+ All Categories
Home > Documents > JMS13_MMREVIEW.pdf

JMS13_MMREVIEW.pdf

Date post: 03-Oct-2015
Category:
Upload: jacobomr1980
View: 6 times
Download: 0 times
Share this document with a friend
Popular Tags:
14
REVIEW Preparation of multicomponent oxides by mechanochemical methods A. F. Fuentes L. Takacs Received: 26 July 2012 / Accepted: 18 September 2012 / Published online: 2 October 2012 Ó Springer Science+Business Media New York 2012 Abstract A large variety of synthesis strategies and processing techniques are currently being used to obtain new multicomponent oxides and/or modify existing ones. Among them, mechanochemical processing has become very popular because it is simple to implement, solvent free, and capable of providing enough volume of the target material in an economically viable manner. The prepara- tion of complex oxides can benefit from mechanochemical methods for two important reasons: First, it is not a dif- fusion-controlled process and thus, high-rate solid state reactions can be promoted between oxides with different physical and chemical properties without using high tem- peratures; secondly, because reactants are processed under non-equilibrium conditions, uncommon metastable phases are frequently obtained featuring flexible crystal structures, small particle size, high concentration of defects, and off-stoichiometry. Furthermore, conversion to the ‘‘true’’ equilibrium phases induced by additional processing (e.g., firing) offers the possibility of isolating fairly stable intermediate states with unusual and desirable properties that are inaccessible for more conventional processing techniques. As oxide particles are hard and brittle, the number of oxide systems prepared by means of mechano- chemical methods grew rapidly only in recent years when more powerful milling devices and abrasion-resistant mill- ing tools became available. This article summarizes recent work carried out in the field; only dry milling of oxides (and occasionally carbonates) in the absence of additives is considered. Some of the main challenges of mechano- chemical processing are also highlighted and discussed. Introduction Multicomponent oxides constitute an important class of inorganic solids exhibiting an extraordinary range of crystal structures and functional properties, which are behind many advances of modern technology. The obser- vation that some of their physical and chemical properties depend not just on intrinsic characteristics such as crystal structure and chemical composition, but are also sensitive to the presence of structural defects has fueled the search for new synthesis strategies and processing techniques aiming to understand and control defect structure. Because of the slow diffusion kinetics in the solid state, most syn- thetic approaches involve a previous dissolution step; however, solid state reactions are gaining attention because of ecological as well as energy reasons. Also, much interest is focused nowadays on identifying and developing sol- vent-free and low-waste procedures, which avoid the use and generation of hazardous substances, while providing the necessary volume of the target material in an eco- nomically viable manner. Several research groups worldwide have turned their attention to mechanochemical processing, encouraged by the simplicity of the process and the moderate cost of the A. F. Fuentes (&) Cinvestav Unidad Saltillo, Carretera Saltillo-Monterrey Km. 13.5, 25900 Ramos Arizpe, Coahuila, Mexico e-mail: [email protected] Present Address: A. F. Fuentes Department of Earth and Environmental Sciences, University of Michigan, Ann Arbor, MI 48109, USA e-mail: [email protected] L. Takacs Department of Physics, University of Maryland Baltimore County, Baltimore, MD 21250, USA e-mail: [email protected] 123 J Mater Sci (2013) 48:598–611 DOI 10.1007/s10853-012-6909-x
Transcript
  • REVIEW

    Preparation of multicomponent oxides by mechanochemicalmethods

    A. F. Fuentes L. Takacs

    Received: 26 July 2012 / Accepted: 18 September 2012 / Published online: 2 October 2012

    Springer Science+Business Media New York 2012

    Abstract A large variety of synthesis strategies and

    processing techniques are currently being used to obtain

    new multicomponent oxides and/or modify existing ones.

    Among them, mechanochemical processing has become

    very popular because it is simple to implement, solvent

    free, and capable of providing enough volume of the target

    material in an economically viable manner. The prepara-

    tion of complex oxides can benefit from mechanochemical

    methods for two important reasons: First, it is not a dif-

    fusion-controlled process and thus, high-rate solid state

    reactions can be promoted between oxides with different

    physical and chemical properties without using high tem-

    peratures; secondly, because reactants are processed under

    non-equilibrium conditions, uncommon metastable phases

    are frequently obtained featuring flexible crystal structures,

    small particle size, high concentration of defects, and

    off-stoichiometry. Furthermore, conversion to the true

    equilibrium phases induced by additional processing (e.g.,

    firing) offers the possibility of isolating fairly stable

    intermediate states with unusual and desirable properties

    that are inaccessible for more conventional processing

    techniques. As oxide particles are hard and brittle, the

    number of oxide systems prepared by means of mechano-

    chemical methods grew rapidly only in recent years when

    more powerful milling devices and abrasion-resistant mill-

    ing tools became available. This article summarizes recent

    work carried out in the field; only dry milling of oxides (and

    occasionally carbonates) in the absence of additives is

    considered. Some of the main challenges of mechano-

    chemical processing are also highlighted and discussed.

    Introduction

    Multicomponent oxides constitute an important class of

    inorganic solids exhibiting an extraordinary range of

    crystal structures and functional properties, which are

    behind many advances of modern technology. The obser-

    vation that some of their physical and chemical properties

    depend not just on intrinsic characteristics such as crystal

    structure and chemical composition, but are also sensitive

    to the presence of structural defects has fueled the search

    for new synthesis strategies and processing techniques

    aiming to understand and control defect structure. Because

    of the slow diffusion kinetics in the solid state, most syn-

    thetic approaches involve a previous dissolution step;

    however, solid state reactions are gaining attention because

    of ecological as well as energy reasons. Also, much interest

    is focused nowadays on identifying and developing sol-

    vent-free and low-waste procedures, which avoid the use

    and generation of hazardous substances, while providing

    the necessary volume of the target material in an eco-

    nomically viable manner.

    Several research groups worldwide have turned their

    attention to mechanochemical processing, encouraged by the

    simplicity of the process and the moderate cost of the

    A. F. Fuentes (&)Cinvestav Unidad Saltillo, Carretera Saltillo-Monterrey

    Km. 13.5, 25900 Ramos Arizpe, Coahuila, Mexico

    e-mail: [email protected]

    Present Address:A. F. Fuentes

    Department of Earth and Environmental Sciences,

    University of Michigan, Ann Arbor, MI 48109, USA

    e-mail: [email protected]

    L. Takacs

    Department of Physics, University of Maryland Baltimore

    County, Baltimore, MD 21250, USA

    e-mail: [email protected]

    123

    J Mater Sci (2013) 48:598611

    DOI 10.1007/s10853-012-6909-x

  • equipment needed to perform it. Even more interesting is the

    possibility to obtain nanocrystalline materials and non-

    equilibrium structures with different properties compared to

    those prepared by conventional processing. Accordingly,

    mechanical milling has become an attractive and powerful

    method to produce new complex oxides or to modify

    existing ones by inducing phase transformations, extended

    solid solutions, disordering, or amorphization. The use of

    mechanical energy to stimulate physical and chemical pro-

    cesses in the solid state is certainly not a new development.

    Traditionally, the preparation of oxide systems used

    mechanical milling mostly for mixing and grinding powders

    and to induce phase transformations; however, the recent

    availability of more powerful milling devices and appro-

    priate milling tools and the advances gained in understand-

    ing the basic phenomena behind mechanochemical reactions

    have triggered renewed interest in the subject with still many

    challenging opportunities remaining for research.

    Interest in the mechanochemistry of complex oxides

    increased substantially during the last ten to fifteen years, as

    evidenced by the large variety of inorganic systems analyzed

    and the increasing number of publications on the subject.

    Unfortunately, the experimental details are often incom-

    plete; thus, direct comparison of the results from different

    laboratories is usually not possible. Moreover, the reaction

    products are rarely characterized completely. These cir-

    cumstances result in limited progress toward understanding

    the fundamental features of mechanochemical effects on

    oxides. Nevertheless, many general trends can be deduced

    from the large body of available empirical data.

    The first review ever published exclusively on the mech-

    anochemistry of complex oxides was written by Zyryanov [1].

    He provides a critical evaluation of the existing literature

    using the crystal structure of the reaction product as the

    organizing principle. Russian researchers have been active in

    this area for decades and Zyryanovs paper provides an

    extensive summary of their work. Therefore, the current paper

    will reference only the latest and most important papers from

    the Russian literature. Another recent review [2] concentrates

    on the mechanochemical synthesis of ferroelectric ceramics

    starting from different precursor materials. It includes both

    completely mechanochemical synthesis and thermal synthesis

    using mechanically activated components. Research in this

    area was triggered by the successful preparation of an

    important ferroelectric Pb(Mg1/3Nb2/3)O3 (PMN) from PbO,

    MgO, and Nb2O5 [3]. If PMN is prepared by conventional

    solid state synthesis, various pyrochlore-type binary lead-

    niobium intermediates form that have poor dielectric prop-

    erties. Using mechanochemical synthesis, single-phase PMN

    could be obtained.

