Optimization of nanostructured oxide
Transcription
Optimization of nanostructured oxide
N° d’ordre Année 2010 Thèse Optimization of nanostructured oxide-based powders by surface modification Présentée devant L’Institut National des Sciences Appliqués de Lyon et il Politecnico di Torino pour obtenir le grade de docteur École doctorale: Matériaux de Lyon Spécialité : Génie des Matériaux par Fernando Lomello Jury BONELLI Barbara FANTOZZI Gilbert LERICHE Anne PALMERO Paola REVERON Helen Chercheur (Politecnico di Torino) Professeur Emérite (INSA de Lyon) Professeur (Université des Valenciennes et du Hainaut Cambrésis) Chercheur (Politecnico di Torino) Chargée de Recherche EPST - CNRS (INSA de Lyon) Laboratoires de recherche Matériaux: Ingénierie et Science (MATEIS) de l’INSA de Lyon Dipartimento di Scienza dei Materiali e Ingegneria Chimica (DISMIC) del Politecnico di Torino INSA Direction de la Recherche 2007-2010 SIGLE ECOLE DOCTORALE - Ecoles Doctorales – Quadriennal NOM ET COORDONNEES DU RESPONSABLE M. Jean Marc LANCELIN Université Claude Bernard Lyon 1 Bât CPE 43 bd du 11 novembre 1918 M. Jean Marc LANCELIN 69622 VILLEURBANNE Cedex Tél : 04.72.43 13 95 Fax : [email protected] Insa : R. GOURDON ELECTRONIQUE, M. Alain NICOLAS ELECTROTECHNIQUE, AUTOMATIQUE Ecole Centrale de Lyon E.E.A. http://www.insa-lyon.fr/eea Bâtiment H9 M. Alain NICOLAS 36 avenue Guy de Collongue 69134 ECULLY Insa : C. PLOSSU Tél : 04.72.18 60 97 Fax : 04 78 43 37 17 [email protected] [email protected] Secrétariat : M. LABOUNE Secrétariat : M.C. HAVGOUDOUKIAN AM. 64.43 – Fax : 64.54 M. Jean-Pierre FLANDROIS EVOLUTION, ECOSYSTEME, MICROBIOLOGIE, MODELISATION E2M2 CNRS UMR 5558 http://biomserv.univ-lyon1.fr/E2M2 Université Claude Bernard Lyon 1 Bât G. Mendel M. Jean-Pierre FLANDROIS 43 bd du 11 novembre 1918 69622 VILLEURBANNE Cédex Insa : H. CHARLES Tél : 04.26 23 59 50 Fax 04 26 23 59 49 06 07 53 89 13 [email protected] M. Didier REVEL INTERDISCIPLINAIRE SCIENCESHôpital Cardiologique de Lyon SANTE EDISS Bâtiment Central 28 Avenue Doyen Lépine Sec : Safia Boudjema M. Didier REVEL 69500 BRON Tél : 04.72.68 49 09 Fax :04 72 35 49 16 Insa : M. LAGARDE [email protected] M. Alain MILLE INFORMATIQUE ET INFOMATHS MATHEMATIQUES Université Claude Bernard Lyon 1 http://infomaths.univ-lyon1.fr LIRIS - INFOMATHS M. Alain MILLE Bâtiment Nautibus 43 bd du 11 novembre 1918 69622 VILLEURBANNE Cedex Secrétariat : C. DAYEYAN Tél : 04.72. 44 82 94 Fax 04 72 43 13 10 [email protected] - [email protected] MATERIAUX DE LYON M. Jean Marc PELLETIER INSA de Lyon Matériaux MATEIS M. Jean Marc PELLETIER Bâtiment Blaise Pascal 7 avenue Jean Capelle 69621 VILLEURBANNE Cédex Secrétariat : C. BERNAVON Tél : 04.72.43 83 18 Fax 04 72 43 85 28 83.85 [email protected] MECANIQUE, ENERGETIQUE, GENIE M. Jean Louis GUYADER CIVIL, ACOUSTIQUE MEGA INSA de Lyon Laboratoire de Vibrations et Acoustique M. Jean Louis GUYADER Bâtiment Antoine de Saint Exupéry 25 bis avenue Jean Capelle Secrétariat : M. LABOUNE 69621 VILLEURBANNE Cedex Tél :04.72.18.71.70 Fax : 04 72 43 72 37 PM : 71.70 –Fax : 87.12 [email protected] ScSo* M. OBADIA Lionel ScSo Université Lyon 2 M. OBADIA Lionel 86 rue Pasteur 69365 LYON Cedex 07 Tél : 04.78.69.72.76 Fax : 04.37.28.04.48 Insa : J.Y. TOUSSAINT [email protected] *ScSo : Histoire, Geographie, Aménagement, Urbanisme, Archéologie, Science politique, Sociologie, Anthropologie CHIMIE CHIMIE DE LYON http://sakura.cpe.fr/ED206 2 Acknowledgements I would like thank specially to Dr. Paola Palmero, Dr. Barbara Bonelli, Prof. Gilbert Fantozzi, Prof. Edoardo Garrone and Prof. Laura Montanaro for giving me the opportunity to work with them and for taking care of my personal development. During the past three years, in both groups I grew up scientifically and as a person, as well. Many people in the three laboratories encouraged me directly and indirectly, as Prof. Jean-Marc Tulliani, with whom I share the same special passion for Ferrari and motorsports. I also want to thank to my colleagues from the laboratories who supported me and helped me in many different situations. I am grateful to Mr. Yann Aman, Mrs. Mira Jaafar, Mrs. Mirella Azar & Mr. Aurélien Pelletant, who made me feel at home while I stayed in Lyon. Also, I do not want forget about Mrs. Liliane Quillot from INSA-Lyon who always found accommodation for me there. Finally, I want to thank my family and my friends for their unconditional support during these three years. 3 INDEX INTRODUCTION SECTION 1: Surface modification and characterization of nanostructured ceramic powders in the attempt of improving sinterability and microstructural tailoring 1. An introduction to ceramic nanopowders 1.1. Definition and properties of nanopowders 1.2. Main issues connected with the use of ceramic nanopowders 1.2.1. Agglomeration and Dispersion 1.3. Green Density of Nanopowders 1.4. Nanosintering 1.4.1. Pore size and its effects during densification behaviour 1.4.2. Grain Growth 1.5. Nanoaluminas 1.5.1. Metastability of Transition Aluminas 1.5.2. Vermicular Growth & Seeding 1.6. Models developed so far to study the transformation kinetics of aluminas 1.7. Sintering kinetics studied by stepwise isothermal dilatometry (SID method) 1.8. Spectroscopic study of the surface of aluminas 1.8.1. Vibrations of the solid 1.8.2. Surface Vibrations 1.8.3. Surface Hydroxyls Groups 1.8.4. The possible role of defective crystal configurations 1.8.5. The basicity of aluminas 1.8.6. The acidity of aluminas 1.8.7. Ammonia adsorptions 1.8.8. The adsorption of carbon monoxide 2. Surface modification of a transition alumina 2.1. Textural characterization of the starting material 2.2. Effect of dispersion on powder granulometry 4 2.2.1. Effect of the pH of the aqueous suspension on the agglomerate size 2.2.2. Comparison among the other dispersion routes 2.2.3. Effect of the dispersion on the specific area 2.2.4. Effect of dispersion on powder composition and evolution 2.2.5. Influence of the dispersion route on the phase evolution 2.2.6. Effect of the dispersion on the kinetics of transformation 2.2.7. Influence of the dispersion route on forming and sintering 2.2.8. Investigation on Sintering Kinetics of Nanotek powders 2.2.9. Influence of the dispersion route on the final microstructure 2.3. Study of the effect of powder dispersion on its surface properties by means of IR 2.3.1. IR spectra of samples outgassed at increasing temperatures 2.3.2. Adsorption of CO at nominal 77 K 2.3.3. Adsorption of CO 2 2.3.4. Adsorption of NH 3 2.4. Conclusions of the first part and perspectives SECTION 2: Nanostructural materials for extreme applications 3. Nanostructured composite materials: elaboration and properties 3.1. Nanocomposites materials: main classification 3.2. Synthesis of composites powders 3.2.1. Physical Methods 3.2.1.1. Vapour condensation methods 3.2.1.2. Spray Pyrolysis 3.2.1.3. Thermochemical/flame decomposition of metalorganic precursors 3.2.2. Chemical Methods 3.2.2.1. Sol-Gel technique 3.2.2.2. Reverse microemulsions/micelles method 3.2.2.3. Precipitation from solutions 3.2.2.4. Chemical synthesis of pre-ceramic polymers coupled with physical processing techniques 3.2.2.5. Colloidal processing route 3.2.2.6. Mechanochemical Synthesis 3.3. Forming and sintering 3.3.1. Dry-Pressing 3.3.2. Slip-Casting 3.3.3. Sintering 5 3.3.3.1. Hot-Pressing 3.3.3.2. Hot Isostatic Pressing 3.3.3.3. Spark-Plasma Sintering 3.3.3.4. Microwave Sintering 3.4. The interest nanocomposites ceramics: the mechanical properties 3.4.1. Role of the second phase on the retention of the matrix grain growth 3.4.2. Hardening 3.4.3. Change in fracture strength 3.4.4. The strength 3.4.4.1. Reduction in processing flaw size 3.4.4.2. Dislocation networks 3.4.4.3. Crack healing 3.4.5. Toughening mechanism 3.4.5.1. Intrinsic Fracture Energy 3.4.5.2. Crack Bowing 3.4.5.3. Average Internal Stresses 3.4.5.4. Toughening by transformation 3.4.6. 3.5. Crack Deflection Creep of particle reinforced composites 3.5.1. Diffusion Creep 3.5.2. Particle reinforced composites 3.6. The Alumina-YAG system: elaboration and properties 3.6.1. Alumina-Yttria phase diagram 3.6.2. Mechanical properties at room temperature and high temperature 4. Elaboration and Characterization of Al 2 O 3 -5 vol.% YAG Nanocomposites 4.1. Characterization of the as-received powders 4.2. Elaboration of Al 2 O 3 -YAG composites 4.2.1. Dispersion and Spray-drying 4.2.2. Thermal Evolution 4.3. Forming and Sintering 4.3.1. Sintering behaviour followed by dilatometric analysis 4.4. Microstructural Characterization 4.5. Mechanical Characterization 4.5.1. Mechanical properties at room temperature 4.5.2. Mechanical properties at high temperature 4.5.2.1. Creep of TM and CR 4.5.2.2. Creep of Y-TM and Y-CR nanocomposites 4.5.3. Activation energy 4.5.4. Microstructural observation after creep tests 6 4.6. Conclusions of the second part and perspectives References Appendix 1: Characterization techniques 1. Laser Granulometry 2. X-ray Diffraction (XRD) 3. Thermal Analysis: DTA-TG 4. Dilatometry 5. Scanning Electron Microscopy/Environmental Scanning Electron Microscopy 6. High Resolution Transmission Electron Microscopy (HR-TEM) 7. Specific Surface Area 8. Young’s Modulus Resonant Frequency Meter: Grindosonic 9. Hardness tests 10. Creep tests 7 Introduction Conventional ceramic materials are widely used today in different fields: from structural to biological applications and in devices such as lasers, semiconductors and piezoelectric components. Such materials include oxides, carbides, nitrides, mixed oxides and composites, as well. Nanoceramics is one of the fields in which nanoscience and nanotechnology have shown a remarkable progress in producing a variety of advanced materials with unique properties and performance, in terms of chemical inertness, strength, hardness and high-temperature stability. Nanoceramics is a term used to refer to ceramic materials fabricated from ultrafine particles, i.e. with grain sizes less than 100 nm in diameter. In this field, a great deal of research has been accomplished in the last 20 years and has resulted in significant outcomes that are of great impact academically as well as industrially. Nowadays, certain types of nanopowders are produced on an industrial scale. However, it is difficult to produce fully dense parts that retain a nanocrystalline grain size. For this reason, most of the research on nanoceramics has been performed on composites with microcrystalline matrices and nanocrystalline second phases. These type of composites have been extensively studied during the last years. The aim of this thesis is to improve the physico-chemical properties of commercial alumina nanopowders (NanoTek® transition alumina and Taimei α-alumina) in order to develop fully dense materials with tailored microstructure. This thesis is divided into four chapters. The first two chapters deal with the study of surface modification of a transition alumina nanopowder and the last two chapters are devoted to the production and mechanical characterization of Al 2 O 3 /Y 3 Al 5 O 12 (Al 2 O 3 / YAG) micro/nanocomposites for extreme applications. In the first chapter, the state-of-art is reported of the main techniques of powders modification by mechanical means, with the aims to lower powder agglomeration and improve sinterability without inducing a relevant grain growth during firing. Some of the most important issues are reviewed such as agglomeration, nanosintering, principal issues regarding transition aluminas, models for the study of the transformation kinetics, the description of the stepwise isothermal dilatometry and finally study of the surface of aluminas by means of IR spectroscopy. The second chapter shows the experimental results of the study of the effect of dispersion on transition aluminas. The following dispersion routes were adopted: ball-milling (in both alumina and zirconia media); magnetic stirring and attrition milling. The role of the dispersion route on agglomeration, phase composition, phase development and surface functionalities was shown. In addition, the best dispersion method was selected in order to develop dense materials with tailored microstructures. This part of the thesis was carried out by means of several physico-chemical characterization techniques (DTA-TG analysis, XRD, BET Specific Surface Area, HR-TEM, SEM, FT-IR spectroscopy) from the LINCE (Laboratorio INgegnerizzazione materiali CEramici) and SCREAM (Surface Chemistry and REActivity of Materials) laboratories. The third chapter summarizes the state-of-art of the existing alumina-based nanocomposites for extreme applications, with particular reference to the Al 2 O 3 / YAG systems. This chapter contains literature reports concerning the synthesis of composite powders, the principal techniques employed for consolidating the nanocomposites and the mechanical characterization at low- and high-temperature. The fourth chapter presents the development of Al 2 O 3 /YAG nanocomposites from production of the composite powder, by means of surface modification of a Taimei α-alumina powder with yttrium chloride, to its consolidation, by means of conventional (natural sintering) and non conventional 8 (Spark Plasma Sintering and Hot Pressing) techniques, up to its microstructural and mechanical characterization at low- and high-temperature, with emphasis on the creep behaviour. This study stems from the co-tutoring between the LINCE (Politecnico di Torino) and MATEIS (INSA-Lyon) laboratories, by exploiting MATEIS expertise in the field of mechanical characterization of materials. 9 “All truths are easy to understand once they are discovered; the point is to discover them” Galileo Galilei 10 CHAPTER 1 An introduction to ceramic nanopowders 11 An introduction to ceramic nanopowders 1. Introduction In order to take advantage of the properties of bulk nanocrystalline materials, the nanometer range powders have to be densified into parts of certain properties, geometry and size. The key to the nanopowders consolidation process is to achieve densification with minimal microstructural coarsening and/or undesirable microstructural grain growth. Attempts to produce and densify nanopowders started as early as in 19681. These efforts were related to MgO to achieve superplastic behaviour. During the 80s, the nanopowders production was initiated on large scale and the attention was focused on nanopowders processing. At the following decade the effort was emphasized to develop reproducible processing methods for manufacturing nanopowders into useful components which retain nanometer properties. During the second part of the 90s many significant advances in the theory of nanosintering permitted to achieve fully dense parts with nanometer grain size (<100 nm). However, the densification of nanopowders created new challenges. Powders agglomeration, contamination and grain coarsening, complicated the production of large and dense parts. Lower temperatures were employed in order to reduce the grain growth hindering the intergranular bonding, thus compromising the expected mechanical strength. The most recent efforts were successful in overcoming these problems, i.e. agglomeration and grain size control. A number of reviews on nanopowders processing and general reviews in sintering issues have been published2. 1.1. Definition and properties of nanopowders Nanopowders are defined as powdered materials with individual particles having sizes under 100 nm. The particles in nanopowders are smaller than the wavelength of the visible light. The tiny size of nanopowders gives them an extremely high surface area to volume ratio2,3. Specific surface area of the particle is related to the average particle size, based on a geometrical consideration. In case of spherical particles, the formula4 is given below. SSAsphere SAsphere msphere 4 r 2 3 .Vsphere .r (1) where SSA is the specific surface area (generally measured in m2/g), SA is the surface area, m is the mass and V is the volume. Generally, the rougher the particle surface, the greater its surface area regardless of the particle size. The equation showed above underestimates the surface associated with the surface texture of the particles. For that reason, it is better to evaluate the SSA by the BET isotherm (the method is explained in the Appendix I). Another characteristic of nanopowders is their incoherent interfaces which introduces misfit between the crystallites. This is the responsible for modifying locally the atomic microstructure by reducing the atomic density and the coordination between the atoms compared with the perfect crystal3. As it is shown in Figure 1.1 a, the reduced density is visible by high resolution electron micrographs. However, the modified coordination it has been studied by means of X-ray diffraction measurements as it is illustrated in Figure 1.1 b. 12 An introduction to ceramic nanopowders Figure 1.1- (a) Atomic structure in the core of grain boundary between two NiO tilted relative to one another by 36.9° about a common [1 0 0] and (b) Coordination number for nanocrystalline Pd (12 nm crystal size) relative to the single crystalline as a function of the interatomic spacing5. The relative coordination number for interatomic spacing are less than one. For instance, the intercept of the curve at 0.95 indicated that about 5% of atoms of Pd occupy non-lattice sites. This fact is though to be the responsible of changing the properties of nanostructured materials compared with single crystals. The third aspect, which differs the nanosized powders from the classical, micronic ones is the morphological metastability related to fine grain size. This metastability in nanocrystalline materials could be divided into three categories: compositional (extended solution ranges), structural or topological (alternate crystal structure or amorphous phases) and morphological (finely divided structures)6. This is particularly true in the case of nanopowders, the retention of the initial metastability is the challenge. However, the influence in some fields are not clear, for example in the mechanical field it has not been proved yet the benefits in keeping the nanoregime11. Although, an arbitrary limit of < 100 nm is often considered in order to keep the properties as it was mentioned before. Some examples of materials are shown in Table 1.I, in which it is provided some information regarding the critical grain size (maximum particle size in which nanoproperties are kept), the structure and their corresponding equilibrium properties. Table 1.I – Physical and structural Metastability Ranges6. 13 An introduction to ceramic nanopowders 1.2. Main issues connected with the use of ceramic nanopowders The potentialities of nanostructured ceramic powders can be affected by some drawbacks, first of all by their inherent agglomeration, due to Van der Waals forces8. Therefore, in order to produce dense materials, with tailored microstructure, it is important to control all the elaboration steps, beginning with the dispersion. 1.2.1. Agglomeration and Dispersion Most nanocrystalline powders are not composed by single nanometer sized particles. Such particles known as crystallites are bonded together to form agglomerates or aggregates as it is illustrated in Figure 1.2. The distinction between the agglomerate and the aggregate is usually based on the degree of bonding. Soft agglomerates in which the particles are bond by weak van der Waals and hard agglomerates or aggregates in which particles show necks between adjacent particles11. Figure 1.2 – Schematic illustration of nanocrystalline and agglomerates/aggregates9. Those agglomerates and/or aggregates, could present two different types of pores; interagglomerate pores (micronic) which coexist with smaller inter-crystallite pores (nanometric). In case of an agglomerated powder, the most common inconvenient is the presence of both intercrystalline and inter-agglomerate pores. If reference to Figure 1.3 is made, we can observe that the pore size distribution of a non-agglomerated powder (curve c) is characterized by a single peak, related to inter-crystalline pores, whose maximum is centred at a size corresponding to few nanometres. On the other hand, the agglomerated powder (curve a) will have two peaks; the first one corresponding to the small inter-crystalline pores and the second due to larger inter-agglomerate pores. 14 An introduction to ceramic nanopowders Figure 1.3 – Examples of agglomeration distribution: A) heavily agglomerated, B) more simply agglomerated and C) the ideal case of agglomeration2. Nowadays, some improvements in the powder de-agglomeration. For example, when TiO 2 powders are produced by the gas phase condensation route, it was recognized the crucial role of the post-synthesis oxidation step of Ti for avoiding the solid state bonding that it is usually formed within the particles. Similarly, wet chemical processes such as co-precipitation were shown to display significantly less agglomeration of crystallites, thus indicating high sinterability12. However, it is preferable to have soft agglomerates rather than hard agglomerates. Soft agglomerates could be broken by low-energy mechanical methods (i.e. ball milling) or by dispersion in a liquid. Hard agglomerates cannot be easily broken and must be removed from the powder13. In ceramic processing, colloidal suspensions have been used in consolidation for improving the green body homogeneity and the microstructural control during sintering. Each colloidal suspension or slurry contain particles with average sizes between 10-3 μm and 1 μm within a continuous fluid matrix, called colloids14. They are being used increasingly in the consolidation of ceramic powders to produce the green body. The ultrafine particles approach and then separate from each other by Brownian motion, and as a result, settling of particles out of solution does not occur. A basic problem is the stability of colloidal suspension. It is also important to define the attractive forces between forces. Attractive van der Waals forces V vdw exist between particles regardless of whether other forces may be involved. If the attractive force is large enough, the particles will collide and stick together, leading to rapid sedimentation of particle clusters, i.e. flocculation or coagulation. Although in principle, the reduction in the attractive force can be also used, the techniques employed to prevent flocculation rely on the introduction of repulsive forces. There are three types of repulsive forces between particles: the repulsion between electrostatic charges V elec (electrostatic stabilization), the repulsion between polymer molecules V ster (steric stabilization), or the combination of the two forces described before (electrosteric stabilization). In fact, as the colloidal stability is ruled by the total potential energy of interaction between particles (V t ) it could be expressed as: Vt Vvdw Velec Vster (2) 15 An introduction to ceramic nanopowders Figure 1.4- Additivity of the van der Waals force between macroscopic bodies15. To determine the van der Waals forces between macroscopic bodies, it is possible to assume that the interaction between one molecule and a macroscopic body is simply the sum of the interactions with all the molecules in the body (Fig. 1.4). For a sphere of radius a at a distance h, V vdw proposed by Hamaker is given by: Vvdw A.a h h h . 1 ln 6.h 2a h a 2a h (3) where A is the constant of Hamaker16, which can be determined experimentally by direct measurement of the surface forces for bodies separated by vacuum, air or liquids. In order to obtain a stable colloidal dispersion, it is necessary to introduce repulsive forces to offset the attractive potential. There are several ways for achieving this, but the most commonly used are: Electrostatic stabilization [Fig. 1.5 a] in which the repulsion between the particles is based on electrostatic charges on the particles. Steric stabilization [Fig. 1.5 b] in which the repulsion is produced by uncharged polymer chains adsorbed onto to the particle surfaces. Electrosteric stabilization [Fig. 1.5 c], consisting of a combination of electrostatic and steric repulsion, achieved by the adsorption of charged polymers (polyelectrolytes) onto the particle surfaces. Figure 1.5 – Schematic representation of (a) electrostatic stabilization for negatively charged particles, (b) steric stabilization, and (c) electrosteric stabilization17. 16 An introduction to ceramic nanopowders 1.3. Green Bodies Forming The initial step in most densification processes is to compact the powders at room temperature, or cold compactation, to form a green body. Final sintering results are largely dictated by the green compact microstructure. A large number of initial point contact, smaller pores in a green density compact and a uniform pore distribution favour higher final density. The agglomeration, as it was mentioned, introduce an inhomogeneous particle packing structure and poor green density. So, optimized and homogeneous green bodies can be potentially densified, since it is possible to sinter at lower temperatures. On the contrary, inhomogeneities such as particle packing limits the achievable final density. Green bodies are frequently yielded by compaction processes; however, it is more critical for nanocrystalline powders than for conventional ones. Compaction involves sliding and rearrangement of particles, elastic compression at particle contact points, plastic yielding form metals or fragmentation for brittle materials. Sliding and rearrangement in nanopowders are severely restricted owing to large frictional forces among powder particles2,18. These forces are a result of mechanical, electrostatic, van der Waals, and surface adsorption phenomena which increase as a particle size decreases. Particles bonded by weak agglomerates are thought to be easily slided into voids of the green body. Besides, hard agglomerates generate large shearing forces necessary to produce neck cleavage before sliding of these particles can take place19. Figure 1.6 – Compactation behaviour of ceramic powders: hard agglomerated (curve 1) and weakly ZrO 2 -17% Y 2 O 3 and agglomerated (curve 2) 5%Y-TZP 24,25 powders. At lower pressures, hard agglomerates formed in ceramics cannot be fractured. To break these agglomerates it is necessary to reach a critical pressure (Py). On a density-pressure plot (see Figure 1.6 for yttria-stabilized zirconia powders25), the transition point Py where the slope change occurs, is interpreted as the strength of the agglomerates. Above, Py the weak agglomerates break and no transition is seen. When the powder is hardly agglomerated, there is no transition point, as shown in Figure 1.6 - curve 2. As a result, particle rearrangement is hindered and lower green densities are likely to be achieved compared to conventional micron size powders (Figure 1.7). In addition, Chen and co-workers18 found an significant improvement in particle rearrangement by cryogenic compaction through liquid nitrogen. 17 An introduction to ceramic nanopowders Figure 1.7– The effect of particle size on green density upon compactation (uniaxial pressing) for 50 nm Si 3 N 4 (squares), 8 nm ZrO 2 (diamond)20, 20 nm γ-Al 2 O 3 (little triangles) and 20 nm γ-Al 2 O 3 (little squares)18,21. Moreover, Mayo2 found that wet processing provides significantly improved packing uniformity of ceramic nanopowders. For instance, a green density of 74% of the theoretical density (TD) was obtained by Rhodes22 for 12.5 nm Y 2 O 3 -doped ZrO 2 by centrifugation from a slip suspension. Techniques that take advantage of wet processing are the centrifugation, the slip casting and the pressure filtration. Homogeneous and closed-packed green bodies yielded by centrifugation22 give rise to higher fired density, at significantly lower temperatures, if compared to dry-pressed bodies, as shown in Figure 1.8 for nano-crystalline zirconia24. Figure 1.8 – Evolution of fired density as a function of the temperature for centrifuged and dry pressed nanocrystalline ZrO 2 24. Chen and co-workers21 developed an equation for calculating the pressure necessary to achieve a certain density on the ground of volume changes. Literature reports that, at room temperature, it is possible to reach a density as high as 75-90% for nanoceramics19. 18 An introduction to ceramic nanopowders Gallas et al.23 obtained very high densities (90%TD) for nanocrystalline Al 2 O 3 by cold isostatic pressing applying a pressure of 5.6 GPa. The major inconvenient in applying large pressures, apart from the equipment limitations, are the residual stresses that caused the fracture during handling2. However, not all the problems are caused by agglomeration. A large number of particle-particle point contacts per unit volume in a nanopowder, represents a source of frictional resistance to compaction of the powder, and thus the total frictional resistance to compaction could be higher. Consequently, for a given applied stress during compaction, particles tend to re-arrange in a in-homogeneous way compared to conventional powders26. The higher pressures, as it was mentioned previously, can cause fracture during handling23. The figure below illustrated how density and residual stress evolved as compaction pressure is increased. Figures 1.9– Evolution of density and stress gradients in powder compact with increasing applied stress2. Cold isostatic pressing (CIP) could be used to decrease the problems related to stress gradients or density gradients, introduced by uniaxial pressing as it is illustrated in Figures 1.9. Nevertheless, there is one limitation as the maximum pressure could not be <550 MPa. 1.4. Nanosintering Ceramics nanopowders have been fully densified by pressureless sintering: some examples are contained in Table 1.II taken from several references6. 19 An introduction to ceramic nanopowders Table 1.II- Densities and Grain Sizes of some nanopowders densified by pressureless sintering6. In case of nanopowders, the driving force is increased thus to reducing the sintering temperature. In fact, the driving force of sintering is related to the reduction of the total interfacial energy. The interfacial energy expressed as γ.A, where γ is the specific surface energy and A the total surface area of the compact. Figures 1.10– Schematic diagram illustrating the three stages of sintering2 As it is illustrated in Figures 1.10, the sintering curve follows three steps. These steps are described in the following lines. First stage: necks are formed by diffusion between adjacent particles. This neck formation provides an increase of the mechanical strength. There is some speculation in case of nanocrystalline ceramics, as fast surface diffusion combined with applied or residual stresses could occur. This phenomena allows small grains to slip along their boundaries and rearrange in a more packing efficient manner. Second stage: is characterized by the formation of a kind of sponge containing interpenetrating tubular pores to the external surface. In this stage of sintering, the density increases from 6 % to around 90% of theoretical density. In the better situation, this tubular pores shrink in radius29. Third stage: the open pores finally shrink. 20 An introduction to ceramic nanopowders The reduction in size of pores during pressureless sintering occurs by diffusion of the internal surface area. The instantaneous driving force for diffusion is assumed to be inversely proportional to pore’s radius of curvature. The expression given for such phenomena is shown below: A A c co exp co 1 rkT r kT (4) where, c o is the equilibrium vacancy concentration in the bulk material, away from the pore, k is the Boltzmann’s constant, Ω is the atomic volume, γ is the surface tension associated with pore/solid interface and A is a geometric constant with a value of 1 for cylindrical pores and 2 for spherical pores, respectively. The densification is favorized in pores with tight curvature as they have vacancy concentration near the pore surface; consequently, they support a higher vacancy concentration gradient between the pore surface and the nearest grain boundary or external surface, and lately the small pores sustain a greater vacancy flux out of the pore by Fick’s first law. Sintering of nanopowders at low temperature was achieved for the first time during the 80s30,31,24. Andrievski32 was one of the first who studied the effect of additives as Y 2 O 3 and in ZrO 2 . By using Y 2 O 3 , it was achieved 98%TD (theoretical density) sintered at 1220°C. In case of the Bi 2 O 3 , it was achieved a density of 93%TD, by sintering at 935°C. In both cases the final grain size was 600-700 nm. The same concept was extended for different nanograined ceramic powders such as Al 2 O 3 , CdO and TiO 2 by pressureless sintering obtaining fully dense samples. In certain studies, it has been published that fast heating rate in a ramp-and-hold sintering protocol may contribute in obtaining better densities than with low-heating rate36. At low temperature, the mechanism is controlled by surface diffusion which consists in neck formation and little densification. Surface diffusion induce prevalently grain growth. However, at high temperatures the lattice diffusion mechanism helps the material densification avoiding the excessive grain growth. In the following Figure 1.11 it is described how the densification rate changes in function of the temperature. Figure 1.11– Diagram of change in densification rate with the temperature30. The rapid rate will avoid the surface diffusion regime limiting the total grain growth. This method could be suitable for nanopowders due to the high specific surface which generally conducts to surface diffusion. This method was successfully applied in thin films. The ZrO 2 -3 mol% Y 2 O 3 was sintered at different heating rates varying from 2 to 200°C/min; the authors found that by increasing the heating rate, the densification is delayed, as shown in Figures 1.12. 21 An introduction to ceramic nanopowders Figure 1.12– Effect of heating rate on densification of nanocrystalline ZrO 2 -3 mol% Y 2 O 3 13. A second drawback of employing fast heating rates is the generation of density gradients inside the sample. In the following Figure 1.13 it is shown an example of the ZrO 2 -3 mol.% Y 2 O 3 sintered at 20°C/min up to 1200°C and the relative microstructure in different positions over the sample. As shown by the images at higher magnification of Figure 1.13, the density slightly increases from the center (top image) to the border (bottom image), indicating the presence of a density gradient. Figure 1.13– SEM micrographs of ZrO 2 -3 mol.% Y 2 O 3 sintered up to 1200°C (20°C/min) for 2 h13. Chu et al.36 studied the effect of the heating rate on coarsening and sintering of pure ZnO. The authors found that the densification strain rate with a coarsening process is characterized by two types of activation energies, one attributed to densifying processes at higher temperatures and the other to non-densifying process at low temperature. Densification strain rate was defined by the authors as 1 d d , where ρ is the density and is the variation of density as a function of time. dt dt 22 An introduction to ceramic nanopowders The authors studied the volumetric strain rate (the linear densification rate is one third of the volumetric rate) as a function of the heating rate and the sintering temperature. The maximum heating rate shifts to higher temperature with increasing the heating rate. For temperatures above 700°C the densification increases approximately linearly with heating rate. However, below this temperature the densification rate increases slowly (Figures 1.14). The authors also confirmed by SEM observations the hypothesis of the inverse proportional relationship between the grain size and heating rate. Figures 1.14– Sintering at different heating rates of ZnO: (A) densification strain rate versus sintering temperature (B) change on densification strain as a function of the relative density36. Similar data was obtained by Wang et al.31 for α-alumina and α alumina doped with zirconia and titania. The result of the study was that sintering behaviour depends on the dopant, as it is shown in Figures 1.15. By doping with TiO 2 there is an anticipation in densification, while in case of adding ZrO 2 the densification is delayed. In the three samples the densification followed grain boundary diffusion for the heating rates involved. This study also permitted to corroborate the effect of heating rate in the densification, as lower heating rates allows to reach higher fired densities. This fact is well shown by the Figures 1.15 for the pure Al 2 O 3 , as well as, the Al 2 O 3 doped with 5 vol.% TiO 2 and ZrO 2 . 23 An introduction to ceramic nanopowders Figures 1.15– Density and densification as a function fo temperature for Al 2 O 3 , Al 2 O 3 -5 vol.%TiO 2 and Al 2 O 3 -5 vol.% ZrO 2 31. Other hypothesis of the sintering mechanism was proposed by Zeng et al.48 The authors indicated that the initial sintering of α-Al 2 O 3 up to a density of 77%TD is dominated by grain boundary diffusion. 1.4.1. Pore Size and its effects during densification behaviour As in conventional powders, full density or rapid sintering of nanosized powders is achieved when the green sample contains a narrow peak in the pore size distribution (explained previously in Section 1.2.1). It is well known that densification is retarded when the pore distribution is wide. The removal of large pores requires high temperatures. Kingery et al.37 developed a thermodynamic model of the pore shrinkage based on curvature considerations. The authors predicted the pore size for the transition between pore shrinkage and pore growth. A complicated event is the separation of large pores from the grain boundaries that occurs in final sintering stages. Brook et al.38 theorized pore-size-grain size map models to designate the regions where such pore separation from grain boundary occurs. When pore-grain-boundary breakaway takes place, the detached pores will no longer benefit from easy transport paths such as grain boundaries. Instead, the transport is produced by lower volume diffusion and, consequently, densification occurs at a slower rate. This is the stage when open porosity breaks down and pores become closed. Since grain boundary migration is no longer restricted by pores, grain coarsening takes place, or pinning of grain boundary migration due to residual pores is no longer effective. The result is a hindered densification at the final sintering stage which avoids the significant grain growth. This prevents high fired density being achieved. If coarsening occurs in late sintering stages, the overall notion of nanograined size is compromised. Experimentally, a strong correlation 24 An introduction to ceramic nanopowders between the closure of open porosity and the onset of exaggerated grain growth was noted in ceramic nanopowders. Figure 1.16 – Densification behaviour: exaggerated grain growth in agglomerated TiO 2 and Al 2 O 3 non-agglomerated powder37,39,40. As it is shown in Figure 1.16 and 1.17, accelerated grain growth occurs in agglomerated titania powders at densities above 90% or when pores become closed41. However, for non-agglomerated powders (case of alumina in the same Figure), no exaggerated grain growth takes places upon densification. Figures 1.17 – Densification behaviour: Effects of dopants on grain growth of TiO 2 41. Mayo et al.42 studied the influence of agglomerate/pore size on the sintering temperature. In Figure 1.18 are shown the results for a nano-TiO 2 . Non-agglomerated powders (curve 1) achieve higher densities at lower temperature (1073 K), as compared to the agglomerated ones (curves 2 and 3). The second titania powder has a lower crystallite size (<16 nm) as compared to the previous one, but it is agglomerated (curve 2): as a consequence, its densification occurs at higher temperature (1173 K). Finally, the third powder is even more agglomerated (curve 3): as expected, high final 25 An introduction to ceramic nanopowders density is reached at a even higher temperature (1373 K). As declared by the Authors, high sintering temperatures for agglomerated powders are the result of having large inter-agglomerate pores. Figure 1.18 – Effect of agglomerate/particle size on the sintering temperature of nano-TiO 2 : agglomerate size (particle size): 1) non-agglomerated (<40 nm), 80 nm (16 nm), 3) 340 nm (8-10 nm)43. Mayo2 developed a modified sintering law that directly accounts for pore size effects on the densification rate: 1 d 1 1 Q n exp 1 dt d r RT (5) where ρ is the density, d is the particle size, n is a constant dependent on the sintering mechanism, r is the pore radius, Q is the activation energy, R is the gas constant, and T is the absolute sintering temperature. This equation predicts that the highest densification rate occurs for the finest pore size. It was found to hold throughout the sintering process and for large pore sizes, at least when pore size distribution is uniform. This relationship has two implications. First, the pore size in addition to grain size should be controlled during sintering. Fast sintering kinetics result when pore sizes are small. Second, the densification rate is dictated by the instantaneous pore size, not only the initial pore size. Therefore, to maintain a fast sintering rate in late sintering stages, the pore should remain small at late sintering stages. Small pore size is critical throughout the control of the final grain size. For this purpose, a small and uniform pore distribution is desired in the green compact. Most often, pore distribution is dictated by the green density value. A high green density with a small pore population is easily reached in non- or weakly agglomerated powders2,25,44. Rhodes22 was one of the first researchers who demonstrated that de-agglomeration powders may reduce sintering temperature from 1500°C to 1100°C. In Mayo’s review2 it is well summarized the progress of producing non-agglomerated ceramic powders. The advantages of a small pore size and narrow size distribution, in reaching high densities, have been studied by many researchers for non-agglomerated γ-Al 2 O 3 10,33. 26 An introduction to ceramic nanopowders 1.4.2. Grain Growth In nanomaterials, grain coarsening is described by an equation similar to that of conventional materials: d n d o n kt (6) where d is the average grain diameter, d o is the initial grain diameter, n is the grain growth exponent, k is an Arrhenius dependence (k=k o exp(-Q/RT)), and t is the time. From theory, n should take on a value of 2 for normal grain growth in pure, single phase system, 3 for grain growth in the solutes, and 4 in the presence of pores. Some researchers as Harmer et al. have determined experimentally K pure materials and composites45. At low temperatures, n values are high or grain, similar to regular materials. Such increased n values are rationalized by restricted grain boundary mobility which is most commonly due to pore or solute effects. The influence of pores on grain size has been well documented theoretically and experimentally in ceramics2,11. Open pores, which are present throughout the second stage of sintering are effective in limiting grain coarsening (Figure 1.10)32. The pinning action of the pores is difficult to predict7. Liu et al.46 found a linear relationship between the inverse of grain size and the pore surface area per unit volume. The experimental verification of this model was performed on porous materials by adding dopants such as MgO in Al 2 O 3 achieving the total pore closure46,47. Another strategy developed to control grain growth are by kinetic and thermodynamic approaches. Restricting grain boundary mobility using particle pinning effect (see on page 15, 3.4.1 Role of the second phase on the retention of the matrix grain growth) is an example of controlled coarsening kinetics. Zeng et al.48 proposed that by avoiding the powder agglomeration on α-Al 2 O 3 and having an average particle size <20 nm, nanoceramics with grain size <100 nm may be produced. The sintering temperature could be done by using i.e. the Herring’s scaling law. The Herring’s scaling law20, made several assumptions based on the particle size and sintering time based on the operational sintering mechanism. The Herring’s scaling law assumes that the time t 1 required to sinter a particle of diameter D 1 to achieve a sintered neck size of X 1 is known. Then the effect of a change in particle size can be predicted. The sintering time t 2 for a particle of size D 2 to reach the same neck size ratio (X 1 /D 1 =X 2 /D 2 ) is given as, t2 D2 t1 D1 m (7) where D 1 and D 2 are the particle sizes, t 1 and t 2 are the sintering time for different particles sizes, and m the dimensionless scaling-law exponent. The m values were determined for different mechanisms as follows: m=1 for viscous flows and plastic flow, m=3 for volume diffusion and m=4 for surface diffusion and grain boundary diffusion. There are still some discrepancies in using the scale law6. By applying the Arrehenius expression for temperature, the sintering temperature on the particle size could be calculated by using the following formula: d Q 1 1 n ln 1 (8) d 2 R T1 T2 27 An introduction to ceramic nanopowders where Q is the activation energy, R is the gas constant, d 1 and d 2 are the different particles sizes and T 1 and T 2 are their respective sintering temperatures. Messing et al.33 calculated the activation energy for volume diffusion in alumina, obtaining 543 kJ/mol close to volume diffusion. 1.5. Nanoaluminas 1.5.1. Metastability of Aluminas The commercial zirconia powder with nanometric dimension is easy available55. However, in case of commercial α-aluminas, it is almost impossible to produce it with crystallites size lower than 80100 nm. This size makes difficult obtaining nanosized grains in the fired materials. The alternative is to use transition alumina (δ-, γ-, θ-) which usually has a crystallite size of around 50 nm. The higher crystallite size of α-aluminas is caused by the calcination temperature required to yield the α-phase (Figure 1.19). Figure 1.19 – Flowchart showing transformation among Al-O compounds49. 1.5.2. Vermicular Growth & Seeding In case of transition aluminas in the form of δ-, γ-, θ-, the transformation in α-phase is generally accompanied by the formation of a vermicular microstructure, consisting of a network of large pores36 (Figure 1.20 (A)). This morphology is induced by relative density change (Δρ/ρ) of about 17% which accompanies γ→α transformation (theoretical densities 3.41 g/cm3 and 3.897 g/cm3 for gamma and alpha aluminas, respectively). As a consequence of the development of such microstructure, the final stage of sintering requires very high firing temperatures to consolidate the material up to the theoretical density, thus inducing a significant grain growth32. 28 An introduction to ceramic nanopowders Figure 1.20– SEM micrograph of γ-Al 2 O 3 sintered for 100 min at 1400°C: (A) un-seeded and 1.25 wt.