. 77, No 6, 77-87 (1994) (in Russian) .
I. V. Alexandrov, R. Z. Valiev. Computer simulation of X-ray diffraction
patterns of nanocrystalline materials. Philosoph. Mag. B 73, 861-872 (1996)
I. V. Alexandrov, R. Z. Valiev. X-ray pattern simulation in textured nanostructured
copper. Nanostruct. Mater. 6, 763-766 (1995)
K. Zhang, I. V. Alexandrov, K. Lu, R. Z. Valiev. Structural characterization
of nanocrystalline copper by means of X-ray diffraction. J. Appl. Phys.
80, 5617-5624 (1996)
R. Z. Valiev, I. V. Alexandrov, W. A. Chiou, R. S. Mishra, A. K. Mukherjee.
Comparative structural studies of nanocrystalline materials processed by
different techniques. Materials Science Forum 235-238, 497-506 (1997)
A. Inoue, H. M. Kimura, K. Sasamori, T. Masumoto. Ultrahigh strength
of rapidly solidified Al 68-x Cr 3 Ce 1 Co x (x = 1, 1.5 and 2 %) alloys containing an icosahedral phase as a main component. Mater. Trans. Japan. Inst.
Met. 35, 85-94 (1994)
A. Inoue. Preparation and novel properties of nanocrystalline and
nanoquasicrystalline alloys. Nanostruct. Mater. 6, 53-64 (1995)
J. Kwarciak, L. Pajak, J. Lelatko. Crystallization kinetics of iron-cobalt-

281

Nanocrystalline Materials

127.

128.

129.

130.

131.

132.

133.
134.

135.

136.
137.

138.
139.
140.
141.
142.

143.

144.

silicon-boron (Fe,Co) 78 Si 9 B 13 glasses. Z. Metallkunde 79, 712-715 (1988)


N. I. Noskova, E. G. Ponomareva, V. A. Lukshina, A. P. Potapov. Effect of
rapid crystallization on the properties of Fe 5 Co 70Si 15B 10. Nanostruct. Mater.
6, 969-972 (1995)
N. I. Noskova, E. G. Ponomareva. Structure, strength, and plasticity of
nanophase Fe 73.5 Cu 1 Nb 3 Si 13.5B 9 alloy: I. Structure. Fiz. Metal. Metalloved.
82, No 5, 163-171 (1996) (in Russian). (Engl. transl.: Phys. Metal. Metallogr.
82, 542-548 (1996))
N. I. Noskova, E. G. Ponomareva, M. M. Myshlyaev. Structure of nanophases
and interfaces in multiphase nanocrystalline Fe 73 Ni 0.5 Cu 1 Nb 3 Si 13.5 B 9 alloy
and nanocrystalline copper. Fiz. Metal. Metalloved. 83, No 5, 73-79 (1997)
(in Russian). (Engl. transl.: Phys. Metal. Metallogr. 83, 511-515 (1997))
N. I. Noskova, E. G. Ponomareva, V. N. Kuznetsov, A. A. Glazer, V. A.
Lukshina, A. P. Potapov. Crystallization of amorphous Pd-Cu-Si alloy in
creep condition. Fiz. Metal. Metalloved. 77, No 5, 89-94 (1994) (in Russian)
N. I. Noskova, E. G. Ponomareva, I. A. Pereturina, V. N. Kuznetsov. Strength
and plasticity of a PdCuSi alloy in the amorphous and nanocrystalline
states. Fiz. Metal. Metalloved. 81, No 1, 163-170 (1996) (in Russian). (Engl.
transl.: Phys. Metal. Metallogr. 81, 110-115 (1996))
A. Inoue, H. M. Kimura, K. Sasamori, T. Masumoto. Structure and mechanical strength of Al-V-Fe melt-spun ribbons containing high volume fraction
of nanoscale amorphous particles. Nanostruct. Mater. 7, 363-382 (1996)
K. Yamauchi, Y. Yoshizawa. Recent development of nanocrystalline soft magnetic
alloys. Nanostruct. Mater. 6, 247-254 (1995)
B. V. Zhalnin, I. B. Kekalo, Yu. A. Skakov, E. V. Shelekhov. Phase transformations and magnetic properties variation during nanocrystalline state
formation in amorphous iron base alloy. Fiz. Metall. Metalloved. 79, No
5, 94-106 (1995) (in Russian)
V. K. Malinovskii, V. N. Novikov, P. P. Parshin, A. P. Sokolov, M. G. Zemlyanov.
Universal form of the low-energy (2 to 10 meV) vibrational spectrum of
glasses. Europhys. Lett. 11, 43-47 (1990)
M. I. Klinger. Low-temperature properties and localized electron states of
glasses. Uspekhi Fiz. Nauk 152, 623-652 (1987) (in Russian)
U. Buchenau, Yu. M. Galperin, V. L. Gurevich, H. R. Shober. Anharmonic
potentials and vibrational localization in glasses. Phys. Rev. B 43, 50395045 (1991)
J. E. Graebner, D. Golding. Phonon localization in aggregates. Phys. Rev.
B 34, 5788-5890 (1986)
V. K. Malinovskii, V. N. Novikov, A. P. Sokolov. On nanostructure of disordered
substances. Uspekhi Fiz. Nauk 163 (5), 119-124 (1993) (in Russian)
V. K. Malinovsky, A. P. Sokolov. The nature of boson peak in Raman scattering
in glasses. Solid State Commun. 57, 757-761 (1986)
V. K. Malinovskii, V. N. Novikov, A. P. Sokolov. Log-normal spectrum of
low-energy vibrational excitations in glasses. Phys. Lett. A 153, 63-66 (1991)
V. K. Malinovsky, V. N. Novikov, A. P. Sokolov, V. G. Dodonov. Low-frequency
Raman scattering on surface vibrational modes of microcrystals. Solid State
Commun. 67, 725-729 (1988)
E. Duval, A. Boukenter, B. Champagnon. Vibration eigenmodes and size
of microcrystallites in glass: observation by very-low-frequency Raman
scattering. Phys. Rev. Lett. 56, 2052-2055 (1986)
C.Yu. Clare, J. J. Freeman. Thermal conductivity and specific heat of glasses.

282

Microstructure of Compacted and Bulk Nanocrystalline Materials

145.

146.

Phys. Rev. B 36, 7620-7624 (1987)


V. K. Malinovsky, V. N. Novikov, A. P. Sokolov. Structural difference of
two glasses of amorphous semiconductors. J. Non-Cryst. Solids 114, 6163 (1989)
V. K. Malinovsky, V. N. Novikov. The nature of the glass transition and
the excess low-energy density of vibrational states in glasses. J. Phys.: Condens.
Matter 4, L139-L144 (1992)

283

Nanocrystalline Materials

+D=FJAH%
7. Effect of the Grain Size and
Interfaces on the Properties of Bulk
Nanomaterials
When interpreting the experimental results, obtained for bulk
nanomaterials, it is important to be able to separate grain boundary
(associated with interfaces) from volume (associated with small
grain size) effects. This problem is far from being solved because
at present investigation of bulk nanomaterials is still in the stage of
collecting experimental results. For this reason, the level of
understanding of the structure and properties of the bulk
nanocrystalline materials is considerably lower in comparison with
isolated nanoparticles. The properties of bulk nanomaterials in
relation to the particle size and the state of the grain boundaries
were discussed in reviews [1, 2].
7.1 MECHANICAL PROPERTIES
Because of the application of bulk nanocrystalline materials in
practice, it has been necessary to investigate in detail their
hardness, strength, elasticity, plasticity and other mechanical
properties.
In the group of the mechanical properties of nanocrystalline
materials, special attention must be given to the very high hardness.
Hardness characterises the resistance of materials to plastic
deformation during indentation of a harder body into it. Diamond is
used generally as an indentor material. When measuring hardness
by the Vickers method, the effects associated with the difference
in the elastic properties of the materials are almost completely
284

Effect of Grain Size and Interfaces on Properties of Bulk Nanomaterials

excluded because the size of the indentation is measured after


removing the stress, i.e. in the absence of elastic loading. The
experimentally measured values of hardness are subjected to the
effect of various secondary factors, such as the non ideal surface
of the material, deviation from the perpendicularity of the surface
of the material and the axis of the indentor, incorrect selection of
the loading time and the load, and also the presence in the material
of pores, voids or free volumes. In the main the hardness of the
material is determined by yield limit y . The grain size has a strong
effect on microhardness; this effect has been studied extensively
on metals, alloys and ceramics with a grain size D greater than
1 m. According to the HallPetch law [3, 4]:

y = 0 + k y D 1/2 ,

(7.1)

where 0 is the internal stress preventing the movement of


dislocations and k y is a constant. At a temperature T/T melt < 0.4
0.5 (T melt is the melting point), hardness H V (Vickers microhardness)
is linked with the yield limit y by empirical relationship H V/ y 3
[5]. This leads to the following size dependence of hardness
H V H 0 + kD 1/2 ,

(7.2)

where H 0 and k are constants.


Deformation is carried out by diffusion sliding, then according to
[6] at a low temperature T/T melt the deformation rate  d /dt is

d
= Bd dif /k B TD 3 ,
dt

(7.3)

where B is a proportionality coefficient; is applied stress, is


the atomic volume, is the thickness of the grain boundary, d dif is
the coefficient of grain boundary diffusion.
Equations (7.1)(7.3) show that a decrease in the grain size
should result in a strong change in the mechanical properties. In
particular, equations (7.1) and (7.2) predict hardening of the
material with a decrease in the grain size D. At the same time,
equation (7.3) shows that at the nanometer size of grain, diffusion
sliding becomes important even at room temperature, greatly
increasing the deformation rate. Thus, the effect of the grain size
on the strength properties of a nanocrystalline material is ambiguous
285

Nanocrystalline Materials

and depends on the ratio between the change of the yield limit and
the deformation rate. In addition, it is important to take into account
the possible increase of the grain boundary diffusion coefficient d dif
with a decrease in the grain size.
At 300 K, the microhardness of nanocrystalline materials is 2
7 times higher than H V of coarse-grained materials.
The literature data on the size dependence of the microhardness
of nanocrystalline materials are relatively contradicting. The effect
of the grain size of the microhardness of nanocrystalline metals
(copper and palladium) was investigated for the first time in [7].
The grain size of Cu and Pd was varied by annealing. The results
show that a decrease in the grain size of coarse-grained copper
from 2510 3 to 510 3 nm is accompanied by an increase in
microhardness. Microhardness of nanocrystalline copper n-Cu
(D ~16 nm) was ~2.5 times higher than that of copper with a grain
size of 510 3 nm, but with a decrease in the grain size of n-Cu from
16 to 8 nm H V decreased by ~25%. The decrease of H V was also
found in the case of a decrease in the grain size of n-Pd from 13
to 7 nm. The decrease of the microhardness of n-Cu and n-Pd was
explained [7] by the considerably higher (in comparison with
coarse-grained metals) deformation rate
 d /dt .
A decrease in the H V value of nanocrystalline alloys NiP,
TiAlNb, TiAl, NbAl 3 with a decrease in the grain size from 60100
to 610 nm was reported in [811]. According to the authors, this
effect was determined by the increase in the contribution of
diffusion mobility in the deformation with decreasing grain size D.
An increase in H V with a decrease of the grain size of n-Cu, nFe and n-Ni was reported in [1215].
An interesting fact was reported by Fougere et al [16]: an
increase or decrease in H V depends on the method of changing the
grain size. An increase in H V with a decrease in grain size D is
observed generally when hardness is measured on a series of asprepared specimens differing in the grain size. If microhardness H V
is measured on a single specimen, that is successively heated to
produce ever-increasing grain sizes, then dependence H V (D) is
different. In this case, an increase in grain size D up to some critical
value leads to a rise in H V ; the following increase in D is
accompanied by decreasing H V and corresponds to the HallPetch
law. For example, initial annealing at 569 K of nanocrystalline
compacted nc-Pd for 6090 min leads to the grain growth and an
increase of hardness by 711%; longer annealing, which leads to
further grain growth, is accompanied by a decrease of hardness
286

Effect of Grain Size and Interfaces on Properties of Bulk Nanomaterials





QP













+'

*3D

QF&X











 


QP
 

Fig. 7.1. Dependence of the microhardness H V of bulk nanocrystalline copper


nc-Cu on D 1/2 (D is the grain size) [16]: the growth of grains of nc-Cu was achieved
by increasing annealing time at 423 K (dashed arrow indicates the direction of
increase of annealing time).

(here and further abbreviation nc means nanomaterials prepared


by compaction of nanocrystalline powders). The microhardness of
nanocrystalline compacted nc-Cu increased by 45% in the first 20
30 minutes of annealing at 423 K, and then decreased (Fig. 7.1)
[16].
The results of measurements of the hardness of nanocrystalline
metals Ag, Cu, Pd, Se, Fe and Ni on different specimens with
different grain sizes were generalised in [17]. According to [17],
if several independently produced specimens of a nanocrystalline
metal have the grains of different sizes, the hardness of these
specimens increases with a decrease of D to 46 nm; the
dependence of H V on D 1/2 is governed by the HallPetch law
(7.2). The hardness of the intermetallic compounds and alloys in
which the grain size D was varied by annealing depends in a more
complicated manner on D: initially an increase in D leads to an
increase in H V and then decreases H V [17]. In other words, in the
case of grain growth as a result of annealing, the HallPetch law
(decrease of H V with increasing D) was fulfilled only at D > 12
20 nm.
According to [18], the decrease in the hardness of nanocrystalline
materials, detected in the range of low values of D, may be a
consequence of the fact that the volume fraction of triple points
287

Nanocrystalline Materials

becomes larger than the volume fraction of the grain boundaries.


According to [17, 19] the contradiction of experimental data for
the size dependence of the hardness of nanomaterials may be a
consequence of different structures of the interfaces. In view of
this the results of studies [20, 21] of the microhardness of
submicrocrystalline Al 98.5 Mg 1.5 alloy are of interest. They show that
even at unchanged grain size D 150 nm, a transition from
completely non-equilibrium interfaces to less non-equilibrium
interfaces, resulted from annealing, strongly affects the
microhardness. According to electron microscopic data, consecutive
annealing of submicrocrystalline alloy Al 98.5 Mg 1.5 (D ~150 nm) at
400 K results in a relaxation of grain boundaries and their gradual
transition to an equilibrium state, although the grain size remains
unchanged. The relaxation of grain boundaries as a result of
annealing was accompanied by a decrease in microhardness from
1.7 to 1.4 GPa.
The ambiguous fulfillment of the HallPetch law was also
detected for nanocrystalline alloys produced by crystallisation of
FeSiB, FeCoSiB, FeCuNbSiB, PdCuSi amorphous
alloys. Crystallisation of these amorphous alloys results in the
precipitation of highly dispersed phases with a grain size of several
nanometers. For example, in FeSiB alloys, the amorphous phase
showed precipitation of the grains of the bcc phase -Fe(Si) and
Fe 2 B boride. In cobalt-rich FeCoSiB alloys, -Co and -Co,
Co 2 Si, Co 2 B (or Fe,Co) 2 B) and -Fe may precipitate in the
amorphous phase [22, 23]. In nanocrystalline FeCuNbSiB
alloys, the amorphous matrix contains distributed particles of the
main precipitated bcc phase -Fe(Si) with a size of 515 nm and
copper clusters [2426]. Particles of Pd, Pd 5 Si and Pd 9 Si 2 with a
size of 5 to 30 nm are detected in nanocrystalline alloys PdCu
Si [27, 28]. The phase composition of the alloys and the grain size
of the precipitated phases depend on the heat treatment and/or
deformation treatment conditions.
The authors of [28] studied the dependence of microhardness of
nanocrystalline alloys, produced by crystallization, on the grain size
of precipitated phases. The grain size was changed mainly by
variation of the annealing temperature. The dependences of the
microhardness H V of nanocrystalline alloys on D 1/2 (D is the size
of grains (particles) of the nanocrystalline precipitated phase)
measured in [28], are represented in Fig. 7.2. It is seen that the
HallPetch law holds for all alloys in the range of D from ~10 to
~100 nm or more. At a grain size of D < 10 nm, the HallPetch
288

Effect of Grain Size and Interfaces on Properties of Bulk Nanomaterials









+'

*3D












 



 

QP



Fig. 7.2. Dependence of microhardness H V on the grain size D of phases, precipitated


in nanocrystalline alloys, produced by crystallisation of amorphous alloys [28]:
(1) Fe 7.35 Cu 1 Nb 3 Si 13.5 B 9 , (2) Fe 81 Si 7 B 12 , (3) Fe 5 Co 70 Si 15 B 10 , and (4) Pd 81 Cu 7 Si 12 .

law is fulfilled only for the Fe 73.5 Cu 1 Nb 3 Si 13.5 B 9 alloy. For other
alloys, a decrease in the grain size from ~10 nm to ~4 nm is
accompanied by a decrease in microhardness. An increase in
microhardness in accordance with the HallPetch law was observed
for nanocrystalline alloy Ni 75 W 25 [29] with a decrease in the mean
grain size from 1000 to ~10 nm; a further decrease of the grain size
resulted in a decrease in microhardness. Starting amorphous alloy
Ni 75 W 25 was produced by electrodeposition, and the grain size was
varied during crystallisation of the amorphous alloy by a means of
annealing at different temperatures [29]. It should be mentioned
that for the nanocrystalline alloys, discussed in [28, 29] (with the
exception of Fe 73.5 Cu 1 Nb 3 Si 13.5 B 9 ) the dependence of microhardness
on the grain size, varied by a means of annealing, was the same
as in [16, 17].
Control of the structure of nanocrystalline and nanoquasicrystalline materials, produced by crystallisation of amorphous
alloys is an efficient means of obtaining high tensile strength
combined with good malleability. In Section 3.3 we considered
briefly the special features of the structure of rapidly solidified Al
CrCeCo amorphous aluminium alloys because of their high tensile
strength f . The high strength of these alloys is determined by the
precipitation of aluminium-coated icosahedral nanoparticles in the
amorphous phase. The authors of [30] investigated the structure and
strength properties of amorphous alloys of the AlVFe system,
289

Nanocrystalline Materials

differing in a high volume content of the nanosised amorphous


particles. The Al 98x V x Fe 2 and Al 96x V 4 Fe x alloys in the form of
ribbon were prepared by a single roller melt-spinning technique at
a temperature from 1373 to 1423 K with a circumferential rate of
40 m s 1 . Rapidly cooled alloys Al 96 V 4 , Al 95 V 3 Fe 2 and Al 93 V 5 Fe 2
contain only the fcc phase, enriched in aluminium and the
icosahedral phase. In contrast to them, the structure of quenched
Al 94 V 4 Fe 2 alloys is a two-phase mixture of non-homogeneously
distributed nanoregions with a size of ~7 nm for the aluminium
reach fcc phase and ~10 nm for the amorphous phase. The volume
fractions of the amorphous and fcc phases in these alloys are
~60% and ~40%. The particles of the amorphous phase are
surrounded by the fcc phase, but there is no distinctive boundary
between the phases. The composition of the amorphous phase is
Al 94 V 4 Fe 2 and that of the Al-rich fcc phase is Al 93 V 5 Fe 2 . The
similar compositions of the phases indicate that the redistribution of
elements between them is repressed during rapid solidification.
Investigation of the dependence of tensile strength f and
microhardness H V on the composition of rapidly quenched alloys
Al 98x V x Fe 2 and Al 96x V 4 Fe x [30] shows that the highest values
f = 13501400 MPa and H V = 0.470.48 MPa correspond to
Al 94 V 4 Fe 2 alloy. On the basis of comparison of the structure of
alloys of different composition and their mechanical properties, the
authors of [30] assumed that the transition from the icosahedral to
the amorphous phase increases f and H V . This means that the
nanoparticles of the amorphous phase act as a hardening phase. In
addition, the value of f tends to increase with a decrease of the
distance between the particles and an increase in the volume
fraction of the hardening amorphous nanoparticles. Differential
thermal analysis and high temperature X-ray analysis of quenched
Al 94 V 4 Fe 2 alloy show that the crystallisation of the amorphous phase
starts at 580 K. During heating the crystallisation takes place as
the amorphous phase icosahedral phase (Al + Al 11 V)
transition.
Thus, the experimental studies [2230] show that the structure
of amorphous alloys (metallic glasses) immediately after producing
and, even more so, after crystallisation is characterised by the
presence of nanoregions, which correspond to different phases. In
other words, these alloys have a nano-heterogeneous structure.
Of special interest are alloys of Fe and Ni containing Al, Ti, Cr,
V, Mo, Co and W in various combinations, and also from 5 to 12
at.% B and from 0 to 7.5 at.% C, P and Si. The additivities of B,
290