    The aim of the present work is to present a summary of

    the mechanochemistry of complex oxides emphasizing

    chemical composition and potential applications, mainly in

    the area of electroceramics. After some general remarks,

    iron oxides, bismuth oxides, titanates, etc. are considered in

    separate sections, followed by a few other examples of

    practical interest. Only dry milling of oxides (and occa-

    sionally carbonates) in the absence of additives is consid-

    ered, except for a short section on soft mechanochemistry.

    General remarks

    It is well known that high-energy ball milling is an efficient

    method to mechanically activate materials and to affect

    chemical reactions [4] and alloying [5]. Mechanochemical

    processing can be carried out in several types of com-

    mercially available milling devices, each operating based

    on very different mechanical principles and offering vari-

    ous capacities and milling efficiencies. Some research

    groups have developed the ability to design and manufac-

    ture their own milling equipment. While the existence of

    various options has many benefits, it substantially com-

    plicates the task of analyzing the literature as comparing

    data published by different authors using different mills is

    often difficult. The highest impact energies are achieved in

    shaker mills and planetary mills, while drum mills, vibra-

    tory mills, and attritors deliver lower energy and a different

    combination of impact, shear, and friction forces (Fig. 1).

    They all cause activation and reactions in powder mixtures

    by continuously exposing fresh surfaces and creating lat-

    tice defects, surface radicals, and broken bonds.

    There are many promising results on the application of

    mechanochemical processing to complex oxides. Never-

    theless, the method has some inherent difficulties when

    compared to its application to alloys, intermetallic com-

    pounds, and many other systems. Most importantly, oxide

    particles are hard and brittle and consequently do not react

    with each other unless a threshold stress is exceeded. In

    practical terms, reactions between oxides require highly

    energetic impacts of the milling balls; thus, only the most

    powerful milling devices are suitable for the preparation of

    mixed oxides. Extending the milling time in a low-energy

    mill is usually not sufficient as weak impacts do not result

    Fig. 1 Main forces acting on powder particles during milling

    J Mater Sci (2013) 48:598611 599

    123

  • in any chemical change. It might also happen during

    milling that a stable state of mechanochemical equilibrium

    is reached at a given volume fraction of the phases

    involved (e.g., reactants and reaction products) with further

    milling being ineffective in driving the reaction forward.

    Contamination from the milling tools is a particularly

    serious problem when hard and abrasive oxides are milled

    at high intensity for a prolonged time. However, several

    steps can be made to reduce the level of contamination.

    The milling container and the balls can be made from a

    material that is less detrimental to the properties of the

    product. Thus, ceramic tools made of alumina are preferred

    to steel and WCCo. If a high density is crucial, balls made

    from zirconia or even hafnia can be used. Also, selecting

    the lowest acceptable milling speed, the shortest possible

    milling time, and optimizing the amount of powder charge

    and the number and size of the milling balls can keep

    contamination at an acceptable level.

    If processing begins with a clean mill, much of the

    contamination occurs during the early stages of the process

    when the tool surfaces are still bare and the collisions

    between them result in significant abrasion. After a few

    minutes, a powder coating develops and the rate of con-

    tamination decreases substantially. This phenomenon can

    be utilized to coat the working surfaces before processing.

    A preliminary batch is milled until a stable coating

    develops. Then, the loose powder in the mill is discarded

    and processing of the final batch is started with the tool

    surfaces already protected by a layer of compatible powder

    [1]. Nevertheless, contamination cannot be avoided com-

    pletely, even if the most careful protocol is followed.

    Finally, as mechanochemical reactions are inherently

    complex, it is difficult to predict whether or not a desired

    reaction product would be formed. Sometimes, even the

    direction of a mechanochemical process seems to contradict

    thermodynamical calculations, or the reaction proceeds

    according to a route that is substantially different from the

    path of conventional solid state reactions. Chemical changes

    caused by mechanical action result from various thermally as

    well as mechanically induced processes and occur simulta-

    neously at many reaction sites [6]. Designing a mathematical

    model capable of describing the events taking place in such

    reaction sites would entail correlating the milling parameters

    with the energy being transferred to the material subjected to

    milling and with the frequency of collisions. Given the

    complex motion of the grinding media, the number of vari-

    ables involved in determining the energetics of a high-energy

    ball mill is very large, as shown in Fig. 2 [7]. Also, the con-

    nection between the macroscopic operation of the mill and the

    local loading of the particles is very complicated. Therefore,

    developing a complete and predictive mathematical model of

    milling-induced chemical reactions is quite unlikely. How-

    ever, some progress has been made in metal systems to

    ascertain whether or not a mechanochemical reaction is pos-

    sible using a given set of milling parameters [5]; available data

    concerning multicomponent oxides are still scarce. A recent

    theory of ultrafast mechanochemical synthesis in MOM0O3oxide systems alleges that large molecular mass of the reac-

    tants, small difference of their Mohs hardness, and large

    enthalpy of the chemical reaction involved are optimal for

    high yield mechanochemical reactions [8, 9].

    In spite of all the difficulties, it would be a mistake to

    overemphasize the problems and ignore the enormous

    potential of mechanochemical processing as a powerful

    method for obtaining multicomponent oxides. The prepa-

    ration of complex oxides can benefit from mechanochem-

    ical methods for two important reasons:

    (1) Conventionally, complex oxides are prepared by solid

    state reactions between oxides, carbonates, and nitrates

    at high temperature (typically over 1000 C). Firingtimes as long as days or even weeks can be necessary to

    insure single-phase products, resulting in a very costly

    procedure.

    Take for example the case of apatite-type lanthanum

    silicates La10-x(SiO4)6O02?y which are promising oxygen

    ion-conducting materials for application as electrolytes in

    intermediate-temperature solid oxide fuel cells [10]. As for

    many silica-based compounds, conventional solid state pro-

    cessing from a mixture of simple oxides requires several days

    of heat treatment at 1400 C to yield the oxygen ion-conducting solid solution [11]. The same material can be

    prepared by ball milling in less than 9 h using the same

    chemicals, a planetary mill at moderate speed, and zirconia

    milling tools [12].

    (2) High-temperature synthesis produces large grains and

    equilibrium phases, but many applications benefit

    Fig. 2 Milling parameters determining the energetics in a planetaryball mill with a cylindrical vial (adapted from Ref. [7])

    600 J Mater Sci (2013) 48:598611

    123

  • from the formation of uncommon metastable phases

    with flexible crystal structures and the small particle

    size, high concentration of defects, and off-stoichi-

    ometric composition obtained by mechanochemical

    methods.

    Mechanochemical processing creates a unique and often

    quite stable defect structure that is difficult to obtain using

    conventional processing. Furthermore, the transformation

    to the true equilibrium phase induced by additional

    processing (e.g., firing) often proceeds through a series of

    transitions involving fairly stable intermediate states with

    unusual physical and chemical properties. For example,

    pyrochlore-type titanates and zirconates form a very

    interesting family of compounds with potential applica-

    tions as electrolytes for solid oxide fuel cells, thermal

    barrier coatings, and matrix materials for the immobiliza-

    tion of high-level radioactive waste. Pyrochlores prepared

    by ball milling from mixtures of simple oxides present a

    high degree of structural disorder that can be reduced by

    thermal treatment in a controlled manner [1315]. That

    way, the properties of the product can be controlled by

    changing structural/microstructural characteristics, while

    keeping the chemical composition constant.

    From the kinetic point of view, solid state reactions

    induced by ball milling can progress in two different ways,

    either (i) gradually extending to a very small volume with

    each collision or (ii) suddenly through a mechanically

    activated exothermic reaction. The latter are termed

    mechanically induced self-sustaining reactions (MSR) and

    are characteristic of highly energetic powder mixtures

    including oxidemetal systems (e.g., Fe2O3Al), refrac-

    tory-forming metalmetalloid mixtures (e.g., TiC), and

    metalchalcogen mixtures (e.g., ZnS) [16]. Chemical

    reactions in these systems are self-sustained, propagate

    throughout the entire powder charge after some activation

    period, and reach completion in a very short time; however,

    additional milling is frequently needed to achieve a fully

    reacted and homogeneous product. On the contrary, grad-

    ual mechanochemical reactions are spatially limited to a

    small volume around the point of collision between balls or

    a ball and the inner wall of the container; thus, reaction

    rates are time dependent, reaching a maximum at an

    intermediate milling time and then decreasing as reaction

    approaches completion or until a steady state of mecha-

    nochemical equilibrium is reached. These are the reactions

    commonly observed in oxide systems.

    Iron oxides and related systems

    The mechanochemical synthesis of ferrites has attracted the

    interest of many research groups for some time now

    because of the strong correlation between magnetic prop-

    erties and structural disorder in AB2O4 spinels (e.g., [17]).