% seeded sample51. Nordahl and Messing51 proposed to seed γ-Al 2 O 3 with α-alumina nuclei: they reported that nucleation of -Al 2 O 3 is favoured55, the vermicular microstrusture is hindered and thus, sintering can occur at lower temperature. Figure 1.20 compares the microstructure of fired alumina samples, starting from un-seeded (A) and -alumina seeded (B) γ-Al 2 O 3 powders. If fired to the same temperature (1400°C), a highly dense microstructure can be obtained only in the case of the seeded sample. On the contrary, the un-seeded powder reached full densification at 1600°C, thus resulting in a coarsened microstructure. The seeding stimulate the transition alumina transformation into α-phase upon heating50. An investigation has been done by Yen et al., who evaluated the activation energy of nucleation by DTA profiles (see Section 1.6). The seeded sample showed a lower activation energy of nucleation. However, the activation energy for the growth stage for both seeded and un-seeded remain inalterable. Palkar et al.53 produced nanostructured alumina by rapid nucleation of Al(OH) 3 derived from the sol-gel process. Since the nucleation rate is higher that the growth rate, the rapid nucleation was successful in avoiding the formation of the vermicular microstructure. As a consequence, full densification was achieved at a very low temperature (1250°C). Legros et al.54 claimed that in transition aluminas the formation of α-alumina occurs by nucleation and growth process during heating. The nucleation steps depends on the density of the nuclei which can result from the number per unit of α-grains present in the green bodies seeded. The authors demonstrated that by increasing the mechanical contacts between transition particles, which are potential nucleation sites, it is possible to promote nucleation, as it is shown in Figure 1.21. Heating rate was also considered by the authors to be important as the nucleation sites increase by increasing the heating rate. So the capacity of the particles rearrangement during the phase-transformation seems to be the key. This process of re-arrangement depends on the compact density, heating rate and amount of α-alumina present in the initial powder. 29 An introduction to ceramic nanopowders Figure 1.21- Schematic representation of rearrangement and coalescence mechanisms responsible for high relative density changes during gamma to alpha transition53. Bowen et al.55 concluded that no discernable densification is observable in γ-Al 2 O 3 before α-Al 2 O 3 transformation. This suggests that once diffusion processes become active enough for densification of γ-Al 2 O 3 , the α-Al 2 O 3 transformation is already energetically favoured. Despite all the efforts, the final grain size in aluminas produced from transition powders range from 600 nm to 1 μm9,10. In the same direction, Wu et al.10 have shown that a rapid heat treatment of a transition alumina leads to achieve a higher final density near the theoretical (Figure 1.22). The reason is the formation of α-alumina seeds during the rapid heat treatment which is the responsible of increasing the initial density of the power. Also, the authors evaluated the effect of dopants as the MgO. As it is illustrated in Figure 1.22, the doped material reached a final density near 80%TD. However, in this particular case the powder preparation was achieved by mechanical milling which, as the authors claimed, was the responsible of introducing some seeding which avoided the grain growth. Figure 1.22 – Relative density vs. temperature for compact formed from (I) As-received Powder, (II) Partially Transformed into α-alumina and (III) MgO-doped 250 ppm powder10. 30 An introduction to ceramic nanopowders As with γ-Al 2 O 3 , Messing et al.33,58-61 attempted to obtain microstructures with retained size using Boehmite (γ-AlOOH) as starting material. For this reason the authors proposed to study the influence of α-alumina (0.1/0.4 μm seeds) in Boehmite (γ-AlOOH) gels. They found an optimal ratio of 5.1013 of seeds per cm3 of bohemite gel, which could be reached with a loading of 1.5 wt.% of αalumina seeds. In this particular case the samples reached a final density of >98%TD after 5 min at 1150°C. It was revealed that this technique was efficient in reducing the α-phase transformation temperature and in refining the final microstructure, as it is shown in Figure 1.23. Figure 1.23– SEM micrograph of (A) un-seeded and (B) γ-AlOOH seeded gels after sintering at 1250°C57. Similar results were found by other researchers60-61. For instance in case of Xie et al.61, compared the α-alumina seeding introduced by doping and by the milling medium. In the particular case of the second method, the authors found a better distribution of seeding in the sample. The same procedure was extended to γ-Al 2 O 3 52,62, obtaining similar results such as the promotion of a higher density. Wang et al.52 studied the benefits of introducing α-seeding on γ-Al 2 O 3 , as the barrier of nucleation is removed so the only remaining barrier is diffusion with limited grain growth. 1.6 Models developed so far to study the transformation kinetics of aluminas Avrami equation describes how solids transform from one phase to another at constant temperature. It can describe the kinetics of crystallization, generally the transformation phase in materials. The equation is known as Johnson-Mehl-Avrami (JMA) equation63, and it is written in the following way: x 1 exp kt n (9) where x is the amount of material transformed at the time t; k and n are constants with respect to t the firing time. This equation could be only used to describe the transformation kinetics of many solid state processes under isothermal conditions. The rate constant k, could be expressed as: k (T ) .exp E / RT (10) where ν is the frequency factor, R is the gas constant, and E is the activation energy associated with the process which is often interpreted as the energy barrier opposing the reaction. The constant k o , most called frequency factor, is a measure of the probability that a molecule having energy E will participate in a reaction. 31 An introduction to ceramic nanopowders Isothermal process has been used for the determination of the activation energies of chemicals reactions which usually takes a long time65. The majority of the published works have been devoted to studying the kinetic parameters of crystal growth and glass devitrification by means of differential thermal analysis (DTA)65-66. In these investigations a continuous isothermal method was employed. The experimental procedure is based in evaluating the x, for instance by means of XRD, in function of temperature and time. The results can be plotted as it is illustrated in Figure 1.24. Figure 1.24 – A series of isothermal α-time curves at different temperatures67. By applying the natural logarithm in Eq.7, it is possible to build a second plot ln[ln(1/(1-x)] versus ln(t) in order to obtain the n and k for the different temperatures. Finally, by constructing a third plot using Eq.8 expressed in logarithm form ln k=ln ν-(E/RT) it is possible to obtain the constant ν and the activation energy E. A second method proposed by Kissinger, has already shown how activation energy and frequency factor could be calculated from DTA experiments by making a number of differential thermal patterns at different heating rates. This method has only value for homogenous reactions68-70. Kissinger compared his results with several minerals of the kaolin group data obtained by conventional isothermal techniques obtaining comparable results. The Kissinger’s method is based in the assumption of Eq.7 and Eq.8. When the temperature is changing with time, then the reaction rate is changing with time. The reaction rate is: dx x x dT dt t T T t dt (11) The rate of change of x with temperature, with the time coordinate fixed (∂x/∂T), is zero, because fixing time also fixes the number and position of the particles constituting the system. The only effect, of an instantaneous change in temperature is in the velocity of thermal motion of particles. The total rate reaction may be expressed as, E dx A 1 x e RT dt (12) 32 An introduction to ceramic nanopowders This expression holds for any value of T, whether constant or variable, so long as x and t are measured at the same instant. When the reaction rate is maximum, its derivative with respect of time is equal to zero, thus the reaction rate is E d dx dx E dT RT Ae 2 dt dt dt RT dt (13) The maximum value of dx/dt occurs at temperature T m defined by Ae E RTm E dT RTm 2 dt (14) Replacing in the previous equation d ln 2 Tm E R 1 d Tm (15) where α is the heating rate dT/dt. If the reaction rate order is zero, the peak occurs when the material is exhausted at T m , the temperature decomposition. Evaluating T m at different heating rates, it is possible to build a plot ln (α/T m 2) versus 1/T m , and directly obtain the transformation energy from the slope. The are some requirements for applying Kissinger method taken from the Nordahl et al.72 Those conditions are described in the following lines. The temperature of the thermocouple is the temperature of the entire sample. The peak temperature of the DTA peak represents the temperature of the maximum reaction rate. The heat capacity remains constant during the test. The reaction is first order. Due to the consistency of reaction order obtained by the various analysis techniques given in Table 1.III, it is assumed that the phase transformation is of the first order. This last point is based on the last densification stage, as for example in Boehmite the final microstructure is supported by a denditric or vermicular growth process. 33 An introduction to ceramic nanopowders Table 1.III- Differential thermal analysis methods for determining kinetic parameters of the θ to α-Al 2 O 3 phase transformation in γ-Al 2 O 3 72. The application of Kissinger’s method was employed in evaluating the effect α-alumina seeding on Boehmite and γ-Al 2 O 3 . The authors found that the addition of α-alumina seeds have a significant effect on the transformation techniques in increasing the final density and reducing the activation energy for the θ to α-Al 2 O 3 phase transformation (Figure 1.25). Figure 1.25 – Determination of the activation energy for θ to α-Al 2 O 3 phase transformation in: (A) seeded Boehmite and (B) γ-Al 2 O 3 72. Kao et al.73 evaluated the effect of seeding in θ-aluminas, studying the phase transformation into αAl 2 O 3 . The authors obtained higher values (650 +/- 50 kJ/mol) by a modified Arrhenius method. They proposed an explanation for diffusion mechanism which happens during θ to α transformation. The oxygen atoms diffuse from the matrix the θ across the boundary during transformation to α phase, and they generate vacancy clusters on the θ→α interface. The vacancy clusters are eliminated by the interphase boundary diffusion (D i ) and by the concentration gradient near the interface. The activation energy of lattice diffusion (D l ) and grain boundary diffusion of (D b ) of oxygen in Al 2 O 3 are 700 +/- 30 and 500 +/- 30 kJ/mol, respectively. Due to the similar activation energies of D i and D b which are lower than D l , the diffusion of oxygen should be the mechanism which controls the transformation into α-phase (Figure 1.28). Thus, the transformation mechanism is based on the structural rearrangement by diffusion of Al 2 O 3 in the lattice. 34 An introduction to ceramic nanopowders They sustain that activation energies given by other researchers might imply other transformation mechanisms. Figure 1.26- Schematic diagram of the possible diffusion route and vacancy concentration profile of the θ→α transformation73. Other researchers as Yang et.al76 claimed that the discrepancy in total activation energies exist ranging from ≈200 to ≈650 kJ/mol. The reason is that there are three coexisting barriers, and consequently is difficult to determinate each of them. Yen et al.84 defined three steps in the θ→α transformation (Figure 1.27) which are explained in the following lines: Θ-crystallites grow to the critical size of phase transformation (d θ →d cθ ). Θ-crystallites increase to 22 nm and d cθ transform to α-nuclei (d cθ →d cα ). α-nuclei of d cα (17 nm) coarsen, exceeding the primary size (d p ) and with crystallite size ≈45 nm then the phase transformation is completed. Figure 1.27 – Schematic description of the growth phenomena of θ→α transformation74-75. It is believed that the activation energy of θ-crystallite growth can be the most significant factor that dominates the discrepancy. The θ→α transformation is mainly composed by nucleation and nucleation + growth processes: related activation energy are 85 and 580 kJ/mol74-75 respectively. The last value is closer to the previously reported values. 35 An introduction to ceramic nanopowders Figure 1.28 – Free energy changes during θ→α transformation as a function of radius r of an alumina crystallite74. Yang et al.76 studied the θ→α transformation of three different samples: the first was the unmodified, raw powder; the second was dispersed by mechanical stirring; the third was dispersed and than uniaxially-pressed. The authors observed that the dispersion was effective in disintegrating the agglomerates; the subsequent application of uniaxial-pressing decreased the inter-crystallite distance and promoted the θ→α transformation, by lowering the related temperature. If the θ→α transformation temperature is lowered, the vermicular growth is also hindered because of retaining the crystallite growth (Figure 1.28) 1.7 Sintering Kinetics studied by Stepwise isothermal Dilatometry Stepwise Isothermal Dilatometry (SID) is a technique which has been proven to be very useful in sintering studies of ceramic powders taken from the original publication of El Sayed et al.77. Compared to conventional dilatometry, where the samples is heated at a constant rate, SID has the advantage that the controlling mechanism and the activation energy can be determined. The characteristic of SID is that the heating of the sample is controlled by the magnitude of the derived signal, i.e. the dimensional change signal dl/dt. Sintering takes place in isothermal steps where the time span for each step depends upon the sintering rates and it can be controlled by the setting of the two threshold values. Generally the setting is chosen according to the experimentation. In Figure 1.29 , it is shown a schematic representation of this technique. 36 An introduction to ceramic nanopowders Figure 1.29- Example of shrinkage and temperature curve recorded during stepwise isothermal dilatometry77. Each isothermal steps in the initial sintering stage the densification can be expressed in the following way. y l n K (T ).t lo K (T ) A. ..D / k .T .r P (16) (17) where lo the initial sample length at the start of sintering, K(T) the Arrhenius constant, D the diffusion coefficient, r the particle radius, γ the surface tension, Ω the surface tension, k the Boltzmann constant and A, n, p constants whose values which depend on the sintering mechanisms. In order to apply this equation it is necessary that shrinkage and time are recorded simultaneously. For initial sintering stage it is impossible to determine the absolute time involved in the single steps, thus obliging to remove the variable time. In this way differentiation of y could be done as it is expressed in the expression below. . y K T / N . y N 1 (18) . where y is the shrinkage rate and N 1/ n . . Plotting ln y in function of ln y as in Figure 1.30, determined for all the points in the curve. For each isothermal step, a slope could be obtained. For each steps could be obtained N and n from . the slope and K(T) from the intercept with the ln y axis. 37 An introduction to ceramic nanopowders Figure 1.30 - Shrinkage rate versus shrinkage for isothermal steps during sintering of CeO 2 77. By using the Arrhenius expression and applying logarithms on both sites. ln K T ln A Q / RT (19) where A is the pre-exponential factor and Q the activation energy. This values can be calculated by plotting the ln K (T ) in function of 1/T, as it is shown in Figure 1.31. Figure 1.31 – Arrhenius constant (K(T)) versus reciprocal temperature 1/T for sintering of CeO 2 77. 1.8 Spectroscopic study of the surface of aluminas In the practical language of catalysis and surface chemistry, the term alumina is normally referred to the so-called “transition aluminas”. As this paragraph it is mainly devoted to the surface and spectroscopic features of aluminas, the attention is focused in most commonly used crystalline cubic transition phase Al 2 O 3 systems and other less common systems like, for instance, amorphous aluminas and hexagonal-type aluminas (χ-Al 2 O 3 and κ-AI 2 O 3 ) are neglected. Besides transition aluminas, also some of the features of Al hydrates and of the corundum phase will be necessarily mentioned every now and then. In fact, the structural and coordinative evolution of the AI oxide system from one extreme to the other will turn 38 An introduction to ceramic nanopowders out to be quite important for the definition and comprehension of some of the surface properties of the transition phases. For these reasons, and to avoid confusion, for all the systems considered here either the complete phase designation (e.g., γ-AlOOH, θ-Al 2 O 3 and α-Al 2 O 3 ) or the current phase name (as instance, boehmite, transition aluminas and corundum) will be used. The complex thermal evolution of the Al hydroxides (Al(OH) 3 ; gibbsite, bayerite and nordstrandite) towards corundum, passing through the monohydrates (boehmite and pseudoboehmite) and the various transition alumina phases, has been thoroughly studied in the sixties by Lippens78. His detailed phase description is considered to be still valid and has been constantly referred to in all more recent studies carried out in the field. As mentioned, catalytic aluminas belong to the group of transitional Al 2 O 3 phases, (meta-) stable in the ca. 750-1370 K range. They are usually further divided in two families79: the so-called lowtemperature transition phases (γ- and η-Al 2 O 3 ) and the high-temperature transition phases (δ- and θ-Al 2 O 3 ). On a crystallographic ground, the irreversible thermal transition from low- to hightemperature transition aluminas has been reported to be a continuous process79, i.e., a mere transition of the order disorder type. In fact, the structural differences between the two families of aluminas are relatively small as all transition aluminas belong to the cubic system and have the nature of defective spinels. On a catalytic ground, the passage from low-temperature to high-temperature transition aluminas is more critical. In fact, the high-temperature transition phases are definitely less active than the low temperature ones. This is not merely due to the lower surface area of the former ones (brought about by the higher order and larger particle size), but must reflect a different population of surface active sites. In normal spinels, in which no inversion phenomena occur (inversion as observed, for instance, in NiAI 2 O 4 ), M2+ cations occupy only one eighth of the tetrahedral sites in the cubic close packed array of oxide ions, whereas Al3+ (or other possible trivalent cations) occupy half of the octahedral sites, so that the general formula is often written as MIV[AI 2 VIO 4 ] or, with reference to the unit cell, M 8 IV [Al 16 IV O 32 ]. The defective nature of the transition alumina spinels derives from the presence in aluminas of only trivalent cations, so that some of the lattice positions occupied by cations in mixed-oxide spinels must remain empty to guarantee electrical neutrality. The overall formula referred to the unit cell is then Al 8 IV [Al 13 VI 1/3 □ 2 2/3 O 32 ]. The square symbol denotes the presence of cationic vacancies with respect to the ideal spinel structure: this notation implies, as suggested by Wilson and McConnell for δ-Al 2 O 3 , that cation vacancies are essentially located in octahedral sites, but vacancies should be more realistically imagined as randomly distributed between tetrahedral and octahedral cavities80. It is also logical to expect some of the vacant cationic positions, imposed by the Al 2 O 3 stoichiometry, to be present also in the surface layer of transition aluminas. This is certainly one more factor, besides the double coordination presented by AI ions, that is bound to be somehow responsible for the complex and variable surface situation typical of the transition alumina systems. 1.8.1 Vibrations of the solid In Figure 1.32 reports the IR spectra of the fundamental modes of several Al oxidic systems; Figure 1.32 (A) is taken from a recent work by Busca et al.81 and concerns KBr-pellets of the phases γ-Al 2 O 3 (a), θ-Al 2 O 3 (b) and α-Al 2 O 3 (c), whereas Figure 1.32 (B) shows, for comparison, original KBr-pellet spectra of δ-AI 2 O 3 (d), η-Al 2 O 3 (e) and γ- AIOOH (f). 39 An introduction to ceramic nanopowders The spectra of Figure 1.32 confirm that: (i) when the sole octahedral coordination of AI ions is present, like in corundum and boehmite (c, f), the spectrum is dominated by the strong stretching mode of AIO 6 octahedra82, centered in the 750-600 cm-1 region. The absorption may be split into more components (two in the case of α-Al 2 O 3 and three in the case of γ-AIOOH), due to lowering of the local symmetry and resolution of degenerate modes, but no strong absorptions are ever observed at > 800 cm-1; (ii) when also the tetrahedral coordination of Al ions is present, like in transition aluminas (a, b, d and e), there is also a strong and broad absorption in the region 900750 cm-1, due to stretching vibrations of a lattice of interlinked tetrahedra82. In particular, in the case of the high temperature spinel transition phases the band of AlO 4 tetrahedra becomes broader, stronger and partly resolved into several sharp components, due to the higher crystalline order achieved. This can be observed with δ-Al 2 O 3 preparation (spectrum d in Figure 1.32) and, even more, with θ-Al 2 O 3 (spectrum b in Figure 1.32 and the spectrum of a fairly pure θ-Al 2 O 3 preparation reported by Tarte 82). Figure 1.32 – IR spectra in the skeletal region of some Al oxide preparations. Section A: γ-Al 2 O 3 (a), θ-Al 2 O 3 (b), α- Al 2 O 3 (c), δ-Al 2 O 3 (d), η- Al 2 O 3 (e) and γ-AlOOH (f)79. 1.8.2 Surface Vibrations Surface-localized vibrational modes have been reported for high-area transitions between different lattices of aluminas79,83,84. When a strong base such as pyridine is adsorbed on highly dehydrated transition aluminas, an appreciable increase of the IR transparency on the high-frequency side of the alumina cutoff is observed. This effect was interpreted on the basis of in situ experiments as due to the relaxation of some Al–O vibrations localized in the surface layer and absorbing in the spectral region around 1000 cm−1 79,83,84. Differential absorbance spectra showed that a discrete weak and broad band centred at about 1050 cm−1 appears it is shown the spectra of dehydrated aluminas at increasingly higher temperatures (Figure 1.34). This band was ascribed to Al–O vibration modes localized in surface defective structures of the following type: 40 An introduction to ceramic nanopowders Figure 1.33 – Surface defective structure in highly dehydrated aluminas93. The surface defect would be created during high-temperature surface dehydroxylation and readily destroyed during hydration of the alumina (the band at about 1050 cm−1 disappeared upon water adsorption84. Later, both Marchese et al.84 and Morterra et al.79 confirmed the identification of a surface-localized Al–O vibration mode on both low- and high-temperature transition aluminas. They showed that no surface chemical changes (like, for instance, the rupture of a straightened Al–O bond) are actually needed to have the Al–O vibration shifted downwards to a lower value and that upon the reversible adsorption of a soft base, like CO, a complex band localized in the 1100–1000 cm−1 region is gradually and reversibly eliminated. Morterra et al.86 inferred that on highly dehydrated aluminas the adsorption of CO does not cause the rupture of bonds, but just polarization by weak σ-coordination onto unsaturated surface cations. Figure 1.34 – Differential absorbance spectra showing the elimination of surface-localized vibrational states upon adsorption of CO at 77 K onto γ-Al 2 O 3 activated at 1023 K79. The complex band at about 1050–1100 cm−1 has been interpreted as being due to spinel AlIV–O stretching modes localized at the surface. Such modes would be shifted upwards with respect to the regular bulk AlIV –O stretching mode(s), which absorb at ca. 850 cm−1, and the shift caused by the surface increase of covalence, e.g. of the surface decrease of the Madelung energy, brought about by crystal truncation and surface dehydration. Upon gas–solid adsorption, ligands are added to the coordination sphere of the surface cations, the overall coordination of surface ions increases, and the covalence of the surface then decreases. As a consequence, the surface localized AlIV –O vibrational states shift downward toward the regular spectral positions of the bulk AlIV –O vibrations and outside of the spectral range observable in the in situ transmission mode. 1.8.3 Surface Hydroxyls Groups Broken bonds at crystal surface become saturated as a result of dissociation of the H 2 O molecules: hydrogen is bonded to an oxygen atom, whereas the hydroxyl group bonds to the metal 41 An introduction to ceramic nanopowders atom. The appearance in the spectra of several bands characteristic of hydroxyl groups, as well as their different spectral features and chemical properties (acidity, reactivity), are due to the exposure of several crystal faces and different kinds of defects on the oxide surface. Oxides which had been heated at only moderate temperatures exhibit several hydroxyls bands, the appearance of which depends on the preparative conditions and pre-treatment temperature. In general, the IR band due to the OH stretching vibration of an hydroxyl group appears in the 38003000 cm-1 range and is strengthened in integrated intensity and broadened if the group is involved in hydrogen-bonding interactions. It is now clear that the spectral features of isolated surface hydroxyl groups depend on the chemical structures of the oxides, and their detailed interpretation conversely enables conclusions to be drawn about the structures of the oxide surfaces and about their active sites. The most generalized treatment of the influence of crystalline structure on the IR spectra of surface OH groups has been reported by Tsyganenko and Filimonov88. A reason for the appearance of several bands due to free surface hydroxyl groups in the IR spectra is the capability of the oxygen atoms of the OH groups to be in contact with several immediate neighboring metal atoms. Hence, the number of the latter should exert a decisive influence on the vibrational frequency, νOH, which is observed. For isolated (free) hydroxyl groups, the oxygen can be bound to one, two, three, etc., metal atoms (types I, II, III, etc) – Figures 1.35. Figures 1.35 – Three types of hydroxyls groups are possible at the surface of transition aluminas93. The formation of a coordination bond is found to diminish the frequency of the OH stretching vibration and hence the bands ascribed to the OH groups I, II and III will be positioned accordingly. In order to establish a model for hydroxyl coverage, an analysis of the structure of specific crystal planes has been carried out88, with the expectation being that the planes appearing during crystal growth from the gas phase would be predominant. The oxygen of the OH group can form bonds with several metal atoms if this is allowed by their mutual arrangements. The oxygen atoms of the surface OH groups always occupy those positions where the O atoms would have been present in the infinite lattice. The number of metal atoms around the oxygen of an OH group is always smaller than the coordination number of oxygen in the lattice, i.e. the maximum number of OH groups for the oxide in question equals the oxygen coordination number minus one. The stretching frequencies of non-associated OH groups are thus determined mainly by their local surroundings, that is, by the number of bound metal atoms and their chemical nature, and to a lesser extent, by their coordination numbers. In accordance with data reported by Tsyganenko89, the dependence of νOH for type I hydroxyls bound to atoms of different metallic elements is not a smooth function of these atoms’ electronegativities or their positions in the Periodic Table. For elements in the Second Period, the maximum occurs for aluminum, while for types II and III hydroxyls, the maximum values shift towards the less acidic metals. The value of νOH increases in going from type III to type I hydroxyls. 42 An introduction to ceramic nanopowders Table 1.IV-Positions of absorption bands of isolated OH groups on alumina and their possible assignments93. The conceptions which describe the surface properties of different aluminas are quite interesting. Peri91 in 1965 first proposed a model of the surface of alumina which was founded on the 43 An introduction to ceramic nanopowders hypothesis that the planes exposed preferentially are those of index (100), thus explaining the absorption bands observed in the Al 2 O 3 spectra of five types of surface hydroxyl groups (Table 1.IV). Although this model does not adequately describe all of the surface properties, it is still of considerable interest. In 1978, Knözinger and Ratnasamy89 proposed a very detailed OH model as an extension of the Peri model91. The basic assumption of this model is that a mixture of low-index crystal planes ((111), (110) and (100)) are exposed on the surface of the crystallites. The relative abundance of different faces is assumed to vary for different aluminas. Five types of OH groups were considered, corresponding to the coordination of the hydroxyl groups, either to tetrahedral or octahedral aluminum anions, a combination of each, or to both (Figure 1.36). Five absorption bands in the region 3800–3700 cm−1 have been assigned to these different types of OH groups. The very important concept to arise from this analysis is that surface OH groups have different net charges as a function of their environments. Figure 1.36- Different surface OH groups on alumina according to the model of Knözinger and Ratnasamy90. For example, a type-III OH with a net positive charge of +0.5 is expected to be the most acidic one. According to this model, it is possible to foresee an increase in the acidity of the Al3+ sites, and in the basicity of the O2− sites, with the temperature of activation of the alumina since the lability of the OH groups is a function of the basicity of the oxide. Quantum chemical calculations are needed in order to analyze the nature of the anion hydroxyl groups. In the literature, there are different viewpoints concerning assignment of the highest frequency band at 3800 cm−1 according to Morterra and Magnacca79, this band belongs to an OH group bound to tetrahedral aluminum, whereas Knözinger and Ratnasamy90 have attributed it to the most basic OH group bound to an octahedral aluminum atom. Calculations of both the charge on the hydrogen atom and the νOH of hydroxyl groups bound with different numbers of aluminum atoms in different coordinations showed that the highest-frequency absorption band in the spectrum of aluminum oxides is due to the OH groups bound to four-coordinated aluminum. These results confirm the conclusion made by Knözinger and Ratnasamy90 that the OH groups bound to octahedral aluminum have less acidity. Qualitative quantum chemical models of the dependence of the acidity and frequency characteristics of surface OH groups on both the number and electronegativity of the metal atom connected with them have been examined by Pelmenshikov et al.91. This model has been used to 44 An introduction to ceramic nanopowders explain the lack of a smooth correlation between the νOH of mono-coordinated OH groups and the electronegativity of the metal atoms in numerous oxides. The dependence of νOH on the electronegativity of the metal atom (X M ) is found to be represented by a curve with a maximum reached when q = 0 in the point of change in the polarity of the OH bond. It is not possible to describe all of the properties of the active sites, in particular, for the OH groups, because of the use of various approximations in the supposed models. For alumina oxides, Busca et al.81,94 and Della Gatta et al.95 have proposed models based on the presence of cation vacancies and the corresponding OH structures are reported in Figures 1.37. The investigation of transition aluminas showed that the IR spectra of Al 2 O 3 are also more complex and contain at least nine quite clearly resolved absorption bands of hydroxyls which are not hydrogen-bonded. The question of the assignment of the bands to free or H-bound OH groups of alumina has been analyzed in detail by Chukin and co-workers93. Trokhimets et al.93 interpreted the observed bands on the assumption that the vibration frequency of the OH groups depends on the coordination number of both aluminum and oxygen. It was suggested that the aluminum coordination number on Al 2 O 3 surfaces can be equal to five, as well as four and six. The different types of hydroxyl group absorptions, their effective charges and suggested assignments are listed in Table 1.IV above. As follows from this table, the values of the effective charges on OHs in groups I, II, and III form three almost nonoverlapping regions. Figures 1.37- Possible OH structures and corresponding νOH frequencies, at the surface of defectcontaining spinel transition aluminas. (Symbols: □, cation vacancy; ν, average value of νOH frequency)79. Detailed data on such spectral manifestations of hydroxyl groups on aluminas are reported elsewhere79,90. Chukin and co-workers93 have presented a model of the surface of γ-Al 2 O 3 , which easily explains both hydration and dehydration of the Al 2 O 3 surface, as well as the reversibility of dehydroxylation and rehydroxylation. On the basis of an examination of the decomposition mechanisms on bohemite–corundum faces, a model for the primary crystal lattice was developed, based on IR spectroscopic data and crystal structure analysis of the oxides and hydroxides of aluminum. A joining of dehydroxylated boemite ‘packets’ during heat treatment is assumed to be accompanied by the formation of Al–O–Al bonds and migration of a portion of the Al3+ cations into the freshly formed octahedral cationic vacancies. Three types of Lewis acid sites and six types of OH groups have been identified on the fully hydroxylated γ-Al 2 O 3 . According to Chukin and Seleznev93, the partially dehydroxylated surface of γ-Al 2 O 3 consists of seven types of electron- 45 An introduction to ceramic nanopowders accepting centers, three types of extra lattice Al3+ cations, two types of Al cus 3+ (coordinatively unsaturated site), and two types of electron-deficient O atoms. The values of proton affinity (PA) for different surface hydroxyls of Al 2 O 3 (Table 1.V), calculated from the shifts of the OH frequencies due to formation of H-bonds with bases9can be used to compare the chemical properties of this oxide surface hydroxyls. Table 1.V- values of Proton affinity for different surface hydroxyls of Al 2 O 3 93. In order to determine the number of proton-containing sites, two approaches have been used. One of these is based on measurements of the number of OH groups directly from the intensities of the νOH bands; the second involves measurement of the number of protonated bases generated. If OH groups are not isolated, the concentration of protonic centers is determined by the intensity of the bands in spectra of protonated probes (ammonia, pyridine, etc.), as, for example, has been done in the case of other materials like supported heteropoly acids (HPAs), sulfates, etc. 1.8.4 The possible role of defective crystal configurations Few last considerations remain to be done on the surface hydroxyls of aluminas and in particular on the OH species absorbing at ca. 3775 cm -1. Its unique behavior deserves here some further comment. (I) The band is present only on (all) transition aluminas, whereas it is totally absent on boehmite (Figure 1.38) and on well crystallized α-Al 2 O 3 79,81. It is thus certainly ascribable to OH groups involving in their coordination sphere AIIV ions and this is either in agreement or compatible with most of the OH models proposed. What the various models proposed do not explain is why the OH band at ca. 3775 cm-1 is by far the sharpest and the most reactive OH species at the surface of aluminas. Figure 1.38- The high wavenumbers region of the OH spectrum of some Al oxides. (a) γ-AlOOH activated at 300 K, (b-e) η-Al 2 O 3 , γ-Al 2 O 3 , θ-Al 2 O 3 and α-Al 2 O 3 activated at 773 K79. 46 An introduction to ceramic nanopowders Busca's model does predict a different activity for this OH species (□-AllV-OH) with respect to its high frequency partner (AIIV- OH; v = 3800 cm-1), but the different activity should be expected to be a higher basicity. What is found in practice is the 3775 cm-1 OH species is more active in respect to all types of molecules and participates more actively in all catalytic reactions involving OH groups. The higher activity of the 3775 cm-1 species reflects mainly a higher accessibility of the OH species to all types of surface probes and this should reflect the possible presence of the OH group in particularly exposed zones of the surface. For this reason, Morterra et al.86 have attributed the OH band at 3775 cm-1 to AIIV-OH groups present in portions of the surface belonging to crystallographically defective configurations. The latter are expected to be quite frequent in porous systems of high surface area and poor crystallinity; it is here recalled that Soled92 showed how the surface area of γ- and η-Al 2 O 3 must be high for structural reasons (and, in fact, it is normally as high as 200-250 m2 g-1). None of the models discussed so far indicates a special accessibility for the OH species responsible for the 3775 cm-1 band. This is possibly so because all models discuss the surface of spinel aluminas only in terms of regular crystal plane terminations (the 'top' termination of particles) and do not consider the large incidence of structural defects in high area porous materials (the 'side' terminations of particles). The 3775 cm-1 OH band has been considered, in all models proposed, as due to a 'regular' OH species, whereas it is probably not. The unique nature of the OH band at 3775 cm-1 is demonstrated also by its lability. It was previously reported by Zecchina et al.98 and it is here confirmed by the spectra in Figure 1.39, that on low-temperature transition aluminas (and especially on η-Al 2 O 3 ) the thermal elimination of the OH band at 3775 cm-1 (at temperatures as high as ca. 1100 K) occurs irreversibly. Figure 1.39- The OH spectral pattern of an η-Al 2 O 3 sample treated in various conditions. (a-c) the starting sample, activated at 673, 773 and 1023 K respectively, (d-f) after activation at 1123 K (almost complete dehydroxylation), the samples was retreated at 300 K, and further activated 673, 773 and 1023 K respectively79. 47 An introduction to ceramic nanopowders Rehydration at ambient temperature of the virtually fully dehydrated alumina, followed by a second dehydration run, yields an OH pattern in which the relative intensities of the various OH species are altered and, in particular, the band at 3775 cm-1 is very scarce and, sometimes, virtually absent. Only if the rehydration process is carried out with water vapor at temperatures as high as 670 K, in a sort of hydrothermal process, does the band at 3775 cm-1 recover most of its original intensity87. This indicates that, during the high temperature dehydration of aluminas, reconstruction effects are operative and that the most defective configurations (i.e., those yielding the most reactive surface sites) tend to be annihilated. The interpretation here proposed for the OH band at 3775 cm-1 is confirmed by the behavior of the high temperature transition aluminas. As mentioned in the introduction section, structurally these aluminas are still defective spinel systems, but characterized by higher crystalline order, larger crystallites and more regular (i.e., sharper) terminations of the particles. In the OH spectral region (curve d of Figure 1.38), the higher order of the high-temperature transition aluminas is reflected by a clear splitting of the OH bands of type II (two components are resolved at ca. 3745 and ca. 3730 cm-1) and sometimes also of the OH bands of type III (two components are often resolved at ca. 3710 and ca. 3680 cm-1 93). The OH species at 3775 cm-1, that was tentatively ascribed to AllV-OH groups in exposed and/or defective crystallographic configurations, is either very weak (curve d of Figure 1.38) or totally missing99. Also the adsorption of CO confirms this: on δ-Al 2 O 3 94 and on θ-Al 2 O 3 96 a band of strongly adsorbed CO, located at ν > 2230 cm-1, is either missing or very weak. 1.8.5 The basicity of aluminas The surface basicity of transition aluminas is quite low: in fact, Al 2 O 3 -based catalysts are fairly important, on a catalytic and on a conventional chemical ground, for their acidity rather than for their basicity. As a consequence, whenever a basic catalyst or catalyst support is needed, either other oxides are resorted to or aluminas are doped with variable amounts of basic elements. For a long time the surface basicity of oxides has been tested by the gas-solid adsorption of CO 2 that seems to be still the routine way to evaluate the basicity of aluminas and of its changes with surface chemical modifications. Detailed results resulting from IR spectroscopic investigations of CO 2 adsorption over different oxides have been presented by Busca and Lorenzelli80. It should be taken into consideration that there are differences in the spectral images (structures of surface carbonates) of CO 2 adsorption on oxidized and reduced samples of the transition-metal oxides. In addition, CO 2 can be decomposed on reduced surfaces of transition-metal oxides, oxidizing the surface and producing carbon monoxide. 48 An introduction to ceramic nanopowders Figure 1.40- Spectra of end-on surface complexes of CO 2 on various oxides (PCO 2 =12 Torr), (a) γ-ALOOH activated at 300 K, (b-c) γ-Al 2 O 3 activated at 373 and 773 K, (d) θ- Al 2 O 3 activated at 1023 K, (e) α-Al 2 O 3 activated at 1023 K79. Because CO 2 molecules interact specifically with the cations, and their spectral characteristics depend on the cation properties, the spectra of adsorbed CO 2 molecules have been used to study the properties of aprotic sites on oxides. The spectrum of CO 2 molecules adsorbed on the highly dehydroxylated surface of alumina–silica gel96 has an absorption band at about 2375 cm−1, which gradually shifts toward higher frequencies as CO 2 molecules are removed by evacuation at room temperature. It is assumed that the adsorbed CO 2 molecule preserves its linearity, and is adsorbed on centers through ion-quadruple interactions. In addition to the main intense absorption band at 2375 cm−1 (Figure 1.40), there are two absorption bands at 2405 and 2355 cm-1 in the spectra. Peri96, when studying the CO 2 adsorption on Al 2 O 3 , these to rotation branches of the absorption bands of the adsorbed molecules. It is assumed, moreover, that the adsorption of CO 2 proceeds selectively on a small fraction of the aprotic Lewis centers, namely, on the so-called α-centers. According to Peri97, the centers which selectively adsorb CO 2 molecules also selectively absorb the molecules of butene, acetylene and HCI. Molecules of H 2 O and NH 3 are adsorbed fairly well on all aprotic centers, including the α-centers. The carbonate–carboxylate structures are characterized by absorption bands in the region 1300– 1900 cm-1 which originated on the surface, not only during the CO 2 adsorption, but also due to adsorption with decomposition of several other molecules85,86. These compounds can be produced during the pre-treatment of the sample in vacuo at high temperatures due to the oxidation of organic contaminants, probably due to vapors from greases and oil of vacuum pumps. Moreover, they are produced during the adsorption of the simple molecules like CO 2 and CO96. Thus, the spectrum of CO 2 adsorbed on alumina has absorption bands at ca. 1750, 1635, 1500 and 1235 cm-1 attributed to carbonate surface structures. The absorption band of molecularly adsorbed CO 2 was observed at ca. 2350 cm-1. Carbon dioxide is strongly chemisorbed by alumina at 298 K: in this case, the spectrum has absorption bands at ca. 1770, 1640, 1480 and 1320 cm-1 93. 49 An introduction to ceramic nanopowders Peri 96,97 established the presence of at least three different types of compounds formed during the adsorption of CO 2 on an alumina surface evacuated at 873 K. The absorption bands in the 1800– 1870 cm-1 region were assigned to molecularly adsorbed CO 2 . An interpretation of the spectra of CO 2 adsorbed on Al 2 O 3 has also been made by Fink93 (Table 1.VI). Surface compounds of the types I–III are assumed to be in equilibrium. A type-II compound predominates on the surface at 233 K, while a type-III compound predominates at 298 K. A compound of type- III remains on the surface evacuated at temperatures of up to 473 K. Adsorption at temperatures above 473 K leads to the formation of a Type-IV compound. Table 1.VI- Types of surface compounds formed during the chemisorption of carbon dioxide on the surface of γ-Al 2 O 3 93. Figure 1.41- IR absorbance spectra of CO 2 adsorbed, in a bent form, on various Al oxidic systems. (a) γ-AlOOH activated at 300 K, (b-d) γ-Al 2 O 3 activated at 300, 773 and 1023 K, (c) θ- Al 2 O 3 activated at 773 K, (f) α-Al 2 O 3 activated at 1023 K79. 50 An introduction to ceramic nanopowders The difference in the behavior of the absorption bands at 1480 and 1400 cm-1 during the adsorption of CO 2 on alumina at temperatures ranging from 293 to 523 K indicates that monodentate carbonate structures are not formed at the surface. The absorption bands at 1650 and 1230 cm−1 were attributed to bidentate carbonate structures. There were also differences in the interaction of adsorbed CO 2 molecules with different types of surface hydroxyl groups of alumina. Figure 1.42 – Schematic representation of the band positions of carbonate-like species at the surface of metal oxides79. The interaction of CO 2 with the surface oxygens of different oxide produces spectra implying the formation of several different types of carbonates in the region 1200–1800 cm-1, which is typical for the stretching vibrations of CO bonds in individual carbonates, frequently, the spectra of the surface compounds are identical to the spectra of bulk carbonates, thus implying the presence of free CO 3 2− ions. IR spectra of carbonates can be used to establish the type (and sometimes the structure) of the carbonates formed on the oxide surfaces91,99. 