Effect of Grain Size and Interfaces on Properties of Bulk Nanomaterials

C, P, and Si support the formation of a glassy (amorphous) state


[3133]. The amorphous ribbons of these alloys are produced by
melt spinning and then the ribbons are consolidated by extrusion or
hot isostatic pressing. During consolidation, the alloys completely
crystallise with the formation of a nanocrystalline structure. For
example, the Ni 59 Mo 29 B 12 and Ni 56 Mo 23 Fe 10 B 11 alloys have a matrix
based on nickel with the uniform distribution of highly dispersed
particles of Ni 2 Mo and Ni 4Mo intermetallics and large (0.52.0 m)
boride particles NiMoB 2 . These alloys have higher fracture
toughness and higher hardness than cutting steels, especially at a
temperature > 600 K. The microstructure and properties of these
alloys depend greatly on the conditions of producing of the bulk
compacted material. Aging of alloys also increase their hardness.
The nanostructure of Ni 81 Si 10 B 9 alloy is similar to the nanostructure
of aluminium-based alloys: the nanoparticles of nickel with a size
of 1020 nm are uniformly distributed in the amorphous metallic
matrix [34]. NiSiB nanocrystalline alloys are stronger than
completely amorphous alloys with the same composition.
Measurements of the microhardness H V of nanomaterials at
300 K after annealing at a series of temperatures T are used to
determine relaxation temperatures. The effect of annealing
temperature on H V and the structure of submicrocrystalline copper
with a mean grain size of 200300 nm was studied in [35] and the
copper with a grain size of 130160 nm was investigated by the
authors of [3638]. The microhardness of starting and
submicrocrystalline copper, studied in [3638], was 1.2 and 1.4 GPa,
respectively. Annealing at T < 400 K was found to have no
appreciable changes in the structure and H V . An abrupt decrease
of microhardness of submicrocrystalline copper by a factor of 2 to
3 is observed after annealing at 425450 K (Fig. 7.3) and is
associated with grain growth, the relaxation of internal stresses and
the lowering of the dislocation density, i.e. with an irreversible
transition from the submicrocrystalline to the coarse-grained state.
The microhardness of submicrocrystalline Pd with a mean grain
size of 150 nm, produced by twisting under a quasi-hydrostatic
pressure, was 2.1 GPa [3841]. Annealing resulted in a decrease
in the microhardness of submicrocrystalline Pd, and a rapid decrease
of H V started after annealing at 475 K (Fig. 7.4). The sudden
decrease of H V by almost a factor of 3 after annealing at 475
600 K is associated, according to electron microscopic examination,
with grain growth and partial annealing of dislocations.
The microhardness of submicrocrystalline titanium with a mean
291

Nanocrystalline Materials


&X

*3D



+'















Fig. 7.3. Dependence of microhardness of submicrocrystalline Cu, measured at


300 K, on annealing temperature T [37].

3G



+'

*3D

















Fig. 7.4. Dependence of the microhardness of submicrocrystalline Pd, measured


by 300 K, on annealing temperature T [40].

grain size of 150 nm, produced by equal-channel angular pressing


with subsequent cold rolling, was 2.9 GPa [42]. This is twice the
microhardness of starting coarse-grained titanium. Annealing of
submicrocrystalline titanium for 1 hour at temperatures of up to
573 K has no effect on the microhardness and the grain size (Fig.
7.5). A larger decrease of H V , which is caused by the
recrystallisation and grain growth, starts at an annealing temperature
292

Effect of Grain Size and Interfaces on Properties of Bulk Nanomaterials

HV (GPa)

2.8

2.4

2.0

1.6

400

600

800

1000

(K)

Fig. 7.5. Dependence of microhardness H V of submicrocrystalline titanium on annealing


temperature T: () values of HV after annealing for 1 h, (o) values of H V after annealing
for 3 hours [42]. Dashed lines indicate the temperatures of the start and the end
of recrystallisation. The accuracy of measurement of microhardness is
+5 %.

higher than 600 K and continues up to 873 K. After annealing at


773 K, the mean grain size increased to 12 m and after annealing
at 973 K the grain size reached 15 m. A further increase of the
annealing temperature to 1145 K resulted in a smooth decrease of
microhardness to 1.5 GPa, which is close to the microhardness of
coarse-grained titanium, equal to 1.31.4 GPa [43]. Longer, 3 hour
annealing at T < 573 K had no effect on microhardness, and
annealing at a temperature from 573 to 873 results in a decrease
in H V (Fig. 7.5). This means that in the temperature range from 573
to 873 K, which corresponds to the recrystallisation region,
submicrocrystalline titanium is in the non-equilibrium state and does
not transfer to the equilibrium state during annealing for 1 hour.
Annealing of specimens of nc-Ag (D 10 nm), prepared by
compaction of nanoclusters, is accompanied by a gradual decrease
in microhardness from 1.5 to 0.30.5 GPa; no sudden drop in H V
is observed [44]. The decrease in H V and increase of the grain size
start after annealing at 370 K. In contrast to nc-Ag, annealing of
the specimens of nanocrystalline oxide nc-MgO (D 10 nm) up to
a temperature of 870 K is not accompanied by any change in
microhardness which is equal to 2.5 0.2 GPa at 870 K [45].
Different behaviour of the microhardness of nanocrystalline metal
(nc-Ag) and nanocrystalline ceramics (nc-MgO) upon annealing
293

Nanocrystalline Materials

QF0J2

+'

*3D

QF$J









Fig. 7.6. Variation of the microhardness H V of nanocrystalline compacted specimens


of silver nc-Ag and magnesium oxide nc-MgO after annealing [44, 45].

(Fig.7.6) indicates the much higher structural stability of


nanocrystalline ceramics and the conservation of grain size in such
a material up to 800900 K. It is interesting to note that annealing
of ncMgO at 870 K < T 1070 K increased microhardness by
~30%.
Microhardness measurements are used as a method of
certification of nanocomposites. In [46], industrial powders of iron
and copper were milled in a ball mill. Milling in argon for 1220
hours resulted in the formation of a Cu 1x Fe x powder mixture with
a mean grain size of 1020 nm. The nanopowder was subjected to
preliminary cold pressing and then to hot pressing at a pressure of
0.64 GPa; hot pressing temperature was 723773 K. The produced
nanocomposite Cu 1 x Fe x (0.15 x 1.0) has homogeneous
distribution of the grains of iron and copper with a size of 2530
and 4530 nm, respectively. The microhardness of nc-Cu and ncFe was 4 times higher than that of coarse-grained copper and iron.
The microhardness of the Cu 1 -x Fe x nanocomposite in the entire
composition range was 4050% higher than the expected additive
microhardness H V of the mechanical mixture. The largest increase
of H V was found for the nanocomposite of 70 % Fe. The similar
increase of H V was observed in CuW and CuNb nanocomposites [47, 48]. According to [46], the increase of H V in the
CuFe nanocomposite is determined by the formation of interfaces
with an elevated dislocation density, since Cu and Fe have different
structures (fcc and bcc, respectively).
294

Effect of Grain Size and Interfaces on Properties of Bulk Nanomaterials

Consideration of the set of the currently available experimental


data for the hardness of metallic and ceramic nanocrystalline
materials shows that the hardness increases with a decrease in the
grain size to some critical value D crit , and at D < D crit hardness
decreases. The factors determining the critical grain size are not
clear. It cannot as yet be reliably confirmed that the values of
hardness, measured on the compacted nanomaterials, correspond to
the hardness of completely dense (pore-free) nanomaterials.
Evidently, the measured hardness is greatly affected by the pores,
microcracks and other special features of the macrostructure,
associated with the method of producing the nanocrystalline
material.
The deviation of the size dependence of the hardness of
nanocrystalline materials from the HallPetch law (7.2) is explained
by various reasons. In particular, it has been assumed that in
crystals at the grain size D < D crit the formation of flat dislocation
arrays is restricted and this should lead to a decrease of hardness
and strength. The authors of [49] assumed that annealing of the
nanomaterial results in relaxation of the grain-boundary structure
and a decrease in the excess grain-boundary energy. It leads to an
anomalous decrease in hardness with a decrease in the grain size.
According to [50], the change of hardness (or yield limit) with the
grain size is described by the dependence H V(D) ~ D n , where, in
contrast to the HallPetch relationship, n 1/2, and changes from
1/4 to 1. Each value of the exponent n corresponds to a specific
mechanism of interaction of dislocations with the grain boundaries.
It is also assumed that the value n depends on the type of crystal
lattice.
A simple phenomenological model, describing the dependence of
hardness on the grain size, is proposed by the authors of [51]. The
model is based on the assumption [52] according to which the
nanomaterial contains an intragranular crystalline phase and an
amorphous grainboundary phase. If the crystalline phase is
governed by the HallPetch law, and the grain-boundary phase has
constant hardness H gb , the hardness H of the nanocrystalline
material may be treated as the superposition of two contributions:

H = v(H 0 + kD 1/2) + (1 v)Hgb ,

(7.4)

where v = (D s) 3 /D 3 is the volume fraction of the crystalline


phase, s is the thickness of the grain boundary phase, H gb = G/12,
and G is the shear modulus of the coarse-grained bulk material.
295

Nanocrystalline Materials

However, dependence (7.4) does not have an extremum and does


not describe the transition from one type of the size dependence
of hardness to another one in the vicinity of D crit .
The combined mechanism of plastic deformation of
nanocrystalline materials is proposed in [53]; plastic shear in large
grains takes place by the dislocation model, and in fine grains by
means of the vacancy mechanism of grain-boundary sliding.
Thus, a significant role in the considered models of plastic
deformation of nanomaterials [4951, 53] is played by grainboundary sliding. The authors of [2], examining the size effect of
the mechanical properties of nanomaterials, made the following
conclusion: plastic deformation of nanocrystalline materials always
starts with grain-boundary microsliding irrespective of the method
of preparation. Microsliding takes place by the same mechanism as
the process of shear deformation in the amorphous state because
the structure of the grain boundaries in nanomaterials can be
described completely or partially by the model of the amorphous
state.
Measurements of the velocities of longitudinal and transverse
ultrasonic waves in submicrocrystalline copper as a function of the
annealing temperature made it possible to estimate the elasticity
modulus E and shear modulus G [54]. The grain size of
submicrocrystalline copper prior to annealing was 200400 nm.
Annealing was carried out in a temperature range of 373623 K
with a step of 2550 K with holding for 1 hour at each temperature.
The values E and G of the starting submicrocrystalline copper were
1015% lower than those of coarse-grained copper. Earlier, the
reduced (by ~30%) value of the elasticity modulus was observed
in nanocrystalline ncPd [12, 55]. At an annealing temperature 423

&X

























296

*3D

*3D



Fig. 7.7. Effect of annealing


temperature T on elasticity
modulus E and shear modulus
G of submicrocrystalline copper
(elasticity moduli were measured
at 300 K) [54].

Effect of Grain Size and Interfaces on Properties of Bulk Nanomaterials

456 K the values of E and G increased in jumps (Fig. 7.7). The


authors of [54] are explained these variations in the elasticity moduli
by a change in the structural state of the grain boundaries: in the
specimens of submicrocrystalline copper with a grain size of ~200
nm the grain boundaries were in non-equilibrium state and had
excess energy. Annealing at T 423 K resulted in relaxation of the
grain boundaries. Using the data [54], the authors of [56, 57]
evaluated the elasticity modulus of the grain boundaries. For an
equilibrium grain boundary with a thickness of 1 nm the elasticity
eq
eq
= 0.16 E and Ggb
= 0.12G . Similar estimates showed
modulus are Egb

that the thickness of non-equilibrium grain boundaries in


submicrocrystalline metals may reach 4 nm.
In [58] it is shown that a decrease in the elasticity modulus of
copper, subjected to severe plastic deformation, is evidently
associated with both the formation of a crystallographic structure
and with the decrease in the grain size. However, for the
nanocrystalline materials with a grain size smaller than 510 nm,
the fraction of atoms, located in the grain boundaries, is very large.
According to [2], these materials have atomic structure which is
closed to amorphous one. The elasticity modulus of amorphous
materials is approximately half the modulus of similar coarsegrained crystalline materials [59]. Therefore, one may expect that
the elastic properties of nanomaterials with a grain size of 510 nm
or smaller should be considerably lower. This is in agreement with
the estimates [60] obtained by the molecular dynamics methods: a
decrease in Youngs modulus E, shear modulus G and uniform
compression modulus K of copper with a decrease in the grain size
to several nanometers is 25, 50 and 80 %, respectively.
The effect of temperature on elasticity modulus E and the yield
limit of submicrocrystalline copper produced by compacting
nanoparticles (D ~ 80 nm) and by the equal-channel angular pressing
(D ~ 100 nm), was studied in [61, 62]. The copper, compacted from
a nanopowder under a pressure of 1 GPa, had a porosity of ~11 %.
The elasticity modulus E of compacted copper increases after
annealing at T 470520 K and rapidly decreases after annealing
at 670 K as a result of grain growth and a simultaneous increase
in the porosity of the compacted copper. As the annealing
temperature is raised, the elasticity modulus of plastically deformed
copper rapidly increases in the 370500 K and 720870 K
temperature intervals. The yield limit decreases with increasing
annealing temperature. The jump of elasticity modulus E after
297

Nanocrystalline Materials

annealing at 370 K was explained by the authors of [61] by the


pinning of grain-boundary dislocations by point defects and by the
lowering of dislocation mobility as a result of a change in the
structure of the interfaces. The second jump was explained by the
disappearance of the grain-boundary phase as a result of grain
growth.
On the whole, the results in [54, 55, 5862] show that the
unusual elastic properties of the submicrocrystalline materials are
determined not as much by the small grain size as by the state of
the interfaces.
Grain refinement is the well-known method of increasing the
strength properties of materials. Studying of the stressstrain
diagrams for compacted specimens of nanocrystalline Pd (D = 5
15 nm) and Cu (D = 2550 nm) [12] shows that the yield limit y
of nanocrystalline metals is 2 to 3 times higher than that of coarsegrained metals. A large increase of y of the submicrocrystalline
alloys of magnesium in comparison with coarse-grained alloys was
detected by the authors of [63]. In nanocrystalline alloys FeSi
B, FeCoSiB, FeCuNbSiB and PdCuSi, produced by
crystallisation from the amorphous state, a decrease in the size D
of precipitated particles of dispersed phases from 3035 to ~10 nm
is accompanied by an increase in the yield limit, whereas with a
further decrease of D the yield limit decreased [28]. The tensile
strength of nanocrystalline metals is 1.5 to 8 times higher than that
for coarse-grained metals [12, 17]. According to [12], the main
reasons for the increase in the strength of nanocrystalline metals
may be high energy of formation and movement of dislocations,
caused by the small grain size or/and high residual stresses. The
changes in the strength properties of the materials, caused by the
transition to the nanocrystalline state, require detailed study. In
particular, this relates to multiphase nanocrystalline alloys, produced
by crystallisation from the amorphous state [64].
An important problem is the damping of vibrations of metallic
materials. The improvement of the damping properties reduces the
harmful effect of cyclic loads, which are the cause of many
accidents and failures; decreases the noise caused by the vibration
of mechanisms; raises the accuracy of measuring devices owing to
the damping of vibrations. Investigations [6569] of the amplitude
dependence of internal friction in submicrocrystalline copper show
that as the size of the crystallites decreases and as the deviation
of grain boundaries from the equilibrium state increases, the internal
friction background and damping properties of the material increase.
298

Effect of Grain Size and Interfaces on Properties of Bulk Nanomaterials

For example, in submicrocrystalline copper with a mean grain size


of ~200 nm, that has been annealed at 423 K, the background level
is 3 to 5 times higher than that in the coarse-grained specimens and
2 to 3 times higher than that in grey cast iron (the background level
for grey cast iron is a conventional boundary for heavy damping).
The temperature, at which internal friction begins sharp rise, in
submicrocrystalline copper decreases by approximately 100 K in
comparison with that of coarse-grained copper. In addition, at
475 K a pronounced maximum appears in the temperature
dependence of internal friction. The internal friction in the
temperature range from 240 to 475 K is approximately 3 times
higher for submicrocrystalline copper than that for coarse-grained
specimens.
All listed features are related to the difference between the
elasticity modulus of the grains and of grain boundaries [70, 71].
The difference in the modulus makes it possible to examine the
submicrocrystalline material as a heterogeneous material for the
propagation of elastic vibrations. Consequently, the submicrocrystalline material shows considerable scattering of elastic
vibrations resulting in an increase in the damping properties. It
should be mentioned that the nanocrystalline and submicrocrystalline
materials combine enhanced strength [16, 61, 72] and damping [65
71] properties. In conventional materials, the strength properties
decrease with increasing damping properties. The effect of the
simultaneous increase of the damping and strength properties,
discovered in submicrocrystalline copper, was confirmed by the
studies of stainless steel [73, 74]. It is shown that the formation of
the submicrocrystalline structure in the steel increases the internal
friction background and the yield limit approximately 4 times.
The superplasticity of ceramic nanomaterials should also be
mentioned. For a long time, superplasticity was only a dream of
materials scientists studying the processes of shaping and
deformation of ceramics. Superplasticity is characterised by an
exceptionally large relative elongation or shrinking of the material
under tensile loading [75]. This phenomenon was demonstrated for
the first time in 1934 on an example of the elongation of SnB alloy
by more than a factor of 20 [76]. The superplasticity of ceramics
was discovered for the first in 1985 on polycrystalline tetragonal
oxides ZrO 2 stabilised with yttrium oxide Y 2 O 3 [77]. Later,
superplasticity was observed in the two-phase composite ceramics
Si 3 N 4 /SiC [78], and in other ceramic materials.
Superplasticity is very important for producing ceramic ware by
299

Nanocrystalline Materials

shaping, solid-phase sintering or hot pressing at relatively low


temperatures. Superplasticity leads to a high accuracy of the
dimensions of ceramic ware of vary complicated shapes with inner
cavities and surfaces with a varying curvature.
According to [79], the superplasticity of ceramics manifests itself
most vividly when the grains are smaller than 1 m. With increasing
temperature the grain size should remains unchanged as long as
possible. For example, in nanocrystalline compacted magnesium
oxide nc-MgO, the grain size remains almost constant when
annealed up to 800900 K [45]. In zirconium oxide ZrO 2 , grain
growth with increasing temperature is suppressed by small additions
of Y 2 O 3. In two-phase ceramics based on silicon nitride and silicon
carbide, the grain growth of the matrix phase is suppressed due to
precipitation of grains of the second phase. Factors increasing the
plasticity of ceramics, also include the high-angle misorientation of
the grain boundaries and the presence of a small amount of an
amorphous grain-boundary phase [80]. In the nanocrystalline state,
some ceramic materials (for example, TiO 2, [81]) become plastically
deformable already at room temperature.
In [82], special features of the behaviour of various mechanical
properties such as microhardness and elasticity modulus with
decreasing grain size were investigated using the statistical model
of an ensemble of grain-boundary defects. The ensemble of defects
of the type of microshear and microcracks in developed deformation
stages has clear features of collective behaviour. The concentration
of these defects is very high and reaches 10 13 10 14 cm 3 .
Therefore, the reason for the appearance of cooperative effects
may be regarded as a thermodynamic reason. At the same time,
each of the elementary defects (grain boundary or microcrack) is
in a general case a thermodynamically non-equilibrium system.
The evolution of ensembles of grain-boundary defects depends
on the distribution of the nuclei of defects and by the interaction
between defects. The solution of the FokkerPlanck equation for
the distribution function of the nuclei of defects leads to a
conclusion on the different behaviour of the dependence of the
tensor component p zz on applied stress zz at different values of
some dimensionless parameter . Analysis [82] shows that there
are three different regions of solutions separated by asymptotics c
and *. These solutions determine qualitatively different reactions
of the polycrystal to the increase of the concentration of defects,
which is caused by loading, in relation to the grain size. According
to [82], the region of the values > * with a stable distribution
300