    Mossbauer spectroscopy provides unique information

    about the local coordination and the magnetic and oxida-

    tion states of iron atoms that, combined with XRD and

    other methods, allows for a comprehensive characterization

    of ferrites [18]. The ideal spinel structure is formed by a

    cubic close-packed array of oxygen atoms in which one-

    eighth of the tetrahedral and half of the octahedral inter-

    stitial sites (referred in the chemical formula as A and B,

    respectively) are occupied by cations. The physical prop-

    erties of spinels are determined by chemical composition

    and cation distribution over the two available positions.

    There are two types of oxides with spinel structure: In

    normal spinels, the tetrahedral and octahedral sites are

    commonly occupied by divalent (e.g., Mg, Ni, Co, etc.) and

    trivalent cations (e.g., Al, Ga, Fe, etc.), respectively, while

    in inverse spinels, the divalent ions occupy octahedral

    sites, whereas half of the trivalent cations are located in the

    tetrahedral sites. In general, inverse spinels have been

    found to present the most valuable soft magnetic proper-

    ties. Many intermediate cation distributions are possible,

    with the degree of inversion often determined by the

    preparation technique. For example, mechanical milling

    can influence cation distribution and consequently the

    magnetic properties.

    Low conversion yields and partial reduction were a

    common concern when using steel milling tools to combine

    simple oxides to form ferrites. However, high yields could

    be achieved when using tungsten carbide (WC) jars and

    balls. For example, NiFe2O4 was obtained by milling the

    appropriate mixture of NiO and a-Fe2O3 with almostcomplete conversion [1921]. Thorough characterization

    revealed that the product consisted of 613-nm grains with

    a non-uniform coreshell structure, where an ordered inner

    core is surrounded by a 12-nm thick disordered grain

    boundary region. The core has the inverse spinel structure

    of bulk NiFe2O4, while the cation distribution in the shell is

    almost random.

    This coreshell configuration is typical of mechanically

    milled nanopowders, but not nanopowders prepared by

    other techniques [22]. It has been observed in MnFe2O4prepared by a milling-induced displacement reaction [23],

    in Ca2SnO4 prepared by reactive milling from the corre-

    sponding elemental oxides [24], and in LiNbO3 and ZrO2nanopowders prepared by milling commercially available

    microcrystalline powders [2527]. But, X-ray absorption

    spectroscopy studies (XAS) have shown that LiNbO3nanoparticles prepared by the solgel method have well-

    ordered microstructures with interfaces similar to the grain

    boundaries in normal bulk solids [25].

    The coreshell structure has a very strong influence on

    the magnetic behavior of mechanochemically prepared

    J Mater Sci (2013) 48:598611 601

    123

  • spinel ferrites [28]. For example, nanoparticulate NiFe2O4presents lower saturation magnetization (*55 %), butenhanced magnetic hardness compared to bulk nickel fer-

    rite, both attributed to parasitic ferromagnetism due to

    spin canting. Firing above 400 C relaxes the materialtoward the equilibrium structure and consequently the

    magnetic properties approach those typical of bulk

    NiFe2O4.

    MgFe2O4 spinel has been prepared by milling a mixture

    of MgO and a-Fe2O3 powders [2931]. Complete conver-sion was achieved when using WC milling tools. In this

    case, the increase of the magnetization is mostly due to

    cation disorder, while spin canting has a secondary effect

    [32]. Interestingly, milling a mixture of CaO and a-Fe2O3under similar conditions produced only 23 % conversion to

    CaFe2O4 [33]. Spinel-type ZnFe2O4, MnFe2O4, and their

    solid solutions have been prepared starting from a-Fe2O3,ZnO, and/or different manganese oxides [34, 35]. Verdier

    et al. [36] investigated the effect of milling conditions

    using a planetary ball mill with separately adjustable

    speeds for the supporting disk and grinding jars. They

    observed different degrees of conversion depending on the

    combination of impact and friction generated by the mill.

    At low impact energy, partial reduction of Fe3? to Fe2?

    took place, resulting in the formation of a wustite-like

    (Fe,Zn)O phase. A similar reduction from Cu2? to Cu?

    could be the reason why attempts to prepare CuFe2O4spinel were unsuccessful [37]. Harris et al. [35] used a

    high-energy shaker mill and steel tools to achieve almost

    complete transformation to MnZn ferrites starting from

    MnO, ZnO, and a-Fe2O3. As expected, the mechano-chemically prepared ferrites showed non-equilibrium cat-

    ion distribution with a significant fraction of Zn atoms

    located at the octahedral site.

    Mixed iron-containing oxides with the perovskite

    structure were also prepared by mechanochemical meth-

    ods. Weakly ferromagnetic LaFeO3 was obtained by

    co-milling La2O3 and either Fe3O4 or a-Fe2O3 using hardenedsteel vials and balls [38]. Several LnFeO3 phases (Ln = Pr,

    Nd and Sm) were synthesized starting from the appropriate

    rare-earth oxide and freshly fired FeOOH [39]. Mechanical

    milling has also been used to prepare multiferroic BiFeO3powders starting from an equimolar mixture of Bi2O3 and

    Fe2O3 with complete conversion reached after milling for

    12 h using WC tools [40, 41]. The as-obtained powders

    consisted of spherical nanoparticles with the coreshell

    structure; the amorphous shell had a thickness of about

    1 nm. The amorphous fraction proved to be highly reactive

    and the authors were able to observe its rapid crystalliza-

    tion under irradiation with electrons in the transmission

    electron microscope [41]. Milling a mixture of Y2O3 and

    Fe2O3 powders generates yttrium iron garnet Y3Fe5O12with YFeO3 perovskite as an intermediate [42].

    Bismuth(III) oxide and related systems

    Bismuth(III) oxide Bi2O3 exists in at least five different

    crystal forms, identified as a, b, c, d, and e-phases. Theform stable at room temperature is the monoclinic a phase.It transforms to the cubic fluorite-type (face-centered)

    d phase at 729 C, which remains stable to the meltingpoint (825 C). Two metastable forms may occur oncooling between 500 and 650 C, the tetragonal b phaseand a body-centered cubic c phase; either can be stabilizedby certain impurities. The e phase has been synthesized byhydrothermal methods; its crystal structure is closely

    related to the a and b forms. Interestingly, the differentpolymorphs of Bi2O3 exhibit very different optical and

    electrical properties. In particular, d-Bi2O3 is a veryeffective oxygen ion-conducting material, and conse-

    quently it is of interest for applications such as gas moni-

    toring and solid oxide fuel cells. As d-Bi2O3 cannot beretained at room temperature by quenching, strategies like

    aliovalent chemical substitution and non-conventional

    powder processing methods have been attempted for this

    purpose.

    Bismuth oxide is considered a good candidate for

    mechanochemical synthesis because of its high molecular

    mass [1], and thus numerous studies have been carried out

    on Bi-containing materials, including attempts at stabiliz-

    ing the high-temperature modifications of Bi2O3 by

    mechanochemical methods. For example, c-Bi2O3 wasobtained by milling the a form with different oxides suchas SiO2, PbO, ZnO, and a-Fe2O3 in a 12:1 Bi-to-dopantcation atomic ratio, using either hardened steel or zirconia

    milling tools [43]. It was shown that contamination from

    the milling tools plays an important role in determining the

    obtained phases. For example, milling pure a-Bi2O3 withzirconia tools resulted in a mixture of a and b-Bi2O3instead of the c form, showing that Zr4? is unable to sta-bilize c or d-Bi2O3, contrasting with the behavior of Fe

    3?.

    Ball milling a a-Bi2O3:HfO2 powder mixture with a 2:3molar ratio produced some b-Bi2O3 already after a shortmilling time [44]. Yet, although the b phase can be con-sidered a version of the d-phase with ordered oxygenvacancies, 50 h of continuous milling were required to

    obtain the d form.Fluorite-type solid solutions based on d-Bi2O3 have also

    been prepared by milling mixtures of different oxides with

    a-Bi2O3 [45]. Thus, milling a-Bi2O3 with CaO, SrO, Y2O3,In2O3, or La2O3 yielded metastable single-phase

    Bi1.6M0.4O3-x powders (M = Ca, Ca0.5Sr0.5, Y, In, La)

    with the fluorite structure for every dopant except indium,

    the ion with the largest size difference compared to Bi3?.

    Formation of a new metastable orthorhombic-distorted

    fluorite structure d0-Bi2O3 has also been detected whenmilling Bi2O3-containing systems [46].

    602 J Mater Sci (2013) 48:598611

    123

  • As the Bi2O3V2O5 system presents a large number of

    stoichiometric compounds and solid solutions with very

    interesting physical and chemical properties, it is also of

    interest for mechanochemical investigations [47]. For

    example, milling mixtures close to the 1:1 Bi2O3:V2O5molar ratio in an AGO-2 mill with steel vials and balls led

    to the highly disordered clinobisvanite (monoclinic) BiVO4as the dominant phase, which is the form stable at room

    temperature. In off-stoichiometric mixtures, several addi-

    tional phases were formed, including cubic and tetragonal

    fluorite-like phases. Monoclinic BiVO4 has also been

    obtained using VO2 as vanadium source [47], although

    prolonged milling yielded either fluorite-type phases or the

    oxygen-deficient Bi2VO5 material, featuring a mixture of

    V5? and V4?.