1.8.6 The acidity of aluminas In the case of transition aluminas, whose surface acidity is by far the most important feature, the IR technique has been used quite extensively with great variety of adsorbing probe molecules. Merits and limits of the various IR-adsorption procedures adopted have been reviewed by several authors. The study of ammonia and pyridine adsorption on oxides by means of infrared spectroscopy is a classical method for identifying both Brønsted and Lewis acid sites. The formation of NH 4 + (PyH+) ions is a criterion of the presence of Brønsted acid centers, while the presence of coordinated ammonia (pyridine) shows that Lewis acid centers are present at the surface. The spectral features of the coordinated ammonia (pyridine) molecule are significantly different from those of ammonia (pyridinium) ions (Table 1.VII). This allows the identification of the formation of such complexes by infrared spectroscopy. The greatest differences are observed in the spectral region of the NH deformation vibration frequencies. 51 An introduction to ceramic nanopowders Table 1.VII- Spectral characteristics of NH 3 molecule93. Figure 1.43– Schematic representation of the spectral range of the 8a-8b, 19a-19b modes for some py-containing systems93. Due to their strong basic properties and proton affinities, both ammonia and pyridine are good probe molecules for establishing the presence of even weak Lewis and Brønsted acid sites on the surface. However, the distinctions that can be made by these two probe molecules are not identical. The first difference between them is their relative sizes, i.e. NH 3 < pyridine. Another difference is their relative basicity: in an aqueous solution, ammonia is a stronger base than pyridine99, the pKa value for ammonia is around 9, while that for pyridine is about 5. However, in the gas phase the basicity of pyridine is significantly greater than that of NH 3 103. 52 An introduction to ceramic nanopowders It could be argued that for metal surfaces 103,104, gas-phase basicity is a more appropriate measure of the basic strength of the adsorbate, as no solvation effects occur during adsorption. For metal oxides, the situation is more complicated. When the adsorption proceeds via coordination to the metal cation, the gas-phase basicity is probably the appropriate one to apply. However, when hydrogen-bonding or adsorption on a Brønsted acid site takes place, neighbouring anions or hydroxyl groups may be involved in the interaction100. 1.8.7 Ammonia adsorption Ammonia can be bound to the oxide surface by (i) a hydrogen-bond, NH· · ·O2−, with a surface oxygen or with the oxygen of a surface hydroxyl group, (ii) a bond between the nitrogen atom and a surface hydroxyl atom, or (iii) a coordination bond with a surface cation (Lewis acid site): Figures 1.44 – Schematic representation of the ammonia molecule bounds on a oxide surface93. Complete proton transfer can occur with the formation of an NH 4 + ion (Figures 1.44) or alternatively ammonia dissociative adsorption with the formation of surface NH 2 and OH groups can also take place (Figures 1.45): Figures 1.45 – Schematic representation of the formation of surface NH 2 93. Each type of surface complex gives information about the types of surface centers through qualitative differences in its spectral manifestations and can also reflect qualitative differences in the properties of the same type of center. The spectral manifestations of adsorbed ammonia have been analyzed in detail by Filimonov and co-workers and Davydov 93. The mechanism of ammonia adsorption depends significantly on the state of the hydroxyl coverage of the oxide surface. In the case of strongly hydroxylated surfaces, the adsorption predominantly occurs on the surface hydroxyl groups via the formation of hydrogenbonds between these OH groups and nitrogen atoms (the band δ s NH 3 in such a form of adsorption appears in the region 1150–1100 cm−1). This form of adsorption is thermally unstable and is also reversed by evacuation at room temperature. The number of coordinatively unsaturated surface cations capable of forming a coordinative bond with ammonia molecules(δ s NH 3 at 1280–1150 cm-1) grows with the dehydroxylation of the oxide surface. The detailed position of the δ s NH 3 band is, in this case, sensitive to the electron-acceptor ability of the cation and increases when the cation electronegativity increases93. The δ s -NH 3 bands can be also observed at lower frequencies (<1150 cm-1). In such cases, it is difficult to distinguish these bands from those of hydrogen-bonded ammonia. It is then necessary to use additional procedures to prove that the relevant species is formed and hence that sites of a corresponding 53 An introduction to ceramic nanopowders nature are present on the surface. The frequencies of different ammonia complexes observed in the spectra of ammonia adsorbed on a series of simple and complex oxides are collected in Table 1.VIII. In the spectra of several oxides, especially those pre-evacuated at high temperatures, bands attributable to surface hydroxyl groups appear and grow in intensity upon the adsorption of ammonia. This shows that some of the ammonia molecules dissociate on the surface of these oxides, forming OH and NH 2 groups, thus revealing the presence of surface acid–base pairs such as Mn+O−2. It should be noted that spectral identification of NH 2 groups by using only the absorption bands of the νNH and δNH types, without taking into account the appearance of the surface hydroxyl groups, is not reliable. Table 1.VIII – Special features of coordinated ammonia and ammonium ion on alumina93. The δNH 2 and νNH 2 values for both metal amides M(NH 2 ) n and surface NH 2 groups occur over wide ranges of 1490–1630 and 3200–3580 cm−1 (Table 1.IX). As shown, the position of δNH 2 of such groups is overlapped by the δ as of coordinated NH 3 . Table 1.IX- Vibrational frequencies of NH 2 groups93. This is why there are difficulties in assigning bands to the vibrations of NH 2 species adsorbed on oxides, and especially on those oxides with strong electron-acceptor sites, because both NH 2 groups and coordinatively bound ammonia can be present on the surface simultaneously. The covalence of the M–N bond increases with increasing metal electronegativity, as is shown by the spectra of the alkali-earth metals. This is accompanied by a simultaneous increase in all frequencies of the NH 2 vibrations. 1.8.8 The adsorption of carbon monoxide The adsorption of CO on numerous oxides has been investigated in order to identify both Lewis acid sites and surface oxides. Some of the data obtained are presented in Table 1.X. To describe the detailed structure of the surface and active surface sites, examination of such systems has been made on the basis of both the crystallographic structure of the oxides and data obtained for the number and properties of hydroxyl groups. Table 1.X- Spectral features of CO adsorbed on the surfaces of the various oxide systems32. 54 An introduction to ceramic nanopowders There are a number of studies in the literature of the interaction of CO with aluminas at different temperatures 79-93. There are no doubts that the strongest centers of the alumina surface indicated by the CO adsorption (the band at 2235 cm−1) are due to Al3+ cus ions 79,90,93. Detailed investigations showed that these centers are involved in the so-called ‘X-centers’, the concentrations of which are small (about 1016 centers m−2). X-centers are acid–base pairs, e.g. Al3+ cut O2−. Support for this assignment comes from the absence of dissociative adsorption of both ammonia and propene upon their adsorption on Al 2 O 3 samples treated with NaOH. The band at 2235 cm-1, characteristic of the strongest centers of the alumina surface, is not obtained for the latter types of sample. The nature of the complexes characterized by the absorption band at 2215 cm−1 which is also observed upon CO adsorption at room temperature, is not so clear. However, they are probably also due to the presence of Al3+ cut ions79. A comparison of the spectra of CO adsorbed on Al 2 O 3 at different temperature is shown below in Figure 1.46. At room temperature, two bands at 2210 and 2235 cm-1 are identified, while as the CO pressure increases, the intense absorption of CO in the gas phase appears in the 2100–2200 cm-1 region so that it is difficult to identify the weakly bound complexes on the surface with the lowerfrequency bands91. On saturation of the surface at low temperature, three new bands appear in the spectrum. One of these belongs to physically adsorbed CO because its position corresponds to that in the gas and liquid states of the molecule: νCO = 2140–2150 cm-1. The 2160 cm-1 band corresponds to CO adsorbed on OH groups through hydrogen-bonding, as confirmed a decrease in the intensity of the original OH groups with the simultaneous appearance of lower-frequency νOH bands. Similar results have been obtained by Zaki and Knözinger 90. Bands in the region above 2180 cm-1 belong to the complexes of CO with Lewis acid sites. Thus, the bands in the 2195–2210 cm-1 (species (CO) A ), 2215–2220 cm−1 (species (CO) B ) and 2235– 2240 cm-1 (species (CO) C ) regions are assigned to the presence of coordinatively unsaturated tetrahedral Al3+ ions. The higher-frequency (CO) C and (CO) B bands have been assigned to two families of sites involving Al3+ cut ions located in crystallographically defective configurations, and the lower-frequency adsorbed species (CO) A to sites involving particularly exposed Al3+ cut ions located in extended areas of regular low-index crystal planes79. Fig. 1.46 – IR spectra of CO adsorbed at 293 K (1) and 173 K (2) after dehydration at 773 K: (1,2) γ-Al 2 O 3 93. It should be pointed out that Morterra and Magnacca107 have suggested that Al3+ cuo (coordinatively unsaturated octahedral site) ions located in octahedral positions cannot be coordinatively unsaturated and adsorb CO. The absence of absorption bands in the IR spectra of CO adsorbed on oxidized α-Fe 2 O 3 91 and α-Al 2 O 3 106,107 provides evidence for this suggestion. However, according to calorimetric data, there is a form of adsorption of CO with a heat of adsorption of around 36 kJ mol−1 in the case of α-Al 2 O 3 . Such a value corresponds to the absorption band at 55 An introduction to ceramic nanopowders about 2170–2180 cm-1 which was not observed for these samples, probably due to their low specific areas (2 m2 g−1) and small number of such active centers. In several studies the existence of a great number of Lewis acid sites, including three-, four- and five-coordinated Al3+ cations, as well as pairs of Lewis acid sites which form upon the removal of a bridge oxygen atom, has been analyzed. Borovkov et al.93 observed two bands in the 2210–2215 cm-1 region and at 2245 cm-1 in the infrared spectra of CO adsorbed on different Al 2 O 3 samples. The high-frequency band is attributed to CO molecules adsorbed on α- or X-centers, which include the Al+ O− acid–base pairs. The concentrations of the different surface centers are important characteristics, and many authors have tried to estimate these by different ways, including IR spectroscopy of adsorbed CO. The Lewis acid site concentration on transition aluminas is quite high, depending on the degree of dehydration, and amounts to 0.015–0.03 sites nm−2. Knözinger and Ratnasamy90 have given value of about 0.1 sites nm−2. According to Morterra and Magnacca79, the concentration of Lewis acid sites on γ-Al 2 O 3 activated at 773 K equals 0.1 molecules nm−2 and 0.23 molecules nm−2 on alumina activated at 1023 K. The distribution among the different sites is 0.14 molecules nm−2 of strong acid Lewis sites on the crystal planes, about 0.06 molecules nm−2 for strong Lewis acid sites in crystal defects, and about 0.03 molecules nm−2 for the strongest and more defective sites. The maximum concentration of pyridine strongly adsorbed onto the Al3+ cut (coordinatively unsaturated tetrahedral site) ions (derived from the band at 1625 cm-1) is three to four times larger than the maximum overall CO uptake obtainable at 300 K, due to the fact that pyridine is a stronger base than CO and can detect also weaker sites. 56 CHAPTER 2 Surface modification of a transition alumina 57 Surface modification of a transition alumina 2. Introduction The request to produce fully dense nanostructured ceramics has received much attention over the past 10 years, focusing the interest on the processing of ultra-fine and low-agglomerated primary particles, required for obtaining highly dense ceramics with tailored microstructures. As it was reviewed in the previous chapter, this requirement becomes particularly strict in the case of transition aluminas, whose metastability has a critical influence on their sintering behaviour. The main purpose of the first part of this thesis is to correlate the chemical-physical surface properties of transition alumina to its sintering behaviour. The final goal is pursued by inducing surface modification by dispersion steps, namely magnetic stirring, ball-milling and attrition milling, for improving the densification during sintering obtaining fully dense ceramics with tailored microstructures. For this aim a transition alumina was employed. The optimization of powder thermal reactivity was studied by means of a synergic action of several chemical-physical characterization techniques and compared to the as-received material. 2.1 Textural characterization of the starting material The development of this work has firstly carried out by using a commercial transition alumina powder (Nanotek® by Nanophase Technologies Corporation, Darien, IL, U.S.A.). This alumina powder is prepared by Physical Vapor Synthesis (PVS) as it was published in literature1. This powder, labeled now as A, is characterized by an average particle size of 47 nm, a specific surface area of 35 m2/g and it is composed by two phases, δ and γ alumina, as it is claimed by the producer2. Intensity (a.u.) The XRD pattern of the as-received powder is presented in Figure 2.1. The XRD pattern reveal that powder is a mixture of transition alumina phases, precisely δ-Al 2 O 3 (ICDD file n°. 040877) and γ-Al 2 O 3 (ICDD file n°. 48-0367). 10 20 30 40 50 60 70 Diffraction Angle (2.) Figure 2.1- XRD pattern of A powder. 58 Surface modification of a transition alumina In order to evaluate the particle average size, the as-received powder was submitted to HR-TEM observation as it is shown in Figure 2.2. 50 nm Figure 2.2- XRD pattern of A powder. The observation allowed to define the primary particle size distribution, reported in the following image Figure 2.3. A mean size of about 37 nm was determined, with a geometrical standard deviation of 20.5 nm after analysing some micrographs with a number of particles ranging from 120-150. The result of this analysis is in good agreement with the result published by Azar et al.3. Percentual Frequency [%] 18 16 14 12 10 8 6 4 2 0 0 20 40 60 80 100 120 Particle Size [nm] Figure 2.3 – Particle size distribution of A powder. Similarly, the powder placed on a graphite disk was submitted to SEM observations, Figure 2.4. The powder is characterized by large agglomerates/aggregates; so, the powders was submitted to several dispersion process, as will be discussed in the following. 100 m Figure 2.4– SEM micrograph of the A powder. 59 Surface modification of a transition alumina The initial granulometry of A powder has been performed by using laser granulometry after dispersing the powder in distilled water in order to obtain a suspension having a dilution factor of 7 vol.%. The instrumental detection range is included between 300 nm and 300 m. The cumulative size distribution by both volume and number, as a function of agglomerate size, it is presented in Figure 2.5. Cumulative frequency (%) 100 80 60 40 20 0 0 5 10 15 20 25 Agglomerate Size [m] Figure 2.5 - Cumulative size distribution by volume (solid line) and by number (dashed line) as a function of agglomerate size of A powder As it is shown, in nanometric powders the presence of a low percentage of agglomerates could strongly modify the volume distribution. In contrast, in the case of number distribution it is not possible to evidence the presence of these agglomerates. Agglomerate sizes, corresponding to (d 10 ) 10, (d 50 ) 50 and (d 90 ) 90% of the cumulative volume distribution, were 1.7, 5.5 and 10.4 μm, respectively. In the case of the number distribution, the agglomerates sizes are <0.3, 0.36, 0.89 μm corresponding to (d 10 ) 10, (d 50 ) 50 and (d 90 ) 90%, respectively. Herewith, the crystallization of the A powder into the stable α-phase was evaluated by means of simultaneous differential thermal analysis up and thermogravimetry. The test was carried out up to 1450°C in static air (heating rate of 10°C/min). In following Figure 2.6 are shown the results from DTA-TG tests. -40 0 -30 -20 DTA [V] 0 -4 10 20 TG [%] -2 -10 1325.3°C -6 30 40 Exo 50 -8 60 200 400 600 800 1000 1200 1400 Temperature [°C] Figure 2.6 - DTA-TG Curves of the A powder: dotted line corresponds to DTA and the solid to TG, respectively. 60 Surface modification of a transition alumina DTA curve exhibited an exothermal signal at about 1325°C, imputable to the phase transformation θ→α, as reported in literature4. TG analysis revealed a limited weight loss of about 2.05 %. BET (Brunauer, Emmer, Teller) test was performed by outgassing at 150°C for 2 h. The test was conducted by N 2 adsorption/desorption isotherm at 77K, i.e. after removal of water and surface contaminant. In Figure 2.7 are shown both curves correspond of Type II isotherms, typical of either non-porous adsorbents or relatively large pores. The BET SSA measured on the A sample was 34.5 m2g-1 in good agreement with the information stated by the producer. 35 A 30 3 -1 Volume (cm g ) 25 20 15 10 5 0 0,0 0,2 0,4 0,6 0,8 1,0 p/p0 Figure 2.7- N 2 adsorption/desorption isotherms at 77 K of A powder out-gassed at 150°C: full and white refer to adsorption and desorption. Sintering behaviour of A was investigated by dilatometric analysis performed on uniaxially pressed bars (at 350 MPa), by heating up to 1500°C (heating rate of 10°C/min), with a soaking time of 3 h at the maximum temperature and cooling down up to 20°C (cooling rate of 20°C/min) Zone II -0,02 -0,0002 -0,04 -0,0004 -0,06 -0,0006 L/L o 0,0000 -0,08 -0,0008 -0,10 -0,0010 -0,12 -0,0012 -0,14 -0,0014 200 400 600 800 1000 1200 1400 derivative signal Zone I 0,00 1600 Temperature [°C] Figure 2.8– Dilatometric (solid line) and derivative (dashed line) curves of A. Finally, in order to complete the characterization of the as-received powder a dilatometric test was performed on uniaxially pressed bars (350 MPa) by heating up to 1500°C (heating rate of 10°C/min) and soaking time of 3 h at maximum temperature. 61 Surface modification of a transition alumina The dilatometric curve (Figure 2.8) shows a typical example of a transition alumina, characterized by two regimes of densification during constant heating, named zone I and II. The zone I is imputable to the transformation into α-phase, associated to a volume decrease and rearrangement phenomena in good agreement with literature data4. The peak on the derivative curve located at about 1147°C indicates the temperature associated to δ→α transformation. The densification step (zone II) is located from 1170°C to the maximum sintering temperature which is associated to the α-alumina densification. A certain disagreement between the derivative peak from dilatometric analysis and the value determined by DTA can be imputed to the nature of the samples, powder compact for the former and powders for the latter. In the derivative curve, the peak related to the temperature of maximum densification rate is not observable. The green density evaluated by weight and geometrical measurements of pressed bar is 1.86 g/cm3 which corresponds to 53.3 % of the theoretical density (taking in consideration that Nanotek3 has a reference value given by the producer of 3.49 g/cm3). The fired density was 2.79 g/cm3 which implies a 70.4 %TD (reference value for α-alumina 3.96 g/cm3). 2.2 Effect of dispersion on powder granulometry In order to reduce the starting agglomerate size, A powder was dispersed by magnetic stirring in pure distilled water, by preparing aqueous suspensions with a solid content of 50%, and maintained under magnetic stirring for 170 h. This powder is labeled as A MS . The pH was measured on the suspension as a function of time. The suspension presented a pH of about 5.1 for the as-dispersed powder and it stabilized to the value of 5.6 after 24 hours of magnetic stirring, remaining almost stable. It is possible to assume that 5.6 is the natural pH of the suspension. After dispersion, the cumulative size distribution by volume of the dispersed samples are collected in Figure 2.9, showing a significant reduction of soft agglomerate size. Cumulative Distribution [%] 100 80 60 40 Samples d50 [µm] A AMS 8,08 0,44 20 0 0 4 8 12 16 20 24 Agglomerate Size [m] Figure 2.9 - Cumulative distribution by volume of: A (solid line, no symbols) and A MS (circles). Insert: d 50 values of the overall distributions After dispersion, agglomerates size corresponding to (d 10 ) 10, (d 50 ) 50, and (d 90 ) 90% of the cumulative volume distribution, were <0.3, 0.44 and 0.75 μm, respectively. In Figure 2.10 is illustrated as a comparison the evolution of d 50 during dispersion up to 170 h. As it is shown d 50 was successfully reduced in 1 order of magnitude from 8.08 μm to 0.44 μm. 62 Surface modification of a transition alumina 9 A 8 7 d50 [m] 6 5 4 3 2 AMS 1 0 0 20 40 60 80 100 120 140 160 180 stirring time (h) Figure 2.10- Evolution of d 50 values as a function of the dispersion time under magnetic stirring 2.2.1 Effect of the pH of the aqueous suspension on the agglomerate size A parallel study has been done in order to study the effect of the pH on de-agglomeration. On the ground of the results reported in literature (Figure 2.11), the Nanotek powder has the isoelectric point at about 9.3. At the isoelectric point, the surface charge of the particles is near zero and the absence of a surface charge implies agglomeration. Figure 2.11 – Zeta potential of the Nanotek powder as a function of pH5. In this context, a stable suspension can be prepared at pH higher that 10 or lower than 8. Taking in consideration the zeta potential curve, pH equal to 4 should be considered as the best condition for the dispersion. Nanotek was dispersed by magnetic stirring, as in the previous case having a solid content of 50 wt%. Diluted HCl was added to decrease the pH up to 4. The pH was monitored and corrected every 24 hours. As it is shown in Figures 2.12 / 2.13 , after 170 h of dispersion agglomerates size 63 Surface modification of a transition alumina corresponding to (d 10 ) 10, (d 50 ) 50 and (d 50 ) 90% in volume, were <0.3, 0.44 and 1.11. If compared to the natural, unmodified pH suspension, no significant improvement was found. Cumulative Distribution [%] 100 80 60 Samples A AMS pH=4 40 d50 [m] 8.08 0.64 20 0 0 4 8 12 16 20 24 Agglomerate Size [m] Figure 2.12 – Cumulative distribution by volume of: A (solid line, no symbols) and A MS pH=4 (circles). Insert: d 50 values of the overall distributions 9 A 8 7 d50 [m] 6 5 4 3 2 AMS pH=4 1 0 0 20 40 60 80 100 120 140 160 180 stirring time (h) Figure 2.13 – Evolution in time of the Agglomerate Size (d 50 ) in function of the dispersion time of A MS pH=4. A second test has been conducted at pH equal to 3, with the aim of evaluating the effect of a lower pH on the powder dispersability. As it is shown in figures 2.14/2.15 , the agglomerate sizes corresponding to (d 10 ) 10, (d 50 ) 50 and (d 90 ) 90 % in volume were <0.3, 0.53 and 0.9 μm. As a conclusion, the modification of the slurry pH from natural value (5.6) to 4 and 3 did not produce any significant effect neither on the final particle size, nor on the dispersion time. On the ground of such results, the following dispersions were carried out on unmodified suspensions, left to their natural pH. 64 Surface modification of a transition alumina Cumulative Distribution [%] 100 80 60 d50 [m] Samples A AMS pH=3 40 8.08 0.53 20 0 0 4 8 12 16 20 24 Agglomerate Size [m] Figure 2.14 – Cumulative distribution by volume of: A (solid line, no symbols) and A MS pH=3 (circles). Insert: d 50 values of the overall distributions 9 A 8 7 d50 [m] 6 5 4 3 2 AMS pH=3 1 0 0 20 40 60 80 100 120 140 160 180 stirring time (h) Figure 2.15 – Evolution in time of the Agglomerate Size (d 50 ) in function of the dispersion time of A MS pH=3. 2.2.2 Comparison among other dispersion routes As in the previous case, powder A was submitted to ball milling with the aim of comparing the effectiveness of such method with the lower energetical magnetic stirring. The angular velocity was optimized by following the literature data6,7. The test were conducted under a nominal angular velocity ω n , using the following formula (1) taken from literature: 65 Surface modification of a transition alumina n 0.6c 25.4 Di where c 42.3 Di (1) Figure 2.16 – Scheme of Ball Mill used for this study6. A PE container of 50 ml. volume was employed to determinate the nominal rotational velocity (see Figure 2.16). The test was followed up by testing both α-alumina (Bitossi-Φ= 2 mm, A BMα ) and zirconia (TSZ TOSOH- Φ= 1,75 mm, A BMz ) spheres (powder to spheres weight ratio of 1:10). The solid content of 50wt.% was selected to follow these tests. After 3 hours of dispersion, similar agglomerates sizes were found for A BMα and A BMz and they were similar to A MS . The results of the cumulative size distribution by volume for both spheres nature are collected in Figure 2.17. Cumulative Distribution [%] 100 80 60 40 20 0 0 4 8 12 Samples d50 [µm] A ABM 8,08 0,50 ABMz 0,58 16 20 24 Agglomerate Size [m] Figure 2.17 - Cumulative distribution by volume of: A (solid line, no symbols), A BMα (triangles) and A BMz (squares). Insert: d 50 values of the overall distributions. A BMα and A BMz powders were submitted to XRD analysis, in order to evidence if powders were contaminated by the milling media. No -alumina or zirconia XRD reflections were found in the dispersed powders, as shown in Figure 2.18: if contaminated, the pollutants amount was so low, not to be detectable by XRD. 66 Surface modification of a transition alumina Intensity (a.u) ABM 10 20 30 40 50 ABMz 60 70 Diffraction Angle (2. Figure 2.18 – XRD patterns of Nanotek after ball milling. The last technique employed to disperse the powder was attrition milling. This dispersion route was selected in order to compare the performance of this higher-energy method among the others. In this particular case, the sample was labelled as A AM . The test was conducted using αalumina (Bitossi-Φ= 2 mm) spheres (powder to spheres weight ratio of 1:10). As in the previous case, the angular velocity was set up following literature data9-10. After 6 h of attrition milling, the cumulative size distribution by volume of the dispersed samples corresponding to (d 10 ) 10, (d 50 ) 50 and (d 90 ) 90% of the cumulative volume distribution, were 0.46, 0.88 and 1.54 μm as it is shown in the figure below (Figure 2.19). Cumulative Distribution [%] 100 80 60 40 Samples d50 [µm] A AAM 8.08 0.88 20 0 0 4 8 12 16 20 24 Agglomerate Size [m] Figure 2.19 - Cumulative distribution by volume of: A (solid line, no symbols) and A AM (squares) The results obtained denoted that the adopted milling conditions have to be optimized. In fact, a lower powders dispersability was yielded if compared to the previous routes. Additionally, some α-alumina seeds contaminated the powder, as determined after analyzing the attrition milled material by XRD. In fact, near transition aluminas, some peaks were ascribed to α-alumina phase (ICDD file n°. 81-2266) as it is shown in Figure 2.20. 67 Surface modification of a transition alumina Figure 2.20 – XRD pattern of Nanotek after attrition milling. As a conclusion, both magnetic stirring and ball-milling were effective methods for powders deagglomeration; moreover, by ball-milling, dispersed suspensions can be prepared in a very limited time. On the contrary, attrition milling carried out in the previously reported conditions was less effective and also produced a significant pollution by the milling media into the ground material. As a consequence, in the following experimentation, this dispersion route was no more considered. 2.2.3 Effect of the dispersion on the specific surface area The samples A MS , A BMα and A BMz were submitted to N 2 adsorption/desorption isotherm and were performed on the samples outgassed at 150°C as in the case of A. 35 A AMS 30 3 -1 Volume (cm g ) 25 20 15 10 5 0 0,0 0,2 0,4 0,6 0,8 1,0 p/p0 Figure 2.21- N 2 adsorption/desorption isotherms at 77 K out-gassed at 150°C, comparison between A and A MS : full and white refer to adsorption and desorption. 68 Surface modification of a transition alumina The Figure 2.21 reports curves with similar shape for both samples, corresponding to Type II isotherms. The measured BET SSA value for A MS were 37.3 m2 g-1, showing a limited effect of magnetic stirring on the specific surface area (SSA of raw powder A was 34.5 m2g-1, as previously reported). The A MS exhibited a small hysteresis loop and may although indicate a small change of inter-particles porosity, probably as a consequence of the prolonged dispersion under magnetic stirring, which should also be responsible for the increase in surface area. These results showed the presence of inter-particle porosity may be excluded. For the ball-milled samples, it was not observed a significant SSA increase as respect to A (Figures 2.22), probably because of the shorter dispersion time in comparison with the A MS . The comparison (Figures 2.22) between A and the ball milled samples of N 2 adsorption/desorption isotherms at 77 K is shown. 35 A ABM 25 3 -1 Volume (cm g ) 30 20 15 10 5 0 0,0 0,2 0,4 0,6 0,8 1,0 0,6 0,8 1,0 p/p0 a) 35 A ABMz 25 3 -1 Volume (cm g ) 30 20 15 10 5 0 0,0 b) 0,2 0,4 p/p0 Figures 2.22- N 2 adsorption/desorption isotherms at 77 K out-gassed at 150°C, comparison between A with (a) A BMα and (b) A BMz : full and white refer to adsorption and desorption. 69 Surface modification of a transition alumina 2.2.4 Effect of dispersion on powder composition and evolution DTA analysis was performed on the three dispersed samples and their thermal behaviour compared with that of sample A.The dispersion process significantly affect the thermal behaviour of the raw powders, at both low (i) and high (ii) temperature regime, as shown in Figure 2.23. DTA signal [20 V] i) At low temperature regime, the dispersed samples show a endothermic signal at about 280°C, which was not present into the raw, A sample. If the TG curves are concerned, it can be observed that the above endothermic signals are accompanied by an abrupt mass loss of about 0.5-1%. The insert of Figure 2.23 shows, for instance, the TG curve of A MS in the 150°-500°C temperature range, in which a mass loss of 0.7% is recorded. 279°C 1325°C 1203°C 282°C 1210°C 286°C 1180°C A AMS ABMz ABM exo 200 400 600 800 1000 1200 1400 Temperature [°C] Figure 2.23- DTA curves of the samples. In the insert, A MS TG curve. Intensity (a.u.) In order to better understand such difference in thermal behavior between A and the dispersed materials, the XRD patterns of A and A MS were compared (Figure 2.24). 15 AMS 20 25 10 20 30 A 40 50 60 70 Diffraction Angle (2 Figure 2.24- XRD pattern of A and A MS . In the insert, XRD pattern of A MS in the 15-25° (2θ) range. As already reported, NanoTek® is a mixture of transition alumina phases, precisely -Al 2 O 3 and - Al 2 O 3 . If A and A MS are compared, diffraction patterns of - and -phases are seen in both spectra; the only differences among them are ascribable to peaks at 2 = 18.2° and 2 = 20.3° in 70 Surface modification of a transition alumina the A MS spectrum (see arrows in Figure 2.24). For a sake of clarity, inset to Figure 2.24 reports a detail of the 15° – 25° 2range, in which their spectra are superimposed: it is shown that the two extra-peaks actually appear after magnetic stirring. The two peaks are very close, in both position and relative intensity, to the (002) and (200) reflections of gibbsite, respectively (ICDD 76-1782). The two polymorphs of aluminium hydroxide Al(OH) 3 , gibbsite and bayerite, can be however discriminated by the diffraction peak at about 18°, occurring at 2 = 18.2° in gibbsite and 18.8° in bayerite. The formation of a crystalline gibbsite Al(OH) 3 phase may thus be inferred. Its formation was imputed to the prolonged magnetic stirring of A in water to produce A MS , as will be better explained in the following. The formation of γ-Al(OH) 3 during water dispersion of transition aluminas was already reported in literature13. This process is described by the following reaction: 1 Al 3 3 H 2O Al2O3 3H 2 2 (1) In this case, the hydrated Al3+ ion is only stable under acidic conditions, so its hydrolysis has to be considered: Al OH 3H Al 3 3H 2O (2) 3 Equations (1) and (2) can be summarized in the following one: 1 3 Al OH Al2O3 H 2O 3 2 2 (3) According to literature12,13, the pH of the alumina suspension plays a key role into the Al(OH) 3 formation. In fact, bayerite precipitation from -aluminas occurred only at pH higher than 4.5-5, as occurs in our investigation. This result is in good agreement with the publication of Pourbaix et al.14, which reports the solubility diagram of alumina and five polymorphs of aluminium hydrates - empirically obtained by knowing the free formation enthalpies - as a function of pH (Figure 2.25). The minimal solubility of many aluminium hydrates and hydroxides was found at pH 5.1, while their solubility significantly increased at both lower and higher pH. 71 Surface modification of a transition alumina Figure 2.25- Influence of pH on the solubility of Al 2 O 3 and the aluminium hydrates at 25°C14. In fact, in samples dispersed for 170 h at acidic conditions (namely at pH 4 and 3) no gibbsite peaks were detected, as it is shown in Figure 2.26. Intensity (a.u.) AMS pH=4 10 20 30 40 50 AMS pH=3 60 70 Diffraction Angle (2.) Figure 2.26- XRD patterns of A MS dispersed at pH 3 and 4 for 170 h On the ground of the above considerations, the endothermic signals of the DTA curves collected in Figure 2.23 can be reasonably associated to the gibbsite thermal decomposition. According to literature12, the dehydration of Al(OH) 3 proceeds with the intermediate formation of boehmite (AlOOH), according to the following reactions: Al (OH )3 AlOOH H 2O AlOOH 1 1 Al2O3 H 2O 2 2 (4) In fact, X. Carrier et al.12 carried out TGA/DTA curves on a transition alumina sample suspended for 168 h at pH 11, under continuous stirring. They found a rather sharp endothermic DTA peak at 250 ◦C concomitant with a weight loss; a broad hump centred at about 550°C was also seen by the Authors on the DTA curve, accompanying a progressive weight loss. It was associated to the dehydration of boehmite (AlOOH) formed from the dehydration of Al(OH) 3 . They same Authors strengthened their founding with TEM analysis, which allowed the observation of large particles (about 200 nm long) besides the small γ–alumina nanometric ones. Electron diffraction patterns allowed the assignation of the gibbsite structure to the large grains, according to their XRD data. In our experimentation, any further DTA endothermic signals, ascribable to the boehmite dehydration, was observed, probably due to the low fraction of gibbsite formed upon stirring. In order to confirm such hypothesis, equation (4) was exploited to determine the gibbsite amount; it was calculated from the weight of water removed at 280°C and normalized to the mass of the sample after completing the dehydration. It was estimated that about 2 wt.% of gibbsite was formed upon hydration of transition alumina. 72 Surface modification of a transition alumina Intensity (a.u) It is interesting to note that magnetic stirring actually plays a role in the formation of the hydroxide phase, since a blank sample prepared by suspending the powder in water for 170 h without stirring did not show any change in the XRD patterns (see Figure 2.27) 10 20 30 40 50 60 70 Diffraction Angle (2.) Figure 2.27: XRD pattern of A powder suspended for 170 h in distilled water, without stirring Furthermore, XRD measurements on powder stirred for increasing times showed that the two peaks appear after 36 h and then their intensities remain almost constant up to 170 h. In the following Figure 2.28, XRD patterns evolution of A MS as a function of the stirring time is reported. Figure 2.28- XRD pattern evolution of A MS as a function of the stirring time Finally, HRTEM was conducted on A and A MS samples (Figure 2.29 for A sample). Both powders are composed by spherical particles of heterogeneous diameters, in the 5 – 100 nm range, forming agglomerates with variable size, while “single particles” were rarely observed in both materials. In both cases, it can be observed that particles are covered by a poorly ordered layer of about 2.5 nm thickness (see insert in Figure 2.29), contrasting with the well crystallizer inner part in which the crystalline lattice fringes could be easily observed. The presence of such 73 Surface modification of a transition alumina amorphous layer, most probably induced by the physical vapour synthesis15 of the nanostructured powders, seems to cover the majority of particles and will be directly in contact with water during dispersion under magnetic stirring and therefore affecting the surface properties of the final material. In addition, From HRTEM analysis it was not possible to detect any gibbsite Al(OH) 3 particles, unlike other works12, reporting the formation of well-shaped and large gibbsite crystals separated from the alumina surface after suspension in water. Figure 2.29- Selected HRTEM picture of sample A; the insert reports a micrograph, taken at a higher magnification, showing the presence of an outer amorphous layer. ii) Figure 2.23 shows a different thermal behaviour also in the high temperature range. It is evident, in fact, the role of the dispersion in lowering the crystallization temperature of the αalumina phase. The sharp exothermal signals were, in fact, associated to the to transformation temperature and they were detected at significantly lower temperatures for the dispersed materials as compared to sample A. Precisely, such transformation signal was detected at 1325°, 1203°, 1180° and 1210°C for A, A MS , A BMα and A BMz , respectively. In addition, if the transformation temperatures of the three dispersed powders are compared, a further difference can be observed. The lowest crystallization temperature was determined for A BMα , the highest to A BMz . As already observed in literature10, such difference can be imputed to a seeding effect due to the milling media, even if undetectable by XRD analysis (see Figure 2.18). In fact, alpha-alumina seeds are able to promote the above transformation11, while zirconia ones have the opposite effect, and delay the α-phase crystallization16. This aspect will be deepened in the following, since a systematic investigation of the phases evolution, as a function of the calcinations temperature, in the as-received and dispersed samples was performed. 2.2.5 Influence of the dispersion route on the phase evolution Sample A, as already reported, is composed by δ-Al 2 O 3 and γ-Al 2 O 3 phases. After calcination at 1100°C for 0.5 h, some traces of θ-Al 2 O 3 (ICDD file n°11-0517) appear. Subsequently, raising the calcination temperature up to 1150°C for 0.5 h, θ-Al 2 O 3 tends to increase its relative abundance. Finally, by increasing the calcination temperature up to 1200°C for 0.5 h it is possible to denote 74 Surface modification of a transition alumina Intensity (a.u) the first traces of α-Al 2 O 3 (ICDD file n°10-0173). In the Figure 2.30 it is shown the thermal evolution of sample A as a function of the calcination temperature. 1200°C 1150°C 10 20 30 40 1100°C 50 60 non-treated 70 Diffraction Angle (2. Figure 2.30– Thermal evolution of sample A The dispersed samples were then submitted to the same calcination temperatures. In the case of A MS (Figure 2.31), the powder exhibited a different thermal evolution if compared with A. By calcining at 1100°C for 0.5 h, the powder is composed by a mixture of δ-Al 2 O 3 , γ-Al 2 O 3 and θAl 2 O 3 phases. At 1150°C, a significant amount of the alpha-phase was detected in the material near θ-Al 2 O 3 ; finally, at 1200°C the powder is mainly composed by well-crystallized α-Al 2 O 3 phase. Intensity (a.u) 1200°C 10 20 30 40 50 1150°C 60 1100°C As-received 70 Diffraction Angle (2. Figure 2.31 – Thermal evolution of A MS The XRD pattern of sample A BMα (Figure 2.32) calcined at 1100°C presents the three phases: δ-Al 2 O 3 , γ-Al 2 O 3 and θ-Al 2 O 3 . By rising the calcinations temperature up to 1150°C the transformation into α-phase is completed. 75 Surface modification of a transition alumina Intensity (a.u) 10 1200°C 1150°C 1100°C As-received 20 30 40 50 60 70 Diffraction Angle (2. Figure 2.32 – Thermal evolution of A BMα . Intensity (a.u.) A BMz (Figure 2.33) calcined at 1100°C is prevalently composed by δ and θ phases. At 1150°C the powder is a mixture of θ and α phases and finally at 1200°C only α-phase appears. 10 20 30 1200°C 1150°C 1100°C 40 50 non-treated 60 70 Diffraction Angle (2.) Figure 2.33 – Thermal evolution of A BMz . For a better comparison between samples, Figure 2.34 collects the XRD patterns of the four materials after calcining at 1150°C for 0.5 h. As it is shown in the figure, the dispersion route has been effective in modifying the high-temperature crystallization path of the Nanotek powder. After calcination at 1150°C for 0.5 h (Figure 2.34), A was composed only by transition phases (namely δ and θ-phases); in A MS , α-Al 2 O 3 was the prevalent phase near θ-Al 2 O 3 ; a pure, well-crystallized α-Al 2 O 3 was detected in A BMα and finally, A BMz was a mixture of transition and α-Al 2 O 3 phases. These data state the role of dispersion in lowering the transformation temperature, according to DTA curves (see Figure 2.23). 76 Intensity (a.u.) Surface modification of a transition alumina 20 30 40 10 ABMz 50 ABM 60 AMS A 70 Diffraction Angle (2) Figure 2.34 - XRD pattern of the samples calcined at 1150°C for 0.5 h. In addition, if the XRD patterns of the three dispersed samples are compared, a further difference can be ascribed. If we consider the un-polluted A MS sample as a reference, -alumina phase is contained in a higher and lower fraction in A BM and A BMz , respectively. So, the anticipation of the temperature transformation occurring in A BMα is an probably related to some seeding (not visible in the XRD pattern) introduced during the milling process which promotes the α-alumina crystallization. On the contrary, a delay in crystallization in A BMz compared with A MS and A BMα is reasonably attributable to zirconia impurities introduced during milling, in agreement with literature16. 2.2.6 Effect of the dispersion on the kinetics of transformation The kinetics of to -alumina phase transformation of A and of the dispersed samples was studied by employing the Kissinger’s method. The Kissinger’s plots are collected in Figure 2.35, while the related activation energy are reported in Table I. The obtained data are discussed in the following, by separately considering the transformation temperature (i) and the related activation energy (ii). i) The comparison among the Kissinger’s plots allows to evidence again the significant role of the dispersion process in lowering the transformation temperature. In addition, if the unpolluted stirred sample, A MS , is kept as a reference, the lowering of the transformation temperature in A BM was more relevant as compared to the delaying effect presented by A BMz . Such behaviour was, once again, imputed to a different effect of alumina and zirconia seeds form the milling media which affect in a different way the transformation process. 77 Surface modification of a transition alumina A10ABMzAMS ABM -ln(dT/dt / Tm2) A A3 0.2 0,00060 0,00065 0,00070 1/Tm Figure 2.35- Determination of the activation energy for transition to α-Al 2 O 3 phase transformation. In insert are the calculated activation energies are reported. Table I - Activation energy (E a ) for θ to α-Al 2 O 3 phase transformation (kJ/mol) Sample A A MS A BMα A BMz A α3 A α10 E a (kJ/mol) 486 498 480 496 502 438 R2 0.9959 0.9946 0.9969 0.9993 0.9892 0.9990 So, in order to discriminate the role of dispersion form that of seeding in affecting the transformation temperature, two new materials were prepared. Samples A α3 and A α10 were produced by flash plunging the sample A into in a tubular furnace kept at 1290°C for 3 and 10 minutes, respectively. The aim of this thermal treatment was to induce the crystallization of certain amounts of α-Al 2 O 3 phase in the powders, but trying to minimize the crystallite growth during the heat treatment. In Figure 2.36, XRD patterns of the above materials are presented: in A α3 the alpha phase was almost undetectable (as in the case of A BM ) while an appreciable Al 2 O 3 XRD peak was indeed observed in A α10 . Figure 2.36- XRD patterns of A α3 and A α10 in the 38-50° range. 78 Surface modification of a transition alumina The comparison among the Kissinger plots of sample A with those of A BM, A 3 and A 10 shows that also seeding lowers the transformation temperature: the higher the -seeds amount, the lower the transformation temperature. So, the more pronounced anticipation occurring in A BM should be reasonably imputed to a synergic effect induced by coupling dispersion and seeding with -phase. On the ground of this hypothesis, also the “delaying” effect of A BMz should be imputed to the seeding effect induced by zirconia milling media, which acts in a negative way on the transformation temperature, as expected on the ground of literature16. ii) Concerning activation energy, also in this case the role of dispersion and seeding should be separately discussed. The activation energy of A was 486 kJ/mol, in a good agreement with literature data17, but it was almost unaffected by the dispersion routes. This result is only in partial agreement with literature data, in which a lower to -Al 2 O 3 transformation temperature due to an effective dispersion process was also reflected into a lower related activation energy17-18. On the contrary, seeding is able to lower the above energy, as shown by A 10 , for which a decrease of about 10%, as compared to A, was determined. However, a similar decrease was not presented by A 3 and A BM , thus suggesting that detectable amounts of -seeds are required for affecting the activation energy of the transformation. 