Effect of Grain Size and Interfaces on Properties of Bulk Nanomaterials




QDQR3G



5 - PRO

FRDUVHJUDLQHG 3G


QDQR&X


FRDUVHJUDLQHG &X










Fig. 7.8. Effect of the nanostructured state on the temperature dependence of heat
capacity C p (T) of copper and palladium [83]: ( F) nanocrystalline compacted ncPd and () coarse-grained palladium Pd; ( l )nanocrystalline compacted nc-Cu and
( ) polycrystalline coarse-grained copper Cu.

of the grain-boundary defects corresponds to a reaction of materials


with a very small grains, i.e. nanocrystalline materials. In particular,
the variation of the grain size in the vicinity of * may be
manifested in the different slope of the HallPetch dependences,
and in a rapid change of the elasticity modulus (see, for example,
Fig. 7.1 and 7.7).
7.2 THERMAL AND ELECTRIC PROPERTIES
The surface and size effects which observe in the phonon spectrum
and in behaviour of heat capacity of nanoparticles, are studied in
detail (see Section 5.3). The theoretical analysis and calorimetric
investigations show that in the temperature range 10 K T D
the heat capacity of nanopowders is from 1.2 to 2 times higher than
that of the coarse-grained bulk materials. The increased heat
capacity of the nanopowders is determined both by the size effect
and by the very large surface area, which introduces an additional
contribution to heat capacity. In contrast to the nanoparticles,
investigations of the heat capacity of nanocrystalline bulk materials
are limited to several studies.
The heat capacity C p of nanocrystalline compacted specimens of
nc-Pd (D = 60 nm) and nc-Cu (D = 8 nm), prepared by compacting
nanoclusters, was measured in the temperature range from 150 to
300 K [83]. The relative density of the ncPd specimens was equal
to 80 % and that of the nc-Cu specimens was equal to 90 % of the
density of pore-free polycrystalline coarse-grained palladium Pd and
301

Nanocrystalline Materials

copper Cu. Measurements revealed that the C p of nc-Pd and ncCu specimens are 2953 % and 911 % higher than the heat
capacity of conventional polycrystalline Pd and Cu, respectively
(Fig. 7.8). When nc-Pd was heated at T = 350 K, an exothermic
effect was observed but the grain size remained unchanged or
increased to 10 nm. The heat capacity of nc-Pd heated to 350 K,
was found to exceed the heat capacity of coarse-grained palladium
by 5 %. The authors of [83] assumed that the observed elevated
heat capacity is caused by the looser structure of the interfaces.
This explanation is not plausible because it has been established
that the structure of the grain boundaries in the compacted
nanomaterials contains free volumes with the size of monovacancy
or divacancy, but the effect of this free volumes is not large enough
in order to explain excess of heat capacity. One of the explanation
of excess could be impurities in palladium.
Indeed, when studying the heat capacity of nc-Pt, the authors of
[84] concluded that at a temperature ~300 K a large part of the
excess heat capacity of the compacted nanomaterials is due to the
excitation of the impurity hydrogen atoms. Impurity hydrogen is
often present in nanomaterials, prepared by condensation of
nanoclusters in an inert gas and by subsequent compacting. For
example, the high solubility of hydrogen in grain boundaries of ncPd is noted in [85, 86]. According to [87], hydrogen dissolves
primarily in the nc-Pd grains rather than in the interfaces.
The low temperature heat capacity of the bulk nanocrystalline
compacted copper nc-Cu with a grain size of 6.0 or 8.5 nm in the
temperature range from 0.06 to 10.0 K is 5 to 10 times higher than
that of coarse-grained copper [88]. The largest increase in heat
capacity was detected in the specimens of nc-Cu with the smaller
grains. The increase in the heat capacity of nc-Cu at T > 1 K may
be caused by the fact that the weakly bonded atoms on the surface
of grains behave as Einsteins linear oscillators and surface
vibrational modes appear in the phonon spectrum. According to
estimate [88], each 6 th to10 th surface atom (depending on the grain
size), is such an oscillator.
In [89], inelastic neutron scattering at 100300 K was used to
study the density of phonon states g( ) in an n-Ni nanopowder, in
a nanocrystalline compacted specimen of nc-Ni with a relative
density of 80 % and in coarse-grained nickel. The grain size in both
n-Ni and nc-Ni was about 10 nm. The most pronounced size effect
is the increase in the density of phonon states g( ) of both n-Ni
and nc-Ni specimens in comparison with coarse-grained nickel in
302

Effect of Grain Size and Interfaces on Properties of Bulk Nanomaterials

the energy range below 14 meV (Fig. 5.12). Calculations carried


out using the data for the density of phonon states show that the
heat capacity of nc-Ni at T 22 K is 1.5 to 2 times higher than
the heat capacity of coarse-grained nickel. According to [89], the
change in the phonon spectrum and high heat capacity of nc-Ni are
caused by the contribution of the grain boundaries with reduced
density of substance. In [89] it is also noted that the excess heat
capacity of the compacted nanomaterials at room temperature is
most likely due to the impurity of hydrogen atoms, whose vibrations
are excited at T 300 K.
Measurements of the temperature dependence of the heat
capacity of compacted specimens of nanocrystalline nickel nc-Ni
with a mean grain size of ~70 nm [90] show that at T 600 K ncNi has higher heat capacity in comparison with coarse-grained
nickel. According to [70, 90], the high heat capacity of nc-Ni is
caused by the contribution of the grain-boundary phase with a
reduced Debye temperature and higher (by 1025 %) heat capacity
in comparison with the coarse-grained metal.
The measurements of the temperature dependence of the heat
capacity of amorphous, nanocrystalline and coarse-grained selenium
Se in the temperature interval from 220 to 500 K [91] showed a
small increase in the heat capacity of nanocrystalline n-Se in
comparison with a coarse-grained Se at T < 375 K. The heat
capacities of amorphous and n-Se coincide within the measurement
error. Bulk nanocrystalline n-Se was prepared by crystallisation from
the amorphous state in order to exclude the effect of distortions of
the structure and also gaseous and other impurities on heat capacity.
Table 7.1 shows the heat capacity of several substances in the
Table 7.1 Comparison of heat capacity C p (J mol 1 K 1 ) of the nanocrystalline,
amorphous and coarse-grained polycrystalline states of different materials [83]
State
nanocrystalline

Material

Pd
Cu
Ru
Ni0.8P0.2
Se

synthesis
method*

crystalline
size D(nm)

Cp

1
1
2
3
3

6
8
15
6
10

37
26
28
23.4
24.5

Amorphous
Cp

Coarse- grained
Cp

27
23.4
24.7

25
24
23
23.2
24.1

Reference

(K)

250
250
250
250
245

83
83
72
92
91

*(1) compaction of ultrafine powders prepared by evaporation, (2) ball milling, (3) crystallization from the
amorphous state

303

Nanocrystalline Materials

nanocrystalline, amorphous and coarse-grained states. A large


difference in heat capacity in comparison with a coarse-grained
state is observed for the specimens prepared by compacting of the
nanopowders. On the other hand, this difference is very small and
does not exceed 2 % for the specimens produced by crystallisation
from the amorphous state. It may be assumed that the main part
of the excess heat capacity of the bulk nanomaterials is determined
by the large area of the interfaces, structural distortions and
impurities. Temperature measurements of the crystal lattice constant
of selenium [93] also made it possible to determine the dependence
of the coefficient of volume thermal expansion, V, on the crystallite
size D. Coefficient V of nanocrystalline selenium increases by
approximately 30 % with a decrease in the crystallite size D from
45 to 10 nm.
To explain the anomalies of low temperature heat capacity, the
authors of [94] proposed a model of a bulk nanocrystalline material
in which all the grains are rhombohedrons and are of the same size.
The model cell consists of 8 such grains (Fig. 7.9). In modeling, the
grain size D, which is determined as the diameter of a spherical
particle with the same number of atoms, was assumed to be equal to
1.1, 2.0 and 2.8 nm. Interatomic interactions were described by the
LennardJones potential. Calculations of the density of vibrational
states, g( ), showed that in comparison with a perfect fcc crystal

2
2

2
1

2
2

2
1

2
1

z
y
x

Fig. 7.9. Idealised 3-dimensional space-filling polycrystal model [94]. The 3Dperiodic simulation cell contains 8 identically shaped rhombohedral grains. There
are two types of grains, which are indicated by 1 and 2. For each type, all the
surfaces of the grains are crystallographically equivalent. The eight grains are connected
by 24 identical asymmetric tilt grain boundaries.

304

Effect of Grain Size and Interfaces on Properties of Bulk Nanomaterials





















7+]

Fig. 7.10. Density of the vibrational states, g(v) for the idealised model nanocrystal
with grains of size D = 1.1 nm (solid line) and for a fcc perfect monocrystal,
which contains 500 atoms (dashed line) [94].


  QP

& ;

N





  QP


  QP













Fig. 7.11. Excess heat capacity C V for the idealised model nanocrystal as a function
of temperature T and grain size D [94].

consisting of 500 atoms, g( ) of the model nanocrystal (D = 1.1 nm)


is extended into both low- and high-frequency regions (Fig. 7.10). The
majority of additional low- and high-frequency vibrational modes are
localised at the grain boundaries. According to the calculations, the
heat capacity of the nanocrystal is higher than that of the ideal fcc
crystal; the difference of their heat capacities C increases with a
decrease in the grain size (Fig. 7.11). The excess heat capacity of the
nanomaterial is determined by low-frequency vibrational modes related
to the grain boundaries. The contribution of high-frequency vibrations
305

Nanocrystalline Materials

to the anomalous increase of the heat capacity of the nanocrystal is


very small.
Theoretical analysis of the internal energy and excess heat
capacity of nanocrystalline materials was also carried out in [95]
using a formalism equivalent to the mean-field approximation.
According to [95], the excess heat capacity C in the low
temperature range is a linear function of temperature; when T J,
the difference C has a wide maximum (J is the energy parameter
describing the interaction of atoms, each of which has two
equilibrium positions).
In the simplest case, according to the Gruneisen equation, the
coefficient of thermal expansion, , is proportional to the heat
capacity C V . Taking this into account, it may be expected that
nanocrystalline material should have a higher coefficient in
comparison with coarse-grained material. In fact, nc-Cu with a
mean crystallite size of 8 nm has a coefficient of thermal expansion
= 3110 6 K 1 , which is twice as large as the value = 16
10 6 K 1 for coarse-grained copper [96, 97].
To detect the effect of the grain boundaries on the coefficient
of thermal expansion, the authors of [98] measured the thermal
expansion of rolled copper foils with a grain size of 17 m and of
polycrystalline copper with a grain size of 19 mm. The thermal
expansion coefficient of the copper foil was higher than that of
coarse-grained copper. According to [98], the high value of is
caused by the fact that the grain boundaries have a considerably
higher thermal expansion coefficient than that of the crystallites: for
grain boundaries gb = (4080)10 6 K 1 , i.e. 2.5 to 5 times higher
than the value of of coarse-grained copper. It should be
mentioned that the copper foil, studied in [98], have been produced
by the method which was similar to the method of production of
such submicrocrystalline materials in which the atoms in the
interfaces have an enhanced mobility.
The size dependence (D) of nanocrystalline alloy Ni 0.8 P 0.2 was
studied in [99]. Nanocrystalline specimens were produced by
crystallisation of a ribbon of Ni 0.8 P 0.2 amorphous alloy at seven
different annealing temperatures from 583 to 693 K. As-crystallised
NiP specimens contain two phases: a solid solution of phosphorous
in Ni with a fcc structure and compound Ni 3 P with a bodycentered tetragonal structure. The mean grain size of the
precipitated phase Ni 3 P in relation to annealing temperature was
7.5127 nm. Measurements show that as the grain size decreases
from 127 to 7.5 nm, the coefficient of linear thermal expansion
306

Effect of Grain Size and Interfaces on Properties of Bulk Nanomaterials



1L3 



 .
















QP

Fig. 7.12. Linear thermal expansion coefficient of nanocrystalline Ni 0.8 P 0.2 alloy
versus the grain size D of the Ni 3 P phase [99]

increases from (15.51.0)10 6 K 1 to (20.71.5)10 6 K 1 (Fig.


7.12). The linear thermal expansion coefficients of coarse-grained
NiP alloy (D 10 m) and NiP amorphous alloy of the same
composition are 13.710 6 and 14.210 6 K 1 , respectively. It is
clear that the value of of the nanocrystalline alloy is higher than
that of the coarse-grained and amorphous alloys.
The authors of [99] represented coefficient nc for the
nanocrystal in the form

nc = in f in + c(1 f in) ,

(7.5)

where in and c are the linear thermal expansion coefficients of


interfaces and crystallites, f in = c/D is the volume fraction of the
interfaces, c = 1.9 is a constant, and D is the crystallite size.
Calculations using the experimental data showed that with a
decrease of D the difference ( in c ) = ( nc c )/f in rapidly
decreases. For example, at D = 100 nm the ratio in / c is equal
to 12.7, i.e. the thermal expansion coefficient for the interfaces is
ten times higher than that for the crystallites. For a nanocrystal
with a grain size of several nanometers in diameter in / c = 1.2.
According to [99], the large decrease of the value of ( in c )
with a decrease in the grain size may be a consequence of the
increase in the density of the interfaces and/or of the decrease of
lattice constant of nanometer crystallites. The latter seems a more
likely reason.
307

Nanocrystalline Materials

N- PRO









QP

Fig. 7.13. Variation of the molar Gibbs free energy G of a binary nanocrystalline
solid solution (alloy) with crystallite size D at fixed p, T, and fixed concentration
of second component, x = 0.05. The dotted line denotes the Gibbs free energy of
the monocrystalline binary solid solution with the same x ; this Gibbs free energy
is independent on the grain size and is equal to 4 kJ mol 1 [102] .

The high heat capacity and thermal expansion coefficient for the
nanocrystalline compacted materials indicate that such nanomaterials
are thermodynamically unstable. It is shown in [100] using nc-Pd
as an example that the structural state of the nanocrystalline
compacted
material
immediately
after
preparation
is
thermodynamically non-equilibrium (see Section 6.1).
General problems of thermodynamics and segregation for
nanocrystalline binary solid solutions (alloys) are discussed in [101
104]. The effect of segregation at the grain boundaries on the
thermodynamic stability of solid solutions, which have a large
segregation heat, is considered in [101, 102]. According to [101,
102], if specific grain-boundary energy in a binary solid solution
is negative, this solid solution is metastable. This occurs only at
rather low temperatures. If < 0, the dependence between the free
Gibbs energy G and the crystallite size D becomes non-monotonic
and exhibits a pronounced minimum. Fig. 7.13 displays a model
G(D) dependence for the polycrystalline solid solution with
the concentration of the second component x = 0.05. For this solid
solution = 1.7 J m 2 . Calculation is performed for a temperature
of 600 K. As is seen, the G(D) dependence (solid line) has a free
energy minimum, which corresponds to the range of D from 10 to
20 nm. The free energy G 4 kJ mol 1 of the monocrystalline solid
solution with the same composition is shown in Fig. 7.13 as a
308

Effect of Grain Size and Interfaces on Properties of Bulk Nanomaterials

dashed line. When the grain size is rather large, the free energy of
a polycrystalline alloy asymptotically tends to the free energy
G 4 kJ mol 1 of the monocrystalline alloy. The results obtained
in [101, 102] reveal that the dependence of the free Gibbs energy
on the solid solution composition changes both quantitatively and
qualitatively when the grain size of a polycrystalline alloy decreases
to a nanometer range.
The problem of thermal stability of bulk nanocrystalline
compacted materials was studied in [105]. In this study, the time
dependence of electromotive force (emf) E of high-purity nc-Pd
with a mean grain size of 11 and 18 nm was measured at 613 K.
For the reference electrode, coarse-grained palladium with a grain
size of ~20 m was used. Measurements show that the emf of the
nanocrystalline specimens is negative, rapidly increases in the first
45 minutes and then increases slowly asymptotically approaching
zero. Because of the large area of the grain boundaries, the
exchange electrochemical reaction in nc-Pd takes place at the grain
boundaries at a high rate; in this case, the emf E of the
nanomaterial is directly related to the thermodynamic characteristics
of grain boundaries by a simple relationship G gb = |z|FE, where
z is the valence of the palladium ion, F is the Faraday constant, and
G gb is the Gibbs energy of the interfaces. Taking this into account,
the negative emf corresponds to the positive Gibbs energy of
nanocrystalline palladium in comparison with coarse-grained
palladium; this means that nc-Pd is thermally unstable at elevated
temperatures. According to [105], the rapid increase of the emf in
the first stage of measurement is caused by the relaxation of
interfaces; the subsequent slow approaching of the emf to zero is
due grain growth. Similar behaviour of emf, associated with the
relaxation of interfaces and grain growth, was observed in
calorimetric measurements of nanocrystalline Pt [106]. The emf
after relaxation of the interfaces was equal to 36, 7 and 4 mV
for nc-Pt with a grain size of 11, 18 and 20 nm, respectively. Thus,
as the grain size decreases, the thermodynamic stability of the
material also decreases. The thermal instability of the nanomaterial
is caused primarily by the non-equilibrium state of the grain
boundaries.
The greatly developed interfaces and the high defect
concentration cause the extensive scattering of current carriers in
nanomaterials. The large increase of the specific electrical
resistivity of nanocrystalline Cu, Pd, Fe, Ni and the different alloys
with a decrease in the grain size was reported by many
309

Nanocrystalline Materials

researchers. Investigation of the temperature dependence of the


electrical resistivivity of compacted nanomaterials is used to
characterise the state of grain boundaries and to determine the
relaxation temperature.
Specific electrical resistivity of nc-Cu (D = 7 nm) in the
temperature range 0 < T 275 K is 7 to 20 times lower than that
of conventional coarse-grained copper [107]. At T 100 K the
specific electrical resistivity of coarse-grained Cu and nc-Cu
increases linearly with increasing temperature, but the temperature
coefficient of resistivity, /T, for nc-Cu is equal to 1710 9
cm K 1 and is higher than /T = 6.610 9 cm K 1 of
conventional copper. Analysis of the experimental dependences (T)
of nanocrystalline and coarse-grained copper showed that the
coefficient of electron scattering, r, at the grain boundaries in
nc-Cu is equal to 0.468 at 100 K and is equal to 0.506 at 275 K,
while for coarse-grained copper r = 0.24, which is lower by a factor
of two. This difference is a consequence of the different width and
structure of the grain boundaries in nanocrystalline and coarsegrained copper. The temperature dependence of the coefficient r
of nc-Cu is caused by a high thermal expansion coefficient of the
grain boundaries: according to [96, 97], gb = 6610 6 K 1 . Kai
[107] believes that the high electrical resistivity and temperature
coefficient /T of nc-Cu are caused primarily by electron
scattering at the grain boundaries. Another reason for the high
electrical resistivity of nc-Cu may be the short mean free path
of electrons: for nc-Cu 4.7 nm, and for coarse-grained copper
44 nm.
Investigation of the resistivity of polycrystalline cobalt films with a
thickness in the range from 2 to 50 nm demonstrated that the value
of is almost independent on temperature, decreases with increasing
film thickness and is higher than that of bulk cobalt [108]. According
to [108], the temperature coefficient of resistivity, which is close to
zero, and the high specific electrical resistivity of nanocrystalline Co
films are a consequence of partial localisation of the electrons when
the grain diameter becomes shorter than the electron mean free path.
Localisation affects the electrical conductivity more strongly than the
increase in the scattering of charge carriers at the interfaces because
it leads to a decrease in the concentration of charge carriers. As a
result, a decrease in the crystallite size increases the degree of
localization, decreases the concentration of charge carriers and, hence,
increases specific electrical resistivity.
The specific electrical resistivity of submicrocrystalline Cu, Ni and
310