    An interesting case is the Bi4V2O11 compound (2:1

    Bi2O3:V2O5 molar ratio) since this oxide is the parent

    compound of the BIMEVOX family of excellent oxygen

    ion-conducting materials [4749]. The Bi4V2O11 phase

    occurs in at least four different crystal forms, of which the

    c form is the most interesting due to its high ionic conduc-tivity. After extended milling of an appropriate Bi2O3V2O5mixture in a vibratory mill, only an amorphous material was

    obtained which crystallized upon annealing at 385 C.However, Shantha et al. [50] were allegedly able to obtain

    30-nm nanoparticles of tetragonal c-Bi4V2O11 by millingstoichiometric mixtures of the elemental oxides using agate

    vials and balls. The inclusion of Si4? ions from the milling

    tools might have aided the formation of Bi4V2O11.

    Partial replacement of V5? by some other metal ions

    (e.g., Cu2?, Ni2?) suppresses the c ? b ? a transition oncooling, allowing the highly conductive c-phase to bestabilized at room temperature. Therefore, ball milling has

    also been used to prepare substituted Bi4V2O11. Zhang

    et al. [51] prepared nanosized (*19 nm) Bi4V1.8Cu0.2O10.7powders by milling appropriate mixtures of V2O5, Bi2O3,

    and CuO using WC vials and balls. Below 600 C, thiscompound presents the highest oxygen ion conductivity

    ever measured (10-2 Scm-1 at 350 C), two orders ofmagnitude higher than any other known solid oxide ion-

    conducting material in the same temperature range.

    Zyryanov and Uvarov [52] were also able to prepare a large

    group of BIMEVOX phases, Bi4V1.8M0.2O11-x (M = V4?,

    Zn2?, Sc3?, Sb3?, In3?) and Bi1.8Pb0.2VO5.4-x, by reactive

    milling starting from the corresponding oxides. The

    as-prepared materials presented a high degree of compo-

    sitional disorder, which helped to stabilize the c phase.Although a fraction of these structural defects relaxed

    during sintering at 723 C, the c phase never transformedto the orthorhombic or monoclinic forms. In other words,

    the ordering of oxygen vacancies observed on cooling in

    materials prepared by conventional synthesis did not occur.

    The Bi2O3GeO2 system attracted attention because of

    its various stoichiometric phases with attractive optical and

    electrical properties. Milling mixtures of GeO2 and Bi2O3with different stoichiometries lead first to a layered Auri-

    villius-type Bi2GeO5 material and then to different phases

    including basically eulytite Bi4Ge3O12 and a cubic sillenite

    phase of composition Bi12GeO20 [53]. All three phases

    contained a high concentration of vacancies and a high

    degree of compositional disorder. In the Bi2O3TiO2 sys-

    tem, several attempts have been made to prepare the

    Aurivillius-type Bi4Ti3O12 compound, perhaps the most

    popular ferroelectric material ever, starting from mixtures

    of Bi2O3 and different titanium sources [5457]. Various

    Bi2O3:Nb2O5 compositions (1:1, 5:3, and 3:1 molar ratios)

    were milled using WC jars and balls, but the only mixed

    oxide identified by XRD was cubic Bi3NbO7 with an

    excess of poorly crystalline Nb2O5 remaining in the mix-

    tures [58]. Brankovic et al. [59] prepared multiferroic

    BiMnO3 powders starting from Bi2O3 and Mn2O3 using

    stainless steel vials and balls. Although obtaining single-

    phase BiMnO3 by traditional solid state reaction of the

    simple oxides requires pressures over 40 kbar, it was

    readily obtained by high-energy ball milling. Apparently,

    the as-prepared powder presents a tetragonal unit cell,

    similar to that obtained when heating the triclinic form

    above 490 C in traditional materials.

    TiO2 and related systems

    Titanium(IV) oxide has three naturally occurring crystal-

    lographic modifications, namely anatase, rutile, and

    brookite. A high-pressure polymorph with the a-PbOstructure is known as srilankite. As titanium dioxide and

    other titanium-containing compounds are very common

    industrial ceramics, their behavior under high-energy ball

    milling attracted significant attention.

    The evolution of TiO2ZrO2 mixtures was studied by

    several groups [6062]. Milling an equimolar mixture of

    monoclinic ZrO2 and anatase TiO2 using zirconia jars and

    balls produced ZrTiO4. The first step was a polymorphic

    transformation of TiO2, followed by the formation of

    mutual and partially amorphized solid solutions, with either

    TiO2 or m-ZrO2 as solute, which evolved to ZrTiO4 on

    further milling [60]. Complete transformation was

    achieved only by firing at temperatures close to 1100 C.When milling mixtures of TiO2 and ZrO2 using WC vial

    and balls, (Zr1-xTix)O2 (0.44 B x B 0.60) solid solutions

    with the orthorhombic srilankite structure were obtained

    [61]. The as-prepared metastable solid solution was stable

    up to 700 C and even to 1100 C in the 0.44 B x B 0.52composition range.

    J Mater Sci (2013) 48:598611 603

    123

  • Zinc titanates are useful dielectric materials that can be

    sintered at relatively low temperature. There are three

    mixed compounds in the TiO2ZnO phase diagram,

    namely Zn2TiO4 (a cubic spinel), ZnTiO3 (hexagonal

    ilmenite type), and Zn2Ti3O8 (also cubic spinel). Manik

    and Pradhan [63] investigated the evolution of an equi-

    molar ZnO-anatase powder mixture while milling with

    hardened chrome steel balls and obtained a mixture of

    Zn2TiO4, ZnTiO3, and a small amount of rutile. Qian et al.

    [64] analyzed the effect of the Ti source (anatase or rutile)

    on the evolution of different ZnOTiO2 mixtures during

    high-energy ball milling with WC vials and balls. Inde-

    pendent of the Ti source, mixtures with a 2:1 molar ratio

    yielded a new cubic phase, identified as Zn4-xTi2?yO8 and

    thought to be either a solid solution or a mixture of two

    similar cubic phases, Zn2TiO4 and Zn2Ti3O8. The ZnTiO3content of the final powder was larger when starting with

    rutile, whereas the use of anatase favored the formation of

    Zn4-xTi2?yO8. The structural similarities between rutile

    and ZnTiO3 on one side and anatase and Zn2TiO4

    Zn2Ti3O8 on the other may explain the preferential

    formation of either ZnTiO3 or Zn4-xTi2?yO8.

    The behavior of alkaline earth oxide-TiO2 mixtures is

    also interesting. Milling BaOTiO2 mixtures in a high-

    energy shaker mill resulted in nanocrystalline BaTiO3powders [6567], both in nitrogen and in air. Although the

    formation of some BaCO3 is expected in air, continuous

    milling decomposes the carbonate, allowing the formation

    of single-phase barium titanate. Milling mixtures of SrO

    and TiO2 led to the formation of SrTiO3 in a broad com-

    position range [68]. The crystalline product was obtained

    embedded in an amorphous matrix. Poorly crystalline

    Sr2TiO4 was observed when milling a (2 SrO):TiO2mixture.

    Nanocrystalline CaTiO3 has also been prepared by

    milling CaO and TiO2 [69]. If rutile was the Ti source,

    CaTiO3 formed, while a mixture of Ca(OH)2, anatase, and

    CaTiO3 was obtained when milling anatase with CaO. The

    presence of CaO seems to inhibit the polymorphic trans-

    formation from anatase to rutile. Brankovic et al. [70]

    reduced the time needed to obtain pure CaTiO3 by opti-

    mizing the experimental conditions. They were also able to

    obtain CaTiO3 by milling CaCO3 and rutile TiO2. In this

    reaction, the decomposition of the carbonate is the first

    step; thus, it is necessary to let CO2 escape from the milling

    vial. Only an amorphous product formed when mixtures of

    MgO and TiO2 were milled, but the crystalline MgTiO3and Mg2TiO4 phases could be obtained by annealing [71].

    Perovskite-type CaTi1-xMnxO3-d was prepared by

    milling appropriate mixtures of CaO, anatase TiO2, and

    Mn2O3. The product formed with no apparent amorphous

    or crystalline intermediate [72]. Another important low-

    temperature-sinterable dielectric material is CaCu3Ti4O12.

    It has the perovskite structure and can be prepared by

    milling stoichiometric mixtures of CuO, TiO2, and either

    CaO, Ca(OH)2 or CaCO3 using stainless steel jars and balls

    [73, 74].

    Mixed oxides of Ti and lanthanides

    Mixed oxides form relatively easily when milling mixtures

    of TiO2 and Ln2O3 (Ln = Y3? and lanthanides) [15, 75,

    76]. There are only two stoichiometric compounds in the

    corresponding phase diagrams, Ln2TiO5 and Ln2Ti2O7.