2.2.7 Influence of the dispersion route on forming and sintering As in the case of A, the dispersed samples were submitted to dilatometric analysis performed on uniaxially pressed bars (at 350 MPa) by heating up to 1500°C (heating rate of 10°C/min) and a soaking time of 3 h at maximum temperature. Firstly, the effect of dispersion of the compactability, under dry pressing, of the powders was investigated. In Table II, their green density, calculated from weight and geometrical measurement of the pressed bars is reported. A slight increment of the green density from the raw powder to the dispersed ones was observed. In addition, A BMz , presented the higher value: if the density of zirconia is concerned (reference value for TSZ is 6.09 g/cm3), such datum can be explained on the ground of a possible seeding from the milling media. Table II- Green density of differently sintered samples Sample Green Density (g/cm3) Green Density (%TD 1 ) A 1.85 46.9 A MS 1.89 47.8 A BMα 1.93 48.9 A BMz 2.04 51.7 A α10 1.91 48.2 As evidenced by the following Figures 2.37, all the dispersed materials present a two-step sintering behaviour, as in the case of sample A. In spite of this, the samples differ into onset sintering temperature and θ→α transformation temperature, (evidenced by the derivative curve: see dotted lines in the following graphics). In order to better evidence their different behaviour, Table III collects the onset sintering (T onset , °C) and the δα transformation (T δα , °C) temperatures. In addition, the total linear shrinkage (ΔL/L 0 ) tot (%), as well as shrinkage percentages recovered during the heating step (ΔL/L 0 ) heating , 1 Referred to the α-Al 2 O 3 theoretical density 3.96 g/cm3 79 Surface modification of a transition alumina shared into two contributions, one associated to Zone I (ΔL/L 0 zone I ), the other related to Zone II (ΔL/L 0 zone II ), and finally the shrinkage due to the isothermal step (ΔL/L 0 ) isothermal , are collected. In some cases, the derivative curve allows to detect a second significant sintering temperature, i.e. the temperature of maximum sintering rate of the α-phase (T α max ). If detectable, this value is again collected in Table IV. 0,0000 0,00 -0,02 -0,02 -0,0002 1420°C -0,0002 -0,04 1420°C -0,0004 -0,0008 -0,12 1135°C -0,0010 -0,14 -0,10 -0,18 -0,0008 -0,12 -0,0010 -0,14 1121°C -0,0012 -0,16 -0,0006 -0,08 o o L/L -0,10 L/L -0,0006 -0,08 -0,0004 -0,06 derivative signal -0,06 derivative signal -0,04 0,0000 0,00 -0,0012 -0,16 -0,18 -0,0014 -0,0014 -0,20 -0,20 200 400 600 a) 800 1000 1200 1400 200 1600 400 600 b) Temperature [°C] 0,0000 0,00 -0,02 800 1000 1200 1400 1600 Temperature [°C] 0,0000 0,00 -0,02 -0,0002 -0,0002 -0,04 -0,0004 -0,0008 -0,12 -0,0010 1145°C -0,14 -0,18 -0,10 -0,0008 1145°C -0,12 -0,0010 -0,14 -0,0012 -0,16 -0,0012 -0,16 o o L/L -0,10 -0,0006 -0,08 L/L -0,0006 -0,08 -0,0004 -0,06 derivative signal -0,06 -0,18 -0,0014 derivative signal -0,04 -0,0014 -0,20 -0,20 200 400 600 800 1000 1200 1400 200 1600 400 600 800 1000 1200 1400 1600 Temperature [°C] Temperature [°C] d) c) Figures 2.37 – Dilatometric (solid line) and derivative (dashed line) curves: (a) A MS , (b) A BMα , (c) A BMz and (d) A α10 . Table III- Total linear shrinkage (ΔL/L 0 ) tot , shrinkage percentages recovered during the heating step (ΔL/L 0 ) heating , associated to Region I (ΔL/L 0 Zone I ), and to Region II (ΔL/L 0 Zone II ), shrinkage during isothermal step (ΔL/L 0 ) isothermal , onset temperature (T onset ), temperature of δα transformation (T δα , °C) Sample ΔL/L 0total (%) ΔL/L 0heating (%) A A MS A BMα A BMz A α10 14.6 17.7 19.4 19.2 17.1 10.9 13.9 17.4 14.4 12.4 ΔL/L 0 ΔL/L 0 ΔL/L 0 Zone I Zone II isothermal (%) 6.5 7.1 7.3 6.5 6.0 (%) 3.1 5.9 9.2 7.0 5.8 (%) 3.0 3.8 2.0 4.8 4.7 T onset (°C) T δα (°C) T α max (°C) 1021 996 996 1030 996 1147 1135 1121 1145 1145 Not detected 1420 1420 Not detected Not detected As it is shown in Table III, the dispersion slightly lowered the T δα , as seen in samples A MS and A BMα . Once again, the lowest transformation temperature detected in A BM was imputed to synergic effects of dispersion and seeding. The effect of the dispersion on T δα in sample A BMz is not observable as seeding introduced by the zirconia medium countered this phenomenon16. 80 Surface modification of a transition alumina Samples present almost the same onset temperature: however, slightly higher values were presented by samples A and A BMz . Similar shrinkages were found in samples during the first sintering step, specially in the case of A BMα and A MS (≈7%). During the second step, all the samples shrank considerably more than sample A which exhibited a total linear shrinkage of about 3 %. However, among them A BMα achieved a larger linear shrinkage during the second step. As a consequence, its densification during the isothermal step was limited - if compared with the other samples (≈2 %). This fact is in good agreement with the position of the second peak on the derivative curve, which represents the maximum densification rate. This peak was only recorded at 1420°C in samples A BMα and A MS . In sample A BMz it was observed a delay in densification during the beginning of the second step, as reported in literature19. Nevertheless, it recovered before isothermal step. During the dwell period, samples A BMz and A α10 presented similar linear shrinkage of about 4.7 %. Conversely, sample A MS in the last step recorded 3.8%. As a conclusion, the sintering behaviour of A BMα shows the combination of the single effects of A MS and A α10 , i.e. dispersion and seeding. Sample A presented the lowest total shrinkage value: this datum coupled with its lowest green density, allowed to explain the lowest fired density (75.3%) of such sample (values are collected in Table IV). A significant increment (of 10-15%) of the final density was achieved by the dispersed materials. Moreover, a similar increment was reached by the un-dispersed, seeded A α10 sample. This datum was of interest, since it allows to explain the highest fired density of A BMα as compared to all the other samples, since it combines the positive effects of both dispersion and seeding. Table IV- Fired density of differently sintered samples Sample Fired Density (g/cm3) Fired Density (%TD1) A 2.98 75.3 A MS 3.25 82.1 A BMα 3.42 86.4 A BMz 3.39 85.6 A 10 3.36 85.8 On the ground of such results, it can be reasonably assumed that a further increment in the final density can be achieved by dispersing A α10 sample. Some attempts were performed but even after several hours of ball-milling (3 h), the powders were still agglomerated. As a future activity, optimized processed powders should be produced by coupling the flash heating step to an effective dispersion process, thus to induce both effects. 2.2.8 Investigation on Sintering Kinetics of Nanotek powders Stepwise isothermal dilatometry (SID) was applied for investigating the sintering kinetic of nanopowders. Shrinkage curves recorded for each isothermal step were fitted polynomially and the shrinkage rate was calculated from shrinkage value. The method is accurately described in chapter one. 81 Surface modification of a transition alumina Three samples were chosen for this study, namely samples A, A MS and A BMα . The selection of these samples was based on the transformation kinetic data, putting emphasis on A BMα . For this aim, dilatometric tests were performed employing bars produced by cold uniaxial pressing at 350 MPa. In this study, the 1100°C-1400°C was chosen as temperature range of study. This first material submitted to the study was A sample. The apparent activation energy (data collected in Table V) for A sample was calculated fitting the values with a linear slope obtaining a value of 690 kJ/mol as it is illustrated in the following Figure 2.38. Q=689.9 kJ/mol 2 R =0.9943 -2 -4 -6 -ln(nK) -8 -10 -12 -14 -16 -18 -4 5,7x10 -4 6,0x10 -4 6,3x10 -4 6,6x10 -4 6,9x10 -4 7,2x10 -4 7,5x10 -1 1/T (K ) Figure 2.38 – Arrhenius constant (K(T)) versus reciprocal temperature 1/T of A sample. Table V - Activation energy of the different samples. SAMPLE E a (kJ/mol) R2 n A A MS A BMα 690 606 450 0.9943 0.9342 0.9899 0.15 0.10 0.18 The second sample analyzed was the A MS . By applying the same procedure, the apparent activation energy was 606 kJ/mol, which is in good agreement with the data obtained regarding the kinetics transformation data20-21. Similarly, the ball-milled sample named A BMα was submitted to the same analysis. In this particular case, as it was expected, A BMα exhibited a lower activation energy (514.5 kJ/mol) referable to the some seeding introduced by the milling media. As reported in literature21, a sharp change of parameter n (from 0.24-0.07 in case of A) in the temperature range (1100-1200°C) was observed in all samples correlated with phase transformations. In literature it was not found relevant information about this method applied on transition aluminas. For this reason many studies attempted to understand the activation energies of single phase constituents. For instance, Wang et al.20 applied this technique to evaluate the kinetics parameters of a macroporous α-Al 2 O 3 . In this particular case, values were calculated in the range 1200-1400°C obtaining an apparent activation energy of 415 kJ/mol and an average exponent n equal to 0.232. 82 Surface modification of a transition alumina A second study made by Holkova et al.21 attempted to study and establish a comparison between the stepwise isothermal dilatometry. For boehmite and α-Al 2 O 3 evaluated by SID, the authors found apparent activation energies of 950.1 kJ/mol and 541.1 kJ/mol, respectively. A similar behaviour was found by the same authors21, by adding boehmite into α-Al 2 O 3 which increased the apparent activation energy to 597.7 kJ/mol. In this context, the effect of introducing some α-Al 2 O 3 milling contamination into the powder may be the principal responsible of reducing the apparent activation energy in sample A BMα . 2.2.9 Influence of the dispersion route on the final microstructure SEM observations were performed in SE mode on the fracture surface of the fired bodies obtained by uniaxial pressing. The corresponding micrographs are reported in Figures 2.39. The SEM analysis revealed a different microstructural development. Sample A (Figure 2.39 a), presented a vermicular microstructure accompanied by inhomogeneous morphologies. It is made of fine grained alumina particles of about ≈0.9 μm entrapping diffuse residual porosity. The role of the dispersion route on the microstructural evolution was evidenced by comparing A MS and the ball-milled samples. Sample A MS (Figure 2.39 b) is less porous compared with A, in spite of its better densification, the ultrafine primary particle size was retained, yielding round-shaped grains of ≈0.77 μm, entrapping mostly intragranular porosity. The microstructure of A BMα (Figure 2.39 c) was highly dense, consisting of well-facetted alumina grains of about 2 μm with a limited residual porosity mostly located, as in the previous case, in intragranular position. A lower porosity was observed in A BMz (Figure 2.39 d), characterized, as reported in literature, by a finer microstructure19. The average size of alumina grain was 1.5 μm. The fracture mode was radically different: in the case of A it is mostly transgranular, while in A MS it is prevalently intergranular; finally A BMα and A BMz presented both inter- and transgranular modes. a) b) 83 Surface modification of a transition alumina c) d) Figures 2.39 – SEM micrographs of the sintered materials: (a) A, (b) A MS , (c) A BMα and (d) A BMz . (SEM observation performed on the fracture surface) 2.3 Study of the effect of powder dispersion on its surface properties by means of IR spectroscopy 2.3.1 IR spectra of samples outgassed at increasing temperatures The species at the surface of A and A MS powders were studied by means of FT-IR spectroscopy. Samples were outgassed at increasing temperatures, namely 150, 350 and 500°C, since the presence of surface gibbsite should give rise not only to peculiar IR bands in the hydroxyls range (3900-3000 cm-1), but also to different adsorbed species and different behaviour towards thermal treatments. In Figures 2.40, FTIR spectra of A outgassed at 150, 350 and 500°C are reported the spectrum of the sample outgassed at 150°C shows abroad absorption band in the stretch range 3900-3000 cm-1 range. Absorbance 4 a 3 150°C 2 350°C 1 500°C 0 3600 3200 2800 2400 2000 1600 1200 Wavenumbers (cm-1) 84 Surface modification of a transition alumina 3675 b 3725 Absorbance 3600 3775 3790 350°C 3680 500°C 3800 3600 3400 3200 3000 -1 Wavenumbers (cm ) Figures 2.40- FT-IR spectra recorded on sample A: (a) outgassed at 150°C, 350°C and 500°C and (b) detail of the hydroxyls spectra outgassed at 350 and 500°C. As it was reviewed in the first chapter, this is typical for a highly hydrated surface caused by the atmospheric moisture. Bands in the 1650-1200 cm-1 range are related to several carbonate-like species, usually observed in transition aluminas23,24 and definitely removed by outgassing at 500°C. In Figure 2.40 b, a detail of hydroxyls spectra of sample A outgassed at 350 and 500°C is reported, showing the presence of different OH groups. According to the model proposed by Knözinger and Ratnasamy, Bands at 3790 and 3775 cm-1 are assigned to I b and I a hydroxyls, i.e. free terminal hydroxyls bonded to octahedral (AlVI) and tetrahedral (AlIV) aluminium ions. The 3725 cm-1 band is assigned to di-bridged free OH group (type II a ) and 3675 cm-1 band to hydroxyls (type III a ), which are tri-bridged among two octahedral (AlVI) and one tetrahedral (AlIV) aluminium ions; band at ≈3600 cm-1, with a tail on the lower wavenumbers side, is assigned H-bonded hydroxyls, which should be eliminated after outgassing at 500°C. Figures 2.41 reports IR spectra of A MS outgassed at 150, 350 and 500°C. Carbonate-like species bands at 1650-1200 cm-1 are stable to thermal treatment at 500°C, indicating that after magnetic stirring the surface present stronger basic sites to which CO 2 may coordinate. 85 Surface modification of a transition alumina Absorbance 4 3 2 150°C 350°C 1 500°C 0 3600 3200 2800 2400 2000 1600 1200 -1 Wavenumbers (cm ) Figure 2.41- FT-IR spectra recorded on sample A MS outgassed at 150°C, 350°C and 500°C. The difference of the magnetic stirring on the surface properties is shown Figures 2.42, in which hydroxyls spectra of two samples outgassed at 150°C (a), 350°C (b) and 500°C (c) are compared. A AMS a Absorbance 3.0 3725 2.5 2.0 3780 1.5 3900 3600 3300 3000 -1 Wavenumbers (cm ) 86 Surface modification of a transition alumina 2.5 3675 b 3725 A AMS 3580 Absorbance 3775 2.0 3790 1.5 3900 3600 3300 3000 -1 Wavenumbers (cm ) 2.5 c A AMS 3725 Absorbance 3775 3680 2.0 3790 1.5 3800 3600 3400 3200 3000 -1 Wavenumbers (cm ) Figures 2.42- Normalized FT-IR spectra recorded on samples A and A MS outgassed at (a) 150°C, (b) 350°C and (c) 500°C. 87 Surface modification of a transition alumina Spectra corresponding to the samples outgassed at 150°C (Figure 2.42 a) are dominated by the broad absorption of H-bonded hydroxyls (below 3600 cm-1), whereas at higher wavenumbers, bands are seen at 3780 and 3725 cm-1 due to free AlVI-OH and AlIV-OH (I a ). In sample A MS an additional band appears at 3330 cm-1 which cannot be imputed to residual water molecules are removed at room temperature. This band is reasonably attributed to OH stretch mode of gibbsite hydroxyls. Gibbsite structural OH groups have been carefully studied by means single-crystal Raman and FTIR methods24, which allow to single our several distinct types of structural OH groups, basically inter-layer and intra-layer hydrogen bonded hydroxyls. With single crystals, authors were able to single out six νOH in both IR and Raman spectra. In this case, only the presence of a new broad absorption was observed at 3330 cm-1. The observed frequency was in this case 3376 cm-1, due to OH-species interacting by H bonding. The stability of OH species was studied by increasing the outgassing temperature assuming the decomposition of Gibbsite below 300°C: after outgassing at 350°C, hydroxyls bands decrease in intensity, due to surface dehydroxylation and the main difference between the two samples is the higher intensity of the band of H-bonded hydroxyls with sample A MS at 3590 cm-1 indicating that more hydroxylated surface formed during magnetic stirring. Figure 2.42 (c) reports hydroxyls spectra of samples outgassed 500°C, the typical surface features of de-hydroxylated transition alumina are expected, with bands at 3790 cm-1 (type I b hydroxyls), 3775 cm-1 (type I a hydroxyls), 3725 cm-1 (type II a hydroxyls) and 3680 cm-1 (type III a hydroxyls), whereas 3580 cm-1 band (H-bonded hydroxyls) is removed. It was confirmed that surface physico-chemical properties were modified by the dispersion route, the main difference being the OH species population due to the presence of Gibbsite on sample A MS . 2.3.2 Adsorption of CO at nominal 77 K Carbon monoxide is widely used as probe molecule to study both Lewis and Brønsted acidic sites at the surface of oxides and zeolites24,27-31. When electrostatic interaction takes place between CO and the adsorbing site, like in this case, a hypsochromic shift occurs, with respect to free CO molecule (2143 cm-1), and characteristic bands are seen in the C≡O stretch region (2250-2050 cm-1)27. Being the interaction very weak, low temperatures are needed and experiments are performed at the nominal temperature of liquid nitrogen. Increasing pressures of CO (in the 0.05–30 mbar range) were dosed, at the nominal temperature of N 2(l) , on the two samples outgassed at 150, 350 and 500°C: Figures 2.43 and 2.44 report normalized difference spectra, obtained by subtraction of bare samples spectra reported in Figures 2.42. CO dosage on sample A outgassed at 150°C (Figure 2.43 a) gives rise to the formation of i) a main band at 2152 cm-1; ii) a weaker band in the 2189 – 2178 cm-1 range and iii) a minor absorption at about 2107 cm-1. The 2152 cm-1 band is due to CO molecules interacting via H-bonding with AlIV-OH species originally absorbing at 3725 cm-1 (in Figure 2.42 b), whereas 3780 cm-1 hydroxyls are very weak acid and do not interact with CO. The weak absorption at 2107 cm-1 is probably related to that at 2152 cm-1 and is assigned to CO molecules adsorbed through the O atom (CO---HO adducts), according to previous work32. The band at 2189 cm-1, shifting to 2178 cm-1 with coverage, is assigned to CO molecules interacting with weak Lewis acidic sites, like 5-coordinate Al3+ or, most probably, coordinatively unsaturated tetrahedral Al3+ of low index crystal planes33-34. 88 Surface modification of a transition alumina Figure 2.43 b reports difference spectra recorded after CO dosage on sample A outgassed at 350°C: two bands are seen at 2197 cm-1, shifting with coverage to 2183 cm-1, and at 2160 cm-1, shifting to 2152 cm-1. The former is assigned to CO adsorbed on Al3+ sites forming an extended phase, the latter to CO molecules H-bonded to hydroxyls with different acidity. With respect to sample A pre-treated at 150°C, the band of CO on Al3+ sites appears more intense and shifted to higher wavenumbers: this is ascribed to surface de-hydroxylation with formation of new stronger Lewis sites (coordinatively unsaturated Al3+ ions). The band of CO interacting with OH species is seen to shift with coverage from 2160 to 2152 cm-1 due to the presence of several hydroxyls with different acidity, as shown in Figure 2.42 b. With sample A outgassed at 500°C (Figure 2.43 c), bands are seen of CO adsorbed on an Al3+ sites forming an extended phase (band at 2198 cm-1 shifting with coverage to 2183 cm-1) and on residual hydroxyls (band at 2160 cm-1 shifting with coverage to 2156 cm-1). The relative intensities of bands due to CO adducts with Al3+ ions and OH species changed with the hydration degree of the surface, since new coordinatively unsaturated sites become available at the expenses of hydroxyls removed at higher temperature. a OH IOH /IAl3+ = 11.5 Absorbance 1,5 1,0 0,5 Al 3+ 0,0 2250 2200 2150 2100 2050 -1 Wavenumbers (cm ) 89 Surface modification of a transition alumina b I /I = 1.6 OH Al3+ Absorbance 1,5 1,0 0,5 0,0 2250 2200 2150 2100 2050 -1 Wavenumbers (cm ) c IOH/IAl3+ = 1.1 Absorbance 1,5 1,0 0,5 0,0 2250 2200 2150 2100 2050 -1 Wavenumbers (cm ) Figures 2.43- FT-IR difference spectra, in the CO stretch region 2250 – 2050 cm-1 recorded after dosing CO on sample A outgassed at 150°C (a); 350°C (b) and 500°C (c). CO equilibrium pressures range: 0.5 - 20 mbar; 90 Surface modification of a transition alumina Figures 2.44 report corresponding spectra recorded in same conditions, i.e. under the same equilibrium CO pressures, on sample A MS pre-treated in the same way. As a whole, the same surface species were observed, i.e. coordinatively unsaturated Al3+ ions and surface hydroxyls, but with the following relevant differences: with A MS outgassed at 150°C (Figure 2.44 a), at low coverage the band of CO H-bonded to hydroxyls is seen at 2154 cm-1, whereas at higher CO equilibrium pressures another component is seen at 2149 cm-1. The difference with respect to the corresponding band on sample A outgassed at the same temperature is better shown in Figure 2.44 b, reporting spectra recorded on samples A and A MS under the same CO pressure (10 mbar). The smaller shift with respect to the free molecule mode (2143 cm-1) indicates that the component at 2149 cm-1 should be related to CO interacting with weaker acidic hydroxyls, like those originally absorbing at 3330 cm-1 (Figure 2.42 a), stemming from the hydroxide phase (gibbsite); normalized spectra recorded under the same CO equilibrium pressures allow to draw some semi-quantitative observation: with A MS , intensities of CO bands are always smaller than with A, indicating a smaller amount of sites actually accessible at the surface. The ratio I OH /I Al3+, reported for each experiment, between intensities of the bands due to CO---HO and CO---Al3+ adducts, respectively, may be used to evaluate the relative abundance of Lewis and Brønsted sites. After treatment at 150°C, I OH /I Al3+ is 11.5 and 10.1 for A and A MS , respectively: this can be explained by the fact that 3330 cm-1 hydroxyls belonging to gibbsite, though abundant, are less acidic and less prone to interact with CO; after treatment at 350°C, I OH /I Al3+ is 1.6 and 2.0 for A and A MS , respectively, since A MS surface is more hydrated, due to prolonged stirring in water. After outgassing at 500°C, I OH /I Al3+ is the same for both materials, due to the formation of the same. IOH/IAl3+ = 10.1 a Absorbance 1,5 1,0 0,5 0,0 2250 2200 2150 2100 2050 -1 Wavenumbers (cm ) 91 Surface modification of a transition alumina 1,5 A A MS b 2149 Absorbance 1,0 0,5 0,0 2200 2150 2100 -1 W avenumbers (cm ) c I /I = 2.0 OH Al3+ Absorbance 1,5 1,0 0,5 0,0 2250 2200 2150 2100 2050 -1 Wavenumbers (cm ) 92 Surface modification of a transition alumina d IOH/IAl3+ = 1.0 Absorbance 1,5 1,0 0,5 0,0 2250 2200 2150 2100 2050 -1 Wavenumbers (cm ) Figures 2.44- FT-IR difference spectra, in the CO stretch region 2250 – 2050 cm-1, recorded after dosing CO (equilibrium pressures in the 0.5 - 20 mbar range) on sample A MS out-gassed at 150°C (a); 350°C (c) and 500°C (d). CO equilibrium pressures range: 0.5 - 20 mbar; difference spectra obtained by subtracting spectra of bare sample reported in Figure 2.42. Section b: comparison of difference spectra recorded at the same CO equilibrium pressure (10 mbar) on samples A and A MS outgassed at 150°C. 2.3.3 Adsorption of CO 2 CO 2 interaction with the surface uncoordinated cations (Lewis acid sites) occurs by σ-change release from one of O lone pair orbitals, so the molecule loses the center of symmetry preserving its linear shape. By increasing the outgassing temperature, the band due to CO 2 adsorbed on Al3+ sites increases in intensity, whereas that due to CO 2 interacting with OH groups decreases. This result is due to progressive surface dehydroxylation and is in agreement with CO adsorption spectra. CO 2 was dosed by increasing the pressure (in the 0.05-40 mbar range), on both samples outgassed at 150, 350 and 500°C. Figure 2.45 reports the comparison between A and A MS outgassed at 150°C after dosing ≈20 mbar CO 2 . Only one band is seen at 2344 cm-1, readily assigned to CO 2 molecules interacting with surface OH groups. 93 Surface modification of a transition alumina 2344 Absorbance 1,5 A AMS 1,0 0,5 0,0 2400 2380 2360 2340 2320 2300 2280 -1 Wavenumber [cm ] Figure 2.45– Comparison of difference spectra recorded at the same CO 2 equilibrium pressure (20 mbar) on samples A and A MS outgassed at 150°C. In Figures 2.46 reports difference spectra recorded after CO 2 dosage on sample A degassed at 350°C. Two bands are seen at 2360 cm-1 and 2345 cm-1. The former is assigned to CO 2 molecules interacting with weak Lewis acidic sites like coordinatively unsaturated tetrahedral Al3+ of low index planes, the latter is assigned to CO 2 molecules interacting via H-bonding with OH species 35-39. 1,2 ab OH Absorbance 1,0 0,8 0,6 3+ Al 0,4 0,2 0,0 2400 2380 2360 2340 2320 2300 2280 -1 Wavenumber (cm ) 94 Surface modification of a transition alumina 0,6 bb OH Absorbance 0,5 3+ Al 0,4 0,3 0,2 0,1 0,0 2400 2380 2360 2340 2320 2300 2280 -1 Wavenumber (cm ) Figures 2.46- FT-IR difference spectra, in the stretch region 2400-2280 cm-1 recorded after dosing CO 2 in sample at A outgassed at: (a) 350°C and (b) 500°C. Figures 2.47 reports spectra recorded after dosing the same CO 2 equilibrium pressures on A MS . 2,0 1,8 ab OH Absorbance 1,6 1,4 1,2 1,0 0,8 3+ Al 0,6 0,4 0,2 0,0 2400 2380 2360 2340 2320 2300 2280 -1 Wavenumber (cm ) 95 Surface modification of a transition alumina 0,25 bb Al Absorbance 0,20 3+ OH 0,15 0,10 0,05 0,00 2400 2380 2360 2340 2320 2300 2280 Wavenumber (cm-1) Figures 2.47- FT-IR difference spectra, in the stretch region 2400-2280 cm-1 recorded after dosing CO 2 in sample at A MS outgassed at: (a) 350°C and (b) 500°C. Figures 2.48 reports normalised difference spectra in the 1900-1400 cm-1 recorded after dosing 20 mbar CO 2 on samples outgassed at 350 and 500°C. The same carbonate species exist in A and A MS . No significant differences were found on strength of the basic sites evaluated by [Δν=III ν as (COO-)- III ν s (COO-)], according to Peri36. This calculation confirms the similarities of dehydrated surfaces in terms of strength of basic sites. 0,35 ab A AMS 0,30 Absorbance 0,25 0,20 0,15 0,10 0,05 0,00 -0,05 -0,10 1900 1800 1700 1600 1500 1400 -1 Wavenumber (cm ) 96 Surface modification of a transition alumina 0,5 bb A AMS Absorbance 0,4 0,3 0,2 0,1 0,0 -0,1 1900 1800 1700 1600 1500 1400 -1 Wavenumber (cm ) Figures 2.48- FT-IR difference spectra, in the stretch region 2000-1280 cm-1 recorded after dosing CO 2 in sample at A MS outgassed at: (a) 350°C and (b) 500°C. The slight more intense carbonate bands observed in sample A MS outgassed at 500°C indicate a slightly higher abundance of basic sites. 2.3.4 Adsorption of NH 3 As reported in Chapter I, NH 3 is very often used as a probe to determine the number and nature of acidic surface sites. NH 3 was dosed by increasing the pressure (in the 0.05-40 mbar range) at room temperature. In the following Figures 2.49/2.50 are shown normalized FT-IR spectra recorded on samples A and A MS recorded after dosing NH 3 outgassed at 150, 350 and 500°C. 3,5 ab NH as Absorbance 3,0 a + 4 NH 3 2,5 2,0 1,5 4000 3500 3000 2500 2000 1500 -1 Wavenumber (cm ) 97 Surface modification of a transition alumina bb as a NH 3 Absorbance 2,0 1,5 1,0 4000 3500 3000 2500 2000 1500 -1 Wavenumber (cm ) 2,0 cb Absorbance as a NH 3 1,5 1,0 4000 3500 3000 2500 2000 1500 -1 Wavenumber (cm ) Figures 2.49- Normalized FT-IR spectra recorded on samples A recorded after dosing NH 3 outgassed at (a) 150°C, (b) 350°C and (c) 500°C (dashed line indicates the final outgassing at room temperature) Figures 2.49 reports the IR spectra after NH 3 dosage on sample A outgassed at 150, 350 and 500°C. As it is shown NH 3 molecules interacts via H-bonding (indicate with an arrow) and leads the formation of δ as NH 3 and the δNH 4 + ion as a result of the interaction with the Lewis and Brønsted sites. As reported in Figures 2.50, a similar behaviour was found in A MS sample leading the formation of the same species. 98 Surface modification of a transition alumina ab 4,0 NH + 4 Absorbance 3,5 a 3,0 as NH 3 2,5 2,0 1,5 4000 3500 3000 2500 2000 1500 -1 Wavenumber (cm ) 3,15 bb 3,10 Absorbance as a NH 3 3,05 3,00 2,95 4000 3500 3000 2500 2000 1500 -1 Wavenumber (cm ) Absorbance 2,4 cb as 2,2 a NH 3 2,0 1,8 1,6 4000 3500 3000 2500 2000 1500 -1 Wavenumber (cm ) Figures 2.50- Normalized FT-IR spectra recorded on samples A MS recorded after dosing NH 3 outgassed at (a) 150°C, (b) 350°C and (c) 500°C (dashed line indicates the final outgassing at room temperature) 99 Surface modification of a transition alumina Table VI reports the values corresponding to the position of Lewis and Brønsted bands with the relative position. Table VI- Position of δ as NH 3 and δNH 4 + in samples A and A MS . Sample A - Degassed 150°C A - Degassed 350°C A - Degassed 500°C A MS - Degassed 150°C A MS - Degassed 350°C A MS - Degassed 500°C Lewis δ as NH 3 (cm-1) 1623 1626 1626 1628 1626 1626 Brønsted δNH 4 + (cm-1) 1460 Not detected Not detected 1404 Not detected Not detected As a comparison, δNH 4 + is slightly shifted in the case of A MS but δ as NH 3 mostly preserves its position in both samples. 100 Surface modification of a transition alumina Conclusion Modification of a nano-crystalline commercial transition alumina was produced by different dispersion routes, as magnetic stirring or ball-milling with the aim of improving the sinterability and optimizing the fired microstructures. The study of the effect has been carried out by complementary techniques in the field of the materials science and surface physico-chemistry. The work gave rise that dispersion is effective in reducing the agglomerate size and modifying the crystallization path at low and high temperature regimes. Moreover, the dispersion induced some physico-chemical surface modification. Stirred sample has a surface more hydrophobic as well as more basic, compared with asreceived sample. Spectroscopic techniques permitted to evidence that stronger basic sites exists on the surface of the stirred sample. In fact, gibbsite phase was determined near transition aluminas in the stirred sample, which decomposes at about 280°C. Such hydroxide did not form on a blank sample suspended in water without stirring. The second conclusion is that dispersion do not modify the activation energy. However, this value was only decreased in the sample in which certain amount of α-phase was introduced by flash heating cycle. Furthermore, dispersed samples showed a better sinterability as higher densities were achieved, as well as, a different microstructural development. This fact permits to avoid the formation of the vermicular microstructure usually found in non-dispersed transition aluminas. In future, optimized powders should be produced by coupling flash heating followed by an effective dispersion process with the purpose of improving the powder sinterability. 101 CHAPTER 3 Nanostructured composite materials: elaboration and properties 102 Nanostructured composite materials: elaboration and properties 3.1. Nanocomposites materials: main classification Ceramic composites classification has been proposed for the first time by Niihara1. They can be divided into two types: microcomposites and nanocomposites. In the microcomposites, micro-size second phases such as particulates, platelets, whiskers and fibers are dispersed at the grain boundaries of the matrix. Nanocomposites can be defined as multiphase solid materials where one of the phases has one, two or three dimensions less than 100 nanometers. Niihara proposed four types, as it is shown in Figure 3.1 Figure 3.1- Classification of ceramics nanocomposites 1,2 As it is drawn in the figures above, the nano-sized particles can be dispersed mainly within the matrix grains (intra-type), at the grain boundaries of the matrix (inter-type) or can occupy both positions (intra/inter-type). The last classification is composed of dispersoid and matrix grains with nanometer-size. The most remarkable advantages of the intra- and intergranular nanocomposites are the improved fracture strength and toughness and reliability at room temperature; in addition, increased hightemperature mechanical properties such as hardness, strength, and creep resistance can be found. On the other hand, the nano/nano composites have the purpose of adding new functions such as machinability and the superplasticity like metals to ceramics. Many different nanocomposites were reported in the literature. The most studied ones were Al 2 O 3 and Si 3 N 4 matrix reinforced by SiC particles. Some properties of alumina-based nanocomposites were reviewed by Kuntz 3, as presented in the following figure. The interest of using these materials occurred after the publication of Niihara1 which showed an important increase of the fracture strength from 350 MPa in Al 2 O 3 up to 1000 MPa in Al 2 O 3 - 5 vol.% SiC. 103 Nanostructured composite materials: elaboration and properties Table 3.I : Different alumina-based composites reviewed by Kuntz3. 3.2. Synthesis of composites powders In principle, any method capable of producing very fine grain size materials can be used to process nanocomposites. Essentially, all of the successful processes have attributes that enable the crystalline phases to nucleate but suppress the growth of nuclei. Table 3.II summarizes the main used processing routes involved in the synthesis of ceramic-based nanocomposites. In addition, the main physical, chemical and mechanical/mechanicochemical routes are reviewed in the following items. 3.2.1. Physical Methods 3.2.1.1. Vapour condensation methods The inert gas condensation technique (Figure 3.2), conceived by Gleiter4, consists of evaporating a metal (by resistive heating, radio-frequency, heating, sputtering, electron beam heating, laser/plasma heating, or ion sputtering) inside a chamber that is evacuated to a very high vacuum of about 10-7 Torr and then backfilled with a low-pressure inert gas, like helium. They generally involve two steps: in the former, a metallic nanophase powder is condensed under an inert 104 Nanostructured composite materials: elaboration and properties convection gas. As a consequence, a supersaturated metal vapour is obtained in the chamber. In the latter, the powder is oxidized by allowing oxygen into the chamber (to produce metal oxide powder). Table 3.II- Main synthesis Methods of Ceramic based Nanocomposites5. A subsequent annealing process at high temperatures is often required to complete the oxidation. The system consists of a vapour source inside a vacuum chamber containing a mixture of an inert gas, usually argon or helium, mixed with another gas, which is selected on the ground depending on the material to prepare. Oxygen is mixed with the inert gas to produce metal oxides. NH 3 is usually used to prepare metal nitrides and an appropriate alkane or alkene, as a source of carbon and it is usually used to prepare metal carbides. Nanoparticles are formed when supersaturation is achieved above the vapor source. A collection surface, usually cooled by liquid nitrogen, is placed above the source. The particles are transported to the surface by a convection current or by a combination of a forced gas flow and a convection current, which is set up by the difference in the temperature between the source and the cold surface. Some improved systems involve a way to scrap the nanoparticles from the cold collection surface so that the particles would fall into a die and a unit where they can be consolidated into pellets. Supersaturated vapor can be achieved by many different vaporization methods. The most common techniques include thermal evaporation, sputtering, and laser methods. A variety of nanoscale metal oxides and metal carbides have been prepared using laser-vaporization techniques. The peak densities of the as-compacted metal samples have been measured with 105 Nanostructured composite materials: elaboration and properties values of about 98.5% of bulk density. However, it has been established that porosity has a important effect on the mechanical strength. Figure 3.2- Schematic drawing of the inert gas condensation technique for production of nanoscale powder 6,7. The advantages of vapor condensation methods include versatility, good performance and highpurity products. On the other hand, they can be employed to produce films and coatings. Furthermore, laser-vaporization techniques allow the production of high-density, directional, and high-speed vapor of any metal within an extremely short time. Despite the success of these methods, they have the disadvantage that the production cost is still high because of low yields. Heating techniques have other disadvantages that include the possibility of reactions between the metal vapors and the heating source materials. 3.2.1.2. Spray pyrolysis This technique is known by several other names including solution aerosol thermolysis (Figure 3.3), evaporative decomposition of solutions, plasma vaporization of solutions, and aerosol decomposition. The starting materials in this process are chemical precursors, usually appropriate salts, in solution, sol, or suspension. The process involves the generation of aerosol droplets by nebulizing or ‘‘atomization’’ of the starting solution, sol, or suspension. The generated droplets undergo evaporation and solute condensation within the droplet and the drying stage. Subsequently, it is followed by a thermolysis stage of the precipitate at higher temperature in order to form a microporous particle. Different techniques for atomization are employed including pressure, two-fluid, electrostatic, and ultrasonic atomizers. These atomizers differ in droplet size (2–15 mm), rate of atomization, and droplet velocity (1–20 m/sec). These factors affect the heating rate and residence time of the droplet during spray pyrolysis which, in turn, affect some of the particle characteristics including particle size. 106 Nanostructured composite materials: elaboration and properties For a specific atomizer, particle characteristics, including particle size distribution, homogeneity, and phase composition depend on the type of precursor, solution concentration, pH, viscosity, and the surface tension. Figure 3.3- Schematic drawing of thermal spray process, showing the different variables involved8. Aqueous solutions are usually used because of their low cost, safety, and the availability of a wide range of water-soluble salts. Metal chloride and nitrate salts are commonly used as precursors because of their high solubility. Precursors that have low solubility or those that may induce impurities, such as acetates that lead to carbon in the products, are not preferred. The advantages of this method include the production of high-purity nanosized particles, homogeneity of the particles as a result of the homogeneity of the original solution, and the fact that each droplet/particle goes through the same reaction conditions. The disadvantages of spray pyrolysis include the need for large amounts of solvents and the difficulty to scale-up the production. The use of large amounts of nonaqueous solvents increases the production expenses because of the high cost of pure solvents and the need for proper disposal. 3.2.1.3. Thermochemical/flame decomposition of metalorganic precursors Flame processes have been widely used to synthesize nanometer-sized particles of ceramic materials. This is another type of gas-condensation technique with the starting material being a liquid chemical precursor. The process is referred to as chemical vapour condensation (CVC). In this process, chemical precursors are vaporized and then oxidized in a combustion process using a fuel-oxidant mixture such as propane–oxygen or methane–air (Figure 3.4). It combines the rapid thermal decomposition of a precursor–carrier gas stream in a reduced pressure environment with thermophoretically driven deposition of the rapidly condensed product particles on a cold substrate. The flame usually provides a high temperature (1200–3000 K), which promotes rapid gas-phase chemical reactions. . 107 Nanostructured composite materials: elaboration and properties Figure 3.4 – (A) Silicon nanoparticles by CVC methods; (B) Pieces of nanocomposites prepared by adding metal salts to a sol before gelation8. A variety of chemical precursors can be used including metal chlorides, such as TiCl 4 to prepare TiO 2 and SiCl 4 to prepare SiO 2 10, metal-alkyl precursors, metal alkoxides, and gaseous metal hydrides, such as silane as a source of silicon for the preparation silica. Chlorides have been the most widely used precursors in the industry and the process is sometimes referred to as the ‘‘chloride process.’’ The high vapor pressure of chlorides and the fact that they can be safely stored and handled make them excellent potential precursors. The disadvantages of using chloride precursors are the formation of acidic gases and contamination of the products with halide residues. Flame processes are used industrially to produce commercial quantities of ceramic particulates, such as silica and titania. This is because of the low cost of production as compared to all other methods. The disadvantage of flame synthesis is that the control of particle size (both primary particle and aggregates size), morphology, and phase composition is difficult and limited. 3.2.2. Chemical Methods 3.2.2.1. Sol–gel technique The sol–gel process is typically used to prepare nanometer-sized particles of metal oxides (Figure 3.5)12. This process is based on the hydrolysis of metal reactive precursors, usually alkoxides in an alcoholic solution, resulting in the corresponding hydroxide. Condensation of the hydroxide by giving off water leads to the formation of a network-like structure. When all hydroxide species are linked, gelation is achieved and a dense porous gel is obtained. The gel is a polymer of a three-dimensional skeleton surrounding interconnected pores. Removal of the solvents and appropriate drying of the gel result in an ultrafine powder of the metal hydroxide. Further heat treatment of the hydroxide leads to the corresponding powder of the metal oxide. As the process starts with a nanosized unit and undergoes reactions on the nanometer scale, it results in nanometer-sized powders. For alkoxides that have low rates of hydrolysis, acid or base catalysts can be used to enhance the process When drying is achieved by evaporation under normal conditions, the gel network shrinks as a result of capillary pressure that occurs and the hydroxide product obtained is referred to as xerogel. However, if supercritical drying is applied using a high-pressure autoclave reactor at temperatures higher than the critical temperatures of solvents, less shrinkage of the gel network occurs as there is no capillary pressure and no liquid–vapor interface, which better protects the porous structure. The hydroxide product obtained is referred to as an aerogel. Aerogel powders usually demonstrate higher porosities and larger specific surface areas than analogous xerogel powders. 108 Nanostructured composite materials: elaboration and properties Sol–gel processes have several advantages over other techniques to synthesize nanopowders of metal oxide ceramics. These include the production of ultrafine porous powders and the homogeneity of the product as a result of homogeneous mixing of the starting materials on the molecular level (Figure 3.5). Figure 3.5- An example of sol-gel processing conditions on film formation8. 3.2.2.2. Reverse microemulsions/micelles method The reverse micelle approach is one of the recent promising routes to nanocrystalline materials including ceramics. Surfactants dissolved in organic solvents form spheroidal aggregates called reverse (or inverse) micelles9. In the presence of water, the polar head groups of the surfactant molecules organize themselves around small water pools (≈100 Å), leading to dispersion of the aqueous phase in the continuous oil phase. Reverse micelles are used to prepare nanoparticles by using a water solution of reactive precursors that can be converted to insoluble nanoparticles. Nanoparticles synthesis inside the micelles can be achieved by different methods including hydrolysis of reactive precursors, such as alkoxides, and precipitation reactions of metal salts. Solvent removal and subsequent calcination lead to the final product. Several parameters, such as the concentration of the reactive precursor in the micelle and the weight percentage of the aqueous phase in the microemulsion, affect the properties, including particle size, particle-size distribution, agglomerate size, and the phases of the final ceramic powders. There are several advantages when this method is applied including the ability to prepare very small particles and the ability to control the particle size. Disadvantages include low production yields and the need to use large amounts of liquids. 3.2.2.3. Precipitation from solutions Precipitation is one of the conventional methods to prepare nanoparticles of metal oxide ceramics. This process involves dissolving a salt precursor, usually chloride, oxychloride or nitrate, such as AlCl 3 to make Al 2 O 3 , Y(NO 3 ) 3 to make Y 2 O 3 , and ZrCl 4 to make ZrO 2 11,12 . The corresponding metal hydroxides are usually obtained as precipitates in water by adding a base solution such as sodium hydroxide or ammonium hydroxide solution. 109 Nanostructured composite materials: elaboration and properties The remaining counter-ions are then washed away and the hydroxide is calcined after filtration and washing to obtain the final oxide powder. This method is useful in preparing ceramic composites of different oxides by co-precipitation of the corresponding hydroxides in the same solution. Solution chemistry is also used to prepare non-oxide ceramics or pre-ceramic precursors that can be converted to ceramics upon pyrolysis. One of the disadvantages of this method is the difficulty in controlling the particle size and size distribution. Very often, fast and uncontrolled precipitation takes place resulting in large particles. 3.2.2.4. Chemical synthesis of pre-ceramic polymers coupled with physical processing techniques This method is based on the use of molecular precursors, which facilitates the synthesis of nanomaterials containing phases of desired compositions. It involves a chemical reaction to prepare an appropriate polymer, which is then converted into ceramic material upon pyrolysis. Figure 3.6- General structural formulas of polycarbosilanes and polysilazanes13. Using chemical reactions to prepare the pre-ceramic polymer not only allows control of phase compositions but also overcomes the limitation of low production yields of the physical methods. This method has been very useful in preparing nonoxide ceramics such as silicon carbide and silicon nitride. The conversion of an organometallic precursor into a ceramic depends on different parameters such as the molecular structure of the precursor and the pyrolysis conditions (temperature, duration, and atmosphere). Metal carbides and metal nitrides have been obtained by pyrolysis of polymers containing the appropriate elements such as Si or Al and C or N. These polymers are called pre-ceramic polymers and are prepared from simpler chemical precursors. A considerable amount of free carbon from the thermolysis process is very often a problem. Silicon carbide (SiC) and silicon nitride (Si 3 N 4 ) are the most studied ceramic materials prepared via this route. They are usually synthesized by the pyrolysis of polycarbosilanes and polysilazanes, for which general structural formulas are shown in Figure 3.6, at temperatures between 1000°C and 1200°C. This method could be employed to prepare Al 2 O 3 /SiC with ultrafine SiC particles using a Si-containing precursor15. The polycarbosilane (Figure 3.6) is coated into a surface-modified alumina powder and pyrolysed at 1500°C to produce ultrafine SiC particles of less than 20 nm. The following flow chart shows the different elaboration methods generally employed for producing Al 2 O 3 /SiC nanocomposites. 