Effect of Grain Size and Interfaces on Properties of Bulk Nanomaterials

Fe, produced by equal-channel angular pressing, was investigated in


[109111]. The mean grain size of submicrocrystalline metals was
100200 nm. At 80 K the specific resistivity of submicrocrystalline
copper was almost twice the value of of coarse-grained copper. The
high electrical resistivity of submicrocrystalline copper is caused by a
high coefficient of electron scattering, r, at non-equilibrium grain
boundaries: in submicrocrystalline copper r = 0.290.32 instead of
r = 0.24 for equilibrium grain boundaries in coarse-grained copper.
According to [112], the increase in coefficient r is due to distortions
of translation symmetry caused by long-range stress fields and by the
dynamically excited state of atoms in the grain-boundary phase.
Annealing at 420470 K results in a rapid decrease in ; with a
further increase in annealing temperature, specific electrical resistivity
slowly decreases [109, 110]. According to the results of
microstructural study, the rapid decrease in as a result of annealing
at 420470 K is caused by the relaxation of grain boundaries and by
their transition from the stressed non-equilibrium state to equilibrium
state. The subsequent slow decrease in is a consequence of grain
growth.
According to the results [111], the specific electrical resistivity
of submicrocrystalline Cu, Ni and Fe at 250 K is 15, 35 and 55 %
higher than that of the same coarse-grained metals; the temperature
coefficients of the electrical resistivity of submicrocrystalline and
coarse-grained Cu, Ni and Fe differ only slightly. In [111],
measurements of the temperature dependences of the thermal emf of
submicrocrystalline copper, nickel and iron were also taken. The
absolute value of the thermal emf of submicrocrystalline metals is
lower than that of coarse-grain metals; for Cu and Fe thermal emf is
positive, and for Ni it is negative. The observed change of the transport
properties of the submicrocrystalline metals with an increase in
temperature from 20 to 270 K was explained by extensive electron
scattering at the grain boundaries [111].
7.3 MAGNETIC PROPERTIES
The effect of the nanocrystalline state on the magnetic properties
of paramagnetics is studied in [3841, 113] on an example of
palladium. Polycrystalline Pd with crystallites several micrometers
in size is characterised by a unique electronic structure, which is
extremely sensitive to ferromagnetic impurities or to the effect of
external pressure. This prompted the authors of [113] to assume
that the formation of the submicrocrystalline structure in palladium
311

Nanocrystalline Materials

may affect the electronic structure and magnetic susceptibility.


Submicrocrystalline Pd was produced from coarse-grained Pd by
severe plastic deformation with using torsion under quasi-hydrostatic
pressure; this resulted in a true logarithmic degree of deformation
e = 7. The density of submicrocrystalline Pd coincided with the
density of starting palladium and did not change after annealing at
temperatures of 300 to 1200 K. This indicates the absence of
porosity of submicrocrystalline Pd. The grain size of
submicrocrystalline Pd, determined by X-ray diffraction and electron
microscopy, was 120150 nm.
Magnetic susceptibility was measured by the Faraday method
with an accuracy of 0.0510 6 cm 3 g 1 in a vacuum of 1.310 3
Pa (10 5 mm Hg). The procedure and sequence of magnetic
measurement is described in Section 5.4 and is shown in Fig. 5.13.
The magnetic susceptibility of starting and submicrocrystalline Pd
did not depend on the magnetic field strength H; this indicates the
absence of ferromagnetic impurities in the specimens. Measured
temperature dependence (T) and annealing dependence (300,T)
of the magnetic susceptibility of starting and submicrocrystalline Pd
in the temperature range from 300 to 1225 K are shown in Fig. 7.14



.  FP  J


3G



















Fig. 7.14. Magnetic susceptibility of submicrocrystalline and coarse-grained palladium


[40]: (1) and (2) are annealing (300, T) and temperature (T) dependences of the
susceptibility of submicrocrystalline Pd, respectively; (3) and (4) are annealing
(300, T) and temperature (T) dependences of the susceptibility of starting coarsegrained palladium, respectively. Annealing dependences (300,T) of susceptibility
(curves 1 and 3) were measured at 300 K after annealing at temperature T and
subsequent cooling to 300 K.

312

Effect of Grain Size and Interfaces on Properties of Bulk Nanomaterials

(measurements of made directly at annealing temperature T after


holding at this temperature for 1 h relate to the temperature
dependence (T); measurements of susceptibility, made at 300 K
after a specimen had been annealed at temperature T and then
cooled to room temperature, relate the annealing dependence
(300,T)).
At T < 825 K, the magnetic susceptibility (300,T) of
submicrocrystalline Pd considerably exceeds the susceptibility of
starting palladium, which is independent of the annealing temperature.
As a result of annealing at 8251025 K, the susceptibility of
submicrocrystalline Pd first drops and then decreases slowly to the
value of corresponding to the starting palladium. The temperature
dependence (T) for submicrocrystalline Pd (curve 2 in Fig. 7.14)
does not contain similar sharp transition. With an increase of
temperature, the susceptibility of submicrocrystalline Pd decreases and
smoothly changes to the dependence (T) of starting palladium;
starting from T = 725 K, the temperature dependences (T) of
submicrocrystalline and starting Pd almost completely coincide.
The main contribution to the susceptibility of polycrystalline
palladium is due to the Pauli spin paramagnetism of conductivity
electrons p [114]. There is strong Stoner enhancement of p by a
factor of 10 to15 [115], which reflects many-particle effects of
electron interaction. Another important feature of palladium is the
presence of a high and narrow (~0.3 eV) peak near the Fermi
energy [115]. This peak is responsible for the high density of
electron states (~2.3 eV 1 atom 1 ) at the Fermi energy. The
narrow density of states peak determines the sensitivity of the
properties of palladium to the Fermi energy.
In the low-temperature region, approximately down to 50 K [116],
the magnetic susceptibility of palladium increases as T 2 and is
described by the Pauli temperature dependence common to fermions

p (T) = p (0)[1 + (1/6) 2 0 k B T 2 ] ,

(7.6)

where p (0) is the magnetic susceptibility at T = 0 K with


allowance for the Stoner enhancement, 0 is a constant which
depends on the density of states on the Fermi level N(E F) and also
its first and second derivatives with respect to energy.
As the temperature is raised, the susceptibility of Pd passes
through a pronounced maximum in the range of 50 to 100 K and
then rapidly decreases [116, 117]. The authors of [3941, 113]
investigated the high-temperature region T 300 K, in which the
313

Nanocrystalline Materials

contribution (7.6) to susceptibility decreases continuously and there


is no maximum on the experimental dependence (T).
Statistical processing results for temperature dependences of the
magnetic susceptibility (T) for starting and submicrocrystalline
palladium, measured in [3941, 113], shows that at T 775 K these
curves are adequately described by the Curie relation
=

C 1 N A B2 p 2
=
,
T 3 k BT

(7.7)

where is the magnetic susceptibility of the mass unit, C is the


Curie constant, N A is the Avogardo number, B is the Bohr
magneton, is the density of the substance, and p is the effective
number of Bohr magnetons per atom. At T 775 K the Curie
constants for starting and submicrocrystalline Pd coincide within the
limits of the measurement error and equal 194510 cm 3 K g 1 . The
effective magnetic moment, calculated from the Curie constant C
using equation (7.7), was eff = p B = 0.44 B . The gradual and
continuous conversation of susceptibility from the Pauli dependence
to the Curie dependence, detected in a wide temperature range from
50 to 775 K, is associated with the thermal excitation of electrons
and smearing of the Fermi level. This conversation may be viewed
as an analogue of the transition in the distribution of electrons from
the FermiDirac statistics to the classical MaxwellBoltzmann
statistics. The susceptibility of Boltzmann electrons is not associated
with the density of states on the Fermi level, but is determined only
to the localised electron magnetic moments being non-compensated.
For this region, the magnetic susceptibility of Pd at high
temperatures is not sensitive to the effects influencing the density
of states. This is confirmed by the coincidence of the temperature
dependences of susceptibility (T) for the starting and
submicrocrystalline palladium at T > 775 K, detected in [3941,
113].
The most interesting result [3941, 113] is the large difference,
by 8 %, in the magnetic susceptibility of submicrocrystalline and
starting coarse-grained palladium, observed at 300 K. This
difference also remains after annealing submicrocrystalline Pd at
T < 825 K. According to [39, 40], the observed difference of
cannot be related with the variation of the volume content of the
grain boundaries and their transition from the stressed nonequilibrium state to the equilibrium state. Indeed, according to the
314

Effect of Grain Size and Interfaces on Properties of Bulk Nanomaterials

electron microscopic data and the results of measurement of


microhardness (see Fig. 7.4), the most considerable grain growth,
the decrease in the density of dislocations and grain boundary
relaxation take place after annealing at T < 800 K. The volume
fraction of the grain boundaries in submicrocrystalline palladium
varies by more than 10 times as a result of annealing in the
temperature region of 300 to 800 K, and the lattice dislocations
density changes by three orders of magnitude. However, this is not
reflected in the behaviour of susceptibility which starts to decrease
only at an annealing temperature of T > 810 K (Fig. 7.14, curve
1).
According to [3941], intragranular vacancy complexes are the
most probable defect type affecting the behaviour of the
susceptibility. In [118] it is found that in nanocrystalline n-Pd
(D = 510 nm) the vacancies agglomerate into complexes, which
are less mobile than monovacancies and may exist up to
temperatures higher than 400 K. The effect of intragranular vacancy
complexes on the magnetic susceptibility of submicrocrystalline
palladium may be a consequence of the change in the density of
electron states at the Fermi energy. As already mentioned, in
palladium the Fermi energy lies on the slope of a very narrow and
high peak of the density of states N(E) [115]. The appearance of
c v vacancies in the palladium lattice leads to emptying of n ec v states
in the conduction band (n e = 10 is the number of electrons in the
conduction band per one palladium atom). If c v vacancies are
introduced the Fermi energy decreases by E v, then the number of
empty states may be represented as
EF

EF Ev

N( E )dE = ne cv

(7.8)

In order to determine the value of E v , the authors of [40] used


the numerical data [115] for N(E F ), N'(E F ) and N''(E F ) and
expansion N(E) in a Taylor series to the terms of the second order.
According to calculation results, the Fermi energy must decrease
by E v = 0.014 eV for the enhancement of susceptibility and, hence,
the density of states by a factor of 1.08 at T = 0 K. (in [39, 40]
the given value E v = 0.14 eV is a misprint). Taking E v value into
account and using equation (7.7) it was found that the vacancy
concentration, that guarantees the required increase in N(E F ) and
315

Nanocrystalline Materials

a decrease of the Fermi energy by E v 0.014 eV, is equal to


0.003 vacancies per atom. This vacancy concentration, 0.3 at.%,
can be achieved quite easily by severe plastic deformation because
according to the equation [119]
c v [exp(e)1]10 4

(7.9)

the concentration of vacancies at e = 7 is considerably higher.


At T > 500 K, the smearing of the density of states near the
Fermi energy by the value ~k B T becomes comparable with the
width of the narrow peak of the density of states near the Fermi
energy. As a consequence of this, the effect of vacancies on the
magnetic susceptibility of Pd at high temperatures is very low. This
explains the absence of a drop in the temperature dependence (T)
of submicrocrystalline palladium.
Thus, the high susceptibility of submicrocrystalline palladium is
associated with the excess concentration of vacancy clusters. The
recovery of the susceptibility of submicrocrystalline palladium to the
values, characteristic of coarse-grained palladium, is caused by the
condensation of vacancies and the annealing of dislocation tangles
at T > 825 K. However, the results [120] on the positron lifetime
in submicrocrystalline palladium contradict to a certain extent the
conclusions [39, 40] on the effect of vacancy clusters on the
magnetic susceptibility of palladium. According to [120], the
positron lifetime spectrum of submicrocrystalline palladium contains
two components with lifetimes of 1 167 ps and 2 280330 ps.
The first component is characterised by a high intensity
(approximately 95 %) and the intensity of the second one is around
5 %. The value of lifetime 1 indicates the trapping and annihilation
positrons in lattice vacancies, the second component with a longer
lifetime 2 corresponds to the annihilation of the positrons in
vacancy clusters with a volume of 612 atoms. According to [120],
the vacancy clusters are observed in submicrocrystalline palladium
only to a temperature of T 455 K, and are annealed at higher
temperatures. Thus, although the results in [120] confirm the
assumption [40] on the presence of vacancy clusters, which affect
the electronenergy spectrum of submicrocrystalline palladium near
the Fermi energy, the problem of the temperature stability of these
complexes has not as yet been completely solved.
In recent years, there has been special interest in the producing
of submicrocrystalline titanium [121125]. It is expected that the
physical properties of submicrocrystalline and coarse-grained
316

Effect of Grain Size and Interfaces on Properties of Bulk Nanomaterials

titanium should be greatly differ, as reported for other substances


[1, 126, 127]. In fact, the application of severe plastic deformation
(equal-channel angular pressing and uniform forging) allows to
reduce the grain size in titanium to several hundreds nanometers
[121125] and allows to detect changes in the mechanical
properties. A large difference of magnetic susceptibility of coarsegrained and submicrocrystalline titanium is observed by the authors
of [42]. The starting material was hot-rolled polycrystalline titanium
with a hcp structure (a = 0.29494 nm, c = 0.46844 nm). The mean
grain size in starting Ti was 15 m, the density 4.505 g cm 3. The
content of impurities in starting titanium was (wt.%): 0.32 Al, 0.18
Fe, 0.12 O, 0.07 C, 0.04 N, and 0.01 H. To analyse the possible
effect of ferromagnetic impurities, reference specimen from nondeformed superpure titanium containing 1.7510 4 wt.% Fe,
110 6 wt.% Co and 1010 4 wt.% Ni was used.
Submicrocrystalline titanium was produced from titanium rod
with a diameter of 40 mm by two-stage processing. The first stage
of severe plastic deformation consisted of 8 consecutive cycles of
warm equal-channel angular pressing at a temperature 670720 K,
the angle between the channel was 90 [124, 128]. After each
cycle, the specimens were rotated through 90 around the
longitudinal axis, and after 4 cycles additionally through
180 around the axis normal to the longitudinal axis. After first
deformation stage, the specimens had a homogeneous structure in
both the longitudinal and cross sections. The mean grain size was
300 nm. In the second stage of processing, the deformed titanium
was subjected to cold rolling in circleovalrhombcircle gauges;
the achieved degree of deformation was 93 % [128]. Rolling was
carried out in the direction parallel to the pressing axis. Rolling
resulted in a decrease in the grain size from 300 nm to 150 nm.
Rollers and setup for equal-channel angular pressing were produced
from a ferromagnetic tool steel. Equal-channel angular pressing and
rolling were carried out in air over a short period of time (the
duration of each cycle did not exceed 30 s) and, consequently,
there was no oxidation or inflow of impurities into the volume of
the specimen as a result of diffusion. In addition, the specimens for
measurements were cut out from the centre of the volume of
deformed titanium so that particles of the roller material could not
penetrate into titanium. It should be mentioned that such cutting of
the specimen does not prevent the transfer of particles of die or
plunger into the specimen during equal-channel angular pressing
because there is constant transfer of the surface of the specimen,
317

Nanocrystalline Materials

which is in contact with die, inside the specimen. Thus, regardless


of the measures taken, a penetration of ferromagnetic impurities into
the paramagnetic specimen cannot be excluded.
Structural certification of the specimens was carried by two
methods: transmission electron microscopy (JEM-100B) of thin foils
and X-ray diffraction in CuKa 1,2 radiation with a determination of
diffraction reflection broadening. Microhardness measurements (with
a load of 100 g) were taken to analyse the thermal stability and
structural transformations during annealing.
Magnetic susceptibility was measured using the same procedure,
which was employed for studying the magnetic susceptibility of
submicrocrystalline palladium and copper [3640] and was
described previously (see also Section 5.4 and Fig. 5.13). Heating,
annealing and cooling of specimens were carried out in situ directly
in magnetic susceptibility balance. Annealing temperature was varied
from 300 to 1043 K in 50 K steps, holding time of the specimen
at annealing temperature was 1 h.
The microstructure of the produced submicrocrystalline titanium
in the cross and longitudinal sections is shown in Fig. 7.15. In the
cross section, the grains are equiaxial and their mean size is 150
nm; in the longitudinal section, grains have a subgrain structure and
are elongated in the rolling direction, which coincides with the
pressing direction.
Both the sections are characterised by the presence of high- and
low-angle grain boundaries and by a large lattice dislocation density
of 10 14 10 15 m 2 . The azimuthal smearing of spot reflections on
electron diffraction patterns indicates the presence of high internal
stresses. Analysis of the broadening of the diffraction reflections

Fig. 7.15. Microstructure of submicrocrystalline titanium produced by equal-channel


angular pressing and cold rolling [42]: transverse (a) and longitudinal (b) sections
in relation to the rolling axis. The rolling axis coincided with the pressing axis
and the titanium rod axis.