    The crystal structure of Ln2TiO5 is related to the mineral

    cuspidine [Ca4(Si2O7)(OH,F)2], while Ln2Ti2O7 titanates

    belong to the pyrochlore family. The ideal pyrochlore

    crystal structure can be visualized as a superstructure of an

    anion-deficient fluorite structure, where a double unit cell

    has 1/8th of the anion positions empty in an ordered way

    and the cations and anions are distributed in four crystal-

    lographically non-equivalent sites. The main difference

    between anion-deficient fluorites and pyrochlores is that

    vacancies are randomly distributed in the anion sublattice

    of the first, whereas ordered in a particular way in the

    second. Many pyrochlores have partially disordered atomic

    arrays and phase transition to a fluorite structure is possible

    by fully disordering the cations and anions in their

    respective sublattices. As pyrochlore-type oxides present

    good high-temperature oxygen ion conductivity, very low

    thermal conductivity, and enhanced radiation resistance,

    they are important materials for possible applications in

    solid oxide fuel cells, thermal barrier coatings, or even as

    immobilization matrices for high-level radioactive waste.

    As in spinel-type oxides, many properties are influenced by

    ion distribution and site occupancy, and thus mechanical

    milling is a promising method for their synthesis.

    Single-phase pyrochlore-type Ln2Ti2O7 (Ln = Gd, Dy,

    Y) titanates were obtained by milling Ln2O3:TiO2 mixtures

    (1:2 molar ratio) using zirconia containers and balls [15]. As

    shown in Fig. 3a and b, the formation of the target materials

    (e.g., Gd2Ti2O7) started with the polymorphic transforma-

    tion of Ln2O3 from its highly symmetric cubic form (C), that

    is stable at room temperature, to the more dense and

    monoclinic B-form. Normally, this transformation takes

    place at high temperature and the transformation tempera-

    ture increases as the size of the lanthanide ion decreases

    [77]. Accordingly, temperatures close to 1850 C are neededto induce the C to B transformation in Dy2O3 compared to

    1200 C for Gd2O3. The thermal stability of C-Y2O3 is evenhigher. The only existing solid state phase transition takes

    place at 2240 C, yielding a fluorite-type Y2O3; monoclinicB-Y2O3 can only be obtained at high pressure. The same C

    to B transformation can be brought about by milling with

    steel tools relatively easily, and the transformation is faster

    604 J Mater Sci (2013) 48:598611

    123

  • for larger cations as anticipated from the behavior upon

    heating [78, 79]. Interestingly, fluorite-type Y2O3 was

    obtained when milling C-Y2O3 with zirconia tools.

    XRD results (Fig. 3b) suggest nucleation and crystalli-

    zation of the rare-earth titanates from an amorphous matrix

    after the amorphization of the initial mixture of oxides. A

    thorough characterization of these pyrochlores prepared by

    mechanical milling suggested that the thermally induced

    transition to a more ordered state proceeds through two

    types of transformations taking place between 750 and

    950 C (Fig. 3c, d); the first one just below 800 C isassociated with an ordering process affecting mostly the

    oxygen sublattice, whereas the second one at slightly

    higher temperatures (differential thermal analysis shows

    generally only a broad exothermic peak) is related to cation

    ordering, grain growth, and the release of strain energy [13,

    14, 80]. Therefore, temperatures close to 1000 C are

    needed to transform these metastable forms into the ther-

    modynamically stable pyrochlores. Previously, disorder in

    these pyrochlore titanates could only be achieved by

    replacing Ti4? by a larger cation, ion irradiation, or high

    pressure.

    Solid solutions of Ln2Ti2-xZrxO7 and Gd2Ti2-xSnxO7with pyrochlore structure were prepared under similar

    experimental conditions [13, 76, 8183]. In these systems,

    disorder generated by milling adds to the chemical disorder

    introduced by substituting Zr4? or Sn4? for Ti4?. A typical

    example is the Gd2Ti2-xZrxO7 system, where anion-deficient

    fluorite-like materials form upon milling, irrespective of the

    zirconium content. This is surprising as both limiting com-

    pounds are pyrochlores and mixed systems formed

    by high-temperature reactions also crystallize with pyroch-

    lore structure. These rather stable materials could

    be progressively ordered by annealing. For example,

    Fig. 3 Mechanochemical synthesis of Gd2Ti2O7 a XRD pattern ofthe starting mixture and b its evolution with milling time. c DTAcurve of the reaction product obtained after milling for 19 h and

    d evolution of its XRD pattern with post-milling thermal treatments.

    Numbers in parenthesis are the Miller indexes of each reflection.Reflections labeled in d are those characterizing the pyrochloresuperstructure and their intensity increases with the degree of

    structural ordering

    J Mater Sci (2013) 48:598611 605

    123

  • Gd2Zr2O7 prepared by ball milling a stoichiometric mixture

    of monoclinic ZrO2 and C-Gd2O3 preserves the fluorite

    structure even after annealing at 1200 C [13], whileGd2Zr2O7 prepared by solid state reaction at high tempera-

    ture has a pyrochlore structure and disorders only above

    1550 C. Intermediate pyrochlore oxides with very unusualcation distributions appear during the annealing of

    mechanochemically prepared powders. Their Gd atoms are

    distributed between the 6- and 8-coordinated positions and

    the Zr atoms are relegated to the larger A site. Similar results

    were obtained in the Dy2Ti2-xZrxO7 system [82]. In general,

    mechanochemical synthesis combined with post-milling heat

    treatments make controlling disorder and crystallite size

    possible; these are the characteristics that determine con-

    ductivity [84], resistance to ion-beam-induced amorphization

    [85], and compressibility in pyrochlores [86].

    Compounds with the Ln2TiO5 composition were also

    obtained by milling the appropriate starting mixtures [75].

    While only one crystal form exists for Ln2Ti2O7 pyroch-

    lores, at least three polymorphs have been identified for

    Ln2TiO5, depending on the annealing temperature and the

    size of the trivalent ion involved. Specifically, as the ionic

    radius of the lanthanide component decreases with the

    increasing atomic number, three structures with increasing

    symmetry are observed: the a- or orthorhombic low-tem-perature form, the hexagonal b- or high-temperature form,and an F-phase, with a cubic fluorite-type structure [87].

    Waring and Schneider [88] reported a reversible transi-

    tion between a and b-Gd2TiO5 at 1712 C, which makespreserving the high-temperature form by quenching difficult

    due to the proximity of the melting point (1765 C). ForDy2TiO5, a relatively narrow stability range from 1330 to

    1650 C has been established for the hexagonal b phase.But, the Ln2TiO5 phases obtained by milling had the high

    temperature and hexagonal forms as a rule, illustrating the

    potential of mechanochemical methods to prepare these

    phases. In addition, powders obtained by ball milling show

    high stability. For example, b-Gd2TiO5 was converted to theorthorhombic low-temperature form only by firing above

    1000 C. This high degree of stability is a general feature ofany ball-milled Ln2TiO5 and Ln2Ti2O7 material.

    Milling a mixture of SnO2 and TiO2 using WC vials and

    balls led to Sn0.5Ti0.5O2 with rutile structure [89], while

    tetragonal PbTiO3 was obtained [90] by milling PbOTiO2mixtures. In the latter case, intermediate Pb3O4 and Ti10O18phases were also observed at the early stages of the milling

    process.

    Rare-earth silicates

    Rare-earth apatite-type silicates have attracted attention

    since the first reports by Nakayama [91, 92] about their

    high oxygen ion conductivity. Their general formula is

    Ln10-x(SiO4)6O02?y (Ln = lanthanide) and their crystal

    structure (hexagonal symmetry) is built of isolated SiO4tetrahedra, with the extra oxide ions (denoted by O0)occupying the center of one-dimensional channels running

    through the structure along the c-axis. These O0 ions areresponsible for the ionic conduction. In addition, there are

    two possible positions for the Ln3? cations, one of them

    coordinated by 7, the other by 9 nearest neighbors. A

    variety of chemical substitutions as well as non-stoichi-

    ometry cation vacancies and/or oxygen excess are possible

    in such a crystal structure, both having an important effect

    on the electrical properties of these materials. The syn-

    thesis of these silicates was successfully attempted by

    milling stoichiometric mixtures of the corresponding ele-

    mental oxides, Ln2O3 (Ln = La3?, Nd3?, Gd3?, and Dy3?)

    and SiO2; ZrO2 containers and balls were used [12, 93, 94].

    Both amorphous and low cristobalite SiO2 were evaluated

    as silicon source in the mechanochemical preparation of

    the La-containing silicate since that is the one showing the

    highest conductivity.