110 Nanostructured composite materials: elaboration and properties Figure 3.7 – Flow chart representing the processing of Al 2 O 3 /SiC nanocomposites by (A) classical powder processing, (B) sol-gel processing and (C) the polymer coating route15. 3.2.2.5. Colloidal processing route Schehl et al.16 developed a new elaboration process for alumina-based nanocomposites (Figure 3.8), known as colloidal processing route. It consists in doping of a commercial, high-purity alumina powder with different alkoxides. Precisely, alumina slurries were dispersed in absolute ethanol. Suitable precursors were then drop-wise added to the alumina slurries. The slurries were first dried under magnetic stirring at 70°C and subsequently in air at 120°C for 24 h in order to eliminate the traces of alcohol. The dried powders were subsequently crushed in a mortar, to remove agglomerates resulting from the drying process. The final phases are yielded by subsequent calcinations processes. When alumina particles are dispersed in ethanol, protons or hydroxyls are adsorbed on the surface of the alumina particles, as it is shown in Figure 3.8. The addition of metal alkoxides to this dispersion causes a substitution reaction between the metal alkoxides and the OH groups on the surface of the alumina, as a result, the surface of the oxide particle is coated with a metal alkoxide. Figure 3.8 – Substitutional reaction between the metal alkoxides and the OH groups on the surface of the alumina16. 111 Nanostructured composite materials: elaboration and properties Palmero et al.17 showed an alternative route which consists in doping commercial alumina with aqueous solution of inorganic salts. 3.2.2.6. Mechanochemical synthesis Mechanochemical synthesis involves mechanical activation of a solid-state reaction. This process has been successfully used to make nanopowders such as Al 2 O 3 and ZrO 2 . Selected precursors (usually a salt and a metal oxide) react under milling and subsequent heating, to form a mixture of dispersed nanoparticles of the desired oxide within a salt. Nanoparticles of Al 2 O 3 (10–20 nm), for example, can be prepared by milling AlCl 3 with CaO14. 2AlCl3 + 3CaO -Al2O3 + 3CaCl2 (1) 3.3. Forming and Sintering 3.3.1. Dry Pressing Dry pressing is a suitable method to produce simple solid shapes and consists of three basic steps: filling the die, compacting the content, and ejecting the pressed solid. Figure 3.9 - Stages in Dry Pressing18. Figure 3.9 shows a schematic diagram of the double action dry-pressing process, in which both the top and bottom punches are movable. When the bottom punch is in the low position, a cavity is formed in the die and this cavity is filled with free flowing powder. Once the cavity has been filled, the powder is pressed in the die. The top punch descends and compresses the powder either to a predetermined volume or to a set pressure. During pressing, the powder particles must flow between the closing punches so that the space between them is uniformly filled. A particle size distribution between 20 and 200 μm is often preferred for dry pressing: a high volume fraction of small particles causes problems with particle flow and also results in the sticking of the punches. The pressures used in dry pressing may be as high as 300 MPa, depending upon material and press type. After pressing, both punches move upward until the bottom punch is level with the top of the die and the top punch is clear of the powder-feeding mechanism. The compact is then ejected, the bottom punch is lowered, and the cycle is repeated. Because the dry-pressing process is so simple and involves low capital equipment costs it is the most widely used high-volume forming process for ceramics. Production rates depend on the size and shape of the part and on the type of press used. 112 Nanostructured composite materials: elaboration and properties 3.3.2. Slip Casting The slip is poured into a mold with the required shape. The porous nature of the mold provides a capillary suction pressure, estimated to be of the order of ≈0.1–0.2 MPa, which draws the liquid from the slurry into the mold. A consolidated layer of solids, referred to as cast, forms on the walls of the mold (Figure 3.10). After a sufficient thickness of the cast is formed, the surplus slip is poured out and the mold and cast are allowed to dry. Normally, the cast shrinks away from the mold during drying and can be easily removed. Once fully dried, the cast is heated to burn out the binder and sintered to produce the final sample. Figure 3.10 - Schematic illustration of the drain-casting process: (a) the mold is filled with slip; the liquid is drained from the mold, forming a compact along the mold walls, (b) after it is possible to remove the sample once the sample is dried, subsequently green ceramic is removed18. Among the various forming procedures, slip casting has received much attention, since it has the main advantage of eliminating the drying step of slurries, which can lead to the formation of hard agglomerates and, consequently affecting achievable green density13. 3.3.3. Sintering A major challenge encountered in the consolidation of ceramic nanopowders is mitigation of grain growth during sintering. Typically, a nanopowder compact experiences rapid densification during the early stages of sintering, driven by the large surface to volume ratio. At this stage, the grain size of the partially sintered material remains small due to the presence of a uniform distribution of nanopores, which act as barriers to grain-boundary migration. However, in the final stages of sintering (>90% of the theoretical density, TD), when the nanopores disappear, it is followed by undesiderable grain growth that occurs, often leading to a micrograined sintered product. In the following, the most exploited sintering methods to consolidate nanopowders into dense, fine composite ceramics are briefly described. 3.3.3.1. Hot Pressing Gao et al.22 produced YAG-Al 2 O 3 composites from commercial alumina powders sintered by HP. In order to obtain fully dense samples, sintering temperature was set to 1600°C for 1 h. The applied pressure, during sintering, was 25 MPa. In Figure 3.11 are shown the microstructures of two composites with different compositions. 113 Nanostructured composite materials: elaboration and properties Figure 3.11 – SEM micrographs of (A) YAG- 10 vol.% Al 2 O 3 and (B) YAG- 45 vol.% Al 2 O 3 22. 3.3.3.2. Hot Isostatic Pressing The hot isostatic pressing apparatus (Figure 3.12) consists of a high-temperature furnace enclosed in a water-cooled autoclave capable of withstanding high internal gas pressures. As pressurization gas, it is widely used argon and/or helium17. HIP has the potential of solving some of the major limitations of hot-pressing. It makes possible net shape forming because the pressure is equally applied from all directions. As a consequence, the material has greater uniformity eliminating the preferred orientation, resulting high-strength and Weibull modulus. Figure 3.12 – Schematic diagram of a pressure vessel with a sample for HIP21. 3.3.3.3. Spark-Plasma Sintering Spark-plasma sintering is a low-pressure sintering method that makes use of a plasma discharge through a powder compact to achieve rapid densification. The discharge is most effective when a DC current is applied in an on-off pulsing mode. It has been suggested that DC pulsing generates: (1) spark plasma,(2) spark impact pressure, (3) Joule heating, and (4) an electrical field diffusion effect. 114 Nanostructured composite materials: elaboration and properties The inventors claimed that the pulsed direct current generates a spark discharge and/or plasma is the responsible of cleaning the surfaces of the samples of CO 2 , H 2 O and OH-. These processes promote the material transfer and make rapid densification of the powder compact possible at low temperature and short holding time. In a typical operation, powders are loaded into a graphite die and heated by passing an electric current through the assembly. The experimental setup and electric field effects are illustrated in Figure 3.13. Figure 3.13 – Scheme of SPS sintering apparatus and mechanisms involved15. The low heat capacity of the graphite die allows rapid heating, thus promoting heat and mass transfer. Hence, SPS rapidly consolidates powders to near TD through the combined actions of rapid heating rate, pressure application, and possibly powder surface cleaning. In most investigations, SPS is carried out under vacuum. Starting with a cold-pressed (≈ 200MPa) powder compact, typical processing parameters are: (1) an applied pressure of <100 MPa, (2) pulse duration of 12 ms and pulse interval 2 ms, and (3) pulse current of ≈2000 A at a maximum voltage of 10 V. Typical heating rates range from 150 to 500°C/min. As Shen et al.22 summarised three reasons that contributes to the rapid densification: The application of a mechanical pressure (as in HP and HIP). The use of rapid heating rates. The use of pulsed direct current, implying that the samples are exposed to an electric field. SPS has been used to consolidate a wide variety of materials, including metals, intermetallics, ceramics, composites, and polymers. As for functionally graded materials and nanocrystalline materials, which are difficult to sinter by conventional methods, the advantage of SPS is more evident. For example, Zhan et al.23 successfully sintered nanocrystalline α-Al 2 O 3 with 5 vol.% carbon nanotubes at 1150°C in 3 min, while conventional hot pressing of α-Al 2 O 3 requires 1500 to 1600°C for 3 to 4 h. Hence, with a lower sintering temperature and shorter sintering time, SPS enables better control of structure and properties of the consolidated material. Gao et al.24 compared the microstructures (Figure 3.14) of YAG-Alumina composites obtained by HP (A) and SPS (B). HP was performed at 1200 K, with a heating rate of 600°C/min and an applied pressure of 40 MPa; SPS indeed was carried up at 1573 K (with a heating rate of 600 K/min) under an applied pressure of 40 MPa. 115 Nanostructured composite materials: elaboration and properties Figure 3.14 – SEM micrographs of YAG- 10 vol.% Al 2 O 3 (A) densified by SPS and (B) HP sintering routes24. As a result, the SPS specimens were fully dense with an average grain size 3 times smaller than by HP as it is illustrated in Figure 3.14. 3.3.3.4. Microwave Sintering Microwave sintering allows the consolidation of the materials with shorter processing time than conventional sintering. The heating flow is generated inside the samples (see Figure 3.15) as a consequence of absorption of electromagnetic waves by electric dipoles, and not derived from external sources as in conventional heating processes. Temperature gradients are also reduced and an overall short sintering time minimizes grain growth. In fact, by rapid heating rate, the low-temperature region in which grain growth occurs during densification can be bypassed. In addition, binder is easily removed and thermal stresses inside the samples are significantly reduced. Since grain boundaries are the primary sites for electric dipoles, microwave sintering appears particularly attractive for the densification of nanocrystalline powders. The method has been applied to ceramic nanoparticles such as TiO 2 25-26 and Al 2 O 3 27-29. Figure 3.15 – Comparison between conventional and microwave sintering techniques26. 116 Nanostructured composite materials: elaboration and properties In these materials, only densities lower than 95% have been achieved, with grain size smaller than 100 nm. For instance, microwave sintering had to be restricted to 1425 K to maintain a nanometer grain size in γ-Al 2 O 3 which fully transformed to α-Al 2 O 3 , reaching a final density of 93%. Tian et al.30 studied microwave sintering of Al 2 O 3 -TiC composites for potential cutting tool applications. Although high densities were achieved, high power levels permit to reduce the sintering temperature up to 1750°C. 3.4. The interest in nanocomposites ceramics: the mechanical properties 3.4.1. Role of the second phase on the retention of the matrix grain growth Second phase fine particles, if uniformly distributed into a micronic matrix, can limit grain growth, by the well known pinning effect32. In 1948, Zener32-33 explained how inclusions retard grain growth and described the dependence of maximum matrix grain size (d max ) as a function on the particle size (r p ) and second phases volume fraction (f): d max 4rp 3f (2) In Figure 3.16 (A) it is schematically shown how shrinking interacts with spherical inclusions. During shrinking, the grains encounter the inclusion and an increasing proportion of the boundary is removed. When N inclusions are simultaneously intersected, the maximum amount of grainboundary is πR2, which corresponds to a decreasing in free energy of πR2γ. If the grain continues shrinking, the grain boundary area adjacent to the inclusions must withdraw up to the third position see Figure 3.16 (A) up to the breaking, regaining its area occupied by the inclusion during bowing. The relative variation of free energy versus the volume fraction is reported in Figure 3.16 (B). Figure 3.16 – (A) Schematic representation of grain boundaries of a tetrahedral grain interacting with spherical inclusions. (B) Energy of the tetrahedral grain versus its grain volume variation33. In such a way, Zener provided the concept for the use of inclusions to hinder the grain growth. For instance, the effect of a SiC inclusion is reported in Figure 3.17, in which it is represented the alumina grain size in function of the sintering time at 1700°C. Retaining grain size effect by the inclusions result more efficient when the second phase content increases. 117 Nanostructured composite materials: elaboration and properties Figure 3.17 – Alumina grain size evolution in function of sintering time and SiC content34. A major grain growth restraint can be achieved by the elaboration of the so-called duplex microstructures. Two immiscible phases, contained in similar volume fractions, can give rise to interconnected structures, so that the growth of each phase is hindered by the other one and longorder interdiffusion is strongly limited31. For instance, in Figure 3.18, the microstructure of Al 2 O 3 /50 vol.% ZrO 2 composite (Figure 2) is compared to those of the single-phase constituents, sintered under the same conditions. A significantly finer microstructure was produced in the composite, compared to the pure-phase materials. Figure 3.18 – SEM micrograph of pure alumina (1), pure Zirconia (3) and of the composite Al 2 O 3 /50 vol.% ZrO 2 (2)31. 3.4.2. Hardening Hardness of materials is one of the most important mechanical properties from an engineering point of view. Hardening in metals is performed by the inhibition of dislocation glide, which can be done by microstructural tailoring. The Hall-Petch35-36 hardening is described by the following relationship H V = H 0 + k H d -1/2 (3) where H is the measured hardness, Ho and k H specific constants - depending on the material and d is the grain size - . The similar concept will be used to explain the strengthening mechanism in the next section. In ceramics, the correlation between hardness and microstructure has not been cleared, although many papers reported this subject37. In addition, hardness in nanoceramics with grain sizes lower than 1 μm has not been examined in detail. 118 Nanostructured composite materials: elaboration and properties Figure 3.19 – Variation of hardness vs. grain size for monolithic alumina50 From both the experimental and the theoretical points of view, it is difficult to estimate the plasticity of hard elastic-plastic material which complicates the understanding of the hardness in ceramics, if compared to metals. Many works report hardness vs. grain size. Data reported by Shen et al.38 (Figure 3.19) for monolithic alumina also showed the effect of the processing parameters on the hardness values. By varying these parameters, it is possible to reduce the grain size and consequently reduce mobility of dislocations in small grains. 3.4.3. Change in fracture mode The addition of nano-size dispersed particles change the fracture mode from intergranular to transgranular, resulting in a strengthening mechanism. The first explanation was given by Awaji et al.39, who attributed the strengthening mechanism in nanocomposites to dislocation activities. In Figure 3.20, the crack paths occurring in monolithic alumina (a) and into an alumina-based nanocomposite (b) are shown. In pure alumina, the residual stresses are formed at the interfaces as a result of the thermal expansion anisotropy. When a crack propagates, it selects the interfaces under the residual traction: the predominant fracture mode is intergranular (Figure 3.20 A). On the contrary, in case of the composite (Figure 3.20 B), the dislocations generated around the dispersed particles relieve the tensile residual stresses in the matrix, and consequently reduce the defect size at the grain boundaries. The dislocations are difficult to move in ceramics at room temperature: they act as stress concentrations sites and create small nanocracks around the main crack tip. So, the change in the fracture mode from intergranular to transgranular was imputed to the reduction of both the defect size along the grain boundaries and the transgranular strength in the matrix. 119 Nanostructured composite materials: elaboration and properties Figure 3.20 – Schematic description of the strengthening mechanism as a function of crack path for monolithic alumina (a) and nanocomposites (b)39. A second explanation, given by Ohji et al.40, lies in the differences in relaxation of the tensile residual stresses around the intergranular and intragranular particles. The internal stresses are relaxed by lattice and grain-boundary diffusion around the intragranular and intergranular particles, respectively. As a consequence, the tangential tension around the intragranular particle of a sintered body is always greater than around the interfacial particles, and in the case of crack, it will propagate intergranularly (Figure 3.21). Figure 3.21 – Schematic representation of internal stresses around the particles and crack propagations40. 120 Nanostructured composite materials: elaboration and properties 3.4.4. The strength The matrix grain size in nanocomposites not only decreases with the decreasing of the sintering temperature but it also decreases with increasing the second-phase content, as reported in section 3.4.1. This feature has to be taken into account for a true comparison of mechanical properties of different nanocomposites. The strength of ceramics (σ), as already reported for hardness (Section 3.4.2), usually follows a Hall-Petch relation35-36: o k .d 1/ 2 (4) where σ is the strength for materials with grain size d, σ o is the friction stress and k is a constant. Chantikul et al.46 carried out an investigation of the correlation between strength and grain size by indentation tests on monolithic alumina with a grain size range from 2 to 80 μm, (Fig 3.22). The authors found that strength follows the Hall-Petch relationship. Figure 3.22 – Variation of fracture strength vs. grain size for monolithic alumina46. Figure 3.23 collects data from different authors of strength of alumina-based nanocomposites as a function of SiC content. The best results were obtained by Niihara1, who proved that only 5 vol.% of nanosized SiC increased the fracture strength of pure alumina from about 300 to 1050 MPa. Subsequent SiC addition lowers the strength at a constant value of 800 MPa, as a consequence of SiC particles agglomeration. Borsa et al.41 observed a slight increase in strength after de addition of SiC in one alumina matrix, in contrast with the literature data 1,2 with show gains of 40%. 121 Nanostructured composite materials: elaboration and properties Figure 3.23 – Strength of Al 2 O 3 /SiC nanocomposites as a function of SiC content: (●) Niihara et al.1 by three-point bend test and Vickers indentation; (□) Borsa et al.41 by four- point bend test; (▲) Zhao et al.2 by four- point bend test and indentation-strength method; (▼) Davidge et al.42 by threepoint bend test and notched beams. Niihara et al. 1 found a further improvement of strength up to 1550 Mpa, by post-annealing at 1300°C for 1 h in air or inert atmosphere (Figure 3.24). Chou et al.43 confirmed this behaviour: they recorded an increase of about 50% of fracture strength of annealed composites as compared to un-annealed ones. Figure 3.24 – Improvement on fracture strength of alumina as a function of SiC content and postannealing treatment1 SPS was employed as a densification technique for Al 2 O 3 /SiC composites by Gao et al.44. The authors studied the influence of the sintering temperature on the fracture strength. Figure 3.25 shows the strength dependence as a function of the sintering temperature. The maximum strength of 1000 MPa is obtained at 1450°C. By increasing the sintering temperature over 1450°C there is a decay of strength due to excessive grain growth. Zhan et al.45 synthesized Al 2 O 3 /SiC composites by an innovative process named RHP (reactive hot pressing). Mullite, aluminium and carbon were used as starting materials; they were heated at 1800°C, increasing the pressure up to 30 MPa. The final product was characterized by a SiC content of 18.25 vol.% calculated from the formula below: 3 3 Al2O3 .2 SiO2 8 Al 6C 13 Al2O3 6 SiC (5) 122 Nanostructured composite materials: elaboration and properties The produced composite material showed a very high fracture strength of 795 (+/- 160) MPa. The Authors suggested that such high value was a consequence of the existence of a non amorphous phase between the two phases, as evidenced by HR-TEM observation (Figure 3.26). Figure 3.25 – Bending Strength versus sintering temperature for Al 2 O 3 /5 vol.% SiC sintered by SPS44. Figure 3.26– High-resolution TEM image of the grain boundary of Al 2 O 3 /SiC nanocomposite45. 3.4.4.1. Reduction in processing flaw size Niihara and Zhao et al.1,3 proposed that strengthening arise due to the refinement of the microstructural scale and/or the compressive stresses around SiC particles which inhibit the grain boundary fracture of that matrix during cooling from the fabrications. This phenomena is the responsible for reducing the critical flaw size. The author found an increase from 6.5 to 13.6 by the dispersion of only 5 vol.% SiC. The Weibull plot is reported in Figure 3.27. This conclusion suggests that the reliability of Al 2 O 3 is improved by the nano-size SiC dispersions into the matrix Al 2 O 3 . 123 Nanostructured composite materials: elaboration and properties Figure 3.27 – Weibull plot for Al 2 O 3 /5 vol.% SiC nanocomposites compared with the monolithic alumina1 3.4.4.2. Dislocation networks Another strengthening mechanism for Al 2 O 3 /SiC nanocomposites based on flaw size reduction has been presented by Niihara. He proposes that strengthening arises due to the refinement of the microstructural scale from the order of the alumina grain size to the order of the interparticle spacing, thus reducing the critical flaw size. The occurrence of subgrain or low angle grain boundaries is widely acknowledged. During cooling down, SiC particles can generate dislocations owing to internal stresses. At high temperatures these dislocations can propagate and form dislocation networks. However, Niihara’s strength and toughness values for the Al 2 O 3 /SiC system of 1 GPa and 4.8 MPa1/2, respectively, yielded a Griffith critical flaw size of 18 μm. A refinement of the microstructure from the average alumina matrix grain size of 1.5 μm to the average interparticle spacing of 200 nm plays a minor role. 3.4.4.3. Crack healing Zhao et al.47 suggest that SiC particles only indirectly influence the strength by enabling the compressive stresses induced by the grinding process. These retained compressive stresses are located on the surface region of the specimens. Another theory is that cracks in nanocomposites can heal during annealing. After annealing at 1300°C in Ar for 2 h the materials behave completely differently. Whereas cracks in alumina grow, cracks in nanocomposites close, thus explaining the strength increase of annealed nanocomposites. 3.4.5. Toughening mechanisms 3.4.5.1. Intrinsic Fracture Energy Toughening and strengthening mechanisms in nanocomposites could be explained by the Griffith’s energy equilibrium and the residual stresses around second-phase particles dispersed in matrix grains. It is well known that many polycrystalline ceramic and ceramic-based composites exhibit R-curve behaviour due to the crack bridging mechanism. The mechanism could be expressed as: 124 Nanostructured composite materials: elaboration and properties K R a K i K R a (6) where K R (∆a) is the fracture toughness of the material exhibiting R-curve behaviour, K i is the intrinsic fracture toughness, ∆K R (∆a) is the intrinsic increase in fracture toughness after certain extension from the initial crack tip, ∆a. An example of cracked surface in polycrystalline ceramics with rising R-curve behaviour is shown in Figure 3.28. Figure 3.28– Schematic drawing of a critical FPZ and the bridging in the wake of a polycrystalline ceramics with the R-curve behaviour39. The comparison between the previous formula and the Figure 3.28, indicates that the intrinsic fracture toughness, K I , is related to the energy required to create the damaged FPZ at the crack tip and that ∆K R is caused by the shielding effects of bridging in a process zone wake. The FPZ (fracture process zone) is the place in which micro-failure mechanisms take place. Such processes include micro cracking, crack deviation and crack branching which all contribute to the fracture energy. So, the first term in equation (6) is a fracture energy for formation of the FPZ at the crack tip, while the second one is the mechanical energy consumed at the bridging. Therefore, it is possible to modify the Griffith-Irwin formula, which can be expressed for mode I extension as: KR2 2. i R E' (7) The equation describes the critical energy release rate after certain crack extension in materials with an R-curve, and γ i (mode I fracture per unit area of the cracked surface to create the critical FPZ size) and γ R (fracture energy per unit area to be consumed at the bridging) are the intrinsic and extrinsic fracture energy per unit area of the cracked surface. Much effort has been done towards increasing the intrinsic energy consumed in the FPZ and the extrinsic energy consumed at the bridging by means of whisker, fiber, and platelet reinforcement in the matrix. The toughening mechanism of nanocomposites is related to the intrinsic fractured energy, γ i , which is formed around the second-phases, and is expected to make many nanocracks from the origins of stress concentrations, such as dislocations in a vicinity of a main crack tip. The nanocracks play a role of expanding the size of FPZ and consequently enhancing the material toughness39. 125 Nanostructured composite materials: elaboration and properties Figure 3.29 – Schematic description of the toughening mechanism for nanocomposites39. Figure 3.29 reports an illustration of the FPZ at a main crack tip in an annealed Al 2 O 3 /SiC nanocomposites. In a matrix grain, sub-grain boundaries with dislocation or dislocations networks are generated around the dispersed SiC particles by means of annealing. When a main crack tip reaches this area, the sessile dislocations in the ceramic matrix at room temperature will operate as nano-crack nuclei in the highly stressed area. The FPZ consequently is expanded obliged by nano-crack formation. Ohji et al.39 proposed a comparison between the behaviour between the monolithic alumina and the composite, for the same initial crack size. The strength of the materials can be determined by the slope of the tangent line from the initial crack length to the respective R-curves (Figure 3.30). As it was reviewed by the authors, the toughening of the monolith, due to grain bridging and pullout, microcracking, etc. requires a crack extension of approximately 300 μm. However, for the nanocomposites, the particle bridging operates inside a grain in the composite and consequently, the fracture resistance increases in the 200-300 μm crack size. Figure 3.30 – Variation of the fracture-resistance in function of the square-root of crack extension for the Al 2 O 3 /SiC and the monolithic alumina40. 126 Nanostructured composite materials: elaboration and properties In Al 2 O 3 /SiC nanocomposites, Niihara 1 was the first in demonstrating the improvement in fracture toughness up to 4.2 MPa.m1/2 corresponding to a 5% of SiC content. Other authors as Davidge 42 found a little improvement, as it is shown in Figure 3.31. Figure 3.31 - Toughness of Al 2 O 3 /SiC nanocomposites as a function of SiC content: (●) Niihara et al.1 by three-point bend test and Vickers indentation; (▼) Davidge et al.42 by three-point bend test and notched beams. 3.4.5.2. Crack Bowing Particles in nanocomposites can also cause local changes in crack velocity. This effect can be described as crack bowing and has been proposed as a mechanism for increasing the fracture toughness of brittle materials. Green49 has developed an analytical expression to characterise numerically the fracture toughness associated with the crack bowing effect. This expression depends on the free interparticle spacing, A, but is independent of particle size. In nanocomposites a crack can rest at SIC particles but it is not clear if crack bowing occurs. Again, a whole crack front has to be investigated in order to study the effect of crack bowing. Pezzotti et al.50 have presented a theoretical approach for modelling toughness and strength in ceramic/ceramic and especially in ceramic/metal nanocomposites. Their model is based on the effect of the crack bowing effect. The authors conclude that, in contrast to metallic inclusions, ceramic nanosized dispersoids are completely ineffective on the material strength. 3.4.5.3. Average Internal Stress Since the microstructures of nanocomposite ceramics are formed during sintering at high temperatures, differences in the thermal expansion coefficients of the matrix (α matrix ) and of the nano-particles (α particles ) cause strains during cooling. These differences in thermal expansion <α*> can be calculated by an integration over temperature. The upper limit is taken as the temperature below which plastic deformation is insignificant (T plastic ) and the lower limit is the room temperature15. * Tplastic particle matrix dT (8) To The thermal expansion misfit stress, σ T , inside a single spherical inclusion in an infinite matrix can be described by the following expression: T * 1 m 1 2 m 2 Em 2E p (9) 127 Nanostructured composite materials: elaboration and properties where E and v are Young’s modulus and Poisson’s ratio of the matrix (m) and the particles (p). The tangential, σ Tt , and the radial, σ Tr , stress distributions in the matrix around the particle are given by: Tt T r 2 x 3 3 r Tr T (10) x where r denotes the radius of the inclusion and x is the radial distance from the inclusion surface. The residual stress around the dispersed particles could be represented using a simplified model consisting of spherical particles within a concentric matrix sphere as it is shown in Figure 3.32 to clarify the dislocation idea. Figure 3.32 – A spherical particle with a glassy phase within a concentric sphere of a matrix grain39. For Al 2 O 3 /SiC the existence of a surrounding layer (glass-phase) on the second-phase particle is considered for analysing the effects of the interlayer on the stress distributions. In Figure 3.33 he residual stresses have the highest values at the particle/matrix boundary and reduce drastically as the distance from the boundary increases. Figure 3.33 – Comparison of stresses around a dispersed particle without an interlayer in matrices of finite and infinite spheres39. 128 Nanostructured composite materials: elaboration and properties 3.4.5.4. Toughening by Transformation Dispersions of zirconia particles can be used to toughen other ceramics. Zirconia-toughened alumina (ZTA) is the most common example and allows the strength and toughness of alumina to be improved while retaining some of its advantages over zirconia (i.e., it almost doubles the stiffness)51. ZTA is usually made by sintering mixtures of powders of alumina and zirconia (10 to 30 vol%) containing some stabilizing oxide. The typical size of the zirconia particles in ZTA is similar to the grain size of tetragonal zirconia polycrystals (TZP), and often the same amounts of stabilizer are used as in TZP. It should be noted, however, that the high stiffness of alumina will constrain the t → m transformation more than in TZP, while the thermal expansion mismatch (≈5 × 10−6 K−1) between the two ceramics, during cooling after sintering, causes tensile residual stresses in the zirconia which will help the transformation of Zirconia. To optimize the processing of transformation toughened material, it is essential to understand the toughening mechanism and how it relates to microstructure. In understanding the origin of transformation toughening, it is instructive to consider the propagation of a preexisting straight crack as it is gradually loaded in fracture mode I crack by the application of an external force. Initially, the crack tip is surrounded by untransformed material, which consists wholly or partly of metastable particles of t-ZrO 2 (Figure 3.34 a). When a tensile load is applied, large stresses arise close to the crack tip and the t-ZrO 2 in the region over which a critical stress (assumed hydrostatic), σT, it is exceeded which cause the expansion of frontal zone of transformation as the stress intensity at the crack tip is increased (Figure 3.34 b). This frontal zone is dilated compared with the surrounding, untransformed material resulting in residual stresses both within the frontal zone and in the untransformed material around it. The residual stress, σ R , in the transformation zone is broadly compressive while the stress outside has both tensile and compressive components and decays rapidly with distance from the particle. As soon as the crack begins to grow the situation changes. The crack tip enters the frontal zone and the compressive stresses within it, act to close the crack faces just behind the tip, producing a strongly negative internal contribution, K T , to the total stress intensity (Figure 3.34 c). 129 Nanostructured composite materials: elaboration and properties Figures 3.34 – Schematic representation of events leading to transformation toughening: (a) unstressed crack in matrix containing t-ZrO 2 particles, (b) stress-induced t→m transformation occurs ahead the crack to form frontal zone of transformation and (c) crack grows into compressed transformation zone which extends ahead of the crack51. The condition for further crack growth remains so that the total stress intensity is equal to the local toughness of the material and thus the externally applied stress intensity required for crack growth increases and is given by: K c TL KT (11) As K T is negative, the macroscopically measured toughness, K ∞c , is greater than TL and the apparent toughness increases. This is the essence of the transformation toughening effect. Many authors have shown how to increase the fracture toughness of alumina. Three decades ago Claussen et al.51 published the first works related to the high fracture toughness of composites in Al 2 O 3 -ZrO 2 system by transformation toughening. This type of material is generally characterized by a micrometric matrix and a second, nanometric phase, excepting the works of Bhaduri et al.52, Kim et al. 53 and Vasylkiv et al.56, which concern nano/nano-composites. Bhaduri et al.52 measured the toughness in order of 8.4 MPa.m1/2 in nano/nano Al 2 O 3 /ZrO 2 composites which higher compared with conventional micro/nano nanocomposites. This result is comparable with the micro/nano composites developed by De Aza et al.51 who measured 6 MPa.m1/2. Schehl et al.16 produced Al 2 O 3 –5 wt. % ZrO 2 composites by the colloidal processing route (see Section 3.2.2.5). They obtained a toughness 7.5 MPa.m1/2, significantly higher than that of monolithic alumina (3.5 MPa.m1/2 1). Zhan et al.54 produced Al 2 O 3 – 20 vol. % ZrO 2 nanocomposites by SPS with grain sizes of 265 nm for alumina and 96 nm for zirconia. The full dense composite showed a fracture toughness of 8.9 MPa.m1/2, significantly higher than that of pure nanocrystalline alumina (3.3 MPa.m1/2), densified by the same route. Vasylkiv et al.56 produced yttrium stabilized tetragonal ZrO 2 and Y-TZP/Al 2 O 3 , with an addition of 0.2 to 0.7 wt.% of alumina. The Y-TZP with 0.35 wt.% Al 2 O 3 content exhibited the maximum toughness of 15.7 MPa.m1/2. In the following Figure 3.35 the results obtained by the authors are reported. 130 Nanostructured composite materials: elaboration and properties Figure 3.35 – Fracture toughness of ZrO 2 /Al 2 O 3 as a function of the Yttria-stabilizer content56 and alumina weight percentage. As in Al 2 O 3 /SiC, Niihara et al.58 have developed a nanocomposite Al 2 O 3 /SiC/Y-TZP. The authors have observed that the incorporation of SiC limits the phase transformation of zirconia. The authors found an improvement in fracture toughness due to the thermal coefficient mismatch between the grains of SiC, Al 2 O 3 and ZrO 2 . Nawa et al.59 have developed Ce-TZP/Al 2 O 3 which possesses a high strength, preserving significant high toughness. They found an optimum result by adding 0.05 mol.% TiO 2 in 10Ce-TZP/30 vol% Al 2 O 3 composite. The researchers obtained a strength of 950 MPa and a toughness of 18.3 MPa.m1/2, which is a good compromise in terms of strength and toughness. 3.4.6. Crack Deflection The strong cohesion between the nano-particle and the matrix enables the crack to make a turn within the nano-particle and to enter the matrix grain. The subsequent transgranular fracture proceeds along the direction to maximize the mode I stress intensity factor. The toughening factor is defined by the following formula, as it was reviewed by Tan et al.60: GC nm m 1 (12) GC where G c nm and G c μm are the critical energy release rates of the nanocomposite ceramics. The comparison between the model and the experimental points, in which the volume fraction (Vf) of the second phase is varied, as it is shown in Figure 3.36. 131 Nanostructured composite materials: elaboration and properties Figure 3.36 – Toughening factor of the nanocomposite ceramics with respect to the volume fraction of the dispersed nanoparticles60. The energy release rate for the tilt crack to advance with an θ angle, can be expressed with the formula below: G 2 2 1 2 K I K II E (13) The occurrence of intergranular or transgranular relies on the relative values of Gθ/G C gb and Gθ/G C la, where G C gb and G C la are the fracture energy of the grain boundary and the fracture energy of the lattice with the absence of the nanoparticles, respectively. In case of mode I loading, it can be defined as a trajectory characteristic angle in the following way: 1/ 4 G gb O 2 arccos C la GC (14) Intergranular fracture occurs when O 0, O while transgranular fracture occurs when O 0, / 2 . The probability of occurring an intergranular fracture f is defined by: f 2 (15) Nano-particles along the grain boundaries may steer the crack into the matrix grains. In Figure 3.37 (A) it is shown the case when the second phase is absent. 132 Nanostructured composite materials: elaboration and properties Figure 3.37– Effect of nanoparticles on grain boundary60 The main crack extends intergranularly along the grain boundary, since the fracture resistance of the grain boundary is lower than that of the grain lattice. In Figure 3.37 (B), nano-particles scatter along the grain boundary. Figure 3.38 – Slice model of crack extension in nanocomposite ceramics60. In which G c ngb is the critical energy release rate for the debonding of a nano-particle. Supposing that the crack tip reaches the nano-particle, the angle formed between the main crack (along x 1 direction) and the matrix/nano-particle interface, is defined as θ. The energy release rate is G A for the crack to advance along x 1 direction, and G B for the crack to advance along the matrix/nanoparticle interface. The following expression describes the relationship between the energy release rate G A for the crack to advance along the x 1 direction and G B which is the energy required to advance along the matrix/nano particle interface. GB GA cos 4 / 2 (16) In this context, transgranular fracture could only occurs when: GA GB (17) nbc la GC GC requiring: GC ngb GC la cos 4 / 2 (18) where G C la is the lattice toughness of the second phase particles in case of crack entering the nanoparticles, and the lattice toughness of Al 2 O 3 in case of crack entering the matrix. 133 Nanostructured composite materials: elaboration and properties A crack encountering a particle-free grain boundary can extent transgranularly by a probability of 1 f . With the presence of a second phase, the probability raises to 1 f f .Vf . The toughening by mechanism induced by the second nanophase could be expressed as: GC la GC gb I fV f GC m (19) which indicates the direct proportionality with V f . Microstructural data showed that cracks within the matrix grains exhibit a wavy path, as it is shown in Figure 3.38. A transgranular crack with the macroscopic direction along x 1 is influenced by nanoparticles. In Figure 3.38 nanoparticles A, B, C and D perturbed the crack path χ (x 1 ) formed in the section π. In this case the lattice resistance to a wavy crack during the transgranular fracture is G C la’. Assuming that the surface energy is the same for a wavy or a flat crack, it is possible to write the following equation: GC la '/ GC la l '/ l (20) where l’ is the arc length of the zigzag crack χ(x 1 ), and l is the projected length of χ(x 1 ) on the x 1 direction. The ratio between the arc length and the projected length of the zigzag crack determines the toughening due to wavy fracture surface. 3.5. Creep Creep is a time-dependent permanent deformation that is often due to diffusion processes rather than dislocation motion. Under high temperature, the magnitude of the creep strain ε vs. time, t, critically depends on the applied stress, σ, and on the temperature, T, according to the general formula: f , T , t (21) Creep curves consist in a plot of strain vs. time; they can be usually separated into three stages: Transient or primary creep: following a spontaneous elastic strain, the creep rate, έ decreases with time; Steady-state or secondary stage: strain increases with time, the creep rate is constant and deformation may continue for a long time. Tertiary stage: a rapid increase in creep rate just before failure. It may be noted that the three stages are not independent because creep is a continuous phenomenon. Depending on the applied stress or the temperature, one stage of the creep may be missing. Under low stresses, the third stage may disappear. In contrast, under high stresses, the secondary stage may be replaced by an inflexion point. The typical creep examples are shown in Figure 3.39. 134 Nanostructured composite materials: elaboration and properties Figure 3.39 – Schematic illustration of creep curves expresses as strain vs. time at constant stress: (a) presence of two stages (1 to 2) at low temperature or low stress, (b) presence of the three creep stages (1 to 3), (c) at high temperature or high stress where the second stage is replaced by and inflexion point62. Generally, the analysis of creep results is only related to the “steady-state” stage. This avoids the problems in defining equation that quantify creep-curves. Most creep curves models predict a dependence of creep rate on the temperature. Most of creep models are based on the general Norton relationship61: . A1.D n . k .T d m (22) where A 1 is a material dependent constant, D the diffusion coefficient, k is Boltztman’s constant, T the absolute temperature, d the grain size, σ is the applied stress, and m and n are the grain size and stress exponents, respectively. The diffusion coefficient is given by: Q D Do .exp RT (23) where D O is a constant, R is the gas constant, and Q is the activation energy, m, n, and Q depend on the creep mechanisms acting during creep. In creep analysis, Q is determined by a plot of ln(έ) vs. 1/T. 3.5.1. Diffusion Creep Diffusion creep refers to the deformation of crystalline solids by the diffusion of vacancies through their crystal lattice. There is no dislocation motion; vacancies diffuse from the grain boundaries located nearly perpendicular to the tensile axis to those located parallel to the tensile axis. This process is named Nabarro-Herring creep63,64 (Figure 3.40 (A)) in which the steady state creep is given by: . D .Gb b . 9.3 l kT d 2 . G (24) 135 Nanostructured composite materials: elaboration and properties A second mechanism, proposed by Coble65, is due to vacancy flow occurring through the grain or the grain boundary (Figure 3.40 (B)). The Coble creep is given by the following expression: Gb b 33.4.Dgb . . . kT b d . 3 . G (25) In both equations, D l and D gb represent the diffusion coefficients in the lattice and at the grain boundary, b is the Burgers’ vector, G is the shear modulus, and δ represents the effective grain boundary thickness. Both models lead to a stress exponent n=1, although the grain size exponent differ (m=2 or 3, respectively). Cannon and Langdon66 found that Coble creep is prevalent when the grain size is small. Furthermore, Coble creep is favoured at low temperatures, because the activation energy for grain boundary is lower. Nabarro-Herring and Coble creep can take place in parallel so that actual creep rates will involve both components and both diffusion coefficients. Generally in ceramics, in the situation in which anions and cations are diffusing creates further complications 62,64. Both mechanisms are presented in the Figure 3.40. Figure 3.40 – Diffusion creep and GBS (Grain Boundary Sliding) models62. 3.5.2. Creep of particle reinforced composites Creep behaviour of nanocomposites, made by a micrometer-sized alumina matrix with nanosized reinforcing particles, has been studied by many researchers. The most remarkable advantage of ceramic nanocomposites lies in their high-temperature strength and the creep resistance. However, mechanical properties at room temperature -as it was reviewed- are also significantly improved. As Alumina/Silicon Carbide is extensively studied, it will be considered at first. Ohji et al.68 have investigated the tensile bending creep of and alumina/17 vol.% SiC at 1200°C and 1100°C in air. Nanocomposites and polycrystalline alumina were prepared by hot pressing of mixed powder with average particle sizes less than 100 nm. In both cases, the chosen sintering temperatures were similar, 1800°C for the nanocomposite and 1300°C for the monolith with the aim of obtaining a similar matrix grain size ≈ 2 μm. In the particular case of nanocomposites, SiC particles were located within and along the alumina grains. The typical tensile creep curves of samples are shown in Figure 3.41. 