318

Effect of Grain Size and Interfaces on Properties of Bulk Nanomaterials


(10 10) and (2020) shows that the size of the regions of coherence

scattering is 7090 nm and the magnitude of elastic distortions,


<e 2 > 1/2 , in the cross section is 0.23 % [129]. X-ray investigation
also shows that as a result of deformation treatment,
submicrocrystalline titanium is characterised by a distinctive
crystallographic texture [130]: the basal plane (0001) is located
preferentially in the plane parallel to the pressing axis (Fig. 7.16).
The density of submicrocrystalline titanium was 0.2 % lower than
that of starting coarse-grained titanium. This indicates the absence
of any extensive porosity, characteristic of bulk specimens of
titanium, produced by compaction of nanopowders [131]. Regardless
of the high dislocation density in submicrocrystalline titanium, it may
be assumed that they introduce a negligibly small free volume into
the specimen. Scanning and transmission electron microscopy
confirm the absence of porosity in titanium. Therefore, reduced
density may be associated either with an increase of the interatomic
distances near the grain boundaries [131] or with vacancies and
vacancy clusters formed as a result of severe plastic deformation.
The results of measurement of the magnetic susceptibility of the
specimen after the first stage of deformation, i.e. after warm
multiple equal-channel angular pressing, are shown in Fig. 7.17. In
accordance with the results obtained by transmission electron
microscopy, the microstructure of the specimen is isotropic, the
grains are almost spherical, and the mean grain size is
approximately 300 nm. The temperature dependence (T) for

Fig. 7.16. Pole figures for three different sections of submicrocrystalline titanium
specimen [42]: (1) direction, which is perpendicular to the rolling direction, (2)
rolling direction. Directions [1 0 1 0] , [0 0 0 1] , and [1 0 11] are normal to the plane
of the figure.
319

Nanocrystalline Materials

VXEPLFURFU\VWDOOLQHVPF7L
7L
VXEPLFURFU\VWDOOLQH
FRDUVHJUDLQHG7L7L
FRDUVHJUDLQHG

 FP J



















L ,
Fig. 7.17. Magnetic susceptibility of starting coarse-grained and submicrocrystalline
titanium [42]: ( l ) and ( ) denote annealing dependences (300, T); ( ) and ( )
denote temperature dependences (T). The temperature range, in which irreversible
processes of structural relaxation and recrystallisation take place, is indicated by
the vertical dashed lines.

starting titanium in the studied temperature range is in good


agreement with literature data [132]. The susceptibility of
submicrocrystalline titanium at room temperature is 5 % higher than
that of starting titanium and is equal to 3.310 6 cm 3 g 1 . The
annealing dependence of susceptibility (300, T) of submicrocrystalline titanium remains unchanged up to a temperature of 673
K. At T > 673 K, a smooth irreversible decrease of susceptibility
to the values corresponding to starting titanium is observed. Similar
transition to lower values of susceptibility in the temperature range
673823 K is detected on the temperature dependence of
susceptibility (T). In the temperature range from 300 to 673 K the
susceptibility of submicrocrystalline titanium exceedes that of
starting titanium by the same value 0.1510 6 cm 3 g 1 . At T > 823
K the temperature dependences (T) of submicrocrystalline and
starting titanium coincide almost, i. e. complete recovery of the
values of magnetic susceptibility takes place.
A different situation is found in the case of the magnetic
susceptibility of submicrocrystalline specimens after the second
stage of plastic deformation, i.e. after rolling. Since rolling results
in anisotropy of the structure, two specimens are investigated. One
specimen is cut along the rolling axis and denotes as smc||-Ti. The
320

Effect of Grain Size and Interfaces on Properties of Bulk Nanomaterials

second specimen, denoted by smc-Ti, is cut across the rolling axis.


It is known that the magnetic susceptibility of titanium single crystal
has an anisotropy. For example, the susceptibility of pure titanium
at room temperature along the [0001] direction is 3.35
10 6 cm 3 g 1 , and across this direction it is 3.0710 6 cm 3 g 1 [133].
As a result of anisotropy, the difference in susceptibility for
different orientations of the specimen reaches 9%, i.e. it is
comparable with the effects found in submicrocrystalline titanium.
Comparison of Figs. 7.17 and 7.18 shows that the values of
susceptibility in the longitudinal and transverse sections of
submicrocrystalline titanium are higher than the susceptibility of
starting coarse-grained titanium. The difference between the
susceptibilities of the longitudinal and transverse directions (Fig.
7.18, smc||-Ti and smc-Ti specimens) corresponds to an anisotropy
of 9 %. This shows that the additional contribution to susceptibility,
appeared as a result of pressing and rolling, does not depend on the
crystallographic direction and anisotropy of the microstructure.
The recovery of susceptibility to the values of for coarsegrained titanium in the specimen without rolling takes place in the
temperature range 670820 K (Fig. 7.17), and for the specimen
with rolling it takes place in the temperature range 727935 K (Fig.
7.18). It should be mentioned that there is no complete recovery
of susceptibility for the rolled specimen. The region of the recovery




VXEPLFURFU\VWDOOLQH VPFB7L
_
VXEPLFURFU\VWDOOLQH VPF__7L

 FP  J

























L ,

Fig. 7.18. Magnetic susceptibility of submicrocrystalline titanium, cut out along


(smc||-Ti) and across (smcTi) rolling axis: (l) and () denote annealing dependences
(300, T); ( ) and ( ) denote temperature dependences (T). The temperature
range of structural relaxation and recrystallisation of submicrocrystalline titanium
is indicated by dashed lines.

321

Nanocrystalline Materials

temperature for the specimens with and without rolling is shifted to


high values in comparison with the recrystallisation temperature
range 573873 K, determined from microhardness measurements
(Fig. 7.5). This means that the variation of the susceptibility of
submicrocrystalline titanium as a result of deformation and
annealing is determined to a greater extent by the changes in the
electronic structure of titanium and is associated only indirectly with
recrystallisation.
In principle, there are two main mechanisms of increasing the
magnetic susceptibility of submicrocrystalline titanium: impurity
mechanism and internal mechanism.
The impurity mechanism may be realised by introducing
ferromagnetic and paramagnetic impurities during plastic
deformation. In addition to this, during plastic deformation the
impurities, presented in starting titanium, may precipitate or be
activated, as in the case of submicrocrystalline copper [37]. The
observed increase of the susceptibility of submicrocrystalline
titanium by 5 % cannot be explained by intensive mass transfer of
the paramagnetic substance from the rollers or die to the specimen.
If it is assumed that the paramagnetic susceptibility of the
transferred substance is twice that of titanium, the observed
increase of may take place when 5 wt.% of this substance is
transferred into the specimen. Since annealing of submicrocrystalline titanium without rolling leads to the disappearance of the
addition to the susceptibility without any change of the mass of the
specimen, this impurity mechanism can be unambiguously excluded.
The submicrocrystalline titanium, subjected to rolling, does not show
any disappearance of the additional susceptibility after annealing.
In this specimen, in addition to the increase in susceptibility as the
result of the change in the structural state, there is also an impurity
contribution to the susceptibility.
Let us consider the internal mechanism of increasing
susceptibility. This mechanism cannot be realised by a means of
changes in the paramagnetic Pauli contribution, as takes place in
submicrocrystalline palladium [113], because electron spectrum of
titanium has no singularities near the Fermi energy. The absence
of this contribution is also confirmed by the slope of the
temperature dependences of susceptibility of submicrocrystalline
specimens (Fig. 7.17 and 7.18); for all cases the slope corresponds
to the Pauli paramagnetic contribution of pure non-deformed
titanium. According to [42], the most probable reason for the
increase of the magnetic susceptibility after plastic deformation is
322

Effect of Grain Size and Interfaces on Properties of Bulk Nanomaterials

the variation of Van-Fleck paramagnetism. Van-Fleck


paramagnetism is independent of temperature and appears as a
result of disruption of the symmetry of electron shells of atoms due
to structural heterogeneity, stresses and distortions of the crystal
lattice. All these factors are found in titanium after plastic
deformation. The recovery of susceptibility at recrystallisation
temperatures when the structural heterogeneity, stresses and
distortions are no longer found in submicrocrystalline titanium, is
indirect confirmation of the relationship of the discussed change of
with Van-Fleck paramagnetism.
The majority of studies into the magnetic properties of bulk
nanocrystalline materials have been carried out on ferromagnetic
metals and alloys.
Saturation magnetisation I s , Curie temperature T C and coercive
force H c of nanocrystalline compacted nickel nc-Ni (D = 10 nm)
and single crystal of nickel are studied in [134]. Magnetic
measurements were taken in the temperature from 5 to 680 K in
magnetic field up to 5.5 T. At T < 300 K the magnetisation of
nc-Ni is lower than I s of the Ni single crystal. Measurements of
the temperature dependence of magnetisation of nc-Ni in a field of
0.17 T show that at temperature of 510545 K magnetisation
rapidly decreased by ~20 %. The authors of [134] assume that this
decrease is associated with the ferromagneticparamagnetic
transition in the interfacial substance and, consequently, the Curie
temperature T C for the grain-boundary phase of nc-Ni is equal to
545 K. Further heating of nc-Ni is accompanied by grain growth
from 15 to 48 nm and leads to a decrease or complete
disappearance of magnetisation at a Curie temperature of 630 K,
corresponding to T C of coarse-grained nickel. In cooling from 650
to 450 K, the transition to the ferromagnetic state takes place at
630 K and then magnetisation smoothly increases without any
special features. The values of magnetisation in cooling are higher
than that of in heating of nc-Ni. According to [134], at T = 0 K
magnetisation of nc-Ni was 0.52 B per atom in contrast to coarsegrained nickel, for which I s (0) = 0.6 B atom 1 . Coercive force H c
of nc-Ni in the temperature range from 100 to 300 K remains
unchanged and equal to ~10 Oe. The most interesting result [134]
is the decrease of the Curie temperature of nc-Ni. Therefore, the
same group of authors investigated later [135] again the magnetic
properties of nc-Ni, paying attention to oxygen impurity. They found
that the previously detected non expected decrease of the
magnetisation of the nc-Ni at 510545 K is due to the presence of
323

Nanocrystalline Materials

oxygen impurity and, consequently, there is no justification for


changes of the Curie temperature of nanocrystalline nickel. In fact,
rough estimation shows that every impurity atom of oxygen
decreases magnetic moment of the Ni particle by the value
corresponding to the magnetic moment of a single atom of crystal
nickel.
The decrease of T C of nc-Ni (D = 70100 nm) by 3040 K in
comparison with conventional coarse-grained nickel was reported
by the authors of [90]. This result was obtained by scanning
calorimetry and by measurements of the temperature dependence
of saturation magnetisation. Measurements of the dependence I s (T)
show that at 300 K the magnetisation of nc-Ni (D = 100 nm) is
~10 % lower than that of coarse-grained nickel (D 1 m).
Compacted specimens of nc-Ni were produced in [90] by pressing
the nanopowder in air and, finally, contained a large amount of
oxygen. Taking into account the results [135] and the high sensitivity
of the magnetic properties of nickel to the oxygen impurity [136],
it may be assumed that the effects found in [90], are a consequence
of the contamination of nanocrystalline nickel with oxygen and are
not associated directly with the nanocrystalline state of nickel.
A small (by ~3 %) decrease of the magnetisation of submicrocrystalline nickel, produced by torsion under quasi-hydrostatic pressure,
was noted in [137]. The grain size of submicrocrystalline Ni was
assumed to be equal to 100200 nm. The temperature dependence
(T) of submicrocrystalline nickel is characteristic of ferromagnetics.
During the first heating of plastically deformed nickel specimens, the
decrease of susceptibility on approaching the Curie temperature was
smooth. In efficiently annealed specimens of submicrocrystalline nickel
the approach to T C was accompanied by a rapid decrease of , the
same as in the case of non-deformed nickel in the ferromagnetic
paramagnetic transition range. After consecutive annealing, the
susceptibility of deformed nickel increases to the values of
corresponding to starting non-deformed nickel. The observed anomalies
of the magnetic properties of submicrocrystalline nickel are explained
by the authors of [137] by the fact that part of the smallest crystallites
of submicrocrystalline nickel are in the superparamagnetic state. In this
case, submicrocrystalline nickel should be treated as a heterogeneous
material whose susceptibility is the superposition of the susceptibilities
of the ferromagnetic and superparamagnetic components.
The explanation proposed in [137] is relatively doubtful. The
authors of [137] assume that the grain boundaries are in the
amorphous state and form a paramagnetic shell around the grains
324

Effect of Grain Size and Interfaces on Properties of Bulk Nanomaterials

of nickel. This shell isolates the grains from each other. However,
the results of a large number of experiments show that the
interfaces even in materials with considerably smaller grains retain
the crystalline structure (see Section 6). In addition to this,
amorphous nickel is in the paramagnetic state only at T > 530 K
and can be hardly an efficient magnetic insulator at a mean grain
boundary width of 3 nm. The superparamagnetic behaviour of Ni
may be observed on particles smaller than 1015 nm [138] (see
also Section 5.4). In submicrocrystalline nickel with a mean grain
size greater than 100 nm, the fraction of grains with D < 10 nm
is negligible. It is shown in [135] that magnetisation of
nanocrystalline nickel, containing impurity oxygen, increases after
annealing. The comparison of the results [135] with the data [137]
on a variation of the magnetic susceptibility of submicrocrystalline
nickel after several consecutive annealing cycles indicate that the
effects, observed in [137], are associated with the presence of
oxygen impurity in submicrocrystalline nickel. Contamination of
nickel with oxygen may have taken place during plastic deformation,
which was carried out in air.
Investigation of the temperature dependence of coercive force
H c of NiCu alloys [139] showed that severe plastic deformation
has no effect on the Curie temperature of nickel, whereas the
coercive force H c of submicrocrystalline NiCu alloy and Ni is
several times higher than H c of the starting alloy and coarsegrained nickel.
Investigations of the microstructure and magnetic hysteresis of
submicrocrystalline Ni and Co [140, 141] corroborate that the
coercive force of the plastically deformed ferromagnetics is several
times higher than that of starting metals. It is shown in [140] that
the annealing of submicrocrystalline nickel at T 470 K decreases
the coercive force with the grain size remaining almost unchanged.
Annealing at higher temperature is accompanied by a simultaneous
decrease of H c and an increase in the grain size. This shows that
the high coercive force of the submicrocrystalline metals and alloys
is due equally to the stressed non-equilibrium state of the interfaces,
on the one hand, and to the small grain size, on the other hand.
Relaxation of the interfaces as a result of annealing and grain
growth decreases the value of H c .
In [142] the structure of the interfaces in nc-Fe (D = 10-15 nm)
was studied by the methods of the magnetic after-effect and
magnetic saturation. The magnetic after-effect is the time
dependence of the magnetic susceptibility after demagnetising.
325

Nanocrystalline Materials

Annealing of nc-Fe at T = 350500 K resulted in irreversible


changes in the magnetic after-effect spectrum; at the same time,
the time dependence of the magnetic after-effect was observed.
According to [142], similar changes are caused by the redistribution
of the atoms and redistribution is associated with a decrease in free
volumes at the interfaces. After annealing nc-Fe at 600 K, the
magnetic moment per single iron atom at a temperature of 5 K,
increased from 2.0 B to 2.2 B , i.e. to the value corresponding to
coarse-grained -Fe. This means that in nc-Fe the local distribution
of the atoms at the interfaces slightly differs from that in coarsegrained iron.
In the last decade, special interest was attracted by
ferromagnetic amorphous alloys (metallic glasses) based on Fe with
additions of Nb, Cu, Si, B and based on Co or FeCo with additions
of Si and B, and also by alloys of the systems FeMC, CoM
C, NiMC (M = Zr, Hf, Nb, Ta). Crystallisation of these
amorphous materials yields nanocrystalline alloys with a grain size
of 8 to 20 nm. Alloys obtained characterise by unique magnetic
properties. Changes in the structure and properties of these and
other metallic alloys, associated with transition from the amorphous
state to the nanocrystalline state, are discussed in particular in
[143]. Crystallisation of the amorphous alloys takes place at a low
mobility of the atoms, which favours to a greater extent the
formation of crystallites than their growth, i.e. it supports the
formation of the nanocrystalline structure.
The nanocrystalline alloys of the FeCuNbSiB system (for
example, Fe 73.5 Cu 1 Nb 3 Si 13.5 B 9 ), which are called FINEMET, are
best known. These alloys are soft magnetic materials with a very
low coercive force, comparable with H c of amorphous alloys based
on cobalt, and by high magnetic saturation close to that in Fe-based
amorphous alloys [24, 144].
The development of the nanostructure in the amorphous alloy
assumes a combination of a high rate of formation of crystallisation
centres and a low rate of growth of such centres. In FeCuNbSiB
alloys, the presence of copper in FeCuNbSiB alloys increases
the number of crystallisation centres and favours their uniform
distribution over the volume, Nb inhibits the growth of grains, and
Si supports the formation of the bcc phase Fe(Si). Annealing of
the amorphous alloy at 740820 K leads to the precipitation of
crystallites of an ordered solid solution -Fe(Si). The crystallites
of -Fe(Si) have a size of 1015 nm, contain up to 1319 at.% of
Si and are separated by a thin layer of the amorphous phase (Fig.
326

Effect of Grain Size and Interfaces on Properties of Bulk Nanomaterials

7.19) [144, 145]. The presence of copper decreases the activation


energy of crystallisation and facilitates the nucleation of the bcc
phase -Fe(Si) [146]. Crystallisation at a higher annealing
temperature leads to the formation of boride phases. Precipitation
of one or another phase depends on the relation between the
annealing temperature and annealing time: the amount of the bcc
phase increases and that of the amorphous phase decreases as the
annealing temperature and annealing time increase. The highest
magnetic permeability p and largest magnetic saturation was found
for alloys with a high content of the bcc phase. These alloys are
prepared by annealing at 780820 K for 1 hour.
In conventional ferromagnetic alloys, the grain growth leads to
a decrease in coercive force. According to [144], the coercive force
for the nanocrystalline alloys of the systems FeCuMSiB (M =
Nb, Ta, W, Mo, Zr, V) is proportional to the square of the grain
size, i.e. H c ~ D 2 . The Fe 73.5 Cu 1 Nb 3 Si 13.5 B 9 alloy with a mean grain
size of ~10 nm has a very low coercive force H c 0.5 A m 1 . The
high sensitivity of magnetic permeability, coercive force, saturation
magnetisation, magnetostriction and other magnetic properties to the
alloy microstructure was the reason for intensive investigations of
the conditions of crystallisation of amorphous alloys [144156].
In [25, 157] it is shown that preliminary deformation (~6%) of
a ribbon of Fe 73.5 Cu 1 Nb 3 Si 13.5 B 9 amorphous alloy by rolling with
subsequent annealing for 1 hour at 813820 K leads to an additional
decrease of the grain size from 810 to 46 nm. Low-temperature
annealing of the alloy at 723 K for 1 hour and subsequent shortterm annealing for 10 seconds at 923 K resulted in the formation
of a nanocrystalline structure with a grain size of 45 nm. The
phase composition of the alloys, produced by these methods, was
the same as after conventional crystallisation at 810820 K.
Other soft magnetic nanocrystalline alloys, produced by
crystallisation from the amorphous state, are also available. Alloys
FeMF, FeMB, FeMN and FeMO (M = Zr, Hf, Nb, Ta,
Ti) (Fig. 7.19) with a mean grain size of 10 nm have a saturation
magnetisation of 1.51.7 T, permeability p = 40005000 and low
(< 10 6) magnetostrictions [144, 158160].
Investigation of FeMB (M = Zr, Hf, Nb) nanocrystalline alloys
showed [161] that their magnetic properties can be improved by an
increase of the heating rate to temperature at which crystallisation
annealing is carried out. For example, the magnetic permeability p
of Fe 90Nb 7 B 3 alloy, annealed for crystallisation at 923 K for 1 hour,
was 2400 and 29000 at a heating rate of 0.008 and 3.3 K s 1 ,
327

Nanocrystalline Materials

Fig. 7.19. Schematic image of microstructure of typical nanocrystalline soft magnetic


alloys produced by crystallisation from the amorphous phase [144].

respectively; the coercive force of this alloy at the same heating


rates was 20 and 5 A m 1 , respectively. The high heating rate
resulted in a narrow size distribution of the grains of the precipitated
highly dispersed bcc phase and the decrease of the mean grain size:
at a heating rate of 0.042 K s 1 the mean grain size of the bcc
phase was 19.8 nm, and at a heating rate of 3.3 K s 1 it was
already 13.3 nm. The coercive force H c of nanocrystalline alloys
Fe 7893 M 511 B 211 (M = Zr, Hf, Nb) with the grain size D < 35 nm
is proportional to the D 5/2 and rapidly increases with increasing
grain size; at 35 D 100 nm coercive force H c 12001300 A
m 1 and is independent of the grain size. Small amounts of alloying
additions to FeMB alloys leads to an additional improvement of
the magnetic properties [161]. For example, the magnetic
permeabiliy of Fe 84 Nb 7 B 9 alloy, annealed at 923 K for 1 hour at a
heating rate of 3.3 K s 1 , was ~34000; the addition of a small
amount of gallium to this alloy and crystallisation in the same
conditions resulted in the formation of Fe 83 Nb 7 B 9 Ga 1
nanocrystalline alloy with a magnetic permeability of 38000 at a
frequency of 1 kHz.
The soft magnetic properties of alloys of FeCuNbSiB and Fe
MB systems (where M = Zr, Hf, Nb or Ta) produced by rapid
quenching of ribbons, are unstable at high temperatures. At the same
time, in recording magnetic heads it is necessary to use thin-film soft
magnetic materials with thermal stability sufficient to retain their
properties during high temperature joining with the substrate. These
requirements are satisfied by nanostructured composite materials of
the FeMC, CoMC and NiMC systems (M is the transition
metal of group IV or V) [143, 159, 162]. Films of amorphous alloys
328