    XRD, infrared, and Raman spectra were used to follow the

    reaction and to establish the time needed to achieve complete

    conversion in each case (Fig. 4). Milling La2O3 with SiO2(amorphous or low cristobalite) in a 4:5 molar ratio yielded a

    product with the chemical formula La9.60(SiO4)6O2.4 after

    short milling. Thus, the characteristic XRD reflections of

    hexagonal La2O3 were completely replaced by the typical

    peaks of an apatite-type silicate after milling for 6 h (Fig. 4a,

    b). After 1 h of milling, the IR spectra still show the char-

    acteristic absorption bands of the SiO bond in silica (e.g.,

    1105, 795, and 484 cm-1). The sample after 3 h of milling is

    a mixture, and for longer milling times (Fig. 4c, d), the

    spectrum is dominated by the peaks characteristic of the SiO

    bond in the silicate (asymmetric stretching modes at 988,

    915, and 881 cm-1) [12]. The conversion time was shorter

    when using amorphous silica as the Si source, e.g., after

    milling for 9 h, the 1105 cm-1 strong absorption band of

    SiO bonds in silica is absent in 4c, but still present when

    starting from low cristobalite SiO2 (4d). As opposed to the

    pyrochlore case already discussed, the starting chemicals and

    the reaction product coexist during the mechanochemical

    synthesis of apatite-type silicates and could be observed by

    XRD after short milling times.

    The same experimental conditions were used to prepare

    silicates with Ln3? = Dy3?, Gd3?, and Nd3?. Interest-

    ingly, previous attempts at preparing lanthanum silicates by

    milling La2O3 with amorphous SiO2 using agate milling

    tools resulted in complete amorphization of the reaction

    mixture, and crystalline lanthanum silicates were only

    obtained after firing at temperatures close to 1000 C [95].This result suggests that the energy transferred to the

    powder particles during milling with agate tools is not

    606 J Mater Sci (2013) 48:598611

    123

  • sufficient to induce the crystallization of the target mate-

    rials. Kharlamova et al. [96] analyzed the possibility of

    using the same method to prepare Fe or Al-doped apatite-

    type lanthanum silicates. Although the apatite-type silicate

    did form on milling in all cases, Fe3? was not incorporated

    into the silicate network when using a-Fe2O3 or FeO(OH)as iron sources. However, milling a mixture of La2O3 with

    either Fe(NO3)3/SiO2 or Fe(HCOO)3 precursors yielded

    single-phase Fe-doped lanthanum silicate. Aluminum

    incorporation seems to be much easier and Al-doped lan-

    thanum silicates with an apatite structure were prepared by

    milling appropriate mixtures of La2O3, SiO2, and Al(OH)3for a very short time. Similar apatite-type germanates,

    which are also very good oxygen ion-conducting materials,

    have been readily obtained by reactive milling [97] starting

    from La2O3 and GeO2.

    ZrO2 and zirconia-based systems

    The milling-induced stabilization of tetragonal and/or cubic

    zirconia, t/c-ZrO2, has been documented for some time now

    [27, 98101]. Research was initially encouraged by reports

    that the room temperature monoclinic modification,

    m-ZrO2, becomes unstable as the crystallite size decreases

    below 30 nm. Conversely, the high temperature and more

    symmetric tetragonal and cubic forms would be stabilized

    when the crystallite dimension is below this critical value.

    However, there was some controversy on whether this phase

    stability inversion takes place due to the lower surface

    energy of the latter forms because of kinetic reasons or even

    as an impurity effect. Interestingly, such phase transitions

    were only observed when milling m-ZrO2 using steel tools,

    but not when using zirconia. According to Karagedov et al.

    Fig. 4 Mechanochemical synthesis of apatite La9.60(SiO4)6O2.4a XRD pattern of the starting mixture and b its evolution withmilling time. Evolution of IR spectra with milling time when using

    La2O3 and amorphous (c) and low cristobalite SiO2 (d) as the silicon

    source. The Miller indexes of the reflections are given in a. Thereflection marked with s is the most intense line of the SiO2 pattern.

    Asterisks in c and d mark the main absorption bands of the SiObands in silica

    J Mater Sci (2013) 48:598611 607

    123

  • [102], the stabilization of t/c-ZrO2 by milling with steel

    tools is the result of two parallel processes: decreasing of the

    particle size and the oxidation and incorporation of iron

    impurities into the ZrO2 crystal network. Annealing the

    milled zirconias suppresses the stabilizing role of Fe and the

    tetragonal or cubic form reverts to the monoclinic phase.

    The influence of reactive milling on phase transforma-

    tion in the ZrO2Y2O3MgO system has been examined by

    Stubicar et al. [103] using WC milling tools. Solid solu-

    tions with either cubic or tetragonal symmetry were

    obtained when milling ZrO2 with Y2O3 and/or MgO, but

    not in binary mixtures of yttria and magnesia.

    Perovskite-type CaZrO3 was obtained by milling an

    equimolar mixture of CaO and monoclinic ZrO2 [104]. Cubic

    ZrO2 appeared after a short milling time, stabilized by the

    presence of CaO, whereas the characteristic reflections of

    CaZrO3 started to emerge at a later stage. The amount of

    orthorhombic CaZrO3 grows steadily upon further milling to

    become the only crystalline phase seen by XRD after 18 h.

    Ball milling has also been used to prepare nanosized powders

    of pure and yttria-doped BaZrO3 starting from BaO2, ZrO2,

    and commercially available zirconias partially stabilized with

    yttria [105]; the milling time needed to achieve complete

    conversion increased as the yttrium content increased. No

    mechanochemical reaction was observed when using BaCO3instead of barium peroxide as the Ba source. Faster reaction

    rates were obtained if the milling vial was opened regularly to

    allow the escape of excess oxygen.

    Some other interesting examples

    The synthesis of perovskite-type LaAlO3 was attempted,

    starting with La2O3 and different sources of Al. No

    mechanochemical reaction was observed when a-Al2O3was the Al source, but a transient alumina obtained by

    heating Al(OH)3 at temperatures between 400 and 800 Cas reactants [106] gave the desired product. Other LnAlO3perovskites (Ln = Nd, Sm, Dy) were also obtained by

    milling the corresponding Ln2O3 oxides with partially

    dehydrated Al(OH)3 as the Al source. Milling mixtures of

    Al2O3 (a or c) and Y2O3 resulted in either an amorphousmaterial or the YAlO3 perovskite, depending on the milling

    conditions [107]. Magnesium aluminate spinel, MgAl2O4,

    has been successfully prepared in a low-energy milling

    device using a stainless steel container and balls, starting

    from mixtures of MgO and c-Al2O3 or AlO(OH) [108].Conversion took place with almost linear kinetics, and

    complete conversion was achieved. The spinel was

    obtained even when a-Al2O3 was used as Al source,although the reaction rate was considerably lower. Milling

    time was greatly reduced by milling a similar MgO plus

    c-Al2O3 starting mixture using more energetic conditions

    [71]. However, attempts to prepare SrAl2O4 and BaAl2O4by milling amorphous Al2O3 and SrCO3 or BaCO3 using

    stainless steel tools were unsuccessful [109]. Other spinel-

    like phases prepared by mechanochemical methods include

    ZnCr2O4 [110] and ASbO4 (A = Fe, V) [111].

    Sodium niobate, NaNbO3, which is an interesting

    material for its phase transitions and electrical behavior,

    has been prepared by milling Na2CO3 and Nb2O5 using

    zirconia vials and balls [112]. The vials had a small aper-

    ture of about 5 mm in the cover to allow the escape of CO2during milling. The formation of orthorhombic NaNbO3goes through an intermediate XRD amorphous phase and

    the reaction is incomplete even after prolonged milling.

    This outcome does not depend strongly on the impact

    energy of the balls [113].

    The perovskite LaMnO3 (LSM) has been prepared by

    milling La2O3 and different Mn sources, namely MnO,

    Mn2O3, or MnO2, in a high-energy shaker mill [114, 115].

    While no crystalline phase was obtained when starting with

    MnO, LaMnO3 was obtained from Mn2O3 and MnO2. Milling

    promoted the reduction of Mn(IV) to Mn(III), but not the

    oxidation of Mn(II) to Mn(III). Consequently, the perovskite

    formed when using MnO2, but not with MnO. The product

    obtained with Mn2O3 and MnO2 was a mixture of the ortho-

    rhombic and cubic polymorphs. Perovskite-type ErMnO3 and

    ErMn0.9Ni0.1O3 have also been obtained by milling stoichi-

    ometric mixtures of Mn2O3, Er2O3, and NiO [116]. The

    as-prepared single-phase materials presented an orthorhombic

    unit cell as opposed to the same material prepared by con-

    ventional solid state reaction which is hexagonal.

    Soft mechanochemistry

    The observation that hydroxides and hydrated compounds

    are much softer than the corresponding anhydrous oxides,

    and consequently react more easily during mechanical

    activation, led to the development of soft mechanochem-

    istry [117]. Mechanochemical reactions between oxides

    require highly energetic milling for an extended length of

    time. The process is slow, energy inefficient, and results in

    substantial contamination of the product. Reacting of the

    softer and more reactive hydroxides requires a much lower

    milling intensity and shorter milling time, saving energy

    and providing cleaner products.