136 Nanostructured composite materials: elaboration and properties Figure 3.41 – Tensile creep behaviour of monolithic Al 2 O 3 (curve 1) and Al 2 O 3 /SiC nanocomposite (curve 2) at 50 MPa and 1473 K68. As it is shown in Figure 3.42, the curve of monolithic alumina (1) exhibited a lifetime around 150 h and creep stain of about 4 % at failure. On the contrary, the nanocomposite (2) had a lifetime of 1120 h (10 times longer) and a failure strain of 0.5 % (eight times longer). Under microstructural observation, the nanocomposite presented less microcracks and cavities. The creep rate of the nanocomposite was reduced in three to four order of magnitude (Figure 3.48). The stress exponent for the monolithic alumina was 2.2 in tension and 2.9 in bending, in good agreement with literature data (Canon and Langdon66). However, in case of the nanocomposites, it was 3.1 in tension and 2.2 in bending, respectively. Figure 3.42 – Stress dependence of the strain rate in tension (curves 1 and 1’) or flexure (curves 2 and 2’) of Al 2 O 3 and Al 2 O 3 /SiC67. A similar study was performed by Descamps et al.69, who investigated the differences of n values between a monolithic alumina and 5 vol.% SiC. The authors found a contrast in n value, being 1.68 for monolithic alumina and 3.68 for the nanocomposites. 137 Nanostructured composite materials: elaboration and properties Figure 3.43 – Creep rate vs. inverse temperature of a monolithic alumina (curve 1) and a nanocomposite (curve 2) at stress of 75 MPa69. The incorporation of SiC dispersed particles leads to an important increase in activation energy under bending stress of 75 MPa. In fact, the activation energy of the nanocomposite was 917 kJ mol-1 in contrast with 466 kJ mol-1 for monolithic alumina (Figure 3.43). A slightly different result was obtained by Thompson et al.70, for Al 2 O 3 /SiC nanocomposites characterized by a matrix average grain size of 2.2 μm and dispersed submicrometer SiC particles of 0.15 μm. During creep test, the nanocomposite exhibited an absence of both the primary and steady-state; only the tertiary stage was recorded (Figure 3.44). Figure 3.44– Strain rate vs. strain for an alumina/SiC nanocomposite at 1473 K at a tensile stress of 100 MPa70. Creep rate was decreased by two/three orders of magnitude, as compared with monolithic alumina with a similar grain size (Figure 3.45). In addition, the effect of SiC morphology was investigated: the particle and whisker-reinforced alumina composites showed a similar activation energy at 100 Mpa, of about ≈840 kJ mol-1. 138 Nanostructured composite materials: elaboration and properties Figure 3.45 –Creep rate vs. inverse temperature for monolithic Al 2 O 3 (curve 1) and Al 2 O 3 – 5 vol.% SiC/ Al 2 O 3 – SiC w nanocomposites (curve 2)70. In spite of these results, the role of the SiC particles is difficult to describe. Ohji et al.68 have noticed a rotation of intergranular particles that “plunge” into alumina matrix in association with grain boundaries sliding and the formation of small cavities around the SiC particles (Fig 3.46). The plunging indicates an increase of particle pinning at grain boundaries which is the responsible for obstructing the grain boundaries sliding phenomena. Figure 3.46 – Microstructure of Al 2 O 3 /17 vol.% silicon carbide nanocomposite after tensile creep at 1573 K and 50 MPa68. Ohji et al. in a second publication71, explained how SiC particles could alter the grain boundary chemistry. The silicon carbide interface is much stronger than the alumina-alumina interface, which contains a glassy phase. Thus, the improvement of the creep resistance may be attributed to the inhibition of vacancy creation and annihilation at the SiC-Al 2 O 3 interface due to the decreasing of the boundary diffusion rate. 139 Nanostructured composite materials: elaboration and properties The reduction of the boundary diffusion rate may promote a growth of isolated spherical cavities leading to a decay in cavity connection and resulting failure. Figure 3.47 – TEM micrograph of a nanocomposite after creep test, showing the presence of plastically deformed grains71. The augmentation of grain boundary cohesion by SiC particles inhibits grain boundary sliding and some grains are forced by the neighbours. Thus, a plastic deformation based on the dislocation motion is illustrated in Figure 3.47. As it was reviewed, the increase in n values, coincides with the creep mechanism involving dislocation glide or climb, thus supporting the assumption that dislocation motion plays a role in creep behaviour. Arzt and Grahle72 proposed that climbing grain-boundary acts as a source for vacancy diffusion processes and they presented a model based on the pinning of these dislocations by boundary-grain particles. The best results for creep resistance for Al 2 O 3 /SiC were obtained by Descamps et al.70 for 10 vol.% SiC. It seems that the addition of nanosized particles is the only condition that improves creep resistance, as it is also important to tailor the microstructure, by locating the particles at grain boundaries. 3.6. The Alumina-YAG system: elaboration and properties 3.6.1. Alumina-Yttria phase diagram The early attempts of Rhodes and Reid73 to sinter transparent lamp envelopes began with an attempt to sinter pure Y 2 O 3 . One deagglomeration procedure was achieved by ball-milling using the highest purity AI 2 O 3 spheres. The researchers found that samples reached higher densities. 140 Nanostructured composite materials: elaboration and properties Figure 3.48 – Y 2 O 3 -Al 2 O 3 phase diagram proposed by Noguchi and Mizuno74. Microprobe results confirmed that they had unintentionally "contaminated" the powder with an excellent sintering aid. An extensive program was launched to take advantage of this accidental finding. The most recent Y 2 O 3 -Al 2 O 3 phase diagram by Noguchi and Mizuno74 (Figure 3.48), shows a solid solution region up to 4 mole % Al 2 O 3 in Y 2 O 3 at 1940°C. This portion of the diagram is dashed to indicate some uncertainty. The initial work to use Al 2 O 3 deliberately as a sintering aid was based on the theory that it should be possible to sinter in the liquid plus Y 2 O 3 field and then drive the Al 2 O 3 back into solid solution by annealing at a lower temperature (1940°C). If this is successful, it would qualify as a transient liquid-phase sintering technique. An extensive research program was carried out with Al 2 O 3 additions from 0.02 to 10.0 mole % and sintering temperatures from 1800 to 2200°C in a H 2 atmosphere. Numerous attempts were made to incorporate three temperature holds. Samples were held at temperatures ranging from 1700 to 1900°C to allow diffusion enough time to form the solid solution; then sintered above the 1940°C eutectic over a range from 2000 to 2200°C and annealed for 1 to 8 h in the range of 1800 to 1940°C to drive the eutectic phase back into solution (Figure 3.48). The researchers did not corroborate after the sintering cycle, the formation of a solid solution. The only effect of the lower temperature held at the end of the cycle was to crystallize the grainboundary liquid phase. 141 Nanostructured composite materials: elaboration and properties Figure 3.49 – Translucent microstructure achieved with 0.5 mole % Al 2 O 3 74. In Figure 3.49 it is shown the microstructure after three-temperature sintering cycles in an attempt to minimize the second phase. The residual grain boundary phase is evident even though the sample's starting composition contained 0.5 mole % Al 2 O 3 . The effectiveness of liquid-phase sintering in pore elimination is also shown. Extensive studies on the effect of powder properties on sintered optical properties were discussed by Palilla et al.73 in conjunction with the development of this material as a lamp envelope. The lamp envelope development work demonstrated that the solubility limit may be on the order of 100 ppm, supporting the earlier Al 2 O 3 -Y 2 O 3 phase diagram of Toropov et al.75, which is shown in Figure 3.50. The Y 2 O 3 -rich end of the diagram shows no solubility for AI 2 0 3 . The cubic to hexagonal transition at ~ 2200°C for pure Y 2 O 3 is not accounted (Figure 3.50). Figure 3.50 – Y 2 O 3 -Al 2 O 3 phase diagram proposed by Toropov et al75. 142 Nanostructured composite materials: elaboration and properties It is still not clear which phases appear during cooling of Al 2 O 3 -Y 2 O 3 mixtures with compositions corresponding to YAG, YAlO 3 (YAP) or Y 4 Al 2 O 9 compounds. The pseudo-binary Al 2 O 3 -Y 2 O 3 system is shown in Figure 3.50 76. Caslavsky et al.76 performed optical differential thermal analysis and reported that it melts if heated to temperatures below 2263 K following the equilibrium phase diagram (Figure 3.51), while it melts cooled down from temperatures above 2263 K followed the metastable phase diagrams. However, in general, the selection of the solidified structure is determined not by the melting temperature before cooling but by a combination of nucleation and the growth velocity of the interface. Figure 3.51 –Al 2 O 3 -rich portion of the phase diagram of the Al 2 O 3 -Y 2 O 3 system76. The maximum melting temperature before cooling significantly affects nucleation of YAG. The melt heated up to temperatures above 2273 K never nucleates above 1973 K. Melts heated above 2273 K are difficult to solidify. Even when the melt is kept at 1993 K for 1.8 ks. solidification does not occur. Caslavsky et al. 76 reported a possibility of two immiscible liquids above 2273 K, because the melt was opaque for the YAG composition. If a miscibility gap exists above 2273 K and the melt is kept above this temperature, coarsening should occur, but this was not the case in Caslavsky’s experiments. Mituzani et al.68 have developed equipment for optical differential thermal analysis which has been used to measure the differential thermal analysis curves of Al 2 O 3 -18.5mol% Y 2 O 3 . However, no exothermic or endothermic heat was detected during heating up to 2473 K. Coordination of oxygen around aluminium may have an important role in the melt68. Fourfold coordination exists in the YAG structure, while only Al-O octahedra exists in the α-Al 2 O 3 and YAP structures. Coordination in the melt may change during heating and affect the nucleation behaviour. For example, the Al 2 O 3 /YAG system appears even when specimens solidified as the metastable Al 2 O 3 /YAP system are remelted and slowly cooled. As it was described above, alkoxide and metalorganic derived precursors have been used to produce oxide phases in the Al 2 O 3 –Y 2 O 3 system. Compositions aimed at garnet frequently produce YAG as the first product of crystallisation 77-78, though occasionally an intermediate hexagonal YAlO 3 served as the crystalline precursor to YAG81-83. These results point out the need to preserve molecular homogeneity during hydrolysis (gelation) or during decomposition of the gel or metalorganic precursor. 143 Nanostructured composite materials: elaboration and properties In situations where segregated phases appear (a.i., Y 2 O 3 , Y 4 Al 2 O 9 , YAlO 3 ) a subsequent reaction is needed at higher temperatures to produce garnet73, except in one instance where in YAG was formed using separate Y and Al precursors in an autoclave at only 300°C80. In Figure 3.52, it is presented the most complete phase diagram found in literature, published by Roth et al85. Figure 3.52 –Al 2 O 3 -Y 2 O 3 system phase diagram with the investigated positions85. 3.6.2. Mechanical properties at room temperature and high temperature Al 2 O 3 -YAG composites are deeply investigated, thanks to the high creep resistance of YAG, that makes such composites suitable for high temperature applications. In addition, the two phases have mutual insolubility, similar thermal expansion coefficients and chemical stability86. Figure 3.53 – Schematic procedure for producing directionally solidified Al 2 O 3 -YAG eutectic composites88. 144 Nanostructured composite materials: elaboration and properties Mah et al.87 studied the processing and the mechanical properties of Al 2 O 3 /YAG eutectic composites prepared by the directionally solidified method, as it is shown in Figure 3.53. The authors found that composites had flexural strength of 373 MPa and a fracture toughness of 4.3 MPa.m1/2 at room temperature. In comparison with monolithic Al 2 O 3 or YAG, the composite has a significantly higher fracture toughness at elevated temperature. Waku et al.88 prepared Al 2 O 3 -YAG nanocomposites by hot-pressing at 1973°C under a pressure of 50 MPa. The nanocomposite has a important flexural strength of 450 MPa, but its strength felt drastically above 1073°C. Vrolijk et al.89 reported a comparison between a monocrystalline and polycrystalline Al 2 O 3 /YAG samples prepared by hot-pressing (see Figure 3.54) showing their mechanical behaviour at a lowand high-temperature regime. It is clearly shown the advantages of producing Al 2 O 3 /YAG eutectic composites, as the flexural strength is maintained almost constant up to the melting point. Figure 3.54 – Flexural Strength of Al 2 O 3 /YAG eutectic composites: comparison between the unidirectionally solidified and the polycrystalline samples89. The production of eutectic composites is quite complex; so, the elaboration of fine-grained polycrystalline Al 2 O 3 -YAG may be a useful solution. The discrepancy regarding the flexural strength was explained by the authors, as the difference of interface between both phases. The polycrystalline exhibits an amorphous layer detected by means of TEM (Figure 3.55 a), while the monocrystalline does not have it (Figure 3.55 b). Although, the reports concerning the mechanical properties of polycrystalline Al 2 O 3 /YAG composites are still very few. Figure 3.55 – High resolution TEM image of the grain boundary between the Al 2 O 3 and YAG: (a) polycrystalline and (b) monocrystalline samples89. Li et al.90 prepared Al 2 O 3 /25 vol.% YAG euthectic composite using the co-precipitation method. The samples were sintered by hot pressing up to 1400°C for 1 h and at 30 MPa. The room 145 Nanostructured composite materials: elaboration and properties temperature fracture strength and fracture toughness of the composites were 611 MPa and 4.53 MPa.m1/2, respectively, thanks to the reduction of the grain size of the matrix and the good dispersion of the YAG phase in both inter and intra-granular positions. Figure 3.56 – Temperature dependence of fracture strength for monolithic alumina and YAG systems91. French et al.91 described the variation of fracture toughness by increasing the temperature, as it is shown in Figure 3.56. The authors mainly compared an Al 2 O 3 /50 vol.% YAG with the single-phase constituents. The nanocomposite exhibited a lower decay of toughness among the others, confirming the benefit of producing this type of materials. Schehl et al.16 prepared Al 2 O 3 / YAG polycrystalline composites by colloidal processing (see Section 3.2.2.5). Materials containing YAG showed an important K IC value of 5.8 MPa.m1/2 compared with the monolithic alumina (4.5 MPa.m1/2). The increase in K IC was attributed to the reaction between YAP and Al 2 O 3 to give YAG which consequently gave rise to a volume increase. This phenomena creates a homogeneous stress field at alumina grain boundaries which is the main responsible of blocking the flaw formation during sintering. The behaviour of yttrium-doped alumina in the high-temperature regime has been a subject of several studies. From these results, it was clarified the beneficial effect of YAG on high temperature deformation, as the creep rate could be lower than two order of magnitude compared to monolithic alumina. The phenomenon was attributed to the segregation of Y3+ at Al 2 O 3 grain boundaries, which blocks the diffusion of ions along the grains boundaries, with a consequent reduction in grain-boundary diffusion and thus decreasing the creep rate. If creep rate is controlled by point defects on their transport along grain boundaries, the strong segregation of Y3+ at grain boundaries is likely to hinder this process. Duong et al.92 investigated the creep behaviour of fine-grained two-phase Al 2 O 3 -YAG over a temperature range 1400°C-1500°C under applied stress between 3 and 20 MPa. The authors prepared two composites. Al 2 O 3 /50% vol.% YAG and Al 2 O 3 /75% vol.% YAG nanocomposites, labelled as AY50 and AY75 respectively, which were sintered in air at 1600°C for up to 120 h. In both cases, microstructure was composed by an average matrix grain size of about 8-10 μm, while YAG had an average grain size of 3 μm in both samples. Figure 3.57 shows the creep rate at 1400°C for composites compared with the single constituents. The Al 2 O 3 /50% vol.% YAG has the lowest strain rate among all the samples. The authors attributed the difference in creep behaviour since AY75 has the finer microstructure among the other materials involved. 146 Nanostructured composite materials: elaboration and properties Figure 3.57 - Stress dependence of the strain rate in tension for Al 2 O 3 and YAG composites at 1400°C92. Other authors, such as French et al.93, have studied the creep behaviour of duplex microstructures. In this particular work, the nanocomposites were produced by mechanical mixture (ball milling) and solid state reaction of the single constituents, finally sintering in air at 1650°C for 2 h. The final microstructure was characterized by an average grain size of ≈ 2 μm. Three extra samples were prepared to be compared with the composite; the Al 2 O 3 , YAG and Al 2 O 3 doped with 100 ppm of yttrium. In Figure 3.58, the creep rate plots for different systems are exposed for an applied stress of 75 MPa. French and co-workers arrived to the conclusion that with only 1000 ppm of Y3+ it is possible to decrease the creep rate by two orders of magnitude. The activation energy for this 1000 ppm yttrium-doped Al 2 O 3 was 698 kJ/mol. Figure 3.58 - Creep rate vs. (a) applied stress and (b) temperature for the single constituents Al 2 O 3 /YAG, Al 2 O 3 doped with Y+3 (1000 ppm) and AY50 nanocomposite91. The most relevant found in literature values regarding to Al 2 O 3 /YAG nanocomposites are summarized in Table I. 147 Nanostructured composite materials: elaboration and properties % vol. YAG σ (MPa) T (°C) n Reference 1500-1600 1100-1400 1400-1500 Grain Size (μm) 2 2 8-10 0 50 50 75 5 ≈5 100 10-152 35-75 10 1 2.6 1.1 94 40-80 30-250 20 1400 1200-1400 1400-1600 ≈1.6 3.6 3-6.3 1.2 1-1.46 1.2 95 93 92 96 97 Table I – Summary of creep tests data for Al 2 O 3 and YAG systems. Schehl et al.16 as it was mentioned before, sintered an alumina-YAG composite in air at 1600°C for 2 h. For this particular nanocomposite, the final microstructure was formed by alumina matrix grains with an average grain size of ≈4 μm, and YAG grain in both inter-intragranular positions of 350 nm. The results obtained by Schehl et al.16 confirmed the positive effect of the dopant. The values obtained by the authors for the nanocomposite and monolithic alumina are summarized in Table II. In order to compare the strain rates, the creep rates measured were normalised with the smallest grain size employing the following formula6. p d . n A . S d AY . (26) . where n is the measured creep rate for Y-TM and Y-CR, d A is the mean grain size of the monolithic alumina and d AY is the mean grain size of the composite. For the calculation of the normalised creep rate, the authors used grain size exponent (p) equal to 3 for grain boundary diffusion controlled creep and 2 for lattice diffusion controlled creep. Sample L (μm) A 5.9 AY 3.56 Strain Rate (s-1) 5.7.10-9 2.0.10-9 Strain Rate for p=2 (s-1) 7.5.10-8 9.9.10-9 Strain Rate for p=3 (s-1) 2.7.10-7 2.2.10-8 Table II – Creep test data obtained by Schehl et al.12 By using p=2 the creep rates have approximately one order of magnitude lower than monolithic alumina. However, employing the p=3 the value is lowered 13 times. Other authors as Satapathy and Chokshi95 produced Al 2 O 3 /YAG nanocomposites by solid-state reaction of alumina/yttria powders during sintering at 1500°C for 6 h. The authors found a similar behaviour as Schehl et al.16 concluding that creep mechanism is controlled the grain boundary diffusion65. However, for normalised creep rates using p=3, the achieved values differ from the previous work, obtaining a decrease of only a 3 times at 1400°C. The authors concluded that over the chosen temperature, the YAG phase does not influence significantly the creep mechanism. The results obtained by the authors are summarized in Figure 3.59. 148 Nanostructured composite materials: elaboration and properties Figure 3.59 - Creep rates and Al 2 O 3 -5 vol.% YAG 95. vs. (a) applied stress and (b) temperature Al 2 O 3 The main goal of a second work, from Schehl and co-workers96 was to extend the information regarding alumina-YAG nanocomposite. The authors arrived to the conclusion that the creep mechanism is lattice diffusion (Nabarro-Herring) by measuring a stress component of nearly 1 at 1200-1300°C. In Figure 3.60 it is shown the comparison of creep rates for different applied stresses between the alumina-YAG composite and the monolithic alumina. Figure 3.60 - Creep rate vs. applied stress of monolithic Al 2 O 3 and Al 2 O 3 ≈5 vol.% YAG96. Activation energies calculated for monolithic alumina were 517 kJ/mol for an applied stress of 50 MPa and 527 kJ/mol 70 MPa, in good agreement with the literature93. However, in the case of the AY nanocomposites the activation energies were 554 kJ/mol for an applied stress of 70 MPa and 595 kJ/mol for an applied stress of 100 MPa, respectively. The normalised creep rates with p=2 are shown in Figure 3.61. The authors explained that YAG particles at grain boundaries avoid the grain boundary sliding during deformation. Due to the high activation energy required to deform YAG nanosized particles, 149 Nanostructured composite materials: elaboration and properties the activation energy associated with the grain boundary diffusion of Al3+ rises to values close to polycrystalline YAG. Figure 3.61 – Normalised creep rates of Al 2 O 3 and Al 2 O 3 -5 vol.% YAG96. A significant improvement of strain rate of about 1 order of magnitude was found by the authors in 1200-1400°C range for AY nanocomposites. 150 CHAPTER 4 Elaboration and Characterization of Al 2 O3 -5 vol.% YAG Nanocomposites 151 Elaboration and Characterization of Al2O3-5 vol.% YAG Nanocomposites 4. Introduction Two commercial α-alumina nanopowders were employed to prepare Al 2 O 3 -5 vol. % YAG nanocomposites. Powders have been doped by using aqueous solutions of yttrium salts and subsequently by a solid state reaction is yielded the YAG phase. The aim of this work is to investigate the effect of different densification routes and forming methods on the microstructural and mechanical properties of Al 2 O 3 -5 vol. % YAG nanocomposites. As it was reviewed in chapter III, a second phase is believed to reinforce the Al 2 O 3 by introducing nano-size dispersoids within matrix grains and the grain boundaries. Nanoparticles are the responsible of modifying the crack propagation behaviour due to the difference in thermal expansion coefficients1,2. YAG is believed to be a suitable candidate as its creep resistance, and its thermal expansion and stability in contact up to 1700°C3. Many researchers explained this improvement as cation ions (Y3+) segregated to alumina grains boundaries reduce the grain boundary diffusivity, thus decreasing the creep rate5-7. Torrecillas et al.7 produced Alumina-YAG composites by the colloidal processing route, obtaining a significant improvement of nearly 2.5 times in creep rate in a range of 1200-1400°C. Recently, Palmero et al.4 successfully exploited the post doping of commercial power by using YCl 3 to prepare Al 2 O 3 -5vol.% YAG. The materials developed were submitted to mechanical characterization correlating the microstructural features to the mechanical properties. 4.1. Characterization of the as-received powders Two commercial α-alumina powders were employed for the production of Al 2 O 3 -5 vol.% YAG composites. TM-DAR TAIMICRON, supplied by Taimei Chemical Co., Japan, is made of pure alpha alumina. It is characterized by a theoretical density of 3.96 g/cm3, a specific surface of 4.5 m2/g and a particle mean size of 350 nm. TM is produced by thermal decomposition of aluminum salt and ammonium bicarbonate as it is claimed by the producer8. XRD analysis performed on the starting material confirmed that only α-alumina (ICDD 81-2266) is present, as it is presented in Figure 4.1. Intensity (a.u.) 10 20 30 40 50 60 70 Diffraction Angle (2.) Figure 4.1 – XRD pattern of the as-received TM-DAR TAIMICRON 152 Elaboration and Characterization of Al2O3-5 vol.% YAG Nanocomposites Similarly, as for Nanotek the initial granulometry of TM-DAR TAIMICRON was evaluated by laser granulometry after dispersing the powder in distilled water in obtain a dilution factor of 7 vol. %. The cumulative distribution by volume and number are presented in Figure 4.2. Cumulative frequency (%) 100 80 60 40 20 0 0 5 10 15 20 25 Agglomerate Size [m] Figure 4.2– Cumulative size distribution by volume (solid line) and by number (dashed line) as a function of agglomerate size of as-received TM-DAR TAIMICRON Laser granulometry lead to determine agglomerates size of 8.6, 29.5 and 66.7 μm, corresponding to (d 10 ) 10, (d 50 ) 50 and (d 90 ) 90% of the cumulative volume distribution. In case of the number distribution, the agglomerates sizes are <0.3, 0.36 and 0.61 μm corresponding to (d 10 ) 10, (d 50 ) 50 and (d 90 ) 90 %, respectively. SEM micrograph of the as-received TAIMEI powder provided by the supplier is shown in Figure 4.3 which confirms that the powder is sub-micrometric. Figure 4.3 – SEM micrograph of TM-DAR TAIMICRON8. The second powder employed in this study the CR1 alumina powder, labelled in this study CR, supplied by Baikowski, France9. Powder CR is characterized by a higher average particle size of about 0.6 μm and a lower specific surface area of 3 m2/g. In the following Figure 4.4, is shown the morphology of the CR1. 153 Elaboration and Characterization of Al2O3-5 vol.% YAG Nanocomposites Figure 4.4- SEM micrograph of as-received CR19. As supplied CR powder was characterized by laser granulometry, yielding agglomerate size of 0.77, 2.36 and 4.57 μm, corresponding to (d 10 ) 10, (d 50 ) 50 and (d 90 ) 90 % of the cumulative volume distribution. In case of the number distribution, the agglomerates sizes are <0.3, 0.4 and 1 μm corresponding to (d 10 ) 10, (d 50 ) 50 and (d 90 ) 90 %, respectively. The results are shown in Figure 4.5. Cumulative frequency (%) 100 80 60 40 20 0 0 10 20 30 40 50 60 70 80 90 100 Agglomerate size [m] Figure 4.5- Cumulative size distribution by volume (solid line) and by number (dashed line) as a function of agglomerate size of as-received CR. 4.2. Elaboration of Al 2 O 3 - 5vol.% YAG composites 4.2.1. Dispersion and Spray Drying The powders were dispersed in distilled water and ball milled for 3 h and 113 h for TM-DAR and CR1 powders, respectively. 154 Elaboration and Characterization of Al2O3-5 vol.% YAG Nanocomposites After dispersion, a significant reduction of the starting agglomerate size was achieved. In the particular case of TM, the agglomerate sizes were <0.3, 0.53 and 1.12 μm corresponding to (d 10 ) 10, (d 50 ) 50 and (d 90 ) 90 % of the cumulative volume distribution. For the second powder CR, after ball milling due to the agglomerates in a larger extent, the values decreased to 0.44, 0.81 and 1.37 μm. In Figures 4.6, the cumulative size distribution by volume of the as-received and dispersed powders are collected. Cumulative frequency (%) 100 80 60 40 20 0 0 10 20 30 40 50 Agglomerate size [m] a) Cumulative frequency (%) 100 80 60 40 20 0 b) 0 10 20 30 40 50 60 70 80 90 100 Agglomerate size [m] Figures 4.6- Cumulative size distribution by volume of the as-received (solid line) and dispersed (squares) TM-DAR (a) and CR1 (b) powders. Thereafter, an aqueous solution of YCl 3 .6H 2 O (Sigma-Aldrich, 99.99%) was added and dispersed in both slurries. Suspensions were maintained under magnetic stirring for 2 h. The slurries were diluted to 4 wt. % and spray-dried. 4.2.2. Thermal Evolution of the powders Doped powders were calcined at different temperatures and submitted to X-ray diffraction. The tests were performed by plunging the powders into a tubular furnace kept at fixed temperature for three minutes. This technique was selected in order to avoid the crystallite growth and the aggregates formation. By this method, it was also prevent the possible elution of yttrium dopant during the slip casting process. 155 Elaboration and Characterization of Al2O3-5 vol.% YAG Nanocomposites The first sample analyzed was doped-TM. As it is presented in Figure 4.7, Yttrium aluminates start to crystallize at 1050°C, with the formation of YAP phase (as a trace), whose intensity slightly increased by increasing the calcination temperature. Starting from 1220°C, also YAG phase, near YAP, was detected. Consequently, in case of doped-TM by increasing the calcination temperature up to 1250°C it is possible to assign the first peaks corresponding to the crystallization of YAG at 1220°C. In this particular case, the sample was labelled as Y-TM. P Y Y Y Y Y P P P P P 1250°C counts (a.u.) 1220°C 1200°C 1150°C 1100°C 1050°C 25 30 35 40 45 50 2 Theta (degrees) Figure 4.7- XRD pattern evolution of Y-TM in the range 22-50° 2θ After high-temperature calcination of Y-TM at 1500°C for 3h, it is possible to obtain a composite composed only by well-crystallized YAG and Al 2 O 3 as it is shown in Figure 4.8. 12000 Intensity (a.u.) 10000 8000 6000 4000 Y 2000 Y Y YY 20 30 Y Y Y Y Y 0 10 40 50 60 70 2 Theta (degrees) Figure 4.8- XRD pattern of Y-TM calcined at 1500°C for 3 h (α= α-Al 2 O 3 , Y= YAG) 156 Elaboration and Characterization of Al2O3-5 vol.% YAG Nanocomposites By “flash” calcining the doped-CR at 1050°C for 3 min, it is possible to observe pure YAG (JCPDF N°33-0040) crystallized near α-Al 2 O 3 . As in the previous case, the treated doped-CR was labelled as Y-CR. In Figure 4.9, the XRD comparison between Y-CR 1050°C and Y-TM 1150°C is presented in the 22-50° (2θ) range. counts (a.u.) Y Y Y Y Y Y Y-CR 1050°C P Y-TM 1150°C 25 30 35 40 45 50 2 Theta (degrees) Figure 4.9- XRD patterns in the 22-50° 2θ of Y-CR “flash” heated at 1050°C for 3 min and Y-TM “flash” heated at 1150°C for 3 min (α= α-Al 2 O 3 , Y= YAG, P = orthorhombic YAlO 3 ) As a conclusion, Y-TM and Y-CR powders differ in crystallization behaviour. In fact, Y-TM “flash” heated at 1150°C for 3 min shows traces of orthorhombic perovskite YAlO 3 (JCPDF N° 70-1677, YAP) and α-Al 2 O 3 (JCPDF N° 46-1212). On the ground of the results obtained, “flash” calcination temperatures were 1150°C for TM-DAR and 1050°C for CR1. 4.3. Forming and Sintering Powder suspension was prepared with a solid loading of 50 wt. % and ball-milled for 24 h. Subsequently, suspensions were cast into slip-casting moulds (SC) - made of Plexiglas which lay on a porous alumina plate - (Figure 4.10) or dry in oven before uniaxial pressing (P). Subsequently, Y-TM and Y-CR were sintered by natural sintering (NS), hot-pressing (HP) and spark plasma sintering (SPS). Figure 4.10 - Slip-casting moulds used to produce the samples. 157 Elaboration and Characterization of Al2O3-5 vol.% YAG Nanocomposites For a sake of clarity, the complete elaboration process is schematically shown in the flow-chart of Figure 4.11. TM-DAR / CR1 dispersed slurries YCl 3 6H 2 0 aqueous solution Doped suspensions Homogenization for 2 h Spray drying Fast calcination at 1050°C (TM-DAR) and 1150°C (CR1) Y-TM and Y-CR powders Uniaxial pressing. Y-TM-P Y-CR-P Slip casting Y-TM-SC Y-CR-SC Natural sintering: NS Hot-Pressing: HP Spark Plasma Sintering: SPS Figure 4.11 - flow-chart of the elaboration process with samples designation as a function of the raw alumina powders and forming methods. Designation for the sintering routes are also reported. In the particular case of NS, it was performed up to 1500°C for 3 h (heating rate of 10°C/min to 1100°C and then 2°C/min up to 1500°C). Besides, HP samples were sintered in pellets of 20 or 40 mm, by heating up to 1450°C for 1 h (heating rate of 10°C/min up to 1100°C and 2°C/min up to 1450°C with a holding time of 1 h) with an applied pressure of 80 MPa. In parallel, SPS samples were sintered by SPS heating up to 1350°C for Y-TM and 1450°C for YCR under an applied pressure of 75 MPa. Heating rates and soaking time in both cases were 154°C/min and 3 min, respectively. This process was followed by a second dwell time at 1134°C (cooling rate of 617°C/min) for 5 min. 4.3.1. Sintering behaviour followed by dilatometric analysis Samples Y-TM and Y-CR were uniaxially pressed in bars at 350 MPa and subsequently submitted to dilatometric analysis, performed up to 1500°C for 3 h (heating rate of 10°C/min to 1100°C and then 2°C/min up to 1500°C). The green bodies reached a density of 2.29 g/cm3 for the pressed Y-TM and 2.25 g/cm3 for the pressed Y-CR, respectively. The green density was calculated from weight and geometrical measurements. 158 Elaboration and Characterization of Al2O3-5 vol.% YAG Nanocomposites In Figures 4.12 the comparison of dilatometric curves of Y-TM and Y-CR are presented. Derivative signal A 1335°C LL 0.05 0 200 400 600 800 1000 1200 1400 1600 Temperature [°C] Derivative signal B 1465°C LL 0.05 0 200 400 600 800 1000 1200 1400 1600 Temperature [°C] Figures 4.12- Dilatometric (solid line) and derivative (dashed line) curves: (A) Y-TM-P and (B) Y-CR-P. For the samples linear shrinkages of 16.9% and 16.4% were recorded for Y-TM and Y-CR. Derivatives curves shows the maximum sintering rate at 1335°C and 1465°C for Y-TM and Y-CR, respectively. After sintering, Y-TM reached almost theoretical density (99.7% TD); while a poor value was still achieved by Y-CR (94.4% TD). 4.4. Microstructural Characterization Sintered samples were submitted to microstructural characterization, by using SEM (Hitachi S2300) and ESEM (FEI XL30 ESEM FEG) microscopy. Microstructural observation was performed on polished surfaces. In order to compare the microstructural features of the composite materials to those of pure alumina samples, TM and CR slip cast bodies were also prepared and pressureless sintered. Their 159 Elaboration and Characterization of Al2O3-5 vol.% YAG Nanocomposites fired microstructures are presented in Figure 4.13: as shown, the two materials presented important differences regarding the grain size. Sample TM, as expected, shows a dense and homogeneous microstructure -with some intergranular porosity- composed by a homogeneous microstructure with and average size of 1.43 μm. On the other hand, CR is characterized by a equiaxed microstructure with a average grain size of 3.05 μm and some intergranular porosity, as well as, intra-granular porosity. This fact was expected as the final density was lower (97.4 %) compared with TM (99.4 %). b) Figures 4.13- Micrographs at different amplifications of polished surfaces: (a) TM-SC and (b) CR-SC sintered by pressureless sintering at 1500°C for 3 h. The microstructural observation performed on Y-TM-P and Y-TM-SC samples (Figures 4.14) revealed that the matrix was composed by micronic alumina grains of about ≈0.8 μm, having both equiaxed and elongated shape. The second phase was mainly located at grain boundaries or triples joints with few intra-granular particles distributed throughout the alumina matrix. These composites exhibited a YAG grain size of about 350 nm and 450 nm for Y-TM-SC and Y-TM-P, respectively. As it is observed in Figures 4.14, the second phase (YAG) is the responsible of avoiding the significant matrix grain growth by the grain boundary pinning effect (see section 3.4.1 – Chapter III). a) b) Figures 4.14- ESEM micrographs of polished surfaces: (a) Y-TM-SC (SE-SEM image) and (b) YTM-P (BSE-ESEM images) pressureless sintered at 1500°C for 3 h. 160 Elaboration and Characterization of Al2O3-5 vol.% YAG Nanocomposites In the following Figures 4.15 are presented the microstructures corresponding to Y-CR-SC and Y-CR-P. As in the previous case, the microstructure of naturally sintered Y-CR samples are formed by equiaxial alumina grains with an average grain size of 0.98 μm and YAG particles within the αalumina grains. In these particular cases, YAG mean size were about 650 nm and 400 nm in case of the slip casted and the pressed samples, respectively. The second phase occupy – as in the case of Y-TM – the grain boundaries or triple points positions. The analysis on polished surfaces revealed the difficulty to corroborate the existence of intra-granular grains. As a comparison, between forming methods, it was only found a slight variation regarding the YAG average grain size between the samples. a) b) Figures 4.15- ESEM micrographs of polished surfaces: (a) Y-CR-SC and (b) Y-CR-P pressureless sintered at 1500°C for 3 h (GSE-ESEM images). Samples Y-CR (Figure 4.15) show higher residual porosity compared with the Y-TM samples. Specially, in sample Y-CR-P it was observed an important inter-granular porosity, as it is shown in Figure 4.15 b. Samples Y-TM and Y-CR pressed/slip-casted were sintered by non-conventional techniques, namely hot-pressing (HP) and Spark Plasma Sintering (SPS) as mentioned before. By using these techniques it was possible to decrease the sintering temperature in the range 1350-1450°C yielding almost theoretical densities. a) b) Figures 4.16- ESEM micrographs of polished surfaces: (a) Y-TM sintered by HP and (b) Y-TM sintered by SPS (GSE-ESEM images). 161 Elaboration and Characterization of Al2O3-5 vol.% YAG Nanocomposites Hot pressing was carried out at the same temperature 1450°C for both Y-TM and Y-CR. However, SPS was performed at different temperatures: 1350°C for Y-TM and 1450°C Y-CR. No significant differences were found between the forming methods in terms of final density and microstructural features as seen in Y-TM NS samples. Only in the slip-casted Y-TM HP a slight improvement in final density of ≈1% was found, with respect to the pressed sample. Figures 4.16 show the micrographs corresponding to Y-TM-P sintered by HP and SPS. Microstructures were finer in both cases compared with pressureless sintered samples with low inter-granular porosity. In this particular case, SPS leads to achieve almost theoretical density (99.4 +/- 0.1%) with a lower sintering temperature and a shorter soaking time. The measured average grain sizes were 650 nm and 430 nm for the alumina and YAG in HP. Whereas, in case of SPS the average grain sizes were 510 nm and 265 nm for the alumina and YAG, respectively. It was difficult to determine accurately by the microstructural observation the position of the YAG particles. Although, YAG particles seem to be located at the alumina matrix, mostly at inter-granular positions (Figure 4.16). a) b) Figures 4.17- ESEM micrographs of polished surfaces: (a) Y-CR sintered by HP and (b) Y-CR sintered by SPS (GSE-ESEM images). Y-CR samples sintered by HP (Figure 4.17 a) show a matrix formed by alumina grains with an average grain size of 0.8 μm and a better distribution of the YAG phase – if compared with Y-CR SPS (Figure 4.17 b) - of about 610 μm located at inter-granular positions. As a comparison, Y-CR sintered by SPS (Figure 4.17 b) exhibits a less fine alumina since the average grain size 1.13 μm. YAG grains present an average grain size of 570 nm. As in the previous case, the second phase was located at inter-granular position. The higher matrix grain size in Y-CR samples sintered by SPS and HP –as expected- is ascribable to the higher mean particle size of CR compared with TM powder. Moreover, YAG average grain size also was slightly higher in Y-CR samples specially among SPS samples. The obtained Y-CR samples have a microstructure comparable to the nanocomposites produced by Gao et al.10. The researchers produced 90 vol.% alumina-10 vol.% YAG sintered by SPS at 1300°C in which YAG phase (500 nm) was homogeneously distributed in the micronic alumina matrix. 162 Elaboration and Characterization of Al2O3-5 vol.% YAG Nanocomposites 4.5. Mechanical characterization 4.5.1. Mechanical properties at room temperature Fired samples were polished up to 0.1 μm–finish. Vickers hardness measurements (HV) was determined by Vickers Testwell FV-700 with a load of 98.1 N (10 kg) held for 10 sec. Hardness was calculated over an average of five measurements. Fired densities were determined by the Archimedes’s method. As theoretical density it was used a value of 3.96 g/cm3 for α-alumina and 4.55 g/cm3 for YAG, resulting a theoretical density for the composite alumina-5 vol. % YAG calculated by the rule of mixtures equal to 3.99 g/cm3. In Table I are exposed the values of hardness and toughness of all the fired specimens. Table I: Fired density (%TD) , Vickers Hardness (GPa) and K IC (MPam1/2) values for slip cast (SC) and pressed (P) Y-TM bodies densified by pressureless sintering (NS), hot-pressing (HP) and spark plasma sintering (SPS) Y-TM Sintering Sample route Fired Density (%TD) Hardness (GPa) K IC (MPa.m1/2) NS-1500 SC 99.7 18.8 +/- 1.13 4.42 NS-1500 P 98.2 16.3 +/- 1.84 5.11 HP-1450 SC 99.8 19.8 +/- 2.94 6.95 HP-1450 P 98.5 18.0 +/- 0.88 6.21 SPS-1350 SC 99.5 19.1 +/- 0.52 5.57 SPS-1350 P 99.4 19.9 +/- 0.5 5.82 Fired Density (%TD) Hardness (GPa) K IC (MPa.m1/2) Y-CR Sintering Sample route NS-1500 SC 94.4 18.7 +/- 0.49 4.7 NS-1500 P 95.6 19.1 +/- 1.73 7.12 HP-1450 SC 99.1 21.4 +/- 1.71 6.03 HP-1450 P 98.9 19.7 +/- 1.10 6.37 SPS-1450 SC 99.7 18.8 +/- 0.81 7.12 SPS-1450 P 99.7 19 +/- 0.64 7.32 As it is shown in Table I, hardness values vary depending the starting material, the forming method and the consolidation method. The Vickers hardness values measured in Y-TM samples range from 16.3 to 19.9 GPa. These values are a consequence of the combination of the effect of average grain size and the final 163 Elaboration and Characterization of Al2O3-5 vol.% YAG Nanocomposites densities on the hardness. The influence of the grain size on the hardness has been discussed in the Chapter III – Section 3.4.211,12. For instance, naturally sintered Y-TM samples were strongly influenced by the final density as no significant differences were found concerning the grain sizes. The same phenomenon was observed in Y-TM HP. Samples Y-TM sintered by SPS presented similar hardness values as microstructure and density were similar in both samples. In Y-CR samples were measured hardness values in an interval of 18.7-21.4 GPa. Similar values were measured in NS samples taking in consideration the standard deviation (+/- 1.73 GPa) in the case of Y-CR-P NS-1500. In Y-CR HP samples were measured the highest hardness value since they present the smallest grain size (≈0.8 μm) among all the samples. In particular, in HP slipcasted sample it was measured the highest value 21.4 GPa since its density was slightly higher if compared with the pressed sample. The lowest value of Y-TM-P NS sample is ascribable to some defects produced during the forming process not visible in the previous microstructural observation. Finally, Y-CR samples sintered by SPS presented as in the previous case similar characteristics with respect to density and final microstructure. The values are satisfactory if compared with the values published by Wang et al.13, who produce high density alumina-25 vol.% YAG composite, made by sub-micronic alumina matrix with a homogeneous distribution of fine YAG grains (100-600 nm). For this study the authors obtained a hardness of 16.14 GPa. Young modulus was determined by the impulse excitation technique. The impulse excitation technique (Grindosonic method) allows to assess the elastic modulus of a bar-shaped sample at ambient temperature, necessary to obtain the fracture toughness values. For the different modes of resonance, the specimen is supported as it is illustrated in the Figure 4.18. The vibration of the bar is recorded by a microphone or a transducer. The fundamental frequency of the bar is determinate by a frequency analyzer. With the aim to corroborate the Young modulus, Grindosonic method was conducted based on the rules C 1259-01 from ASTM15, employing bar-shape samples of 40 mm diameter pellets. Figure 4.18- Rectangular specimen for Out-of-Plane Flexure7. mf 2 L3 E 0.9465 f 3 .T1 b t where T1 1 6.585 1 0.0752 0.8109 2 t / L 0.868 t / L 2 (1) 4 4 8.340 1 0.2023 2.173 2 t / L 2 1.000 6.338 1 0.1408 1.536 2 t / L where: E= Young’s modulus, Pa ; 164 Elaboration and Characterization of Al2O3-5 vol.% YAG Nanocomposites M=mass of the bar, g ; b= width of the bar, mm ; L= length of the bar, mm ; t= thickness of the bar, mm; f f = fundamental resonant frequency of bar in flexure, Hz; T 1 = correction factor for fundamental flexural mode to account for finite thickness of bar; μ= Poisson’s ration. Young modulus evaluated in both samples yielded a value of 419.2 GPa for the slip-casted Y-TM and 415.5 GPa for the slip-casted Y-CR. As a comparison, the elastic modulus, given by the rule of mixtures was also calculated being 409.2 GPa, in good agreement with the experimental value. Fracture toughness was estimated using the indentation method using the Anstis’ formula14, as shown below. K IC A. E / H . P / c 2 / 3 1/ 2 (2) where c is the length of the crack from the center of the impression; A is a geometric constant equal to 0.016; P the change expressed in Newton; H the hardness value; E the Young’s modulus. In the following Figures 4.19 , the Y-TM-NS fracture surfaces are presented. As it is shown Y-TM-P fracture surface exhibits inter-granular fracture, whereas in the case the of Y-TM-SC is mostly inter-granular with some trans-granular areas (indicated with arrows). a) b) Figures 4.19 - ESEM micrographs on fracture surface of: (a) Y-TM-SC and (b) Y-TM-P by pressureless sintering at 1500°C for 3h. As in the previous case, Figures 4.20 present the microstructures of Y-CR NS samples. Y-CR-SC fracture surface is mainly inter-granular. Fracture surface of sample Y-CR-P shows a mixture of trans-granular (indicated with arrows) and inter-granular modes. 165 Elaboration and Characterization of Al2O3-5 vol.% YAG Nanocomposites a) b) Figures 4.20 - ESEM micrographs on fracture surface of: (a) Y-CR-SC (b) Y-CR-P by pressureless sintering at 1500°C for 3 h. ESEM observation on Y-TM samples sintered by non-conventional routes revealed that fracture was -as in the previous case,- a mixture of trans-granular (indicated with arrows) and inter-granular as shown in Figures 4.21 . a) b) Figures 4.21- ESEM micrographs on fracture surface of: (a) Y-TM-P by HP 1450°C, (b) Y-TM-P by SPS 1350°C (GSE images). Finally, fracture surfaces of samples Y-CR sintered by HP and SPS are shown in Figures 4.22. The fracture surfaces reveal that both samples prevalently presented trans-granular fracture (indicated with arrows) with some inter-granular zones. 166 Elaboration and Characterization of Al2O3-5 vol.