Effect of Grain Size and Interfaces on Properties of Bulk Nanomaterials

are deposited by spraying and then crystallised at ~700 K. The


obtained nanocrystalline metallic matrix consists of Fe grains with a
size of ~10 nm and nanoparticles of MC carbide with a size of
~14 nm are distributed in matrix (mainly in triple points) [163]. Most
attention has been given to the alloys of the FeTaC system, which
have high thermal stability and retain the nanocrystalline structure up
to 1000 K. For comparison, it should be mentioned that in
nanocrystalline nickel, which does not contain carbide, grain growth
already starts at ~350 K [164]. The high thermal stability of the
nanostructure is caused by the pinning of the grain boundaries of Fe by nanoparticles of tantalum carbide TaC. In crystallisation of Fe
TaC alloys crystallites of -Fe are form in the first stage; during
growth of these crystallites, tantalum and carbon are precipitated into
the amorphous phase and form (at the equiatomic ratio) dispersed
nanosized precipitates of stoichiometric carbide TaC. If there is an
excess of tantalum or carbon, other compounds may form. According
to [165], the best soft-magnetic properties were obtained for
Fe 81.4 Ta 8.3 C 10.3 alloy.
Crystallisation of the amorphous alloys makes it possible to
produce not only soft-magnetic nanomaterials but also hardmagnetic nanocrystalline materials with high coercive force. In
[157] it is shown that annealing of amorphous soft-magnetic (H c
40 A m 1 ) alloys Fe 81 Si 7 B 12 and Fe 60 Cr 18 Ni 7 Si x B 15x (x = 3 or 5) for
1 hour at 823 K increases H c by a factor of 125 to 700.
Crystallisation for 1 hour at 873 K or rapid crystallisation for 10
seconds at 923 K of the amorphous soft-magnetic alloy
Fe 5 Co 70 Si 15 B 10 with H c < 1 A m 1 makes it possible to produce
nanocrystalline alloys with a mean grain size of 50200 nm and
H c = 3200 A m 1 , or with a mean grain size of 1550 nm and
H c = 8800 A m 1 , respectively [23, 166]. These nanocrystalline
alloys have a high residual magnetization. Increase in the coercive
force is associated with the precipitation of highly dispersed
crystalline phases; cubic -Co phase has the highest coercive force.
The high-coercivity state of the alloy, produced by rapid
crystallisation, is thermally stable and remains unchanged after
annealing at 673 K. According to [23, 166] the increase of the
coercive force of a rapid crystallised alloy in comparison with that
of a slowly crystallised alloy is a consequence of the precipitation
of anisotropic single-domain bcc particles of -Fe with high
saturation magnetisation, on the one hand, and of a decrease in the
grain size of -Co, on the other hand. If the grain size is smaller
than the width of the magnetic domain wall, the increase in the
329

Nanocrystalline Materials

residual magnetisation of the nanocrystalline alloy may be caused


by exchange interaction between the magnetic moments of the
grains.
References
1.

2.

3.

4.
5.
6.
7.

8.
9.
10.

11.
12.
13.

14.

15.

16.
17.

A. I. Gusev. Effects of the nanocrystalline state in solids. Uspekhi Fiz.


Nauk 168, 55-83 (1998) (in Russian). (Eng. transl.: Physics - Uspekhi 41,
49-76 (1998))
R. A. Andrievski, A. M. Glezer. Size effects in nanocrystalline materials:
I. Structure charac-teristics, thermodynamics, phase equilibria, and transport phenomena. Fiz. Metal. Metalloved. 88, No 1, 50-73 (1999); Size effects
in nanocrystalline materials: II. Mechanical and physical properties. Fiz.
Metal. Metalloved. 89, No 1, 91-1 12 (2000) (in Russian)
E. O. Hall. The deformation and aging of mild steel. II. Characteristics of
the Lders deformation. Proc. Phys. Soc. (London) B 64, 742-747 (1951);
The deformation and aging of mild steel. III. Discussion of results. Proc.
Phys. Soc. (London) B 64, 747-753 (1951)
N. J. Petch. The cleavage strength of polycrystals. J. Iron Steel Inst. 174,
25-28 (1953)
D. Tabor. The Hardness of Metals (Oxford University Press, London 1951)
175 pp.
R. L. Coble. A model for boundary diffusion controlled creep in polycrystalline
materials. J. Appl. Phys. 34, 1679-1682 (1963)
A. H. Chokshi, A. Rosen, J. Karch, H. Gleiter. On the validity of the HallPetch relationship in nanocrystalline materials. Scripta Metall. 23, 16791683 (1989)
K. Lu, W. D. Wei, J. T. Wang. Microhardness and fracture properties of
nanocrystalline NiP alloy. Scripta Metall. Mater. 24, 23 19-2323 (1990)
T. Christman, M. Jam. Processing and consolidation of bulk nanocrystalline
titanium aluminide. Scripta Metall. Mater. 25, 767-772 (1991)
H. Chang, H. J. Hofler,C. J. Altstetter, R. S. Averback. Synthesis, processing
and properties of nanophase titanium aluminide TiA1. Scripta Metall. Mater.
25, 1161-1 166 (1991)
K. Kim, K. Okazaki. Nanocrystalline consolidation of MA powders by EDC.
Mater. Sci. Forum 88-90, 553-560 (1992)
G. W. Neiman, J. R. Weertman, R. W. Siegel. Mechanical behavior of
nanocrystalline Cu and Pd. J. Mater. Res. 6, 1012-1027 (1991)
J. S. C. Jang, C. C. Koch. The HallPetch relationshiop in nanocrystalline iron produced by ball milling. Scripta Metall. Mater. 24, 1599-1604
(1990)
S. K. Ganapathi, M. Aindow, H. L. Fraser, D. A. Rigney. A comparative
study of the nanocrystalline material produced by sliding wear and inert
gas condensation. Mater. Res. Soc. Symp. Proc. 206, 593-598 (1991)
D. Hughes, S. D. Smith, C. S. Pande, H. R. Johnson, R. W. Armstrong.
Hall-Petch strenghening for the microhardness of twelve nanometer grain
diameter electrodeposited nickel. Scripta Metall. 20, 93-97 (1986)
G. E. Fougere, J. R. Weertman, R. W. Siegel. On the hardening and and softening
of nanocrystalline materials. Nanostruct. Mater. 3, 379-384 (1993)
R. W. Siegel, G. E. Fougere. Mechanical properties of nanophase metals.
330

Effect of Grain Size and Interfaces on Properties of Bulk Nanomaterials

18.
19.
20.

21.

22.
23.

24.

25.

26.

27.

28.

29.

30.

31.
32.

33.

34.

Nanostruct. Mater. 6, 205-216 (1995)


C. Suryanarayana, D. Mukhopadhyay, S. N. Patankar, F. H. Froes. Grain
size effects in nanocrystalline materials. J. Mater. Res. 7, 2114-2118 (1992)
H. Hahn, K. A. Padmanabhan. Mechanical response of nanostructured materials.
Nanostruct. Mater. 6, 191-200 (1995)
V. M. Segal. Methods of investigation of strain-deformed state in processes
of plastic change of shape. Ph. D. Thesis (Physico-Technical Institute of
Acad. Sci. of Belorus. SSR, Minsk 1974) 30 pp. (in Russian)
R. Z. Valiev, R. Chmelik, F. Bordeaux, G. Kapelski, B. Baudelet. The HallPetch relation in submicrograined Al1.5 % Mg alloy. Scripta Metall. Mater.
27, 855-860 (1992)
J. Kwarciak, L. Pajak, J. Lelatko. Crystallization kinetics of iron-cobaltsilicon-boron (Fe,Co) 78 Si 9 B 13 glasses. Z. Metallkunde 79, 712-715 (1988)
N. I. Noskova, E. G. Ponomareva, V. A. Lukshina, A. P. Potapov. Effect
of rapid crystallization on the properties of Fe 5 Co 70 Si 15 B 10 . Nanostruct.
Mater. 6, 969-972 (1995)
Y. Yoshizawa, S. Oguma, K. Yamauchi. New iron-based soft-magnetic alloys composed of ultrafine grain structure. J. Appl. Phys. 64, 6044-6046
(1988)
N. I. Noskova, E. G. Ponomareva. Structure, strength, and plasticity of
nanophase Fe 73 . 5 Cu 1 Nb 3 Si 13 . 5 B 9 alloy: I. Structure. Fiz. Metal. Metalloved.
82, No 5, 163-171 (1996) (in Russian). (Eng. transl.: Phys. Metal. Metallogr.
82, 542-548 (1996))
N. I. Noskova, E. G. Ponomareva, M. M. Myshlyaev. Structure of nanophases
and interfaces in multiphase nanocrystalline Fe 73 Ni 05 Cu 1 Nb 3 Si 135B 9 alloy and
nanocrystalline copper. Fiz. Metal. Metalloved. 83, No 5, 73-79 (1997)
(in Russian). (Eng. transl.: Phys. Metal. Metallogr. 83, 511-515 (1997))
N. I. Noskova, E. G. Ponomareva, V. N. Kuznetsov, A. A. Glazer, V. A.
Lukshina, A. P. Potapov. Crystallization of amorphous PdCuSi alloy
in creep condition. Fiz. Metal. Metalloved. 77, No 5, 89-94 (1994) (in Russian)
N. I. Noskova, E. G. Ponomareva, I. A. Pereturina, V. N. Kuznetsov. Strength
and plasticity of a PdCuSi alloy in the amorphous and nanocrystalline
states. Fiz. Metal. Metalloved. 81, No 1, 163-170 (1996) (in Russian). (Engl.
transl.: Phys. Metal. Metallogr. 81, 110-115 (1996))
T. Yamasaki, P. SchioBmacher, K. Ehrlich, Y. Ogino. Nanocrystallization
and mechanical properties of an amorphous electrodeposited Ni 75 W 25 alloy. Materials Science Forum 269-272, 975-980 (1998)
A. Inoue, H. M. Kimura, K. Sasamori, T. Masumoto. Structure and mechanical strength of AlVFe melt-spun ribbons containing high volume fraction
of nanoscale amorphous particles. Nanostruct. Mater. 7, 363-382 (1996)
R. Ray. High strength microcrystalline alloys prepared by devitrification
of metallic glass. J. Mater. Sci. 16, 2924-2927 (1981)
5. K. Das, K. Okazaki, C. M. Adam. Applications of rapid solidification
processing to high-temperature alloy design. In: High Temperature Alloys
- Theory and Design. Ed. J. O. Stiegler (TMS, Warrendale 1985) pp.451471
L. Arnberg, E. Larsson, S. Savage, A. Inoue, S. Yamaguchi, M. Kikuchi. New
heat-resistant tool materials produced from devitrified amorphous FeCr
MoCB and FeCrMoCV powders. Mater. Sci. Engineer. A 133, 288291 (1991)
A. Inoue, T. Shibata, T. Masumoto. Increase in mechanical strength of Ni-

331

Nanocrystalline Materials

35.

36.

37.
38.

39.

40.
41.

42.

43.
44.

45.

46.
47.
48.
49.
50.
51.

Si-B amorphous alloys by dispersion of nanoscale fcc-Ni particles. Mater.


Trans. Japan. Inst. Met. 33, 491-496 (1992)
N. A. Akhmadeev, R. Z. Valiev, V. I. Kopylov, R. R. Mulyukov. Formation of submicron-grained structure in copper and nickel with the use of
severe shearing strain. Metally No 5, 96-101 (1992) (in Russian)
A. A. Rempel, A. I. Gusev, S.Z. Nazarova, R. R. Mulyukov. Imputity
superparamagnetism in plastically deformed copper. Doklady Akad. Nauk
347, 750-754 (1996) (in Russian). (Engl. transl.: Physics - Doklady 4l, 152156 (1996))
A. A. Rempel, S. Z. Nazarova, A. I. Gusev. Iron nanoparticles in severeplastic-deformed copper. J. Nanoparticle Research 1, 485-490 (1999)
A. A. Rempel, S. Z. Nazarova, A. I. Gusev. Intrinsic and extrinsic defects
in palladium and copper after severe plastic deformation. In: Structure and
properties of nanocrystalline materials. Eds. N. I. Noskova and G. G. Taluts
(Ural Division of the Russ. Acad. Sci., Yekaterinburg 1999) pp.265-278
A. A. Rempel, A. I. Gusev, R. R. Mulyukov, N. M. Amirkhanov. Microstructure and properties of palladium subjected to severe plastic deformation.
Metallofizika i Noveishie Tekhnologii 18, 14-22 (1996) (in Russian)
A. A. Rempel, A. I. Gusev. Magnetic susceptibility of palladium subjected
to severe plastic deformation. Phys. Stat. Sol. (b) 196, 251-260 (1996)
A. A. Rempel, A. I. Gusev, R. R. Mulyukov, N. M. Amirkhanov. Microstructure, microhardness and magnetic susceptibility of submicrocrystalline
palladium. Nanostruct. Mater. 7, 667- 674(1996)
V. V. Stolyarov, S. Z. Nazarova, A. A. Kilmametov, A. A. Rempel, A. I.
Gusev. Structure peculiarities and magnetic susceptibility of ultrafine-grained
titanium produced by equal-channel angular pressing with subsequent rolling.
In: The Problems of Nanocrystalline Materials. Eds. V. V. Ustinov and N.
I. Noskova (Ural Division of the Russian Academy of Sciences, Yekater-inburg
2002) pp.409-419 (in Russian)
A. A. Ivanko. Hardness (Naukova Dumka, Kiev 1968) 128 pp. (in Russian)
Y. Ishida, H. Ichinose, T. Kizuka, K. Suenaga. High-resolution electron
microscopy of inter-faces in nanocrystalline materials. Nanostruct. Mater.
6, 115-124 (1995)
T. Kizuka, Y. Nakagami, T. Ohata, I. Kanazawa, I. Ichinose, H. Murakami,
Y. Ishida. Structure and thermal stability of nanocrystalline silver studied
by transmission electron microscopy and positron annihilation spectroscopy.
Philosoph. Mag. A 69, 551-563 (1994)
L. He, E. Ma. Processing and microhardness of bulk CuFe nanocomposites.
Nanostruct. Mater. 7, 327-340 (1996)
T. Haubold, V. Gertsman. On the structure and properties of nanostructured
coppertungsten alloys. Nanostruct. Mater. 1, 303-312 (1992)
R. L. Holtz, V. Provenzano. Enhanced microhardness of copper-niobium
nanocomposites. Nanostruct. Mater. 4, 241-254 (1994)
K. Lu, M. L. Sui. An explanation to the abnormal Hall-Petch relation in
nanocrystalline materials. Scripta Metall. Mater. 28, 1465-1470 (1993)
T. Christman. Grain Boundary strengthening exponent in conventional and
ultrafine microstructures. Scripta Metall. Mater. 28, 1495-1500 (1993)
J. E. Carsley, J. Ning, W. W. Milligan, S. A. Hackney, E. C. Aifantis. A
simple, mixtures-based model for the grain size dependence of strength in
nanophase metals. Nanostruct. Mater. 5, 441-448 (1995)

332

Effect of Grain Size and Interfaces on Properties of Bulk Nanomaterials


52.
53.
54.

55.

56.

57.

58.
59.
60.

61.

62.

63.
64.

65.

66.

67.

68.

69.

H. Gleiter. Nanocrystalline materials. Progr. Mater. Sci. 33, 223-315 (1989)


R. A. Masumura, P. M. Hazzledine, C. S. Pande. Yield stress of fine grained
materials. Acta Mater. 46, 4527-4534 (1998)
N. A. Akhmadeev, R. Z. Valiev, N. P. Kobelev, R. R. Mulyukov, Ya. M.
Soifer. Elastic properties of copper with submicrocrystalline structure. Fiz.
Tverd. Tela 34, 3155-3160 (1992)
D. Korn, A. Morsch, R. Birringer, W. Arnold, H. Gleiter. Measurements
of the elastic con-stants, the specific heat, and the entropy of grain boundaries
by means of ultrafine grained materials. J. de Physique - Colloque C5 49,
Suppl. No 5, C5-769 C5-779 (1988)
K. Lu, J. T. Wang, W. D. Wei. Thermal expansion and specific heat capacity of nanocrystalline NiP alloy. Scripta Metal. Mater. 25, 619-623
(1991)
A. Inoue, H. M. Kimura, K. Sasamori, T. Masumoto. Ultrahigh strength
of rapidly solidified A1 68xCr 3Ce 1 Co x (x = 1, 1.5 and 2 %) alloys containing
an icosahedral phase as a main component. Mater. Trans. Japan. Inst. Met.
35, 85-94 (1994)
S. R. Agnew, J. R. Weertman. The influence of texture on the elastic properties
of ultrafine-grain copper. Mater. Sci. Engineer. A 242, 174-180 (1998)
A. M. Glezer, B. V. Molotilov. Structure and Mechanical Properties of
Amorphous Alloys (Metallurgiya, Moscow 1992) 208 pp. (in Russian)
S. R. Philipot, D. Wolf, H. Gleiter. Molecular-dynamics study of the synthesis
and characterization of a fully dense, three-dimensional nanocrystalhine material.
J. Appl. Phys. 78, 847-861 (1995)
A. B. Lebedev, Yu. A. Burenkov, V. I. Kopylov, V. P. Filonenko, A. E. Romanov,
V. G. Gryaznov. Recovery of Young modulus at annealing of polycrystalline
copper with ultrafine grain. Fiz. Tverd. Tela 38, 1775-1783 (1996) (in Russian)
A. B. Lebedev, S. A. Pulnev, V. I. Kopylov, Yu. A. Burenkov, V. V. Vetrov,
O. V. Vylegzhanin. Thermal stability of submicrocrystalline copper and Cu:
ZrO 2 composite. Scripta Mater. 35, 1077-1081 (1996)
R. Z. Vahiev, N. A. Krasilnikov, N. K. Tsenev. Plastic deformation of alloys with submicron-grain structure. Mater.Sci. Engineer. A 137, 35-40 (1991)
N. I. Noskova, E. G. Ponomareva, I. A. Pereturina. Structure, strength and
plasticity of nano-phase Fe 73~5 Cu 1 Nb 3 Si 13~5B 9 alloy. II. Strength and plasticity.
Fiz. Metal. Metalloved. 82, No 6, 116-121 (1996)
R. R. Mulyukov, N. A. Akhmadeev, S. B. Mikhailov, R. Z. Valiev. Strain
amplitude dependence of internal friction and strength of submicrometergrained copper. Mater. Sci. Engineer. A171, 143-149 (1993)
N. A. Akhmadeev, N. P. Kobelev, R. R. Mulyukov, Ya. M. Soifer, R. Z.
Valiev. The effect of heat treatment on the elastic and dissipative properties of copper with the submicrocrystalline structure. Acta Metal. Mater.
41, 1041-1046 (1993)
R. R. Mulyukov, N. A. Akhmadeev, R. Z. Valiev, V. I. Kopylov, S. B. Mikhailov.
Amplitude dependence of internal friction and strength of submicrocrystalline
copper. Metallofizika 15, 50-59 (1993)
R. R. Mulyukov, H.-E. Schaefer, M. Weller, D. A. Salimonenko. Internal
friction and superplasticity of submicrocrystalline metals. Materials Science Forum 170-172, 159-164 (1994)
R. R. Mulyukov, M. Weller, R. Z. Valiev, Th. Gessmann, H.-E. Schaefer.
Internal friction and shear modulus in submicrograined Cu. Nanostruct. Mater.
6, 577-580 (1995)

333

Nanocrystalline Materials
70.