    Of course, the product of milling hydroxides is not the

    desired final product, but a precursor, a mixed, often

    amorphous oxide-hydroxide that has to be annealed to

    obtain the final product. The benefits of the milling step are

    the formation of an intimately mixed precursor and a sub-

    stantial decrease in the annealing temperature and time

    compared to the conventional solid state reaction of the

    oxides. During heat treatment, the unusual defect structures

    608 J Mater Sci (2013) 48:598611

    123

  • are lost and it is difficult to achieve the formation of

    metastable phases. Consequently, if the objective of milling

    a mixture of oxide powders is to prepare an unusual meta-

    stable phase with unique defect structureas is often the

    casethe utility of soft mechanochemistry is limited. But,

    if the goal is only to obtain a certain mixed compound as a

    uniform single phase quickly and without much contami-

    nation, the combination of low-energy milling of hydrox-

    ides combined with annealing is an attractive option.

    Conclusions

    Undoubtedly, mechanical milling has become a powerful

    powder processing method for the room temperature syn-

    thesis of a number of important multicomponent oxides.

    Recently, research activity in the field has grown steadily,

    fueled by the development of new and more powerful mills

    and the availability of new materials for milling media

    (vials and balls) with increasing impact energy and resis-

    tance to abrasion. Contamination from the milling tools

    might be a particularly serious problem when hard and

    abrasive oxides are milled at high intensity for a prolonged

    time, although several steps can be made to reduce the

    effects of contamination. Many multicomponent oxides

    have been prepared featuring a broad variety of crystal

    structures (spinels, pyrochlores, fluorites, apatites, per-

    ovskites, etc.) and chemical composition. In some cases,

    nucleation and crystallization of the reaction product takes

    place from an amorphous matrix containing the reactants,

    whereas in some others, both reactants and reaction prod-

    ucts exist simultaneously during milling. The as-prepared

    phases are frequently metastable, featuring a defect struc-

    ture/microstructure that is difficult to obtain by means of

    any other powder processing method. For example, some

    authors have shown that mechanochemically prepared

    oxides consist of grains presenting a non-uniform coreshell

    structure, where an ordered inner core is surrounded by a

    12-nm thick disordered grain boundary region. This core

    shell configuration is typical of mechanically milled

    nanopowders, but not of nanopowders prepared by other

    techniques. In most cases, these metastable oxides show

    rather good thermal stability; thus, heat treatments at high tem-

    peratures can be used to adjust the defect structure without

    inducing transition to more stable crystallographic phases.

    References

    1. Zyryanov VV (2008) Russ Chem Rev 77:105

    2. Kong LB, Zhang TS, Ma J, Boey F (2008) Prog Mater Sci

    53:207

    3. Wang J, Junmin X, Dongmei W, Weibeng N (1999) Solid State

    Ion 124:271

    4. Delogu F, Mulas G (eds) (2010) Experimental and theoretical

    studies in modern mechanochemistry. Transworld Research

    Network, Trivandrum

    5. Suryanarayana C (2008) Rev Adv Mater Sci 18:203211

    6. Takacs L (2012) Acta Phys Pol A 121:711

    7. Rojac T, Kosec M, Malic B, Holc J (2006) J Eur Ceram Soc

    26:3711

    8. Zyryanov VV (2005) Inorg Mater 41:378

    9. Zyryanov VV (1998) Inorg Mater 34:1290

    10. Slater PR, Sansom JEH, Tolchard JR (2004) Chem Rec 4:373

    11. Christensen AN, Hazell RG, Hewatt AW (1997) Acta Chem

    Scand 51:37

    12. Rodrguez-Reyna E, Fuentes AF, Maczka M, Hanuza J,

    Boulahya K, Amador U (2006) J Solid State Chem 179:522

    13. Moreno KJ, Fuentes AF, Maczka M, Hanuza J, Amador U

    (2006) J Solid State Chem 179:3805

    14. Hayun S, Tran TB, Lian J, Fuentes AF, Navrotsky A (2012)

    Acta Mater 60:4303

    15. Fuentes AF, Boulahya K, Maczka M, Hanuza J, Amador U

    (2005) Solid State Sci 7:343

    16. Takacs L (2002) Prog Mater Sci 47:355

    17. Lefelshtel N, Nadiv S, Lin IJ, Zimmels Y (1978) Powder

    Technol 20:211

    18. Campbell SJ, Kaczmarek WA (1996) In: Long GJ, Grandjean F

    (eds) Mossbauer spectroscopy applied to magnetism and mate-

    rials science, vol 2. Plenum Press, New York

    19. Yang H, Zhang X, Ao W, Qiu G (2004) Mater Res Bull 39:833

    20. Bid S, Sahu P, Pradhan SK (2007) Physica E 39:175

    21. Sepelak V, Bergmann I, Feldhoff A, Heitjans P, Krumeich F,

    Menzel D, Litterst FJ, Campbell SJ, Becker KD (2007) J Phys

    Chem C 111:5026

    22. Chadwick AV, Rush G (2002) In: Knauth P, Schoonmann J

    (eds) Nanocrystalline metals and oxides: selected properties and

    applications. Kluwer Academic Publishers, Dordrecht

    23. Muroi M, Street R, McCormick PG, Amighian J (2001) Phys

    Rev B 63:184414

    24. Sepelak V, Becker KD, Bergmann I, Suzuki S, Indris S, Feldhoff

    A, Heitjans P, Grey CP (2009) Chem Mater 21:2518

    25. Chadwick AV, Pooley MJ, Savin SLP (2005) Physica Status

    Solidi C 2:302

    26. Heitjans P, Masoud M, Feldhoff A, Wilkening M (2007) Fara-

    day Discuss 134:67

    27. Chadwick AV, Pooley MJ, Rammutla KE, Savin SLP, Rougier

    A (2003) J Phys Condens Matter 15:431

    28. Sepelak V, Heitjans P, Becker KD (2007) J Therm Anal Cal

    90:93

    29. Pradhan SK, Bid S, Gateshki M, Petkov V (2005) Mater Chem

    Phys 93:224

    30. Bergmann I, Sepelak V, Becker KD (2006) Solid State Ion

    177:1865

    31. Sepelak V, Feeldhoff A, Heitjans P, Krumeich F, Menzel D,

    Litterst FJ, Bergmann I, Becker KD (2006) Chem Mater

    18:3057

    32. Sepelak V, Bergmann I, Kipp S, Becker KD (2005) Z Anorg

    Allg Chem 631:993

    33. Berchmans LJ, Myndyk M, Da Silva KL, Feldhoff A, Sburt J,

    Heitjans P, Becker KD, Sepelak V (2010) J Alloy Compd

    500:68

    34. Nachbaur V, Taubel G, Verdier T, Jean M, Juraszek J, Houbet D

    (2009) J Alloy Compd 473:303

    35. Harris VG, Fatemi DJ, Cross JO, Carpenter EE, Browning VM,

    Kirkland JP, Mohan A, Long GJ (2003) J Appl Phys 94:496

    36. Verdier T, Nivoix V, Jean M, Hannoyer B (2004) J Mater Sci

    39. doi:10.1023/B:JMSC.0000039201.41114.34

    37. Jiang JS, Yang XL, Gao L, Guo JK (2005) Mater Sci Eng A

    392:179

    J Mater Sci (2013) 48:598611 609

    123

  • 38. Cristobal AA, Botta PM, Bercoff PG, Porto Lopez JM (2009)