% YAG Nanocomposites a) b) Figures 4.22- ESEM micrographs on fracture surface of: (a) Y-CR-P by HP 1450°C, (b) Y-CR-P by SPS 1450°C (GSE images). Y-TM samples densified by natural sintering presented different fracture modes. As seen in Y-TM-P (Figure 4.19 (b)) some transgranular areas were detected. This fact may be the responsible of increasing the fracture toughness from 4.42 MPa.m1/2 of the SC sample to 5.11 MPa.m1/2 in case of P material. Nanocomposites, as reported in literature, show a clear change in the fracture mode from inter-granular to trans-granular. In the case of a monolithic material, since the fracture resistance of the grain boundary is lower compared with the one at the grain lattice, the crack will propagate inter-granularly along the grain boundaries. However, the toughening effect of the nanoparticles at the grain boundaries may steer the tip crack into the grain. As a result, the fracture becomes then trans-granular increasing the toughness of the composite. Once the tip crack is inside the grain, the propagation can be pinned by intragranular particles, producing the zig-zag path (see section 3.4.6 – Crack deflection). The same phenomenon was observed in naturally sintered Y-CR samples in which an important increase of toughness occurs reaching 7.12 MPa.m1/2 in the case of sample Y-CR-P. Y-TM samples sintered by non-conventional sintering HP and SPS exhibit similar fractures path. The toughness differences observed among the samples are reasonably imputable to the difference in matrix average grain size: 0.65 μm versus 0.51 μm in the case of HP and SPS, respectively. It is well known that toughness increases linearly in function of d1/2 as reported by Swain17. This fact may explain the disparities in non-conventional sintered samples regarding the toughness values. In order to clarify this fact, in Figure 4.23 it is shown the influence of the grain size on toughness of monolithic alumina and alumina nanocomposites reinforced with SiC whiskers. 167 Elaboration and Characterization of Al2O3-5 vol.% YAG Nanocomposites Figure 4.23 – Influence of the grain size on the fracture toughness of an alumina and alumina nanocomposites reinforced with 10 and 20 % of SiC whiskers17. The same phenomenon was observed in Y-CR non-conventional sintered samples as fracture modes in HP and SPS were similar. In this particular case, Y-CR SPS exhibited a higher toughness value 7.32 MPa.m1/2 compared with 6.37 MPa.m1/2 (HP) again attributable to the higher matrix average grain size 1.13 μm (SPS) against 0.8 μm (HP). This statement is based, as fracture modes in both samples were similar. As a comparison with literature, it was found a range of toughness between 3 to 5.8 MPa.m1/2 for YAG reinforced alumina nanocomposites for 5-50 vol.% 5,13,18. 4.5.2. Mechanical properties at high temperature Creep tests for alumina and doped aluminas were performed in a 4-point-bending fixture at the temperature of 1200°C and an applied stress of 100 MPa. Samples were tested under an applied stress of 100 MPa due to the limited number of samples. The heating rate of the furnace was 300°C/h in order to avoid significant thermal gradients. The specimens were parallelepipeds with dimensions of about 3 x 4 x 32 mm3, in which the tensile face of all the specimens was polished with diamond paste down to 3 μm and the edges were chamfered (about 45°) in order to avoid the influence of microcracks during creep. Natural sintered bar-shaped samples were produced by slip-casting, whereas less-conventional sintered samples were produced by uniaxially pressing, as the forming method. In this particular case it did not influence the final properties in terms of microstructure and final density. A four-point bending fixture was used with an inner/outer span of 30 and 15 mm respectively. The flexural stress on the tensile face of the specimen was calculated employing the following expression: 3P L L' 2bw2 (3) where P is the applied load, L the outer span, L’ the inner span, b the sample width and w the sample height. To determine the creep strain, it was used the method described in literature19. This method is based on the assumption that creep strain ε can be calculated from the deflection y at the center of the specimen and the deflection, whereas there is no major cracking in the specimen and the deflection y is small compared with the inner span L’. 168 Elaboration and Characterization of Al2O3-5 vol.% YAG Nanocomposites K (n). y (4) where K ( n) 2w n 2 L L ' L L ' n 1 The constant K(n), in addition to its dependence on n, is also a function of L and L’. Hollemberg et al.7 have shown that for (L/L’) values close to 2, K(n) are almost insensitive to n. . The steady-state creep rate is defined by the following equation: . A n Q exp d RT (5) p where A: a material constant; σ: the applied stress; n: the stress exponent; d: the grain size; p: the grain size exponent; Q: the activation energy for creep; R: the gas constant; T: the absolute temperature. 4.5.2.1. Creep of TM and CR The creep curves were obtained for TM and CR and they are shown in Figure 4.24 - sintered at 1500°C with a soaking time of 3 h – in order to compare the effect of the second phase on the creep behaviour. Both curves exhibited two stages: transient and steady-state. The strain rate has been determined by the evaluation of the slope of the creep curve on the steady-state stage. 0,9 0,8 0,7 Strain (%) 0,6 0,5 0,4 TM CR 0,3 0,2 0,1 0,0 0 2 4 6 8 10 12 14 16 Time (h) Figure 4.24- Creep curves of TM and CR at 100 MPa and 1200°C. The instantaneous strain rate in function of the time is shown in the Figure 4.25. In these curves it is possible to corroborate that an almost constant strain rate as observed in the second state. A 169 Elaboration and Characterization of Al2O3-5 vol.% YAG Nanocomposites deformation (ε) under 1% was found in both samples being 0.79% in TM and 0.69% for CR, respectively. Strain Rate (1/s) 1E-6 1E-7 1E-8 TM CR 1E-9 0 2 4 6 8 10 12 14 Time (h) Figure 4.25- Strain rate of TM and CR at 100 MPa and 1200°C. The strain rate calculated from the slope of the curve deformation vs. time has been 1.04.10-8 s-1 in TM and 1.61.10-8 s-1 in CR, respectively. The discrepancy observed in terms of strain rate has been imputed to the effect of the porosity on the creep behaviour as reported by Langdon21. The author has analyzed the effect of the porosity on the creep rate in aluminas. The author proposed the following theoretical formula to explain the effect of the porosity. P 1 m n . Gb b 1 1 P A.D. . . . n kT d G 1 P 2 / 3 1 n (6) where A is a dimensionless constant, D is the coefficient of diffusion, G is the shear modulus, b is the Burgers vector, k is the Boltzmann’s constant, T is the temperature in Kelvin, d is the grain size, m is the grain size exponent, n is the stress exponent, β is a constant equal to -4 +/- 0.07 for alumina when porosity varies between 0.3 and 16 % and P is the porosity expressed in %. By applying the previous equation it is possible to plot a graph, in which it is calculated the creep rate as a function of the porosity and the stress exponent, as it is shown in Figure 4.26. Figure 4.26 – Creep rate as a function of porosity for n values from 1 to 521. 170 Elaboration and Characterization of Al2O3-5 vol.% YAG Nanocomposites Considering that the differences regarding the porosity were ≈5% TD and n values reported in literature varied from 1 to 1.5 (see Section 3.6.2 - Chapter III). In theory, the strain rate would increase from 4 to 50 %. The difference between the strain rates of TM and CR was 53.84 %. The empirical value is higher than the expected. Theoretically, the increment caused by the porosity, as expected, should be lower as CR presented a higher average grain size (3.05 μm in CR versus 1.43 μm in TM). 4.5.2.2. Creep of Y-TM and Y-CR nanocomposites As for TM and CR, the same procedure was extended to the composites carrying test until the rupture -as for both aluminas-. In Figure 4.27 (a) it is shown the evolution of the strain for Y-TM as a function of time for a given stress of 100 MPa at 1200°C. The test gave rise to the contained deformation under creep of about ≈ 0.26% (in the case of Y-TM HP), which was significantly lower if compared to the monolithic TM. In fact, especially in case of Y-TM SPS after 55 h, deformation is reduced in ≈7 times in comparison with TM. 0,30 0,25 Strain (%) 0,20 0,15 0,10 0,05 Y-TM NS-1500 Y-TM HP-1450 Y-TM SPS-1350 0,00 0 5 10 15 20 25 30 35 40 45 50 55 60 Time (h) a) 1E-6 Strain rate (1/s) 1E-7 1E-8 1E-9 1E-10 Y-TM NS-1500 Y-TM HP-1450 Y-TM SPS-1350 1E-11 0 b) 5 10 15 20 25 30 35 40 45 50 Time (h) Figures 4.27- (a) Creep curves and (b) Strain Rate of Y-TM. 171 Elaboration and Characterization of Al2O3-5 vol.% YAG Nanocomposites In Figure 4.27 (b) the strain rates calculated in the steady-state corresponding to Y-TM samples are shown. The strain rates έ n were 1.8.10-7 s-1 for Y-TM NS, 4.2.10-10 s-1 for Y-TM HP and 1.35.10-9 s-1 for Y-TM SPS, respectively. As it can be seen, in the Figure 4.27 (b) Y-TM sintered by HP has the lowest strain rate among the samples of about έ=4.2.10-10 s-1. No significant differences were found to explain the disparity among the results, since grain sizes and densities in Y-TM samples are comparable. Not normalized strain rates were successfully reduced ≈25 and ≈8 times in samples: Y-TM HP and Y-TM SPS in contrast to TM. 0,7 0,6 Strain (%) 0,5 0,4 0,3 0,2 0,1 Y-CR NS-1500 Y-CR HP-1450 0,0 Y-CR SPS-1450 0 5 10 15 20 25 30 35 40 45 50 55 Time (h) a) Strain rate (1/s) 1E-7 1E-8 Y-CR NS-1500 Y-CR HP-1450 Y-CR SPS-1450 1E-9 0 b) 5 10 15 20 25 30 35 40 45 50 55 Time (h) Figures 4.28- (a) Creep curves and (b) Strain Rate of Y-CR samples. 172 Elaboration and Characterization of Al2O3-5 vol.% YAG Nanocomposites Subsequently, the procedure was extended to Y-CR samples. In the case of Y-CR, the strain in all the cases is substantially reduced ≈5 times up to 0.134% in the case of Y-CR HP in comparison with CR sample. Only in sample Y-CR SPS, it was measured a little improvement of ε=0.63%. As it is shown in Figure 4.28 b, an important strain rate reduction was found in the natural sintered sample, namely Y-CR NS. This phenomenon is ascribable to the average grain size (0.98 μm) and a good distribution of the second phase within the alumina grains. Strain rates measured on the samples showed in Figure 4.28 (b) were 2.12.10-9 s-1 for Y-CR NS, 4.47.10-9 s-1 for Y-CR HP and 5.44.10-8 s-1 for Y-CR SPS, respectively. As in the previous case, non-normalized strain rates denoted a reduction of ≈8 and ≈4 times in samples Y-CR NS and Y-CR HP with respect to CR. For the sake of clarity, all the values regarding the grain size and strain rates of the different samples are contained in Tables II/III. In order to establish a comparison among the samples, the normalized strain rate was calculated using the criterion taken from literature5,7,18. The normalized strain rate έ n can be calculated by the following equation, taking into account the differences in grain sizes. The formula used for calculation is shown below. P 1 . n . B d A / dB . (7) Where έ n is the measured creep rate for Y-TM and Y-CR, d A is the mean size of sample A and d b is the mean grain size of sample B. Two comparisons have been done: the former, considering the average matrix grain size of the pure alumina, labelled as d A (έ nA ) and the latter employing for the calculation, the smallest matrix alumina grain size in the case of the composite (έ nY ). The second criterion has been chosen with the aim of establishing the advantage of introducing a second phase into the alumina matrix. The grain size exponents p are 2 and 3 for lattice diffusion controlled creep and for grain boundary diffusion controlled creep, as reported in literature. It was assumed that creep is controlled by lattice diffusion controlled creep, for this reason creep rates were normalized using an exponent of a grain size equal to 2 18,23-24. In Table II it is shown the comparison in terms of normalized strain rates among the TM specimens. Table II- Strain rates of TM samples and normalized strain rates of the sintered samples at 1200°C with an applied stress of 100 MPa. Sample TM NS-1500 Y-TM NS-1500 Y-TM HP-1450 Y-TM SPS-1350 Strain rate (s-1) -8 1.04.10 1.8.10-7 4.2.10-10 1.35.10-9 Fired Density 99.4 99.8 98.5 99.4 Grain Size Al 2 O 3 (μm) 1.43 0.77 0.65 0.51 Grain Size YAG (μm) 0.45 0.43 0.26 . nA (s-1) 1.04.10-8 5.22.10-8 8.68.10-11 1.72.10-10 . nY (s-1) 8.18.10-8 4.10.10-7 6.82.10-10 1.35.10-09 For a better understanding, values corresponding to the different strain rates were plotted in Figure 4.29. Non important differences were found among the samples, since grain sizes are comparable -as it was mentioned before-. 173 Elaboration and Characterization of Al2O3-5 vol.% YAG Nanocomposites Strain Rate (1/s) 1E-7 1E-8 1E-9 TM NS-1500 Y-TM NS-1500 Y-TM HP-1450 Y-TM SPS-1350 Samples a) Strain Rate (1/s) 1E-8 1E-9 1E-10 TM NS-1500 Y-TM NS-1500 Y-TM HP-1450 Y-TM SPS-1350 Samples b) Strain rate (1/s) 1E-7 1E-8 1E-9 TM NS-1500 Y-TM NS-1500 Y-TM HP-1450 Y-TM SPS-1350 Samples c) Figures 4.29- (a) Creep rates of TM at 1200°C under a stress of 100 MPa, (b) normalized creep rates with έ nA and (c) normalized creep rates with έ nY of different samples. On one hand, taking the έ nA criterion for Y-TM samples, creep rate is increased by a factor of 5 in Y-TM NS with respect to TM. This fact is difficult to explain, if the homogeneous microstructure and the very high final density is considered. Probably, it could be imputed to the presence of nondetected defects during the samples elaboration by slip-casting. Finally, in samples sintered by 174 Elaboration and Characterization of Al2O3-5 vol.% YAG Nanocomposites non-conventional sintering routes it was observed a clear improvement, since they present the lowest creep rate, reduced by a factor of 119 and 60 in Y-TM HP and Y-TM SPS, respectively. On the other hand, employing the second criterion έ nY –i.e. using the matrix grain size of Y-TM SPS -, the creep rate is doubled compared with HP in comparison with TM. Finally in comparison with Y-TM NS, the creep strain is increased by a factor of 303 and in the case of TM is lowered 60 times. Table III-Strain rates of CR samples and normalized strain rates of the sintered samples at 1200°C with an applied stress of 100 MPa. Sample Strain rate (s-1) CR NS-1500 Y-CR NS-1500 Y-CR HP-1450 Y-CR SPS-1450 1.61.10-8 2.12.10-9 4.47.10-9 5.44.10-8 Fired Density 97.4 95.6 98.9 99.7 Grain Size Al 2 O 3 (μm) 3.05 0.98 0.8 1.13 Grain Size YAG (μm) 0.65 0.61 0.57 . nA (s-1) 1.61.10-8 2.19.10-10 3.08.10-10 7.47.10-9 . nY (s-1) 2.34.10-7 3.18.10-9 4.47.10-9 1.09.10-7 Strain rate (1/s) 1E-6 1E-7 1E-8 CR NS-1500 Y-CR HP-1450 Y-CR NS-1500 Y-CR SPS-1450 Samples a) Strain rate (1/s) 1E-8 1E-9 1E-10 CR NS-1500 b) Y-CR NS-1500 Y-CR HP-1450 Y-CR SPS-1450 Samples 175 Elaboration and Characterization of Al2O3-5 vol.% YAG Nanocomposites Strain rate (1/s) 1E-7 1E-8 1E-9 CR NS-1500 c) Y-CR HP-1450 Y-CR NS-1500 Y-CR SPS-1450 Samples Figures 4.30- (a) Creep rates of CR at 1200°C under a stress of 100 MPa, (b) normalized creep . . rates with nA and (c) normalized creep rates with nY of different samples. The έ nA criterion in CR samples, showed that creep rate is reduced by a factor of 74 and 53 in Y-CR NS and Y-CR HP in comparison with CR. However, in the Y-CR SPS it is only lowered by a factor of 2. The second criterion έ nY (employing Y-CR HP for the calculation) resulted a creep strain increased by a factor of 1.4 in case of the Y-CR NS sample. Although, creep strain is reduced by a factor of 52 and 24 with respect to and Y-CR SPS, respectively. Sample Y-CR NS present a higher matrix average grain size compared with Y-TM which is the responsible for conferring a better creep behaviour (see equation 5). In this context, the second phase distribution, as well as the matrix grain size, seem to be the key for improving the creep resistance in composites. These results are particularly in good agreement with the work published by Yoshida et al.6, who doped high-purity alumina powder (TM-DAR, Taimei) with 0.045 mol% Y 2 O 3 . The authors attributed the reduction in creep rate, 200 times lower than the un-doped alumina for an applied stress of 50 MPa, to the segregation of Y3+ cations to alumina grain boundaries. This segregation is the responsible of restricting the grain boundary diffusion of Al3+ ions. In comparison with the literature, Schehl et al.18 found a non-normalized strain rate of 2.0.10-9 s-1 for alumina-YAG composites in the same experimental condition, giving importance to the values obtained. A comparison among samples using the criterion of normalization with έ nA is shown in Figure 4.31. Monolithic aluminas showed a similar creep behaviour, taking into account the differences in terms of grain sizes and porosity in TM and CR. The importance of a higher matrix grain size has been evident in Y-CR NS sample, as it exhibited a better creep behaviour compared with Y-TM NS. Non-conventional sintered samples exhibited different creep behaviour. This fact is imputed to some heterogeneous distribution of the second phase in some areas of Y-CR samples. 176 Elaboration and Characterization of Al2O3-5 vol.% YAG Nanocomposites Strain Rate (1/s) 1E-8 Y-TM Y-CR 1E-9 1E-10 NS Y HP Y NS Y SPS Figure 4.31- Comparison among samples. For the sake of clarity, in Figures 4.31/4.32 are shown the comparison between HP and SPS samples. From the analysis of the Figure 4.31 (at different magnification), it is possible to observe a slightly different distribution of the second phase in HP samples. Samples Y-CR sintered by HP exhibited an irregular YAG distribution compared with SPS. a) b) Figures 4.31- ESEM micrographs on polished surfaces of: (a) Y-TM HP-1450 and (b) of Y-CR HP-1450 (GSE-ESEM images) This phenomenon is more evident in samples sintered by SPS. As it is shown in Figure 4.32 (b), in this particular Y-CR SPS sample’s region, only few points of YAG phase were found (indicated with arrows). This fact may explain the differences in terms of έ nA normalized strain rates (being ≈44 times higher in sample Y-CR- see Figure 4.30), in spite of the higher matrix grain size of Y-CR sample. 177 Elaboration and Characterization of Al2O3-5 vol.% YAG Nanocomposites a) b) Figures 4.32- ESEM micrographs on polished surfaces of: (a) Y-TM SPS-1350 and (b) of Y-CR SPS-1450 (GSE-ESEM images) 4.5.3. Activation energy The activation energy of the nanocomposites was determined in the temperature range of 1200-1400°C with an applied stress of 100 MPa for the most promising material in terms of creep resistance, operational time, second phase distribution and elaboration process, the Y-CR NS-1500. The activation energy has been determined by the slope of ln(ε)=f(1/T) obtained at constant applied stress of 100 MPa. In the following Figure 4.33 it is shown ln έ-1/T graph. -6 Strain Rate (1/s) 1x10 Q=644 kJ/mol -7 1x10 -8 1x10 -9 1x10 6,0 6,2 6,4 6,6 6,8 -4 1/RT x 10 (1/K) Figure 4.33 – Steady-state creep rate vs. the reciprocal of the absolute temperature of the Y-CR NS-1500 composite. 178 Elaboration and Characterization of Al2O3-5 vol.% YAG Nanocomposites The activation energy for the Y-CR NS-1500 was 644.4 kJ/mol in good agreement with the results found literature, i.e. it was measured for alumina 630 kJ/mol23. A comparable result was found in literature, French et al.5 obtained 698 kJ/mol for a 1000 ppm yttrium-doped Al 2 O 3 with an applied stress of 75 MPa in the range of 1125-1350°C. Yoshida et al.6 obtained 830 kJ/mol for Al 2 O 3 doped with 0.045% Y 2 O 3 for an applied stress of 100 MPa in the range of 1250-1350°C. Similarly, Torrecillas et al.7 produced alumina-YAG nanocomposites by the colloidal processing route (the process is shown in detail in Chapter III) found an activation energy of 595 kJ/mol for an applied stress of 100 MPa. It is known that activation energy for Al3+ grain boundary diffusion is 420 kJ/mol, whereas for Al3+ lattice diffusion the activation energy is 578 kJ/mol, measured in the 1200-1400°C range19-21. In creep of monolithic aluminas a combination of both mechanisms is supposed to coexist, in which lattice diffusion is supposed to be the dominant (Nabarro-Herring)19-21. In composites, as it was reviewed by some researchers, YAG particles formed at grain boundaries may inhibit the grain boundary sliding7. This phenomenon is caused by a combination of high activation energy necessary to deform these nanoparticles and the absence of defects in YAG nanoparticles raise the total activation energy closer to lattice diffusion of Al3+ in YAG nanoparticles7. 4.5.4. Microstructural observation after creep tests In Figures 4.34 it is shown the fracture surface after creep measurements. The micrograph gives rise to a non-significant grain growth occurred during creep and mainly inter-granular fracture occurred on samples. The second conclusion was that no cavitation was observed in the samples. a) b) Figures 4.34- SEM micrographs of fracture surfaces: (a) Y-TM-SC and (b) Y-CR-SC (GSE-ESEM images). 179 Elaboration and Characterization of Al2O3-5 vol.% YAG Nanocomposites Conclusion of the second part and perspectives The developed Al 2 O 3 -YAG composites showed good mechanical properties (hardness, toughness, and Young modulus) at room temperature. In both Y-CR and Y-TM, samples sintered by HP and SPS presented the highest hardness values (≈20 GPa) ascribable to the higher density measured. Toughness values were strongly influenced by the matrix average grain size, since higher values were found in samples having higher grain size reaching ≈7 MPa.m1/2. This phenomenon was corroborated specially in Y-CR sintered by NS and by SPS which have higher average grain size 0.98 and 1.13 μm, respectively. The creep behaviour in monolithic TM and CR was ruled by the differences in terms of porosity, as it was reviewed in the literature. The influence of grain size was evidenced in the naturally sintered materials, since differences were measured involving the porosity (higher in CR). Samples sintered by non-conventional routes presented differences caused by an inhomogeneous distribution of the second phase, as revealed by SEM observation. This phenomenon was confirmed by Archimedes’ measurements, as not substantial differences were found concerning the fired densities. These results in non-conventional sintered samples showed that it is necessary to optimize the second phase distribution, as some discrepancies exist in terms of strain rates. This fact could permit to clarify this phenomenon. However, good results were obtained in the composites with respect to creep behaviour which proved that YAG formation along/inside grain boundaries improved the creep resistance, as it is believed that Y3+ reduces the grain boundary diffusivity. Further studies have to be performed on samples with the aim of completing the mechanical characterization at high temperature. It is necessary to obtain stress exponent values by extending the same approach to more stresses and temperatures and consequently clarify the creep mechanism for a correct normalization. 180 References 181 References Chapter I [1] MORGAN, P.E.D., Superplasticity in Ceramics, in: Ultrafine-Grained Ceramics. J.J. Burke, L.R. Norman, and V. Weiss, editors., Syracuse University Press, Syracuse, 1968. [2] MAYO, M., Processing of Nanocrystalline Ceramics from Ultrafine Particles. International Material Reviews, 1996, vol. 41, pp. 85-115. [3] Azo Nanotechnology [on-line]. Available at: http://www.azo.com/ (consulted 20.11.2009) [4] CHOE, G.M., KURIHARA, L.K., Chemical Synthesis of Processing of Nanostructured Powders and Films. In: KOCH, C.C., Nanostructured Materials: processing, Properties and Applications. William Andrew Publishing, New York, 2002. [5] GLEITER, H., Nanostructured Materials: State of Art and Perspectives. NanoStructured Materials, 1995, vol. 6, pp. 3-14. [6] GROZA, J.R., Nanosintering. Nanostructured Materials, 1999, Vol. 12, pp. 987-992. [7] LANGE, F.F. Powder processing science and technology for increased reliability. Journal of the American Ceramic Society, 1989, vol. 72, pp. 3-15. [8] NIIHARA, L., New Design Concept of Structural “Ceramic Nanocomposites”. The Centennial Memorial Issue of the Ceramic Society of Japan, 1991, vol. 99, n°10, pp. 974-982. [9] BOWEN, P., CARRY, C., LUXEMBOURG, D., HOFMANN, H., Colloidal processing and sintering of nanosized transition aluminas. Powder Technology, 2005, vol. 157, pp. 100-107. [10] WU, S.J., DE JONGHE, L.C., Sintering of Nanophase γ-Al2O3 powder. Journal of the American Ceramic Society, 1996, vol. 79, pp. 2207-2211. [11] ZHAN, G.D, KUNTZ, J., WAN, J., GARAY, J., MUKHERJEE, A.K., A novel processing to developed a Dense Nanocrystalline Alumina Matrix (<100 nm) Nanocomposite Material. Journal of the American Ceramic Society, 2002, vol. 86, vol. 1, pp. 200-202. [12] MEYERS, M.A., MISHRA, A., BENSON, D.J., Mechanical properties of nanocrystalline materials. Progress in Materials Science, 2006, vol. 51, pp. 427-556. [13] CHEN, D.J., MAYO, M.J., Densification and grain growth of ultrafine 3mol% Y 2 O 3 -ZrO 2 ceramics. Nanostructured materials, 1993, vol. 2, n°5, pp. 469-478. [14] LEWIS, J.A., Colloidal processing of ceramics. Journal of American Ceramic Society, 2000, vol. 83, pp. 2341-2359. [15] RAHAMAN, M.N., Ceramic Processing and Sintering. Marcel Dekker, New York, 2003. [16] HAMACKER, H.C., The London-van der Waals attraction between spherical particles. Physica, 1937, vol. 4, pp. 1058-1072. [17] NAPPER, D.H., Polymeric Stabilization of Colloidal Dispersions. Academic Pres, New Yok, 1983. 182 References [18] CHEN, W., PECHENIK, A., DAPKUNAS, S.J., PIERMARINI, G.J., MALGHAN, S.G., Novel Equipment for the Study of Compactation of Fine Powders. Journal of the American Ceramic Society, 1995, Vol. 77, pp. 1005-1010. [19] FERKEL, H, HELLMIG, R.J., Effect of nanopowders deagglomeration on densities of nanocrystalline ceramic green bodies and their sintering behavior. Nanostructured Materials, 1999, vol. 11, n° 5, pp. 617-622. [20] HERRING, C., Effect of Change of scale on Sintering Phenomena. Journal of Applied Physics, 1950, vol. 21, pp. 301-303. [21] CHEN, W., MALGHAN, S.G., Investigations of Compactation Equations for Powders. Powder technology. Powder technology, 1994, vol. 81, pp. 75-81. [22] RHODES, W.H., Agglomerate and Particle Size Effects on Sintering Yttria-Stabilized Zirconia. Journal American Ceramic Society, 1981, vol. 64, pp. 19-22. [23] GALLAS, M.R., ROSA, A.R., COSTA, T.H., DA JORNADA, J.A.H., High Pressure Compaction of Nanosized Ceramic Powder. Journal Material Research, 1997, vol. 12, pp. 764-768. [24] SAGEL-RANSIJN, C.D., WINNBUST, A.J.A., KERKWIJK, B., BURGGRAF, A.J., VEWWEIJ, H., Production of Defect-Poor Nanostructured Ceramics of Yttria-Zirconia. Journal of the European ceramic Society, 1997, vol. 17, pp. 831-841. [25] VAN DE GRAAF, M.A.C.G., TERMAAT, J.H.H., BURGGRAAF, A.J., Microstructure and Sintering Kinetics of Highly Reactive ZrO 2 -Y 2 O 3 Ceramics. Journal of Materials Science, 1985, vol. 20, pp. 1407-1418. [26] BOSCHI, A.O., GILBERT, E., in: Advanced Ceramic Processing and Technology vol.I, ed. J. G.P. Binner, 71-93, New Jersey, 1990. [27] RHODES, W.H., Agglomerate and Particle Size Effects on Sintering Yttria-Stabilized Zirconia. Journal of the American Ceramic Society, 1981, vol. 64, pp. 19-22. [28] CHEN, P.L., CHEN, I.W., Sintering of fine oxide powders: II Sintering Mechanisms. Journal of the American Ceramic Society, 1997, vol. 80, pp. 637-645. [29] COBLE, R.L., Sintering Crystalline Solids: I. Intermediate and Final State Diffusion Models. Journal of Applied Physics, 1961, Vol. 32, n° 5, pp. 787-792. [30] BAUMGARTNER, C.E., Fast Firing and Conventional Sintering of Lead Zirconate Titanate Ceramics. Journal of the American Ceramic Society, vol, 71, n°7, pp. 350-353. [31] WANG, J., RAJ, R., Estimate of the Activation Energies for Boundary Diffusion from rate-controlled sintering of pure alumina, and alumina doped with Zirconia and Titania. Journal of the American Ceramic Society, 1990, vol. 73, n°5, pp. 1172-1175. [32] ANDRIEVSKI, R.A., Compactation and Sintering of Ultrafine Powders. International Powder Metallurgy, 1994, vol. 30, pp. 59-66 [33] MESSING, G.L., KUMAGAI, M., Low-Temperature Sintering of α-Alumina Seeded Bohemite Gels. American Ceramic Bulletin, 1994, vol. 73, pp. 88-91. 183 References [34] YEH, T.S., SACKS, M.D., Low-Temperature Sintering of Aluminium Oxide. Journal of the American Ceramic Society, 1998, vol. 71, n°10, pp. 841-844. [35] KIM, B.C., LEE, J.H., KIM, J.J., IKEGAMI, T., Rapid rate of nanocrystalline indium tin oxide ceramics. Materials Letters, 2002, vol. 52, pp. 114-119. [36] CHU, M.Y., RAHAMAN, M.N., DE JONGHE, L.C., Effect of Heating Rate on Sintering and Coarsening. Journal of the American Ceramic Society, 1991, vol. 74, n°6, pp. 1217-1725. [37] KINGERY, W.D., BOWEN, H.K., UHLMANN, D.R., Introduction to Ceramics, 2nd ed., Wiley, New York, 1976. [38] BROOK, R.J., Pore-Grain Boundary Interactions and Grain Growth. Journal of the American Ceramic Society, 1969, vol. 52, pp. 56-57. [39] AVERBACK, R.S., HOFLER, H.J., TAO, R., Processing of Nano-grained Materials. Materials Science Engineering A, 1993, vol. 166, pp. 169-177. [40] CAMERON, C.P., RAJ, R., Grain-Growth Transition during Sintering of Colloidally prepared Alumina Powder Compacts. Journal of the American Ceramic Society, vol. 71, n° 12, 1988. [41] HAHN, H., AVERBACK, R.S., HOFLER, H.J., LOGAS, J., Sintering and Deformation of Nanocrystalline Ceramics, in: Cluster and Cluster-Assembled Materials. Materials Research Society, Pittsburgh,1991. [42] MAYO, M., HAGUE, D.C., CHEN, D.J., Processing nanocrystalline Ceramics for Applications in: Superplasticity. Materials Science Engineering A, 1993, vol. 166, pp. 145-159. [43] GROZA, J.R., DOWDING, R.J., Nanoparticles materials densification. NanoStructured Materials, 1996, vol. 7, n° 7, pp. 749-768. [44] AHN, J.P., HUH, M.Y., PARK, J.K., Effect of Green density on Subsequent densification and Grain Growth of Nanophase SnO 2 Powder During Sintering. Nanostructured Materials, 1997, vol. 8, pp. 637-643. [45] HARMER, M.P., CHAN, H.M., MILLER, G.A., Unique opportunities for microstructural engineering with duplex and laminar ceramic composites. Journal of the American Ceramic Society, 1992, vol. 75, n° 7, pp. 1715-1728. [46] LIU, Y., PATTERSON, B.R., Grain Growth Inhibition by Porosity. Acta Metallurgica et Materialia, 1993, vol. 41, pp. 2651-2656. [47] CHEN, P.L., CHEN, I.W., Sintering of fine oxide powders: I Microstructural evolution. Journal of the American Ceramic Society, 1996, vol. 79, pp. 3129-3141. [48] ZENG, W., GAO, L., GUI, L., GUO, J., Sintering kinetics of α-Al 2 O 3 powder. Ceramics International, 1999, vol. 25, pp. 723-726. [49] RODRIGUEZ, J.A., FERNANDEZ-GARCIA, M., Synthesis, Properties and Applications of Oxide Materials. Ed. John Wiley & Sons, New Jersey, 2007. [50] YAN, M.F., RHODES, W.W., Low Temperature Sintering of TiO 2 . Material Science and Engineering, 1983, vol. 61, pp. 59-66. 184 References [51] NORDAHL, C.S., MESSING, G.L., Sintering of α-Al 2 O 3 -seeded nanocrystalline γ-Al 2 O 3 . Journal of the European ceramic Society, 2002, vol. 22, pp. 415-422. [52] WANG, Y., SURYANARAYANA, C., AN, L., Phase transformation in nanometer-sized γ-Alumina by Mechanical Milling. Journal of the American Ceramic Society, 2005, vol. 88, n°3, pp. 780-783. [53] PALKAR, V.R., THAPA, D., MULTANI, M.S., MALGHAN, S.G., Densification of nanostructured alumina assisted by rapid nucleation of α-alumina. Materials Letters, 1998, vol. 36, pp. 235-239. [54] LEGROS, C., CARRY, C., BOWEN, P., HOFMANN, H., Sintering of a Transition Alumina: Effects of Phase Transformation, Powder Characteristics and Thermal Cycle. Journal of the American Ceramic Society, 1999, vol. 19, pp. 1967-1978. [55] BOWEN, P, CARRY, C., From powders to sintered pieces: forming, transformations and sintering of nanostructured ceramic oxides. Powder Technology, 2002, vol. 128, pp. 248-255. [56] YEN, F.S., LO, H.S., WEN, H.L., YANG, R.J., θ- TO α-phase transformation subsystem induced by α-Al 2 O 3 –seeding in boehmite-derived nano-sized alumina powders. Journal of Crystal Growth, 2003, vol. 249, pp. 293-293. [57] MESSING, G.L., KUMAGAI, M., SHELLEMAN, R.A., MC ARDLE, J.L., Seeded transformation microstructural control in ceramics. In: HENCH, L.L., ULRICH, D.D. (Eds.), Science of Ceramic Chemical Processing, John Wiley & Sons, New York, 1985, pp. 259-271. [58] KUMAGAI, M., MESSING, G.L., Enhanced densification of Boehmite Sol-Gel by Alpha Alumina Seeded. Journal of the American Ceramic Society, 1984, vol. 67, pp. 230-231. [59] KUMAGAI, M., MESSING, G.L., Control transformation and sintering of Bohemite Sol-Gel by alpha-alumina seeded. Journal of the American Ceramic Society, 1985, vol. 68, pp.500-505. [60] KANO, J., SAEKI, S., TANJO, M., SAITO, F., Reduction in α-Transformation Temperature of Gibbsite and Boehmite by Powder Mixing with Seed. Journal of Chemical Engineering of Japan, 2001, vol. 34, n°5, pp. 595-599. [61] XIE, Z.P., LU, J.W., GAO, L.C., LI, W.C., XU, L.H., WANG, X.D., Influence of different seeds on transformation of aluminium hydroxides and morphology of alumina grains by hot-pressing. Materials and Design, 2003, vol. 24, pp. 209-214. [62] PACHULA, M.L.,YING, J.Y., Mechanical synthesis of nanocrystalline α-Al 2 O 3 seed for enhanced transformation kinetics. Nanostructured Materials, 1997, vol. 9, pp. 161-164. [63] LO, J.R., TSENG, T.Y., Phase development and activation energy of the Y 2 O 3 -Al 2 O 3 system by a modified sol-gel process. Materials Chemistry and Physics, 1998, vol. 58, pp. 56-62. [64] KAO, H.C., WEI, W.C., Kinetics and Microstructural Evolution of Heterogeneous Transformation of θ-Alumina to α-Alumina. Journal of the American Ceramic society, 2000, vol. 83, n°2, pp. 362-368. [65] MALEK, J., The applicability of Johnson-Mehl-Avrami model in the thermal analysis of the crystallization kinetics of glasses. Themochimica Acta, 1995, vol. 267, pp. 61-73. 185 References [66] YAN, Z.J., DANG, S.E., WANG, X.H., LIAN, P.X., Applicability of Johnson-Mehl-Avrami model to crystallization of Zr 60 Al 15 Ni 25 bulk amorphous alloy. Transaction of Nonferreous Metallurgy Society of China, 2008, vol. 18, pp. 138-144. [67] BROWN, M.E., Introduction to Thermal Analysis: Techniques and Applications. Chapman and Hall ed., London, 1988. [68] KISSINGER, H.E., Variation of Peak Temperature with Heating Rate in Differential Thermal Analysis. Journal of Research of the National Bureau of Standards, 1956, vol. 57, n°4, 217221. [69] KISSINGER, H.E., Reaction Kinetics in Differential Thermal Analysis. Analytical Chemistry, 1957, vol. 29, N°11, pp. 1702-1706. [70] AUGIS, J.A., BENNETT, J.E., Calculation of the Avrami parameters for heterogeneous solid state reactions using a modification of the Kissinger Method. Journal of the Thermal Analysis, 1978, vol. 13, pp. 283-292. [71] KERR, M.C.., REED, J.S., Comparative grinding kinetics and grinding energy during ball milling and attrition milling. American Ceramic Society Bulletin, 1992, vol. 71, n°12, pp. 1809-1815. [72] NORDAHL, C.S., MESSING, G.L., Thermal analysis of phase transformation kinetics in α-Al 2 O 3 seeded Bohemite and γ-Al 2 O 3 . Themochimica Acta, 1998, vol. 318, pp. 187-199. [73] KAO, H.C., WEI, W.C., Kinetics and Microstructural Evolution of Heterogeneous Transformation of θ-Alumina to α-Alumina. Journal of the American Ceramic Society, 2000, vol. 83, n°2, pp. 362-368. [74] YEN, F.S., CHANG, J.L., YU, P.C., Growth Thermodynamics and Kinetics of Nano-sized Alpha-Al2O3 Powders. Key Engineering Materials, 2003, vol. 249, pp. 179-182. [75] YU, P.C., YEN, F.S., Preparation of Nano-sized Alpha-Alumina Crystallites Using Theta- Theta Alumina Composite Powders. Key Engineering Materials, 2003, vol. 249, pp. 175-178. [76] YANG, R.J., YEN, F.S., LIN, S.M., CHEN, C.C., Microstructure-controlled effects on temperature reduction of crystallite formation. Journal of Crystal Growth, 2007, vol. 299, pp. 429-435. [77] EL SAYED, M., TOFT SORENSEN, O., Initial sintering stage kinetics of CeO 2 studied by Stepwise Isothermal Dilatometry, Report Risø R-518, February 1985. [78] LIPPENS, B.C., STEGGERDA, J.J., In: LINSEN, B.G. (Editor), Physical and Chemical Aspects of adsorbents and Catalysts, Academic Press, New York, 1970, p. 171. [79] MORTERRA, C., MAGNACCA, G., A case study: surface chemistry and surface structure of catalytic aluminas, as studied by vibrational spectroscopy of adsorbed species. Catalysis Today, 1997, vol. 27, pp. 497-532. [80] WILSON, S.J., MCDONNELL, J.D.C., A kinetic study of the system γ-AlOOH/ Al 2 O 3 . Journal of the Solid State Chemistry, 1980, vol. 34, pp. 315-322. [81] BUSCA, G., LORENZELLI, V., RAMIS, G., WILLEY, R.J., Surface sites on the spinel-type and corundum-type metal oxide powders. Langmuir, 1993, vol. 9, pp. 1492-1499. 186 References [82] TARTE, P., Infra-red spectra of inorganic aluminates and characteristic vibrational frequencies of AlO 4 Tetrahedra and AlO 6 Octahedra. Spectrochimica Acta Part A, 1967, vol. 23, pp. 2127-2143. [83] KNOZINGER, H., KRIETENBRINK, H., MULLER, H.D., SCHULTZ, M., Cooperative effects in surface chemical and heterogeneously catalysed reaction on metal oxides, in Proceedings of the 6th International Congress on Catalysis, London, 1976, BOND, G.C., WELLS, P.B., TOMPKINS, F.S. (Eds.), The Royal Society of Chemistry, London, 1977, vol. 1, pp. 183-194. [84] LAVALLEY, J.C., BENAISSA, M., IR spectroscopic evidence for surface vibrational modes formed upon dehydroxylation of alumina. Journal of the Chemical Society, Chemical Communications, 1984, pp. 908-909. [85] MARCHESE, L., BORDIGA, S., COLUCCIA, S., MARTRA, G., ZECCHINA, A., Structure of the surface sites of δ-Al 2 O 3 as determined by high-resolution transmission electron microscopy, computer modelling and IR spectroscopy of adsorbed CO. Journal of the Chemical Society: Faraday Transactions I, 1993, vol. 89, pp. 3483-3489. [86] MORTERRA, C., BOLIS, V., MAGNACCA, G., IR spectroscopic and microcalorimetric characterization of Lewis acid sites on (transition-phase) Al 2 O 3 using adsorbed CO. Langmuir, 1994, vol. 10, pp. 1812-1824. [87] FIERRO, J.L.G., Spectroscopic Characterisation of Heterogeneous Catalysts, Part A: Method of surface Analysis, Part B: Chemisorption of Probe-Molecules, Elsevier, Amsterdam, 1990. [88] TSYGANENKO, A.A., FILIMONOV, V.N., Infrared spectra of surface hydroxyls groups and crystalline structures of oxides. Journal of Molecular Structure, 1973, vol. 19, pp. 579-589. [89] TSYGANENKO, A.A., Structure and properties of hydroxylated surfaces of oxides. Reaction Kinetics and Catalysis Letters, 1993, vol. 50, pp. 33-38. [90] KNOZINGER, H., RATNASAMY, C., Catalytic Aluminas: Surface Models and Characterization of Surface Sites. Catalysis Reviews: Science and Engineering, 1978, vol. 17, pp. 31-70. [91] PERI, J.B., IR and gravimetric study of the surface hydration of γ-alumina. A model for the surface of γ-alumina. IR study of adsorption of Ammonia on dry γ-alumina. Journal of Physical Chemistry, 1965, vol. 69, pp. 220-221. [92] KAWAKASHI, H., YOSHIDA, S., Theoretical approach to the basicity of alkaline-earth metal oxide catalysis. Journal of the Chemical Society. Faraday Transactions 2, 1984, vol. 80, pp. 921-932. [93] DAVYDOV, A., Molecular Spectroscopy of Oxide Catalyst Surfaces. John Wiley & Sons Ltd. The Atrium, England, 2003. [94] BUSCA, G., LORENZELLI, V., ESCRIBANO, V. S., GUIDETTI, R., FT-IR study of the surface properties of the spinels NiAl 2 O 4 in relation to those of the transition aluminas. Journal of Catalysis, 1991, vol. 131, pp. 167-177. [95] DELLA GATTA, G., FUBINI, B., GHIOTTI, G., MORTERRA, C., Chemisorption of CO on various transition aluminas. Journal of Catalysis, 1976, vol. 43, pp. 90-98. [96] PERI, J.B., IR study of the adsorption of CO 2 , HCl and other molecules on acid sites on silicaalumina and γ-alumina. Journal of Physical Chemistry, vol. 70, pp. 3168-3179. 187 References [97] PERI, J.B., IR study of OH and NH 2 groups on the surface of a dry silica aerogel. Journal of Physical Chemistry, vol. 70, pp. 2937-2945. [98] ZECCHINA, A., COLUCCIA, S., MORTERRA, C., Infrared spectra of molecules adsorbed on oxide surfaces. Applied Spectroscopy Rewiew, 1985, vol. 21, pp. 259-310. [99] LITTLE, L.H., Infrared Spectra of Adsorbed Molecules. Academic Press, London, 1966. [100] DIXIT, L., RAO, P.T., Spectroscopy in the measurements of acidic-basic properties solids. Applied Spectroscopy Review, 1996, vol. 31, pp. 369-472. [101] PARRY, E.P, An infrared study of pyridine adsorbed on acidic solids. Characterization of surface acidity. Journal of Catalysis, 1963, vol. 2, pp. 371-379. [102] TAAGELPERA, M., HENDERSON, W.G., BROWNLEE, R.T.C., BEAUCHAMP, J.L., HOLTZ, D., TAFT, R.W. Gas phase basicities and pyridine subtituent effects. Journal of the American Chemical Society, 1972, vol. 94, pp. 1369-1374. [103] JENSEN, W.B., the Lewis Acid-Base Concepts. An Overview. Wiley, New York, 1980. [104] STAIR, P.C., The concept of Lewis acids and bases applied to the surfaces. Journal of the Chemical Society, 1972, vol. 104, pp. 4044-4052. [105] ZAKI, M.I., KNOZINGER, H., CO- a low temperature IR probe for the characterization of hydroxyls group properties on metal oxide surfaces. Material Chemistry and Physics, 1987, vol. 17, pp. 201-215. [106] ZECCHINA, A., SCARANO, D., BORDIGA, S., RICCHIARDI, G., SPOTO, G., GEOBALDO, F., IR studies of CO and NO adsorbed on well characterized oxide single microcrystals. Catalysis Today, 1996, vol. 27, pp. 403-435. [107] MORTERRA, C., MAGNACCA, G., DEL FAVERO, N., IR study of CO adsorption at 77 K on α-Al 2 O 3 . Langmuir, 1993, vol. 9, pp. 642-645. Chapter II [1] FERKEL, H., HELLMIG, R.J., Effect on the nanopowders deagglomeration on the densities of nanocrystalline ceramic green bodies and their sintering behaviour. Nanostructural materials, 1999, vol. 11, n°5, pp. 617-622. [2] Nanophase Technologies Corporation [on-line]. Available at: http://www.nanophase.com/ (consulted 24.11.2009) [3] AZAR, M., PALMERO, P., LOMBARDI, M., GARNIER, V., MONTANARO, L., FANTOZZI, G., CHEVALIER, J., Effect of initial particle packing on the sintering of nanostructured transition alumina. Journal of the European Ceramic Society, 2008, vol. 28, pp. 1121-1128. [4] LEGROS, C., CARRY, C., BOWEN, P., HOFMANN, H., Sintering of a Transition Alumina: Effects of phase Transformation, Powder Characteristics and Thermal Cycle. Journal of the European Ceramic Society, 1999, vol. 19, pp. 1967-1978. 188 References [5] CHEN, T., SOMASUNDARAN, P., Preparation of novel core-shell nanocomposite particle by controlled polymer bridging. Journal of the American Ceramic Society, 1998, vol. 81, pp. 140144. [6] HAUSSONNE, J.M., CARRY, C., BOWEN, P., BARTON,, J., Traité des matériaux: céramiques et verres – Principes et techniques d’elaboration, vol. 16, Lausanne: Polytechniques et Universitaires Romandes, 2005, pp. 815; Steitz B., Broyage et atomisation, http://mxsg3.epfl.ch/ltp/cours/LTP_TPCeramiques TP1 Broyage.pdf [7] DOCKTOR, T., WEYMOUTH, P., BARSON, T., Proper Techniques for the Ball-Milling of Ceramic Glazes. American Ceramic Society Bulletin, 1994, Vol. 73, n°1, pp. 54-58. [8] STANLEY, D.A., SADLER, L.Y., BROOKS, D.R., SCHWARTZ, M.A., Attrition Milling of Ceramic Oxides. Ceramic Bulletin, 1974, Vol. 53, n°11, pp. 813-829. [9] SADLER, L.Y., STANLEY, D.A., BROOKS, D.R., Attrition Mill Operating Characteristics. Powder Technology, 1975, Vol. 12, pp. 19-28. [10] RABE, T., LINKE, D., Attrition Milling of Silicon Nitride Powder under conditions for Minimal Impurity Pickup. Ceramics International, 1992, Vol. 12, pp. 161-166. [11] NORDAHL, C.S., MESSING, G.L., Thermal analysis of phase transformation kinetics in α-Al 2 O 3 seeded Bohemite and γ-Al 2 O 3 . Themochimica Acta, 1998, vol. 318, pp. 187-199. [12] CARRIER, X., MARCEAU, E., LAMBERT, J.F., CHE, M., Transformation of γ-alumina in aqueous suspensions 1. Alumina chemical weathering studied as a function of pH. Journal of Colloid and Interface Science, 2007, vol. 308, pp. 429-437. [13] ROELOFS, F., VOGELBERGER, W., Dissolution kinetics of nanodispersed γ-alumina in aqueous solution at different pH: Unusual kinetic size effect and formation of a new phase. Journal of Colloid and Interface Science, 2006, vol. 303, pp. 450-459. [14] POURBAIX, M., ZOUBOUV, N., VAN MUYLDER, J., Atlas d’équilibres électrochimiques. Centre Belge d’étude de la Corrosion, Gauthier-Villars & Editeurs-Imprimeur-Libraire, Paris, 1963. [15] HOFMEISTER, H., KODDERITZSCH, P., DUTTA, J., Structure of nanometersized silicon particles prepared by various gas phase processes. Journal of Non-Crystalline Solids, 1998, vol. 232, pp. 182-187. [16] DJURICIC, B., PICKERING, P., GLAUDE, P., MCGARRY, D., TAMBUYSER, P., Thermal stability of transition phases in zirconia-doped alumina. Journal of Materials Science, 1997, vol. 32, pp. 589-601. [17] YANG, R-J, YEN, F.S., LIN, S.M., CHEN, C.C., Microstructure-controlled effects on temperature reduction of α-Al 2 O 3 crystallite formation. Journal of Crystal Growth, 2007, pp. 429-435. [18] NORDAHL, C.S., MESSING, G.L., Thermal analysis of phase transformation kinetics in Al 2 O 3 seeded boehmite and γ-Al 2 O 3 . Thermochimica Acta, 1998, vol. 318, pp. 187-199. α- [19] BOWEN, P., CARRY, C., LUXEMBOURG, D., HOFMANN, H., Colloidal processing and sintering of nanosized transition aluminas. Powder Technology, 2005, vol. 157, pp. 100-107. 189 References [20] WANG, H., LIU, X., CHEN, F., HENG, G., SORENSEN, T.O., Kinetic and Mechanism of a Sintering Process for Macroporous Alumina Ceramics by Extrusion. Journal of American Ceramic Society, 1998, vol. 81, n°3, pp. 781-784. [21] HOLKOVA, Z., PACH, L., KOVAR, V., SVETIK, S., Kinetic Study of Al2O3 Sintering by Dilatometry Ceramics. Ceramics-Silikaty, 2003, vol. 47, n°1, pp. 13-19. [22] WANG, Y., SURYANARAYANA, C., AN, L., Phase transformation in nanometer-sized γ-Alumina by Mechanical Milling. Journal of the American Ceramic Society, 2005, vol. 88, n°3, pp. 780-783. [23] PACHULA, M.L., YING, J.Y., Mechanical synthesis of nanocrystalline α-Al 2 O 3 seed for enhanced transformation kinetics. Nanostructured Materials, 1997, vol. 9, pp. 161-164. [24] MORTERRA, C., MAGNACCA, G., A case study: surface chemistry and surface structure of catalytic aluminas, as studied by vibrational spectroscopy of adsorbed species. Catalysis Today, 1997, vol. 27, pp. 497-532. [25] IORDAN, A., ZAKI, M.I., KAPPENSTEIN, C., Interfacial chemistry in the preparation of catalytic potassium-modified aluminas. Journal of the Chemical Society.: Faraday Transactions, 1993, vol. 89, n°14, pp. 2527-2536. [26] WANG, S.L., JOHNSON, C.T., Assignment of the structural OH stretching bands of gibbsite. American Mineralogist, 2000, vol. 85, pp. 739. [27] ZECCHINA, A., OTERO AREAN, C., Diatomic molecular probes for mid-IR studies of zeolites. Chemical Society Reviews, 1996, vol. 25, pp. 187-197. [28] BONELLI, B., ONIDA, B., CHEN, J.D., GALARNEAU, A., DI RENZO, F., FAJULA, F., GARRONE, E., Spectroscopic chracterization of the strength and stability of the acidic sites of Al-rich microporous micelle-templated silicates. Microporous and Mesoporous Materials, 2004, vol. 67, pp. 95-106. [29] LERCHER, J.A., GRUNDLING, C., EDER-MIRTH, G., Infrared studies of the surface acidity of oxides and zeolites using adsorbed molecules. Catalysis Today, 1996, vol. 27, pp. 353-376. [30] CAIRON, O., CHEVREAU, T., LAVALLEY, J.C., Brønsted acidity of extraframework debris in steamed Y zeolites from FTIR study of CO adsorption. Journal of the Chemical Society Faraday Transactions, 1998, vol. 94, pp. 3039. [31] HADJIIVANOV, K.I., VAYSSILOV, G.N., Characterization of oxide surfaces and zeolites by carbon monoxide as an IR probe molecules. Advanced catalysis, 2002, vol. 47, pp. 307-511. [32] AREAN, C.O., MANOILOVA, O.V., TSYGANENKO, A.A., PALOMINO, G.T., MENTRUIT, M.P., GEOBALDO, F., GARRONE, E., Thermodynamics of hydrogen bonding between CO and the supercage Brønsted acid sites of H-Y zeolite – Studies from variable temperature IR spectrometry. European Journal of Inorganic Chemistry, 2001, vol. 7, pp. 1739-1743. [33] MORTERRA, C., BOLIS, V., MAGNACCA, G., IR spectroscopic and microcalorimetric characterization of Lewis acid sites on (transition-phase) Al 2 O 3 using adsorbed CO. Langmuir, 1994, vol. 10, pp. 1812-1824. [34] BOLIS, V., CERRATI, G., MAGNACCA, G., MORTERRA, C., Surface acidity of metal oxides. Combined microcalirometric and IR-spectroscopic studies of variously dehydrated systems. Thermochimica Acta, 1998, vol. 321, pp. 63. 