71.

72.

73.

74.

75.

76.
77.
78.
79.

80.

81.
82.

83.

84.
85.
86.

87.

88.

R. Z. Valiev, A. V. Korznikov, R. R. Mulyukov. Structure and properties


of metallic materials with submicrocrystalline structure. Fiz. Metall. Metalloved.
73, No 4, 70-86 (1992) (in Russian)
R. Z. Valiev, A. V. Korznikov, R. R. Mulyukov. Structure and properties
of ultrafine-grained materials produced by severe plastic deformation. Mater.
Sci. Engineer. A 168, 141-148 (1993)
E. Hellstern, H. J. Fecht, F. Thu, W. L. Johnson. Structural and thermodynamic properties of heavily mechanically deformed Ru and AlRu. J. Appl.
Phys. 65, 305-310 (1989)
R. R. Mulyukov, S. B. Mikhailov, R. G. Zaripova, D. A. Salimonenko. Damping
properties of l8Cr10Ni stainless steel with submicrocrystalline structure.
Mater. Res. Bull. 31, 639-645 (1996)
R. R. Mulyukov, S. B. Mikhailov, R. G. Zaripova, D. A. Salimonenko. The
amplitude dependence of damping of 18-10 type stainless steel with
submicrocrystalline structure. Materials Science Forum 225-227, 787-792
(1996)
F. Wakai F. Superplasticity of ceramics. In: Ceramics Today - Tomorrows
Ceramics (Mater. Sci. Monogr. 66A, part A (1991)) Ed. P. Vincenzini (Elsevier
Sci. Publ., Amsterdam 1991) pp.61-76
C. E. Pearson. The viscous properties of extruded eutectic alloys of leadtin and bismuthtin. J. Inst. Metals 54, 111-124 (1934)
F. Wakai, S. Sakagushi, Y. Matsuno. Superplasticity of yttria-stabilized
tetragonal zirconia polyctystals. Advanced Ceram. Mater. 1, 259-263 (1986)
F. Wakai, S. Kodama, S. Sakagushi, N. Murayama, K. Izaki, K. Niihara.
A superplastic covalent crystal composite. Nature 344, 421-423 (1990)
O. D. Sherby, O. A. Ruano. Synthesis and characteristics of superplastic
alloys. In: Superplas-tic Forming of Structural Alloys. Eds. N. E. Paton,
C. H. Hamilton (Metall. Soc. of AIME, Warrendale 1982). pp.241-254
J.-G. Wang, R. Raj. Influence of hydrostatic pressure and humidity on
superplastic ductility of the f3-spodumene glass-ceramics. J. Amer. Ceram.
Soc. 67, 385-390 (1984)
J. Karch, R. Birringer, H. Gleiter. Ceramics ductile at low temperatures.
Nature 330, 556-558 (1987)
O. B. Naimark. Nanocrystalline state as a topological transition in an ensemble
of grain-boundary defects. Fiz. Metall. Metalloved. 84, 5-21 (1997) (in Russian).
(Engl. Transl.: Phys. Metal. Metallogr. 84, 327-337 (1997))
J. Rupp, R. Birringer. Enhanced specific-heat-capacity (c p ) measurements
(150-300 K) of nanometer-sized crystalline materials. Phys. Rev. B 15, 78887890 (1987)
A. Tschope, R. Birringer. On the origin of enhanced specific heat in
nanocrystalline platinum. Phil. Mag. B 68, 223-229 (1993)
T. Mtschele, R. Kirchheim. Segregation and diffusion of hydrogen in grain
boundaries of pal-ladium. Scripta Met. 21, 13-140 (1987)
U. Stuhr, H. Wipf, T. J. Udovic, J. Weismuller, H. Gleiter. Inelastic neutron scattering study of hydrogen in nanocrustalline Pd. Nanostruct. Mater.
6, 555-558 (1995)
J. A. Eastman, L. J. Thompson, B. J. Kestel. Narrowing of the palladiumhydrogen miscibility gap in nanocrystalline palladium. Phys. Rev. B 48,
84-92 (1993)
V. Novotny, P. P. M. Meincke, J. H. P. Watson. Effect of size and surface on the specific heat of small lead particles. Phys. Rev. Lett. 28,

334

Effect of Grain Size and Interfaces on Properties of Bulk Nanomaterials

89.
90.

91.

92.
93.

94.

95.
96.

97.
98.
99.
100.

101.
102.
103.
104.
105.
106.
107.

108.

109.

901-903 (1972)
J. Trampenau, K. Bauszus, W. Petry, U. Herr. Vibrational behavior of
nanocrustalline Ni. Nanostruct. Mater. 6, 551-554 (1995)
R. Z. Valiev, R. R. Mulyukov, Kh. Ya. Mulyukov, L. I. Trusov, V. I. Novikov.
Curie temperature and saruration magnetization of nickel with submicrongrained structure. Pisma v ZhTF 15, 78-8 1 (1989) (in Russian)
N. X. Sun, K. Lu. Heat capacity comparison among the nanocrystalline,
amorphous and coarse-grained polycrystalline states in elemental selenium.
Phys. Rev. B 54, 6058-6061 (1996)
K. Lu, R. Luck, B. Predel. Investigation of the heat capacities of Ni
20 % P in different states. Z. Metallkunde 84, 740-743 (1993)
Y. H. Zhao, K. Lu. Grain size dependence of thermal properties of
nanocrystalline elemental selenium studied by means of X-ray diffraction.
Phys. Rev. B 56, 14330-14337 (1997)
J. Wang, D. Wolf, S. R. Philipot, H. Gleiter. Phonon-induced anomalous
specific heat of a model nanoctystal bu computer simulation. Nanostruct.
Mater. 6, 747-750 (1995)
R. Pirc, A. Holz. Specific heat of nanostructured materials. Nanostruct.
Mater. 6, 755-758 (1995)
R. Birringer, H. Gleiter. Nanocrystalline materials. In: Encyclopedia of Material
Science and Engineering. Suppl. Vol.1. Ed. R. W. Cahn (Pergamon Press,
Oxford 1988) pp.339-349
Xing Thu. Structural investigation of nanocrystalline materials. PhD Thesis.
University of SaarbrUcken, Germany, Saarbrcken 1986. 77 pp.
H. J. Klam, H. Hahn, H. Gleiter. The thermal expansion of grain boundaries. Acta Metall. 35, 2101-2104 (1987)
M. L. Sui, K. Lu. Thermal expansion behavior of nanocrystalline NiP alloys
of different grain size. Nanostruct. Mater. 6, 65 1-654 (1995)
J. Weissmller, J. Lffler, M. Kleber. Atomic structure of nanocrystalline
metals studied by diffraction techniques and EXAFS. Nanostruct. Mater.
6, 105-114 (1995)
J. Weissmller. Alloy effects in nanostructures. Nanostruct. Mater. 3, 261272 (1993)
J. Weissmller. Alloy thermodynamics in nanocrystalline structures. J. Mater.
Res. 9, 4-9 (1994)
D. L. Beke. Cs. Cserhti, I. A. Szab. Segregation and phase separation
in nanophase materials. Nanostruct. Mater. 9, 665-668 (1997)
Cs. Cserhti, I. A. Szab, D. L. Beke. Size effect in surface segregation.
J. Appl. Phys. 83, 3021-3027 (1998)
F. Gartner, R. Bormann, R. Birringer, A. Tschope. Thermodynamic stability
of nanocrystalline palladium. Scripta Mater. 35, 805-8 10 (1996)
A. Tschpe, R. Birringer, H. Gleiter. Calorimetric measurements of the thermal
relaxation in nanocrystalline platinum. J. Appl. Phys. 71, 539 1-5394 (1992)
H. Y. Kai. Nanocrystalline materials: A study of their preparation and characterization. PhD Thesis. Universiteit van Amsterdam, Netherlands, Amsterdam 1993. 114 pp.
G. I. Frolov, V. S. Thigalov, A. I. Polskii, V. G. Pozdnyakov. Study of electrical
conductivity of nanocrystalline films of cobalt. Fiz. Tverd. Tela 38, 12081213 (1996) (in Russian)
R. K. Islamgaliev, N. A. Akhmadeev, R. R. Mulyukov, R. Z. Valiev. Influence of submicrocrystalline state on electrical resistivity of copper.

335

Nanocrystalline Materials

110.

111.
112.

113.

114.

115.
116.
117.

118.

119.
120.

121.

122.

123.

124.

125.

126.

Metallofizika 12, 317-320 (1990) (in Russian)


R. K. Islamgaliev, N. A. Akhmadeev, R. R. Mulyukov, R. Z. Valiev. Grain
boundary influence on the electrical resistance of submicron grained copper. Phys. Stat. Sol. (a) 118, K27-K29 (1990)
K. Pekala, M. Pekala. Low temperature transport properties of nanocrystalline
Cu, Fe and Ni. Nanostruct. Mater. 6, 819-822 (1995)
R. R. Mulyukov. Structure and properties of submicrocrystalline metals
produced by severe plastic deformation. PhD Thesis. Institute of Steel and
Alloys, Moscow 1997. 31 pp. (in Russian)
A. A. Rempel, A. I. Gusev, S. Z. Nazarova, R. R. Mulyukov. Magnetic
susceptibility of palladium subjected to plastic deformation. Doklady Akad.
Nauk 345, 330-333 (1995) (in Russian). (Engl. transl.: Physics - Doklady
40, 570-573 (1995))
K. L. Liu, A. H. Macdonald, J. M. Daams, S. H. Vosko, D. D. Koelling.
Spin density func-tional theory of the temperature-dependent spin susceptibility
by Pd and Pt. J. Magn. Magn Mater. 12, 43-47 (1979)
F. M. Mueller, A. J. Freeman, J. O. Dimmock, A. M. Furdyna. Electronic
structure of palladium. Phys. Rev. B 1, 4617-4635 (1970)
C. J. Kriessman, H. B. Callen. The magnetic susceptibility of the transition elements. Phys. Rev. 94, 837-844 (1954)
W. D. Weiss, R. Kohlhaas. Uber die Temperaturabhangigkeit der
Atomsuszeptibilitat von Ruthenium, Rhodium und Palladium sowie Osmium,
Iridium und Platin zwischen 80 und 1850 K. Z. Angew. Phys. 23, 175179 (1967)
H.-E. Schaefer, W. Eckert, O. Stritzke, R. Wrschum, W. Templ. The microscopic structure of nanocrystalline materials. In: Positron Annihilation.
Eds. L. Dorikens-Vanpraet, M. Dorikens and D. Segers (World Scientific
Publ.Comp., Singapore 1989) pp.79-85
Physical Metallurgy (4th edition). Eds. R. W. Cahn, P. Haasen. (North-Holland,
Amsterdam 1996) 2888 pp (in 3 vol.)
R. Wrschum, A. Kbler, S. Gruss, P. Scharwaechter, W. Frank, R. Z. Valiev,
R. R. Mulyukov, H.-E. Schaefer. Tracer diffusion and crystallite growth
in ultra-fine-grained Pd prepared by severe plastic deformation. Annales
de Chimie. Sci. des Mater. 21, 471-482 (1996)
S. P. Malysheva, R. M. Galeyev, G. A. Salishchev, R. R. Mulyukov,
O. R. Valiakhmetov. Influence of large plastic deformation and recrystallization
annealing on titanium density. Fiz. Metall. Metalloved. 82, No 2, 117-120
(1996) (in Russian)
G. A. Salishchev, R. M. Galeyev, S. P. Malysheva, S. B. Mikhailov, M.
M. Myshlyaev. Change of elastic modulus at annealing of submicrocrystalline
titanium. Fiz. Metall. Metalloved. 85, No 3, 178-181 (1998) (in Russian)
V. V. Stolyarov, Y. T. Zhu, T. C. Lowe, R. K. Islamgaliev, R. Z. Valiev. A
two step SPD processing of ultrafine-grained titanium. Nanostruct. Mater.
11, 947-954 (1999)
V. V. Stolyarov, Y. T. Zhu, I. V. Alexandrov, T. C. Lowe, R. Z. Valiev. Influence
of ECAP routes on the microstructure and properties of pure Ti. Mater.
Sci. Engineer. A 299, 59-67 (2001)
G. A. Salishchev, S. Yu. Mironov. Influence of grain size on mechanical
properties of technically pure titanium. Izv. Vuzov. Fizika No 6, 28-32 (2001)
(in Russian)
A. I. Gusev. Nanocrystalline Materials:Preparation and Properties (Ural

336

Effect of Grain Size and Interfaces on Properties of Bulk Nanomaterials

127.
128.

129.

130.

131.

132.
133.

134.
135.

136.
137.

138.

139.

140.

141.

142.

Division of the Russ. Acad. Sci., Yekaterinburg 1998) 200 pp. (in Russian)
A. I. Gusev, A. A. Rempel. Nanocrystalline Materials (Nauka-Fizmatlit,
Moscow 2000) 224 pp. (in Russian)
V. V. Stolyarov, Y. T. Zhu, I. V. Alexandrov, T. C. Lowe, R. Z. Valiev.
Microstructure and properties of ultrafine-grained pure titanium processed
by equal channel angular pressing and cold deformation. J. Nanosci. Nanotechn.
1, 237-242 (2001)
A. A. Kilmametov, V. V. Stolyarov, L. O. Shestakova, I. V. Alexandrov.
Structural peculiarities of nanostructured titanium produced by severe plastic
deformation. In: Structure and properties of nanocrystalline materials. Eds.
N. I. Noskova, G. G. Taluts (Ural Division of the Russ. Acad. Sci.,
Yekaterinburg 1999) pp.185-195 (in Russian)
V. V. Stolyarov, L. O. Shestakova, A. I. Zharikov, V. V. Latysh, R. Z. Valiev,
Y. T. Zhu, T. C. Lowe. Mechanical properties of nanostructured titanium
alloys processed using severe plastic deformation. In: Proc. of 9 th Conf.
Titanium-99. V.1. Eds. I. V. Gorynin and S. S. Ushkov (Nauka, Moscow
2000) pp.466-472 (in Russian)
V. V. Stolyarov, Y. T. Zhu, T. C. Lowe, R. K. Islamgaliev, R. Z. Valiev.
Processing, microstructure and mechanical properties of nanocrystalline Ti
and TiTiO 2 nanocomposites processed by SPTS consolidation. Mater. Sci.
Engineer. A 282, 78-85 (2000)
C. F. Squire, A. R. Kaufmann. Magnetic susceptibility of titanium and zirconium.
J. Chem. Phys. 9, 673-677 (1941)
Materials Properties Handbook: Titanium Alloys. Eds. R. Boyer, E. W.
Collings, G. Welsch (ASM International, Metals Park, Ohio 1994) 1169 pp;
U. Zwicker. Titan und Titanlegierungen (Springer-Verlag, Berlin-New York
1974) 717 pp.
H.-E. Schaefer, H. Kisker, H. Kronmller, R. Wrschum. Magnetic properties of nanocrystalline nickel. Nanostruct. Mater. 1, 523-529 (1992)
H. Kisker, T. Gessmann, R. Wrschum, H. Kronmller, H.-E. Schaefer. Magnetic
properties of high purity nanocrystalline nickel. Nanostruct. Mater. 6, 925928 (1995)
H. Kronmller. Theory of nucleation fields in inhomogeneous ferromagnets.
Phys. Stat. Sol. (b) 144, 385-396 (1987)
A. V. Korolev, A. I. Deryagin, V. A. Zavalishin, R. I. Kuznetsov. Peculiarities of magnetic state of highly deformed polycrystalline ultrafine-grained
nickel. Fiz. Metall. Metalloved. 68, 672-678 (1989) (in Russian)
A. Hahn. Untersuchungen an kleien superparamagnetischen Nickel krnern
zur Frage der Temperaturabhngigkeit der spontanen Magnetisierung. Ann.
Physik 11, 277-309 (1963)
A. V. Korolev, E. G. Gerasimov, B. I. Teitel, N. N. Snchegoleva, V. P. Pilugin,
R. I. Kuznetsoy. Peculiarities of magnetic state of plastically deformed NiCu alloys. Fiz. Metall. Metalloved. 70, No 11, 98-102 (1990) (in Russian)
Kh. Ya. Mulyukov, G. F. Korznikova, A. Z. Abdulov, R. Z. Valiev. Magnetic hysteresis prop-erties of submicron grained nickel and their variation upon annealing. J. Magn. Magn. Mater. 89, 207-213 (1990)
Kh. Ya. Mulyukov, G. F. Korznikova, R. Z. Valiev. Microstructure and magnetic
properties of submicron grained cobalt after large plastic deformation and
their variation during annealing. Phys. Stat. Sol. (a) 125, 609-614 (1991)
H. Kronmller, M. Hirscher, R. Reisser, R. Wrschum, H.-E. Schaefer.
Magnetic after-effect and approach to ferromagnetic saturation in nanoc-

337

Nanocrystalline Materials

143.

144.
145.

146.
147.

148.

149.

150.

151.

152.

153.
154.

155.

156.

157.

rystalline iron. J. Magn. Magn. Mater. 146, 117-122 (1995)


A. L. Greer. Changes in structure and properties associated with the transition
from the amorphous to the nanocrystalline state. In: Nanostrucrured Materials:
Science and Technology. Eds. G. M. Chow, N. I. Noskova. (Kluwer Academic Publishers, Netherlands, Dordrecht 1998) pp. 143-162
K. Yamauchi, Y. Yoshizawa. Recent development of nanocrystalline soft magnetic
alloys. Nanostruct. Mater. 6, 247-254 (1995)
B. V. Zhalnin, I. B. Kekalo, Yu. A. Skakov, B. V. Shelekhov. Phase transformations and mag-netic properties variation during nanocrystalline state
formation in amorphous iron base alloy. Fiz. Metall. Metalloved. 79, No
5, 94-106 (1995) (in Russian)
N. Wang, F. Zhu, P. Haasen. Twinned structure of Fe 2 B in an annealed
Fe 73.5Cu 1 Nb 3 Si 13.5 B 9 soft magnetic alloy. Phil. Mag. Lett. 64, 157-162 (1991)
N. I. Noskova, V. V. Serikov, A. A. Glazer, N. M. Kleynerman, A. P. Potapov.
Electron-microscope and Mssbauer studies of structure of Fe 73.5Cu 1Nb 3Si13.5B9
alloy in nanocrystalline state. Fiz. Metall. Metalloved. 74, No 7, 80-86
(1992) (in Russian)
A. A. Glazer, V. A. Lukshina, A. P. Potapov, N. I. Noskova. Nanocrystalline
alloy Fe 73.5 Cu 1 Nb 3 Si 13.5 B 9 prepared from amorphous state by quick crystallization at elevated temperatures. Fiz. Metall. Metalloved. 74, No 8, 96100 (1992) (in Russian)
K. Hono, K. Hiraga, Q. Wang, A. Inoue, T. Sakurai. The microstructure evolution
of a Fe 73.5Si 13.5B 9 Nb 3 Cu 1 nanocrystalline soft magnetic materials. Acta Metall.
Mater. 40, 2137-2147 (1992)
V. V. Sosnin, O. M. Zhigalina, A. L. Mironov, V. V. Sadchikov, Z. D. Kolev.
Study of crystallization process of amorphous alloys on base of FeCu
NbSiB system (Finemet type). I. Variation of crystallization activation
energy of FeCuNbSiB alloys in dependence on component concentration.
Fiz. Metall. Metalloved. 78, No 2, 140-144 (1994) (in Russian)
M. Baricco, C. Antonione, P. Allia, P. Tiberto, F. Vinai. Nanocrystalline
phase formation in amorphous Fe 73.5 Cu 1 Nb 3 Si 13.5 B 9 submitted to conventional annealing and Joule heating. Mater. Sci. Engineer. A 179-180, 572576 (1994)
N. Clavaguera, T. Pradell, Jie Thu, M. T. Clavaguera-Mora. Thermodynamic
and kinetic factors controlling the formation of nanocrystalline FeCuNbSiB
materials. Nanostruct. Mater. 6, 453-456 (1995)
C. F. Conde, A. Conde. Nanocrystallization in Fe 73.5Cu 1 Nb 3 (Si,B) 22.5 alloys:
influence of the Si/B content. Nanostruct. Mater. 6, 457-460 (1995)
S. Surinach, A. Otero, M. D. Baro, A. M. Tonejc, D. Bagovic.
Nanocrystallization of amorphous FeCuNbSiB based alloys. Nanostruct.
Mater. 6, 461-464(1995)
T. Graf, G. Hampel, J. Korus, J. Hesse, G. Herzer. Influence of Nb concentration on structure and crystallization onset of amorphous Fe(Cu,Nb)SiB
Finemet alloys. Nanostruct. Mater. 6, 469-472 (1995)
N. I. Noskova, E. G. Ponomareva, A. A. Glazer, V. A. Lukshina, A. P. Potapov.
Influence of preliminary deformation and low temperature annealing on the
size of nanocrystallites of Fe 73.5 Cu 1 Nb 3 Si 13.5 B 9 alloy obtained at crystallization of amorphous ribbon. Fiz. Metall. Metalloved. 76, No 5, 171-173
(1993) (in Russian)
N. I. Noskova, N. F. Vildanova, A. P. Potapov, A. A. Glazer. Influence
of deformation and annealing on structure and properties of amorphous alloys.