    Mater Res Bull 44:1036

    39. Zhang Q, Saito F (2001) J Mater Sci 36. doi:10.1023/A:1017

    520806922

    40. Szafraniak I, Poomska M, Hilczer B, Pietraszko A, Kepinski L

    (2007) J Eur Ceram Soc 27:4399

    41. Da Silva KL, Menzel D, Feldhoff A, Kubel C, Bruns M, Pae-

    sano A Jr, Duvel A, Wilkening M, Ghafari M, Hahn H, Litterst

    FJ, Heitjans P, Becker KD, Sepelak V (2011) J Phys Chem C

    115:7209

    42. Mergen A, Qureshi A (2009) J Alloy Compd 478:741

    43. Poleti D, Karanovic L, Zdujic M, Jovalekic C, Brankovic Z

    (2004) Solid State Sci 6:239

    44. Zdujic M, Poleti D, Jovalekic C, Karanovic L (2009) J Serb

    Chem Soc 74:1401

    45. Zyryanov VV, Uvarov NF (2004) Inorg Mater 40:729

    46. Zyryanov VV (2003) Inorg Mater 39:1163

    47. Zyryanov VV (2005) Inorg Mater 41:156

    48. Ricote J, Pardo L, Castro A, Millan P (2001) J Solid State Chem

    160:54

    49. Castro A, Millan P, Ricote J, Pardo L (2000) J Mater Chem

    10:767

    50. Shantha K, Subanna GN, Varma KBR (1999) J Solid State

    Chem 142:41

    51. Zhang TS, Ma J, Kong LB (2006) J Mater Res 21:71

    52. Zyryanov VV, Uvarov NF (2005) Inorg Mater 41:281

    53. Zyryanov VV, Smirnov VI, Ivanovskaya MI (2005) Inorg Mater

    41:618

    54. Zdujic M, Poleti D, Jovalekic C, Karanovic L (2006) J Non

    Cryst Solids 352:3058

    55. Kong LB, Ma J, Zhu W, Tan OK (2001) Mater Lett 51:108

    56. Han K, Ko T (2009) J Alloy Compd 473:490

    57. Lazarevic Z, Stojanovic B, Romcevic M, Mitric M, Jovalekic C,

    Romcevic N (2008) J Alloy Compd 453:499

    58. Zhou D, Wang H, Yao X, Pang L-X, Zhou H-F (2008) J Am

    Ceram Soc 91:139

    59. Brankovic Z, Stanojevic ZM, Mancic L, Vukotic V, Bernik S,

    Brankovic G (2010) J Eur Ceram Soc 30:277

    60. Gajovic A, Furic K, Music S, Djerdj I, Tonejc A, Tonejc AM,

    Su D, Schlogl R (2006) J Am Ceram Soc 89:2196

    61. Kong LB, Ma J, Zhu W, Tan OK (2002) J Alloy Compd 335:290

    62. Dutta H, Lee Y-C, Pradhan SK (2007) Physica E 36:17

    63. Manik SK, Pradhan SK (2006) Physica E 33:69

    64. Qian D, Gerward L, Jiang JZ (2004) J Mater Sci 39. doi:

    10.1023/B:JMSC.0000039251.24875.ee

    65. Xue J, Wang J, Wan D (2000) J Am Ceram Soc 83:232

    66. Stojanovic BD, Jovalekic C, Vukotic V, Simoes AZ, Varela JA

    (2005) Ferroelectrics 319:65

    67. Stojanovic BD, Simoes AZ, Paiva-Santos CO, Jovalekic C,

    Mitic VV, Varela JA (2005) J Eur Ceram Soc 25:1985

    68. Hungra T, Hungra A-B, Castro A (2004) J Solid State Chem

    177:1559

    69. Berry FJ, Wynn P, Jiang J, Mrup S (2001) J Mater Sci 36.

    doi:10.1023/A:1017940809565

    70. Brankovic G, Vukotic V, Brankovic Z, Varela JA (2007) J Eur

    Ceram Soc 27:729

    71. Stubicar N, Tonejc A, Stubicar M (2004) J Alloy Compd

    370:296

    72. Goncalves P, Canales-Vazquez J, Figueiredo FM (2008) Bol

    Soc Esp Ceram V 47:233

    73. Almeida AFL, de Oliveira RS, Goes JC, Sasaki JM, Souza Filo

    AG, Mendes Filo J, Sombra ASB (2002) Mater Sci Eng B

    96:275

    74. Manik SK, Pradhan SK (2006) Physica E 33:160

    75. Garcia-Martinez G, Martinez-Gonzalez LG, Escalante-Garcia JI,

    Fuentes AF (2005) Powder Technol 152:72

    76. Moreno KJ, Silva Rodrigo R, Fuentes AF (2005) J Alloy Compd

    390:230

    77. Coutures J, Sibieude F, Foex M (1976) J Solid State Chem

    17:377

    78. Begin-Colin S, Le Caer G, Zandona M, Bouzy E, Malaman B

    (1995) J Alloy Compd 227:157

    79. Michel D, Mazerolles L, Berthet P, Gaffet E (1995) Eur J Solid

    State Inorg Chem 32:673

    80. Sanjuan ML, Guglieri C, Daz-Moreno S, Aquilanti G, Fuentes

    AF, Olivi L, Chaboy J (2011) Phys Rev B 84:104207

    81. Moreno KJ, Guevara-Liceaga MA, Fuentes AF, Garca-Barrio-

    canal J, Leon C, Santamara J (2006) J Solid State Chem

    179:928

    82. Moreno KJ, Fuentes AF, Amador U, Santamara J, Leon C

    (2007) J Non Cryst Solids 323:3947

    83. Moreno KJ, Fuentes AF, Garca-Barriocanal J, Leon C, Santa-

    mara J (2006) J Solid State Chem 179:323

    84. Moreno KJ, Fuentes AF, Hanuza J, Maczka M, Amador U,

    Santamara J, Leon C (2007) Phys Rev B 75:18430

    85. Zhang J, Lian J, Zhang F, Wang J, Fuentes AF, Ewing RC

    (2010) J Phys Chem C 114:11810

    86. Zhang FX, Lian J, Zhang JM, Moreno KJ, Fuentes AF, Wang Z,

    Ewing RC (2010) J Alloy Compd 494:34

    87. Sukhanov GE, Guseva KN, Mamsurova LG, Shcherbakova LG

    (1981) Inorg Mater 17:759

    88. Waring JL, Schneider SJ (1965) J Res Natl Bur Stand 69A:255

    89. Kong LB, Ma J, Huang H (2002) J Alloy Compd 336:315

    90. Szafraniak-Wiza I, Hilezer B, Pietraszko A, Talik E (2008)

    J Electroceram 20:21

    91. Nakayama S, Kageyama T, Aono H, Sadaoka Y (1995) J Mater

    Chem 5:1801

    92. Nakayama S, Aono H, Sadaoka Y (1995) Chem Lett 24:431

    93. Martnez-Gonzalez LG, Rodrguez-Reyna E, Moreno KJ, Es-

    calante-Garca JI, Fuentes AF (2009) J Alloy Compd 476:710

    94. Fuentes AF, Rodrguez-Reyna E, Martnez-Gonzalez LG,

    Maczka M, Hanuza J, Amador U (2007) Solid State Ion

    177:1869

    95. Tzvetkov G, Minkova N (2000) J Mater Sci 35. doi:

    10.1023/A:1004705332191

    96. Kharlamova T, Pavlova S, Sadykov V, Chaikina M, Krieger T,

    Ishchenko A, Pavlyukhin Y, Petrov S, Argirusis C (2010) Eur J

    Inorg Chem 4:589

    97. Rodrguez-Reyna E, Fuentes AF, Maczka M, Hanuza J,

    Boulahya K, Amador U (2006) Solid State Sci 8:169

    98. Bailey E, Lewis D, Librant ZM, Porter L (1974) Trans J Br

    Ceram Soc 71:25

    99. Gajovic A, Furic K, Stefanic G, Music S (2005) J Mol Struct

    744747:127

    100. Gateshki M, Petkov V, Williams G, Pradhan SK, Ren Y (2005)

    Phys Rev B 71:224107

    101. Stefanic G, Music S, Gajovic A (2006) Mater Res Bull 41:764

    102. Karagedov GR, Shatskaya SS, Lyakhov NZ (2007) J Mater Sci

    42. doi:10.1007/s10853-006-1245-7

    103. Stubicar N, Bermanec V, Stubicar M, Popovic D, Kaysser WA

    (2004) J Alloy Compd 379:216

    104. Manik SK, Pradhan SK (2005) J Appl Cryst 38:291

    105. Antunes I, Brandao A, Figueiredo FM, Frade JR, Gracio J, Fagg

    DP (2009) J Solid State Chem 182:2149

    106. Zhang Q, Saito F (2000) J Am Ceram Soc 83:439

    107. Alkebro J, Begin-Colin S, Mocellin A, Warren R (2002) J Solid

    State Chem 164:88

    108. Domanski D, Urretavizcaya G, Castro FJ, Gennari F (2004)

    J Am Ceram Soc 87:2020

    109. Chen G, Niu D (2006) J Alloy Compd 413:319

    110. Marinkovic ZV, Mancic L, Vulic P, Milosevic O (2005) J Eur

    Ceram Soc 25:2081

    610 J Mater Sci (2013) 48:598611

    123

  • 111. Berry FJ, Ren X (2004) J Mater Sci 39. doi:

    10.1023/B:JMSC.0000013873.04501.52

    112. Rojac T, Masson O, Guinebretie`re R, Kosec M, Malic B, Holc J

    (2007) J Eur Ceram Soc 27:2265

    113. Rojac T, Kosec M, Malic B, Holc J (2008) J Am Ceram Soc

    91:1559

    114. Cortes-Escobedo CA, Sanchez de Jesus F, Bolarn Miro AM,

    Munoz-Saldana J (2007) Physica Status Solidi C 4:4054

    115. Bolarn AM, Sanchez F, Ponce A, Martnez EE (2007) Mater

    Sci Eng A 454455:69

    116. Moure A, Hungra T, Castro A, Galy J, Pena O, Tartan J, Moure

    C (2010) Chem Mater 22:2908

    117. Avvakumov E, Senna M, Kosova N (2001) Soft mechano-

    chemical synthesis. Kluwer Academic Publishers, Dordrecht

    J Mater Sci (2013) 48:598611 611

    123

    Preparation of multicomponent oxides by mechanochemical methodsAbstractIntroductionGeneral remarksIron oxides and related systemsBismuth(III) oxide and related systemsTiO2 and related systemsMixed oxides of Ti and lanthanidesRare-earth silicatesZrO2 and zirconia-based systemsSome other interesting examplesSoft mechanochemistryConclusionsReferences