338

Effect of Grain Size and Interfaces on Properties of Bulk Nanomaterials

158.
159.

160.

161.

162.

163.

164.

165.

166.

Fiz. Metall. Metalloved. 73, No 2, 102-110 (1992) (in Russian)


A. Inoue. Preparation and novel properties of nanocrystalline and
nanoquasicrystalline alloys. Nanostruct. Mater. 6, 53-64 (1995)
N. Hasegawa, M. Saito. Soft magnetic properties of microcrystalline Fe
MC (M = V, Nb, Ta) films with high thermal stability. Nippon Oyo Jiki
Gakkaishi (J. Magn. Soc. Japan.) 14, 313-318 (1990) (in Japanese)
N. Taneko, Y. Shimada, K. Fukamichi, C. Miyakawa. Fabrication of ironzirconium-nitrogen films with very low coercivity. Japan. J. Appl. Phys.
(Part 2) 30, No 2A, L195-L197 (1991)
A. Makino, A. Inoue, T. Masumoto. Soft magnetic properties of nanocrystalline FeMB (M Zr, Hf, Nb) alloys with high magnetization. Nanostruct.
Mater. 6, 985-988 (1995)
N. Hasegawa, M. Saito. Soft magnetic properties of microcrystalline Co
MC (M group IVAVIA elements) films with high thermal stability. Nippon
Kinzoku Gakkaishi (J. Japan. Inst. Met.) 54, 1270-1278 (1990) (in Japanese)
N. Hasegawa, N. Kataoka, K. Hiraga, H. Fujimori. Microstructure of carbidedispersed nanocrystalline FeTaC films. Mater. Trans. Japan. Inst. Met.
33, 632-635 (1992)
S. C. Mehta, D. A. Smith, U. Erb. Study of grain growth in electrodeposited
nanocrystalline nickel1.2 wt.% phosphorus alloy. Mater. Sci. Engineer. A
204, 227-232 (1995)
N. Hasegawa, A. Makino, N. Kataoka, H. Fujimori, A. P. Tsai, A. Inoue,
T. Masumoto. Crystallization process of amorphous FeTaC alloy films
and thermal stability of the resultant soft-magnetic nanocrystalline state.
Mater. Trans. Japan. Inst. Met. 36, 952-961 (1995)
A. A. Glazer, N. I. Noskova, V. A. Lukshina, A. P. Potapov, R. I. Tagirov,
E. G. Ponomareva. Influence of fast crystallization of amorphous Fe5Co70Si15B10
alloy on their magnetic properties. Fiz. Metall. Metalloved. 76, No 2, 171174 (1993) (in Russian)

339

Nanocrystalline Materials

Chapter 8

8. Conclusions
Three groups of technologies have ensured the scientific and
technical advances in the first half of the XXI century: electronics
and computer technology, biotechnology and nanotechnology. It is
believed that the development of electronics and computer
technology will reach the maximum activity in the years 20102015,
and the contribution of biotechnology, which started in 19681973,
will be largest in the period from 2025 to 2035, whereas
nanotechnology will become the main driving force of scientific and
technical advances in 20452055.
The essence of nanotechnology is the ability to work at the
atomic, molecular and supramolecular levels, in the length scale of
about 1100 nm range, in order to create, manipulate and use
materials, devices and systems that have novel properties and
functions because of the small scale of their structures [1]. In some
situations, the length scale under which the novel phenomena and
properties develop may be less than 1 nm or larger than 100 nm.
Control of matter on the nanoscale means tailoring the fundamental
structure, properties, processes and functions exactly on the scale
where the basic properties of solids are defined. Nanotechnology
includes integration of nanoscale structures into larger material
components, systems and architectures that could be used in most
industries, healthcare systems and environment. At present, the
development of nanotechnology and the production and application
of nanomaterials have not yet reached a maximum.
Prior to 1990, there were no specialised scientific journals on
nanomaterials and nanotechnology. At present, approximately 20
international scientific journals are concerned exclusively with
nanomaterials and nanotechnology. Among them are such journals
as Fullerene Science and Technology, International Journal of
Nanoscience, Journal of Micromechanics and Microengineering,
340

Conclusions

350

12000

300
Publications

10000

250

8000

200

6000

150

4000

100
Patents
50

2000

Number o

Number of patent applications at EPO

14000

Number of pu

Number of publications in SCI

Journal of Nanoparticles Research, Journal of Nanoscience and


Technology, Journal of Metastable and Nanocrystalline Materials,
Journal of Physical Chemistry B, NanoLetters, Nanostructured
Materials, Nanotechnology, Physica E: Low-dimensional Systems
and Nanostructures. In addition to these, articles on nanomaterials
and nanotechnology are being published by all materials science
journals, journals for condensed matter physics and for colloid
chemistry. In addition to the number of specialised scientific and
technical journals, two obvious indicators of the interest in
nanomaterials and nanotechnology are the number of published
scientific studies and the number of patents. The first is an efficient
indicator of scientific activity, and the latter indicates the possibility
of application of scientific results. Figure 8.1 shows the change in
the number of publications and patents on nanomaterials and
nanotechnology between 1980 and 1998 [2]. The data on the worldwide number of publications have been extracted from the database
of the Science Citation Index (SCI), and the nanopatents are those
recorded by the European Patent Office (EPO) in Munich. As
indicated by Fig. 8.1, in the period from 1981 to 1985 the number
of articles was small but gradually increased from year to year. A
large increase in the number of publications was recorded in 1986,
and in the period from 1989 to 1998 the number of studies published

Fig. 8.1. Annual number of publications and patents in nanomaterials and nanotechnology
from 1981 to 1998 world-wide [2]. The number of publications includes all
nanotechnology-related articles published world-wide and covered by the Science
Citation Index (SCI) database. The nanopatents are those filed at the European
Patent Office (EPO).

341

Nanocrystalline Materials

every year increased from 1000 to 12000. A large volume of the


cumulated experimental data on nanomaterials and nanotechnology
is used as a reliable base for generalisation of the fundamental and
applied results in review articles and monographs (see, for example
[311]).
A clear confirmation of the interest in nanomaterials and
nanotechnology is also the dynamics of growth of financial
investment into this complex branch of scientific and applied
developments. In the last 1520 years of the 20 th Century, 340
million dollars was spent throughout the world in the development
of new processes and materials associated with the nano concept.
In addition to this, as a result of the development of nanomaterials
and nanotechnology, additional investment was made into
electronics, chemical pharmaceutics, catalysis, aerospace industry
and toolmaking worth 300, 180, 100, 70 and 2 million dollars
respectively. Thus, the total expenditure on scientific and applied
investigations into nanomaterials already at the end of the twentieth
century exceeded 1 billion dollars. In the USA in 2000, 2001, 2002
and 2003, government investment in scientific research programmes
into nanomaterials and nanotechnologies was 270, 424, 697 and 774
million dollars, and the investment by the venture innovative capital
reached 500, 800, 1000 and 1200 million dollars, respectively. In the
USA in 2004, government investment in scientific investigations into
nanomaterials should reach 847 million dollars. National programmes
on nanomaterials in other countries are highly extensive: in 2002 in
Japan, the appropriate expenditure was 800 million dollars, in the
European Union countries 300 million dollars, and in the rest of the
world 400 million dollars. It should be mentioned that there has been
a gradual displacement of scientific and applied interest from the
development of nanomaterials to the construction of devices and
systems utilising the nanocrystalline state effects. The range of
investigations into biological nanomaterials and bionanotechnology
is being expanded continuously.
The broader perspective of qualitative improvements, which
nanotechnology will bring in the society, cannot be underestimated.
Nanoscale science and engineering promise restructuring almost all
industries towards the next industrial revolution, reshaping our
intellectual comprehension of nature, and assuring the quality of
life. Nanotechnology is seen as an emerging technology of the
present century because of the importance of the control of matter
at nanoscale on almost all technologies from information and
medicine to manufacturing and environment.
342

Conclusions

The book which you have just read is concerned mainly with
nanocrystalline solids. New investigations in the last decade have
greatly expanded the knowledge of the effects associated with the
size of grains (crystallites) in solids. For a long time, main attention
has been concentrated on examination of small particles and
nanoclusters, which properties are intermediate between the
properties of isolated atoms and the polycrystalline solids. A
development of the methods of producing bulk materials with an
unusually fine-grained structure, in which the grains are of the
nanometer sizes, has enabled transfer to examining the structure
and properties of the solid in the nanocrystalline state. At present,
the main methods of producing bulk nanocrystalline materials are
[6, 7, 11]: compacting of isolated nanoclusters, produced by
evaporation and condensation; deposition from colloid solutions or
decomposition of precursors; crystallisation of amorphous alloys;
severe plastic deformation; ordering of strongly non-stoichiometric
compounds and solid solutions [7, 1228]. Each of these methods
has advantages and disadvantages and none of them is universal
because they can be used most efficiently for a certain range of
substances.
Analysis of the currently available experimental results shows
that an important role in the nanocrystalline solid is played not only
by the grain size (as in case of the isolated nanoparticles) but also
by the structure of the interfaces (grain boundaries). In fact, grain
boundaries in bulk nanomaterials, produced by different methods, is
characterised by large differences. For example, in nanomaterials
produced by severe deformation, the grain boundaries are
characterised by a high dislocation density, and in nanomaterials
produced by crystallisation from the amorphous state, the grain
boundaries may be quasi-amorphous or may have a greatly distorted
crystalline structure. All these special features must be taken into
account in the interpretation of the properties of bulk nanomaterials.
The effect of the interface on the structure and properties is
especially strong in nanomaterials produced by compacting or severe
plastic deformation. In these materials, immediately after production,
the interfaces are in the non-equilibrium stressed state and have
excess energy. The relaxation of non-equilibrium interfaces in
nanocrystalline metals and alloys takes place spontaneously even at
room temperature and in most cases is accompanied by grain
growth [22, 29]. Ceramic oxide nanomaterials are more stable than
metallic ones; their structure and grain size may remain almost
constant even after annealing at 600800 K [30].
343

Nanocrystalline Materials

The properties of nanocrystalline metals and alloys, especially


those produced by compacting nanoclusters, are very sensitive to
the oxygen impurity. The unusually large area of the interfaces
results in high chemical activity of nanocrystalline metals. A large
part of the unexpected results, obtained in the period up to 1992,
after subsequent examination proved to be the consequence of
contamination of nanocrystalline metals by impurity oxygen; in the
case of nanocrystalline Pd, also contamination by hydrogen is
possible.
Of special importance for theoretical comprehension of the
experimental results, obtained on isolated nanoparticles and bulk
nanocrystalline materials, is the separation of the surface, grainboundary (associated with the interfaces) and volume (associated
with the size of particles, crystallites, or grains) effects. At present,
this problem is far from being solved. On the whole, the
understanding and explanation of the structure and properties of
isolated nanoparticles is considerably higher in comparison with bulk
nanocrystalline materials. This is associated with the fact that
dispersed systems and nanoclusters have already been studied since
the beginning of the XX th century, whereas bulk nanomaterials
became a subject of investigation only after 1982. Undoubtedly, the
active investigations of bulk nanocrystalline materials will make it
possible to eliminate the existing delay in the next 1015 years.
Analysing the state of material science investigations, it is
possible to define four stages of the life of materials: proposal
of a concept, intensive investigations, increase in the production
volume and decrease in the production volume. Evidently, bulk
nanocrystalline materials are in the stage of intensive investigations.
It may be expected that in the near future the intensity of studying
the nanocrystalline materials will increase. The most important
directions will evidently be the deep examination of the
microstructure, separation of effects determined by the particle size
and the interfaces, determination of the conditions of stabilisation
of the nanostructure ensuring the retention of the properties at
elevated temperatures, development of models for adequate
theoretical description of the nanocrystalline state.
The new stable nanocrystalline materials will be developed on the
basis of multi-component systems and not only on the basis of metals.
Compounds of metals with oxygen, nitrogen and carbon, having a high
melting point and high thermal stability, will become the main
components of nanocrystalline materials in the near future. Using of
these compounds will make it possible to develop nanomaterials
344

Conclusions

characterised by stable operation and no changes in their properties


throughout the entire servicelife. In particular, the oxides, nitrides and
carbides of metals are expected to manifest their best properties in the
world of nanomaterials.
Detailed investigations of nanocrystalline materials will result in
the development of such new sciences as physics and chemistry of
nanocrystalline solids so that it will be possible to construct a strong
bridge between nanomaterials and nanotechnology.
References
1.
2.

3.
4.
5.
6.
7.
8.
9.
10.
11.

12.
13.

14.
15.
16.

17.

M. C. Roco. Towards a US national nanotechnology initiative. J. Nanoparticle


Researh 1, 435-438 (1999)
R. Compa, A. Hullmann. Forecasting the development of nanotechnology
with the help of science and technology indicators. Nanotechnology 13,
243-247 (2002)
K. E. Drexler. Nanosystems: Molecular Machinery, Manufacturing, and
Computation (Wiley, New York 1992) 518 pp.
A. ten Wolde. Nanotechnology: Towards a Molecular Construction Kit (New
World Ventures, Boston 1998) 357 pp.
Nanostructured Materials. Ed. J. Yi-Ru Ying. (Academic Press, New York
2001) 350 pp.
A. I. Gusev. Nanocrystalline Materials: Preparation and Properties (Ural
Division of the Russ. Acad. Sci., Yekaterinburg 1998) 200 pp. (in Russian)
A. I. Gusev, A. A. Rempel. Nanocrystalline Materials (Nauka-Fizmatlit, Moscow
2000) 224 pp. (in Russian)
Nanomaterials: Synthesis, Properties and Applications. Eds. A. S. Edelstein,
R. C. Cammarata. (The Johns Hopkins University, Baltimor 1998) 620 pp.
M. J. Mayo. Processing of nanocrystalline ceramics from ultrafine particles. Int. Mater. Rev. 41, 85-115 (1996)
H. Gleiter. Nanostructured materials: basic concepts and microstructure.
Acta Mater. 48, 1-29 (2000)
A. I. Gusev. Effects of the nanocrystalline state in solids. Uspekhi Fiz.
Nauk 168, 55-83 (1998) (in Russian). (Engl. transl.: Physics - Uspekhi 41,
49-76 (1998))
H. Gleiter. Materials with ultrafine microstructure: retrospectives and perspectives. Nanostruct. Mater. 1, 1-19 (1992)
R. Birringer, H. Gleiter. Nanocrystalline materials. In: Encyclopedia of Material
Science and Engineering. Suppl. Vol.1. Ed. R. W. Cahn (Pergamon Press,
Oxford 1988) pp.339-349
R. W. Siegel. Cluster assembled nanophase materials. Ann. Rev. Mater.
Sci. 21, 559-578 (1991)
R. W. Siegel. Nanostructured materials mind over matter. Nanostruct. Mater.
3, 1-18 (1993)
H.-E. Schaefer. Interfaces and physical properties of nanostructurd solids.
In: Mechanical Properties and Deformation Behavior of Materials Having
Ultrafine Microstructure. Eds. M. A. Nastasi, D. M. Parkin, H. Gleiter. (Kluwer
Academic Press, Netherlands, Dordrecht 1993) pp.81-106
B. H. Kear, P. R. Strutt. Chemical processing and applications for nanostructured
345

Nanocrystalline Materials

18.
19.

20.
21.

22.

23.

24.
25.
26.
27.

28.

29.

30.

materials. Nanostruct. Mater. 6, 227-236 (1995)


H. Gleiter. Nanocrystalline materials. Progr. Mater. Sci. 33, 223-315 (1989)
Y. Yoshizawa, S. Oguma, K. Yamauchi. New iron-based soft-magnetic alloys composed of ultrafine grain structure. J. Appl. Phys. 64, 6044-6046
(1988)
A. Inoue. Preparation and novel properties of nanocrystalline and
nanoquasicrystalline alloys. Nanostruct. Mater. 6, 53-64 (1995)
R. Z. Valiev, O. A. Kaibyshev, R. I. Kuznetsov, R. Sh. Musalimov, N. K.
Tsenev. The low-temperature superplasticity of metallic materials. Doklady
AN SSSR. 301, 864-866 (1988)
R. Z. Valiev, A. V. Korznikov, R. R. Mulyukov. Structure and properties
of metallic materials with submicrocrystalline structure. Fiz. Metall. Metalloved.
73, No 4, 70-86 (1992) (in Russian)
V. M. Segal, V. I. Reznikov, V. I. Kopylov, D. A. Pavlik, V. F. Malyshev.
Processes of Plastic Formation of Structure of Metals (Nauka i Tekhnika,
Minsk 1994) 232 pp. (in Russian)
H. Gleiter. Nanostructured materials: state of art and perspectives. Nanostruct.
Mater. 6, 3-14 (1995)
R. Z. Valiev. Approach to nanostructured solids through the studies of
submicron grained polycrystals. Nanostruct. Mater. 6, 73-82 (1995)
K. Yamauchi, Y. Yoshizawa. Recent development of nanocrystalline soft magnetic
alloys. Nanostruct. Mater. 6, 247-254 (1995)
A. A. Rempel, A. I. Gusev. Nanostructure and atomic ordering in vanadium
carbide. Pisma v ZhETF 69, 436-442 (1999) (in Russian). (Engl. transl.:
JETP Letters 69, 472-478 (1999))
A. I. Gusev, A. A. Tulin, V. N. lipatnikov, A. A. Rempel. Nanostructure of
dispersed and bulk nonstoichiometric vanadium carbide. Zh. Obsh. Khimii
72, 1067-1076 (2002) (in Russian). (English transl.: Russ. J. General Chem.
72, 985-993 (2002))
J. Weissmller, J. Lffler, M. Kleber. Atomic structure of nanocrystalline
metals studied by diffraction techniques and EXAFS. Nanostruct.Mater. 6,
105-114 (1995)
Y. Ishida, H. Ichinose, T. Kizuka, K. Suenaga. High-resolution electron
microscopy of interfaces in nanocrystalline materials. Nanostruct.Mater.
6, 115-124 (1995)

346

You might also like