190 References [35] DAVYDOV, A., Molecular Spectroscopy of Oxide Catalyst Surfaces. John Wiley & Sons Ltd. The Atrium, England, 2003. [36] PERI, J.B., Infrared Study of Adsorption of Carbon Dioxide, Hydrogen Chloride, and other molecules on Acid Sites on Dry Silica-Alumina and γ-Alumina. Journal of physical Chemistry, 1966, vol. 70, pp. 3168-3179. [37] MORTERRA, C., MAGNACCA, G., A case study: surface chemistry and surface structure of catalytic aluminas, as studied by vibrational spectroscopy of adsorbed species. Catalysis Today, 1997, vol. 27, pp. 497-532. [38] DAVYDOV, A.A., SHEPIT’KO, M.L., BUDNEVA, A.A., IR-Spectroscopic Methods for Studying the Basic Centers of the Surface of Oxides. Analysis of IR Spectra of Adsorbed ProbeMolecule CO 2 as a Method for Studying Basic center at the Surface of Oxides. Kinetics and catalysis, 1994, vol. 35, n°2, pp. 272-278. [39] LAVALLEY, J.C., Infrared spectrometric studies of the surface basicity of metal oxide and zeolites using adsorbed probe molecules. Catalysis Today, 1996, vol. 27, pp. 377-401. [40] ZECCHINA, A., MARCHESE, L., BORDIGA, S., PAZE, C., GIANOTTI, E., Vibrational Spectroscopy of NH4+ Ions in Zeolitic Materials: An IR Study. Journal of the Physical Chemistry B, 1997, vol. 101, pp. 10128-10135. Chapter III [1] NIIHARA, K., New Design Concept of Structural “Ceramic Nanocomposites”. The Centennial Memorial Issue of the Ceramic Society of Japan, 1991, vol. 99, n°10, pp. 974-982. [2] ZHAO, J., STEARNS, L., HARMER, M.P., CHAN, H.M., MILLER, G.A., Mechanical Behavior of Alumina-Silicon Carbide “Nanocomposites”. Journal American Ceramic Society, 1993, vol. 76, n°2, pp. 503-510. [3] KUNTZ, J.D., ZHAN, G.D., MUKHERJEE, A.K., Nanocrystalline-Matrix Ceramic Composites for Improved Fracture Toughness. MRS Bulletin, 2004, vol. 1, pp. 22-27. [4] GLEITER, M., Nanocrystalline Materials. Advanced Structural and Functional Materials, 1991, pp. 1-37. [5] BHADURI, S., BHADURI, S.B., Recent Developments in Ceramic Nanocomposites . Journal of the Mineral, Metals and Materials Society, 2007, vol. 50, pp. 1543-1551. [6] MORRIS, D.G., Mechanical behavior of nanostructured materials. Transaction Technical Publication, 1998. [7] LEE, K.M., LEE, D.J., AHN, H., XRD and TEM studied on tin oxide (II) nanoparticles prepared by inert gas condensation. Material Letters, 2004, vol. 58, pp. 3122-3125. [8] AJAYAN, P.M., SCHADLER, L.S., BRAUN, P.V., Nanocomposite Science and Technology. WILEY, Weinheim, 2003. [9] HUANG, K.L., YIN, L.G., LIU, S.Q., LI, C.J., Preparation and formation mechanism of Al 2 O 3 nanoparticles by reverse microemulsion. Transaction of Nonferreous Metallurgy Society China, 2007, vol. 17, pp. 633-637. 191 References [10] KIM, J., LEE, J.W., PARK, B.Y., CHOI, C.J., Characteristics of Fe/SiO 2 nanocomposite powders by the chemical vapor condensation process. Journal of Alloys and Compounds, 2008, vol. 449, pp. 258-260. [11] HASSANZADEH-TABRIZI, S.A., TAHERI-NASSAJ, E., Economical synthesis of Al 2 O 3 nanopowder using a precipitation method. Materials Letter, 2009, vol. 63, pp. 2274-2276. [12] KOCH, C.C., Nanostructured Materials: Processing, Properties and Potential Applications. William Andrew Publishing, New York, 2002. [13] SCHWARZ, J.A., CONTESCU, C.L., PUTYERA, K., Dekker Encyclopedia of Nanoscience and Nanotechnology. Taylor & Francis, New York, 2004. [14] DING, J., MSUZUKI, T., MCCORNICK, P.G., Ultrafine alumina particles prepared by mechanochemical/thermal processing. Journal of the American Society, 1996, vol. 79, n°11, pp. 2956-2958. [15] STERNITZKE, M., Review: Structural Ceramic Nanocomposites. Journal of the European Ceramic Society, 1997, vol. 17, pp. 1061-1082. [16] SCHEHL, M., DIAZ, L.A., TORRECILLAS, R., Alumina nanocomposites from powder-alkoxide mixtures. Acta Materialia, 2002, vol. 50, pp. 1125-1139. [17] PALMERO, P., NAGLIERI, V., CHEVALIER, J., FANTOZZI, G., MONTANARO, L., Aluminabased nanocomposites obtained by doping with inorganic salt solutions: Applications to immiscible and reactive systems. Journal of the European Ceramic Society, 2009, Vol. 29, pp. 59-66. [18] CARTER, B.C., NORTON, M.G., Ceramic Materials: Science and Engineering. Springer, Berlin, 2007. [19] COBLE, R.L., Sintering Crystalline Solids: I. Intermediate and Final State Diffusion Models. Journal of Applied Physics, 1961, Vol. 32, n° 5, pp. 787-792. [20] GROZA, J.R., Materials Processing Handbook. CRC Press, London, 2007. [21] RAHAMAN, M.N., Ceramic Processing and Sintering. Marcel Dekker, New York, 2003. [22] SHEN, Z., JOHNSSON, M., ZHAO, Z., NYGREN, M., Spark Plasma Sintering of Alumina. Journal of the American Ceramic Society, 2002, Vol. 85, n°8, pp. 1921-1927. [23] ZHAN, G.D., KUNTZ, J.D., WAN, J., MUKHERJEE, A.K., Single-wall carbon nanotubes as attractive toughening agents in alumina-based nanocomposites. Nature Materials, 2003, vol. 2, pp. 38-42. [24] GAO, L., SHEN, Z., MIYAMOTO, H., NYGREN, M., Superfast Densification of Oxide/Oxide Ceramic Composites. Journal of the American Ceramic Society, 1999, vol. 82, n°4, pp. 10611063. [25] BENGISU, M., Engineering Ceramics, Springer, Berlin, 2001. [26] LEWIS, D., RAYNE, R.J., BENDER, B.A., KURIHARA, L.K., CHOW, G.M., FLITFLET, A., KINCAID, A., BRUCE, R. Conventional and High frequency Microwave Processing of Nanophase Ceramic Materials, Nanostructured Materials, 1997, vol. 9, pp. 97-100. 192 References [27] SUTTON, W.H., Microwave processing of ceramic materials. American Ceramic Society Bulletin, 1989, vol. 68, n°2, pp. 376-389. [28] BYKOV, Y., GUSEV, S., EREMEEY, A., HOLOPSTEV, V., MALYGIN, N., PIRARUNAS, S., SOROKIN, A., SHUROV, A., Sintering of Nanophase Oxide Ceramics by using Millimeter Wave Radiation. Nanostructured Materials, 1995, vol. 6, pp. 855-858. [29] FREIM, J., MCKITTRICK, J., KATZ, J., SICKAFUS, K., Microwave Sintering of Nanocrystalline α-Al 2 O 3 . Nanostructured Materials, 1994, vol. 4, pp. 371-385. [30] TIAN, Y.L., JOHNSON, D.L., BRODWIN, M.E., Microwave Sintering of Alumina-TiC Composites. In: Ceramic Transaction, Vol. 1B, ceramic Powder Science II. Ed. By Messing, G.L., FULLER, E.R., HAUSNER, H., American Ceramic Society, Westerville, OH, 1988. [31] FRENCH, J.D., HARMER, M.P., CHAN, H.M. et al., Coarsening-resistant dual-phase interpenetrating microstructures. Journal of the American Ceramic Society, 1990, vol. 73, n°8, pp. 2508-2510. [32] ZENER, C., Kindly quoted by C.S. Smith. Transaction of the Metallurgical Society of AIME, 1949, vol. 175, n° 15. [33] LANGE, F.F., Powder Processing Science and Technology for Increased Reliability. Journal of the American Ceramic Society, 1989, vol. 72, n°1, pp. 3-15. [34] STEARNS, L., HARMER, M.P., Particle-inhibited grain growth in Al 2 O 3 -SiC: I Experimental results. Journal of the American Ceramic Society, 1996, vol. 79, n°12, pp. 3013-3019. [35] HALL, E.O.H., The deformation and ageing of mild steel: III Discussion of Results Proc. Phys. Soc. B, B64, 1951, pp. 747-753. [36] PETCH, N.J., The cleavage strength of polycrystals. J. Iron Steel Inst., 174, 1953, pp. 25-28. [37] RICE, R.W., WU, C.C., BORCHANLT, F., Hardness-Grain Size Relation in Ceramics. Journal of the American Ceramic Society, 1994, vol. 77, n°10, pp. 2539-2553. [38] SHEN, Z., JOHNSSON, M., ZHAO, Z., NYGREN, M., Spark Plasma Sintering of Alumina. Journal of the American Ceramic Society, 2002, vol. 85, n°8, pp. 1921-1927. [39] AWAJI, H., CHOI, S.M., Review: Ceramic-based nanocomposites. Res. Devel. Mat. Sci. Eng., 2002, vol. 1, pp. 585-597. [40] OHJI, T., KUSUNOSE, T., NIIHARA, K., Threshold Stress in Creep of Alumina-Silicon Carbide Nanocomposites. Journal of the American Ceramic Society, 1998, vol. 81, n°10, pp. 27132716. [41] BORSA, C.E., JONES, N.M., BROOK, R.J., TODD, R.I., Influence of Processing on the Microstructural development and flexure strength of Al2O3/SiC nanocomposites. Journal of the European Ceramic Society, 1997, Vol. 17, pp. 865-872. [42] DAVIDGE, R.W., BROOK, R.J., CAMBIER, F., Fabrication, properties, and modelling of engineering ceramics reinforced with nanoparticles of silicon carbide. British Ceram. Trans, 1997, Vol. 96, pp. 121-127. 193 References [43] CHOU, I.A., CHAN, H.M., HARMER, M.P., Effect of annealing environment on the crack healing and mechanical behaviour of silicon carbide-reinforced alumina nanocomposites. Journal of the American Ceramic Society, 1998, Vol. 81, n°5, pp. 1203-1208. [44] GAO, L., WANG, H.Z., HONG, J.S., MIYAMOTO, H., MIYAMOTO, K., NISHIKAWA, Y., TORRE, S.D.D.L., Mechanical Properties and Microstructure of Nano-SiC-Al 2 O 3 Composites Densified by Spark Plasma Sintering. Journal of the European Ceramic Society, 1999, vol. 19, pp. 609-613. [45] ZHAN, G.J., YANG, J.F., ANDO, M., OHJI, T. Reactive Hot Pressing of Alumina-Silicon Carbide Nanocomposites. Journal of the American Ceramic Society, 2004, vol. 87, n°2, pp. 299-301. [46] CHANTIKUL, P., BENNISON, S.J., LAWN, B.R., Role of Grain Size in the Strength and RCurve Properties of Alumina. Journal of the American Ceramic Society, 1990, vol. 73, n°8, pp. 2419-2427. [47] ZHAO, J., STERNS, L.C., HARMER, M.P., CHAN, H.M., MILLER, G.A., COOK, R.E., Mechanical behavior of aluminas-silicon carbide nanocomposites. Journal of the American Ceramic Society, 1993, vol. 76, pp. 503-510. [48] DERBY, B., Ceramic nanocomposites: mechanical properties. Current Opinion in solid State & Materials Science, 1998, vol. 3, pp. 490-495. [49] GREEN, D.J., Fracture toughness predictions for crack bowing in brittle particulate composites. Journal American Ceramic Society, 1983, vol. 66, C4-C5. [50] PEZZOTTI, G., NISHIDA, T., SAKAI, M., Physical limitations of the inherent toughness and strength in ceramic-ceramic and ceramic-metal nanocomposites. Journal of the Japanese Ceramic Society, 1995, vol. 103, pp. 901-909. [51] CLAUSSEN, N., Fracture Toughness of AI 2 O 3 with an Unstabilized ZrO 2 Dispersed Phase. Journal of the American Ceramic Society, 1976, Vol. 59, N°1, pp. 49-51. [52] BHADURI, S., BHADURI; S.B., Enhanced Low Temperature Toughness of Al 2 O 3 -ZrO 2 Nano/Nano Composites. NanoStructured Materials, 1997, vol. 8, n°6, pp. 755-763. [53] KIM, B.N., HIRAGA, K., SAKKA, Y., JANG, B.K, Effect of cavitations on superplastic flow of 10% Zirconia-dispersed alumina. Scripta Materialia, 2001, Vol. 45, pp. 61-67. [54] ZHAN, G.D., KUNTZ, J., GARAY, J., MUKHERJEE, A.K., A Novel Processing Route to Develop a Dense Nanocrystalline Alumina Matrix (<100 nm) Nanocomposite Material. Journal of the European Ceramic Society, 2002, vol. 86, n°1, pp. 200-202. [55] SADANGI, R.K., SHUKLA, V., KEAR, B.H., Processing and properties of ZrO 2 (3Y 2 O 3 )-Al 2 O 3 Nanocomposites. International Journal of Refractory Metals & Hard Materials, 2005, vol. 23, pp. 363-368. [56] VASYLKIV, O., SAKKA, Y., SKIROKHOD, V. V., High-Toughness Tetragonal Zirconia/AluminaCeramics. Key Engineering Materials, 2006, vol. 317-318, pp. 615-618. [57] De AZA, A.H., CHEVALIER, J., FANTOZZI, G., SCHEHL, M., TORECILLAS, R., Crack growth resistance of alumina, Zirconia and Zirconia toughened alumina ceramics for joint prostheses. Biomaterials, 2002, vol. 23, pp. 937-945. 194 References [58] NIIHARA; K:; UNAL, N., NAKAHIRA; A:, Mechanical properties of (Y-TZP)-alumina-silicon carbide nanocomposites and the phase stability of y-TZP particles in it. Ceram. Soc. Japan, 1991, pp. 404-417. [59] NAWA, M., BAMBA, N., SEKINO, T., NIIHARA, K., The effect of TiO 2 Addition on Strengthening and Toughening in intragranular Type of 12Ce-TZP/Al 2 O 3 Nanocomposites. Journal of the American Ceramic Society, 1998, vol. 18, pp. 209-219. [60] TAN, H., YANG, W., Toughening mechanisms of nano-composite ceramics. Mechanics of Materials, 1998, vol. 30, pp. 111-123. [61] NORTON, F.H., Models for Creep of Fibrous Composite Materials. Materials Forum, 1929, vol. 11, pp. 133-139. [62] FANTOZZI, G., CHEVALIER, J., OLAGNON, C., CHERMANT, J.L, Creep of ceramic matrix composites. In: Warren R (eds) Comprehensive Composite Materials, 2000, Vol. 4 Carbon/Carbon. Cement and Ceramic Matrix Composites. Elsevier, Oxford, UK, Chapter 7. [63] NABARRO, F.R.N., Deformation of Crystals by Motion of Single ions. Proceedings of the Bristol Conference in Strength of Solids. The Physical Society, 1948, pp. 75-90. [64] HERRING, C., Diffusional Viscosity of a Polycrystalline Solid. Journal of Applied Physics, 1950, vol. 21., pp. 437-445. [65] COBLE, R.L., A Model of Boundary Diffusion Controlled Creep in Polycrystalline Materials. Journal of Applied Physics, 1963, vol. 34, pp. 1679-1682. [66] CANNON, W.R., LANGDON, T.G., Review, Creep of Ceramics, Part 1 Mechanical Characteristics. Journal of Materials Science, 1983, vol. 18, pp. 1-50. [67] CARTER, C.B., NORTON, M.G., Ceramic Materials: Science and Engineering. Springer, New York, 2007. [68] OHJI, T., NAKAHIRA, A., HIRANO, T., NIIHARA, K., Tensile Creep Behavior of Alumina/Silicon Carbide Nanocomposite. Journal of the American Ceramic Society, 1994, vol. 77, n°12, pp. 3259-3262. [69] DESCAMPS, P., O’SULLIVAN, D., POORTEMAN, M., DESCAMPS, J.C., LERICHE, A., CAMBIER, F., Creep Behaviour of Al 2 O 3 -SiC Nanocomposites. Journal of the European Ceramic Society, 1999, vol. 19, pp. 2475-2485. [70] THOMPSON, A.M., CHAN, H.M., HARMER, M.P., Tensile Creep of Alumina-Silicon Carbide “Nanocomposites”. Journal of the American Ceramic Society, 1997, vol. 80, n°9, pp. 22212228. [71] OHJI, T., HIRANO, T., NAKAHIRA, A., NIIHARA, K., Particle/Matrix Interface and Its Role in Creep Inhibition in Alumina/Silicon Carbide. Journal of the American Ceramic Society, 1996, vol. 79, n°1, pp. 33-45. [72] ARZT, E., GRAHLE, P., High temperature creep behaviour of oxide dispersion strengthened NiAl intermetallics. Acta materialia, 1998, vol. 46, n°8, pp. 2717-2727. [73] ALPER, A.M., Phase Diagrams in Advanced Ceramics. Academic Press, London, 1995. 195 References [74] NOGUCHI, T., MIZUNO, M., Phase diagrams for Ceramics 1975 Supplement. Levin and Mc Murdie Eds., pp. 132, American Ceramic Society, Westerville OH, 1975. [75] TOROPOV, N.A., BONDAR, I.A., GALAKHOV, F.Y. VINOGRADOVA, Phase diagrams for Ceramics 1969 Supplement, Levin et al. Eds., Westerville OH, 1969. [76] CASLAVSKY, R.L., VIECHNICKI, D.J., Melting behaviour and metastability of yttrium aluminium (YAG) and Al 2 O 3 determined by optical differential thermal analysis. Journal Material Science, 1980, vol. 15, pp. 1709-1719. [77] LIU, Y., ZHANG, Z.F., KING, B., HALLORAN, J., LAINE, R.M., Synthesis of Yttrium Aluminium Garnet from Yttrium and Aluminium Isobutyrate Precursors. Journal of the American Ceramic Society, 1996, vol. 79, n°2, pp. 385-394. [78] LIU, Y., ZHANG, Z.F., HALLORAN, J., LAINE, R.M., Yttrium Aluminium Garnet Fibers from Metalorganic Precursors. Journal of the American Ceramic Society, 1998, vol. 81, n°3, pp. 1029-1032. [79] GOWDA, G., Synthesis of yttrium aluminates by the sol-gel process. Journal of Material Science Letters, 1986, vol. 5, n°10, pp. 1029-1032. [80] KAMAT, R.V., PILLAI, K.T., VAIDYA, V.N., SOOD, D.D., Synthesis of yttrium aluminium garnet by the gel entrapment technique using hexamine. Materials Chemistry and Physics, 1996, vol. 46, n°1, pp. 67-71. [81] HAY, R.S., Phase transformation and microstructure evolution in sol-gel derived yttrium aluminium garnet films. Journal of Materials Research, 1993, vol. 8, pp. 578-604. [82] YAMAGUCHI, K., TAKEOKA, K., HIROTA, K., TAKANO, H., HAYASHIDA, A., Formation of Alkoxy-Derived Yttrium Aluminium Oxides. Journal Materials Science, 1992, vol. 27, pp. 12611264. [83] YAMAGUCHI, O., TAKEOKA, K., HAYASHIDA, A., Formation of alkoxyde-derived Y 3 Al 5 O 12 . Journal of Materials Science Letters, 1991, vol. 10, pp. 101. [84] INOUE, M., OTSU, H., KONINAMI, H., INUI, T., Synthesis of yttrium aluminium garnet by glycothermal method. Journal of the American Ceramic Society, 1991, vol. 74, pp. 1452-1454. [85] ROTH, R.S., Phase Equilibrium Diagrams: Phase Diagrams for Ceramists, Vol. XI. The American Ceramic Society, Westville, OH, 1995, pp.107. [86] KELLER, K., MAH, T., PARTHASARATHY, T., Processing and mechanical properties of polycrystalline Y3Al5O12. Ceram. Eng. Sci. Proc., 1990, Vol. 1, pp. 1122-1133. [87] MAH, T., PARTHASARATHY, T., MATSON, L., Processing and mechanical properties of Al 2 O 3 /Al 3 Al 5 O 12 (YAG) eutectic composite. Ceram. Eng. Sci. Proc., 1990, Vol. 11, pp. 16171633. [88] WAKU, Y., NAKAGAWA, N., WAKAMOTO, T., OHTSUBO, H., SHIMIZU, K., KOHTOKU, Y., The Creep and Thermal Stability Characteristics of Unidirectionally Solidified Al 2 O 3 /YAG Eutectic Composite. J. Mater. Sci., 1998, Vol. 33, pp. 1217. [89] VROLIJK, J.W., WILLEMS, J.W., METSELAAR, R., Coprecipitation of yttrium and aluminum hydroxides for preparation of yttrium aluminum garnet. Journal of the European Ceramic Society., 1990, Vol. 6, pp. 47–51. 196 References [90] LI, W.Q., GAO, L., Processing, Microstructure and Mechanical Properties of 25 vol.% YAGAl 2 O 3 nanocomposites. Nanostructured Materials, 1999, Vol. 11, N°8, pp. 1073-1080. [91] FRENCH, J.D., CHAN, H.M., HARMER, M.P., MILLER, G.A., High-Temperature Fracture Toughness of Duplex Microstructures. Journal of the American Ceramic Society, 1996, Vol. 79, n°1, pp. 58-64. [92] DUONG, H., WOLFENSTINE, J., Creep Behavior of Fine Grained Two-Phase Al 2 O 3 -YAG Materials, Mater. Sci. Eng., 1993, A 173, pp. 173-179. [93] FRENCH, J.D., ZHAO, ,J., HARMER, M.P., CHAN, H.M., MILLER, G.A., Creep of duplex microstructures. Journal of the American Ceramic Society, 1994, vol. 77, pp. 2857-2865. [94] KOTTADA, R.S., CHOKSHI, A.H., The High-Temperature Tensile and Compressive Deformation of Magnesia Doped Alumina. Acta Materalia, 2000, vol. 48, pp. 3905-3915. [95] SATAPATHY, L.N., CHOKSHI, A.H., Microstructural Development and Creep Deformation in an Alumina- 5 % Yttrium Aluminium Garnet Composite. Journal of the American Ceramic Society, 2005, vol. 88, n°10, pp. 2848-2854. [96] TORRECILLAS, R., SCHEHL, M, DIAZ, L.A., Creep behaviour of alumina-mullite-zirconia nanocomposites obtained by a colloidal processing route. Journal of the European Ceramic Society, 2007, vol. 27, pp. 143-150. [97] PARTHASARATHY, T.A., MAH, T., KELLER, K., Creep Mechanisms of Polycrystalline Yttrium Aluminium Garnet. Journal of the American Ceramic Society, 1992, vol. 75, pp. 1756-175. Chapter IV [1] NIIHARA, K., New Design Concept of Structural “Ceramic Nanocomposites”. The Centennial Memorial Issue of the Ceramic Society of Japan, 1991, vol. 99, n°10, pp. 974-982. [2] AWAJI, H., CHOI, S.M., Review: Ceramic-based nanocomposites. Res. Devel. Mat. Sci. Eng., 2002, vol. 1, pp. 585-597. [3] KELLER, K., MAH, T., PARTHASARATHY, T.: Processing and mechanical properties of polycrystalline properties of polycrystalline Y 3 Al 5 O 12 . Ceramic Engineering and Science Proceedings, 1990, 11, pp. 1122-1133. [4] PALMERO, P., NAGLIERI, V., CHEVALIER, J., FANTOZZI, G., MONTANARO, L., Aluminabased nanocomposites obtained by doping with inorganic salt solutions: Applications to immiscible and reactive systems. Journal of the European Ceramic Society, 2009, Vol. 29, pp. 59-66. [5] FRENCH, J.D., ZHAO, J., HARMER, M.P., CHAN, H.M., MILLER, G.A., Creep duplex microstructures. Journal of the American Ceramic Society, 1994, vol. 77, pp. 2857-2865. [6] YOSHIDA, H., IKUHARA, Y., SAKUMA, T., High-temperature creep resistance in rare-earthdoped fine-grained Al2O3. Journal of Materials research, 1998, vol. 13, pp. 2597-2601. [7] TORRECILLAS, R., SCHEHL, M., DIAZ, L.A., MENENDEZ, J.L., MOYA, J.S., Creep behaviour of alumina/YAG nanocomposites obtained by a colloidal processing route. Journal of the European Ceramic Society, 2007, vol. 27, pp. 143-150. 197 References [8] Taimei Chemical Co., LTD. [on-line]. Available at: http://taimei-chem.co.jp/ (consulted 24.11.2009) [9] Baikowski Chimie 24.11.2009) [on-line]. Available at: http://www.baikowskichimie.com/ (consulted [10] GAO, L., SHEN, Z., MIYAMOTO, H., NYGREN, M., Superfast Densification of Oxide-Oxide Ceramic Composites. Journal of the American Ceramic Society, 1999, vol. 82, n°4, pp. 10611061. [11] HALL, E.O.H., The deformation and ageing of mild steel: III Discussion of Results Proc. Phys. Soc. B, B64, 1951, pp. 747-753. [12] PETCH, N.J., The cleavage strength of polycrystals. J. Iron Steel Inst., 174, 1953, pp. 25-28. [13] WANG, H., GAO, L., Preparation and microstructure of polycrystalline Al 2 O 3 -YAG composites. Ceramics International 27 (2001), 721-723. [14] ANSTIS, G.R., CHANTIKUL, P., LAWN, B.R., MARSHAL, D.B., A critical evaluation of indentation techniques for measuring fracture toughness. I. Direct crack measurements. Journal of the American Ceramic Society, 1981, vol. 64, pp. 533-538. [15] ASTM Designation: C 1259-01. Standard Test Method for Dynamic Young’s Modulus, Sher Modulus, and Poisson’s Ration for Advanced Ceramics by Impulse Excitation of Vibration. ASTM Committee on Advance Ceramics, 2001. [16] RICE, R.W., Mechanical Properties of ceramics and Composites: grain and Particle effects. Marcel Dekker, Inc., New York. Basel, 2000. [17] SWAIN, M.V., Structure and properties of Ceramics, Vol. 11, Materials Science and Technology, Cahn R.W. Haasen P., Kramer, E.J, VCH Weinheim, 1994. [18] SCHEHL, M., DIAZ, L.A., TORRECILLAS, R., Alumina nanocomposites from powder–alkoxide mixtures. Acta Materialia 50 (2002) 1125–1139. [19] HOLLEMBERG, G.W., TERWINLLINGER, G.R., GORDON, R.S., “Calculation of stresses and strains in four point bending creep tests”, Journal of the American Ceramic Society, 1971, vol. 54, pp. 196-199. [20] SATAPATHY, L.N., CHOKSHI, A.H., Microstructural development and creep deformation in an alumina-5% yttrium aluminium garnet composite. Journal of the American Ceramic Society, 2005, vol. 88, pp. 2848-2854. [21] LANGDON, T.G., Dependence of Creep Rate on porosity. Journal of the American Ceramic Society, 1962, vol. 45, pp. 630-631. [22] NABARRO, F.R.N., Deformation of Crystals by Motion of Single ions. Proceedings of the Bristol Conference in Strength of Solids. The Physical Society, 1948, pp. 75-90. [23] HERRING, C., Diffusional Viscosity of a Polycrystalline Solid. Journal of Applied Physics, 1950, vol. 21., pp. 437-445. [24] COBLE, R.L., A Model of Boundary Diffusion Controlled Creep in Polycrystalline Materials. Journal of Applied Physics, 1963, vol. 34, pp. 1679-1682. 198 References [25] CHEVALIER, J., OLAGNON, C., FANTOZZI, G., Creep behaviour of alumina, zirconia and zirconia toughened alumina. Journal of the European Ceramic Society, 1997, vol. 17, pp. 859-864. 199 Appendix A 200 Appendix A1 1. Laser Granulometry Figure A1 – Granulometer Fritsch model Analysette 22 Compact. For the development of this PhD thesis, a laser particle size analyser Fritsch model Analysette 22 Compact (Figure A1) was employed to evaluate the particle size distribution. The granulometer is suitable for wet measurements of granulometric powder with particle size ranging from 0.31 to 300.74 μm range. Laser granulometry is a technique based on the light diffraction. It is based on the Fraunhofer theory, using the following hypotheses: Spherical particles are considered to be non porous and non opaque. The diameter of the particles is higher compared with laser wavelength. The particles are in constant motion. The particles diffract light efficiently, regardless of their size. When a laser beam sheds light on a particle, diffraction patterns can be observed. The intensity of the diffracted radiation and the deviation differ according to the size of the particles. The larger the particle, more light will deviate and the weaker its deviation angle, in relation to the propagation will be. The first step is to dilute the sample. The apparatus measure the background in order to record the diffraction phenomena caused by the solvent. Subsequently, the sample solution is injected into the measuring cell, each particle that passes through the radiation beam deviates the light which is analysed by detectors. 2. X-ray diffraction (XRD) The most widespread use of X-ray powder diffraction is for the identification of crystalline compounds by their diffraction pattern. This experimental method is employed to determine the structure of the crystalline materials and to measure the crystalline size by the Scherrer method (explained in chapter I). XRD diffraction analysis were performed by using a diffractometer Philips PW 1710, with a Cu K α radiation (λ=1.5405600 Å). In Figure A2 it is illustrated the instrument in detail. 201 Appendix A1 Figure A2 – Schematic representation of the diffractometer Philips PW 1710. In X-ray powder diffractometry, X-rays are generated within a sealed tube under vacuum. A current is applied which heats a filament within the tube, the higher the current, the greater the number of electrons emitted from the filament. A high voltage, typically 15-60 kilovolts, is applied within the tube. This high voltage accelerates the electrons, which then hit a target, commonly made of copper. When these electrons hit the target, X-rays are produced. The wavelength of these x-rays, is characteristic of that target. These X-rays are collimated and directed onto the sample, which has been ground to a fine powder (typically to produce particle sizes of less than 10 microns). A detector detects the X-ray signal; the signal is then processed either by a microprocessor or electronically, converting the signal to a count rate, changing the angle between the X-ray source, the sample and the detector at a controlled rate by means of a goniometer. When an X-ray beam hits a sample and is diffracted, we can measure the distances between the planes of the atoms by applying Bragg's Law. Bragg's Law is n. 2.d .sin , where the integer n is the order of the diffracted beam, 1 is the wavelength of the incident X-ray beam, d is the distance between adjacent planes of atoms (the d-spacings), and θ is the angle of incidence of the X-ray beam. 202 Appendix A1 Figure A3 – Illustration of the goniometer and angles involved. Since we know λ and we can measure θ, we can calculate the d-spacings. The geometry of an XRD unit is designed to accommodate this measurement (Figure A3).The characteristic set of d-spacing generated in a typical X-ray scan provides a unique pattern of the material. When properly interpreted, by comparison with standard reference patterns from ICDD (International Centre for Diffraction Data) files, this "fingerprint" allows for identification of the material. 3. Thermal analysis: DTA-TG Differential thermal analysis involves heating or cooling a test sample and an inert reference under identical conditions, while recording any temperature difference between the sample and reference. This differential temperature is then plotted against time, or against temperature. Changes in the sample which lead to the absorption or evolution of heat can be detected compared to the inert reference. Differential temperatures can also arise between two inert samples when their response to the applied heat treatment is not identical (Figure A4). DTA can therefore be used to study thermal properties and phase changes which do not lead to a change in enthalpy. The baseline of the DTA curve should then exhibit discontinuities at the transition temperatures and the slope of the curve at any point will depend on the microstructural constitution at that temperature. Figure A4- Schematic illustration of DTA-TG cell. The area under a DTA peak can be imputed to the enthalpy change and it is not affected by the heat capacity of the sample. DTA may be defined formally as a technique for recording the 203 Appendix A1 difference in temperature between a substance and a reference material against either time or temperature as the two specimens are subjected to identical temperature regimes in an environment heated or cooled at a controlled rate. Simultaneously, thermogravimetric (TG) is also carried out, which follows the mass loss of the sample during heating as a result of decomposition, desorption and/or dehydration phenomena. In our particular case, the tests were carried out by using a system of simultaneous DTA-TG Netzsch STA 409C which also provides the possibility of employing the machine as DSC-TG. Pictures of the device are shown below in Figure A5. Figure A5- Thermal analysis device (DTA-TG) 4. Dilatometry A dilatometer Netzsch 402E was employed with the aim to study the sintering behaviour and improving the sintering conditions in order to achieve the highest density. The temperature limit of the machine is 1550°C. Ceramic powders were pressed into bars, introduced into the sample holder and submitted to a controlled thermal cycle. During the test, the length variation is measured as a function of the time. In this particular case, the displacement is measured by a transducer as it is shown in Figure A6. This data permits to calculate the coefficient of thermal expansion by the following equation. L f Li Lo .T L (1) Lo .T 204 Appendix A1 where ΔT is the variation of temperature, L o and L f are the initial and final length of the sample. Figure A6- Netzsch 402E dilatometer. 5. Scanning Electron Microscopy/Environmental Scanning Electron Microscopy The SEM is an instrument that produces a largely magnified image by using electrons instead of light to form an image. A beam of electrons is produced at the top of the microscope by an electron gun. The electron beam follows a vertical path through the microscope, which is held within a vacuum. The beam travels through electromagnetic fields and lenses, which focus the beam down toward the sample. Once the beam hits the sample, electrons and X-rays are ejected from the sample. Detectors collect these X-rays, backscattered electrons, and secondary electrons and convert them into a signal that is sent to a screen similar to a television screen. This produces the final image. In our case, a microscope Hitachi S2300 has been used. Samples are coated by gold sputtering in order to increase the electrical conductivity. The second technique employed to observe the surface of the samples, was the environmental scanning electron microscope (ESEM). An ESEM microscope used was a FEI XL30 ESEM FEG. This technique permits the image of wet systems with no prior specimen preparation. Additionally, sample environment can be dynamically 205 Appendix A1 altered, hydration and dehydration processes can be followed inside the sample chamber. A schematic representation of the microscope is shown in Figure A7. Fig A7- ESEM microscopy schematic diagram. 6. High Resolution Transmission Electron Microscopy (HR-TEM) Transmission Electron Microscopy (TEM) is a technique where an electron beam interacts and passes through a specimen. The electrons are emitted by a source and are focused and magnified by a system of magnetic lenses. The geometry of TEM is shown in Figure A8. The electron beam is confined by the two condenser lenses which also control the brightness of the beam, passes the condenser aperture and hits the sample surface. The electrons that are elastically scattered consist of transmitted beams, which pass through the objective lens. In this PhD thesis, HRTEM (High Resolution Transmission Electron Microscopy) pictures were collected on a JEOL 3010-UHR instrument, operated at 300kV and equipped with a 2kx2k pixel Ultrascan 100 camera. 206 Appendix A1 Figure A8- TEM microscopy schematic diagram. The objective lens forms the image display and the following apertures, the objective and the selected area aperture are used to choose the elastically scattered electrons that will form the image of the microscope. Finally, the beam goes to the magnifying system that consisted of three lenses, the first and second intermediate lenses control the magnification of the image and the projector lens. The operation of TEM requires an ultra high vacuum and a high voltage. The first step is to find the electron beam, so the lights of the room must be turned off. Through a sequence of buttons and adjustments of focus and brightness of the beam, we can adjust the settings of the microscope so that by shifting the sample holder we find the thin area of the sample. Then tilting of the sample begins by rotating the holder. Different types of images are obtained in TEM, using the apertures properly and the different types of electrons. As a result, diffraction patterns are shown because of the scattered electrons. If the unscattered beam is selected, we obtain the Bright Field Image. Dark Field Images are attained if diffracted beams are selected by the objective aperture. Also in TEM, analysis is done with EDX (Energy Dispersive X-ray), EELS (Electron Energy Loss Spectrum), EFTEM (Energy Filtered Transmission Electron Microscopy),etc. In transmission microscopy, we can actually see the specimen’s structure and its atomic columns, thus compositional and crystallographic information is achieved. However, it is a very expensive technique, expertise is needed and the sample preparation phase is too difficult so that very thin samples are achieved. The first step is to decide whether the sample is useful to be observed and in which view, plan or cross-section. Due to the strong interaction between electrons and matter, the specimens have to be rather thin, less than 100nm. This is achieved with several methods, depending on the material. In general, mechanical thinning is used to thin and polish the sample. Then it is glued with epoxy glue on a really small and round holder. Whereas TEM data come from the edges of a hole in the centre of the specimen, in sample preparation, the hole is created by the method of ion thinning. 207 Appendix A1 Ion thinning is a method where a specimen is usually irradiated with beams of Ar ions, and after a period of time, a hole is created. To minimize the damage created during focus ion beam milling, the embedded sample can first be coated with a metal deposition layer. Consequently, sample preparation is a precise and a severe procedure, which may affect the results of the microscopic analysis and study. TEM provides accurate measurements and studies in different types of materials, since observations are in atomic scale in HRTEM. This is due to technology that reduces the errors and corrects more and more the interferences in formed images. In order to improve the results of TEM, ultra high vacuum with no vibrations is needed and this fact results in the production of different types of pumps such as mechanical pumps, oil diffusion pumps, ion getter pumps and cooled stage. Higher voltage up to 3MV and small probe size were developed, and methods to assure monochromaticity and coherency of the electrons. This is a way to avoid chromatic aberration and «spherical aberration», the most usual errors in electron microscopy. Nowadays, high resolution transmission electron microscopes offer resolutions up to 0.1 nm at 300kV and probe diameters up to 0.34nm. Thus, future trends include the use of ultrahigh vacuum TEM instruments for surface studies and computerized data acquisition for quantitative image analysis. 7. Fourier Transformed Infrared Device (FT-IR) For FT-IR (Fourier Transform Infra-Red) measurements, powder samples were pressed into selfsupporting wafers. Spectra were collected at a resolution of 2 cm-1, on a Bruker FTIR Equinox 55 spectrophotometer equipped with a MCT detector. The schematic representation of the apparatus is shown in Figure A9. Figure A9- Bruker FTIR Equinox 55 spectrophotometer. Fourier Transform Infrared (FT-IR) spectrometry was developed, in order to overcome the limitations encountered with dispersive instruments. The main difficulty was the slow scanning process. A method for measuring all of the infrared frequencies simultaneously, rather than individually, was needed. 208 Appendix A1 A solution was developed and it exploited a very simple optical device called an interferometer. The interferometer produces a unique type of signal which has all of the infrared frequencies “encoded” into it. The signal can be measured very quickly, usually in the order of one second or so. Thus, the time element per sample is reduced to a matter of a few seconds rather than several minutes. The schematic diagram is shown in Figure A10. Figure A10- The normal instrumental process. Most interferometers employ a beamsplitter which takes the incoming infrared beam and divides it into two optical beams. One beam reflects off of a flat mirror which is fixed in place. The other beam reflects off from a flat mirror which is on a mechanism that allows this mirror to move a very short distance (typically a few millimeters) away from the beamsplitter. The two beams reflect off from their respective mirrors and are recombined when they meet back at the beamsplitter. Since the path that one beam travels is a fixed length and the other is constantly changing as its mirror moves, the signal which exits the interferometer is the result of these two beams “interfering” with each other. The resulting signal is called an interferogram which has the unique property that every data point (a function of the moving mirror position) which makes up the signal that has information about every infrared frequency which comes from the source. This means that as the interferogram is measured, all frequencies are being measured simultaneously. Thus, the use of the interferometer results in extremely fast measurements. Since the analyst requires a frequency spectrum (a plot of the intensity at each individual frequency) in order to make an identification, the measured interferogram signal can not be interpreted directly. A means of “decoding” the individual frequencies is required. This can be accomplished via a wellknown mathematical technique called the Fourier transformation. This transformation is performed by the computer which then presents the user with the desired spectral information for analysis. 8. Specific Surface Area BET (Brunauer, Emmett, Teller) specific surface areas (SSA) were measured by means of N 2 adsorption/desorption isotherms at 77 K performed by using a Quantachrome Autosorb 1C instrument (Figure A11). The gas adsorption method permits to measure the amount of gas adsorbed on the surface of a powder sample as a function of the pressure of the adsorbate gas, and it is used to determine the specific surface area of a powder sample. The concept theory is an extension of the Langmuir theory, which is a theory for monolayer molecular adsorption, to multilayer adsorption with the following hypotheses: Gas molecules physically adsorb on a solid in layers infinitely. 209 Appendix A1 There is no interaction between each adsorption layer. Langmuir theory can be applied to each layer. Taking into account these conditions, the resulting BET equation is shown below. 1 C 1 P 1 P VmC Po VmC Va o 1 P (2) where P is the partial vapor pressure of adsorbate gas in equilibrium, P o is the saturated pressure of the adsorbate gas, V a is the volume of the gas adsorbed at equilibrium, V m is the volume of the gas adsorbed in a monolayer and C is a dimensionless constant related to the enthalpy of adsorption and condensation of the adsorbate gas. Figure A11 - Quantachrome Autosorb 1C Instrument. The specific surface area, SSA, is determined from V m , the volume of gas adsorbed in a monolayer on the sample. S Vm .N .a m (3) where S is the specific surface area, N is the Avogadro constant, a is the effective cross-sectional area of one adsorbate molecule and m is the mass of the test powder. 210 Appendix A1 Figure A11 – Schematic Diagram of the device. Test powder is placed in a sample contained with a known volume (Figure A11) and the volume of gas adsorbed is determined from the change in the pressure associated with the adsorption of gas on the surface of the sample powder. Initially, a pre-treatment is performed in order to remove gases or vapours, that have been physically adsorbed onto the sample surface, by outgassing the sample under reduced pressure. After pre-treatment is completed, the sample container should be precisely weighed with the sample and the mass of the tared contained measured previously should be subtracted, to obtain the true mass of the powder. A fixed quantity of the adsorbate gas is introduced intro the sample container surrounded by liquid nitrogen. The pressure decreases until gas/solid adsorption reaches to a new equilibrium. The volume of gas adsorbed is calculated from the difference between the volume of the un-adsorbed gas and the volume remaining in the void volume. For a multiple point BET analysis, it is possible to calculate the specific surface area by repeating the measurements under equilibrium pressure of the adsorbate gas in the range of 0.05 to 0.30. 9. Young's Modulus Resonant Frequency Meter: Grindosonic The Young’s modulus was determined the Grindosonic method, which allows measurement of the resonant frequency of a ceramic sample at room temperature from which it can be obtained the Young's modulus of the sample as described in ASTM C885. The machine is shown in Figure A12. The method consists on inducing a vibration on a parallelepiped cylinder along its longest dimension, a detector transforms these mechanical vibrations into an electrical signal that can be registered. Then, assuming a value of Poisson’s constant it can be obtained Young's modulus of the material. The Young's modulus is an intrinsic property of materials of extreme importance in the design of materials and facilities and that affects other material properties such as resistance to thermal shock. 211 Appendix A1 Figure A12 – Grindosonic machine. 10. Hardness Tests Figure A13 - Vickers Testwell FV-700 The indentation tests were performed using the Vickers Testwell FV-700 device (Figure A13), which has the possibility of varying the charge from 0.3 to 30 kg. The charge was fixed in 98.1N (10 kg). Five indentations were performed in each samples. The toughness was determined by indentation tests by using Anstis’ formula, as it is shown below. 1/ 2 K IC E P A. . 2 / 3 H c c : length of the crack from the center of the impression A : geometric constant equal to 0.016. P : change expressed in Newton - in our case equal to 98.1N. H : Hardness value obtained directly from the machine. E : Young’s Modulus. In our case it was determined by the volume fraction law shown below. E f Al2O3 .E Al2O3 fYAG .EYAG 212 Appendix A1 d 2c Figure A14 - Optical microscope’s observation of an impression. 11. Creep Tests Creep resistance tests were carried out in a 4-point bending fixture in a lever arm machine (Figure A15 (A)). Samples were fixed in an alumina holder with distances between the different points of 28 and 14 mm (Figure A15 (B)). Bar-shaped samples were parallelepipeds of 32 x 4 x 3 mm3. In order to avoid surface flaws the edges were chamfered at about 45°. Additionally, samples were polished with diamond paste up to 3 μm. Inside the sample container the samples were centred with a bar-shaped block with the aim of assuring the parallelism of the faces. The maximum load allowed for the test is 500 N. Whereas, the maximum temperature of the machine is 1400°C. Thermocouples inside the chamber have a precision of +/- 3 degrees. Deformation was recorded by a LVDT (Linear variable differential transformer) sensor with a resolution of 1 μm. The heating rate for all the test was 300°C/min in order to homogenise the temperature of the chamber. During tests two variables were simultaneously recorded: the displacement and the temperature as a function of the time. The machine is presented in the Figure A15. 213 Appendix A1 Figure A15- Creep machine (A) and sample holder (B). 214 FOLIO ADMINISTRATIF THESE SOUTENUE DEVANT L'INSTITUT NATIONAL DES SCIENCES APPLIQUEES DE LYON NOM : LOMELLO DATE de SOUTENANCE : 20 mai 2010 (avec précision du nom de jeune fille, le cas échéant) Prénoms : FERNANDO TITRE : Optimization of nanostructured oxide-based powders by Surface modification NATURE : Doctorat Numéro d'ordre : 2010-ISAL Ecole doctorale : École Doctorale Matériaux de Lyon Spécialité : Génie des Matériaux Cote B.I.U. - Lyon : T 50/210/19 / et bis CLASSE : RESUME: Ce travail a été consacré d’une part à l’étude de l'effet de la dispersion dans une alumine de transition commerciale et d’autre part à l’étude des propriétés mécaniques à basse et haute température des nanocomposites Al 2 O 3 -5 vol. % YAG. Pour étudier l’effet de la dispersion dans l’alumine de transition, différentes techniques de caractérisation appartenant à la physico-chimie des surfaces et à la science des matériaux ont été utilisées comme l’analyse thermogravimétrique et l’analyse thermique différentielle (ATD-TG), la diffraction des rayons X (DRX), l’adsorption d’azote (B.E.T.), la microscopie électronique en transmission (MET) et l’infrarouge à transformée de Fourier (IR-TF). En particulier, les alumines de transition présentent des phases métastables qui subissent des transformations pendant le frittage et provoquent la formation d’une structure vermiculaire avec de larges porosités. La densité finale et la microstructure ont été améliorées grâce à une dispersion efficace de la poudre initialement agglomérée qui permet un réarrangement des particules et facilite la transformation vers la phase alpha. L’étude de l’influence de la dispersion sur la cinétique de transformation (Méthode de Kissinger) et la cinétique de frittage (Méthode SID) a été développée. Dans la deuxième partie de la thèse, le travail a été centré sur la production des nanocomposites Al 2 O 3 -5 vol. % YAG à partir de deux alumines commerciales frittées naturellement et par des méthodes non-conventionnelles comme le pressage à chaud (HP) et le spark plasma sintering (SPS). La caractérisation mécanique à température ambiante (dureté, ténacité, module d’élasticité) a été corrélée à une étude microstructurale (ESEM). Des valeurs intéressantes de dureté et de ténacité ont été mesurées dans les échantillons frittés par SPS et HP, environ 20 GPa et 7 MPa.m1/2, respectivement. Pour la caractérisation à haute température, les essais de fluage ont été conduits sous air en flexion 4 points à 1200°C sous une contrainte de 100 MPa. Les résultats montrent que les propriétés mécaniques à haute température dépendent fortement de la répartition de la deuxième phase dans la matrice d’alumine. Dans tous les cas, les résultats obtenus sont intéressants. MOTS-CLES : Alumine, Dispersion, Transformation de phase, Frittage, Microstructure, YAG, Nanocomposites, Fluage Laboratoire (s) de recherche : Dipartimento di Scienza dei Materiali e Ingegneria Chimica (DISMIC) Matériaux: Ingénierie et Science (MATEIS) Directeur de thèse: FANTOZZI Gilbert, PALMERO Paola Président de jury: Composition du jury : BONELLI Barbara, FANTOZZI Gilbert, LERICHE Anne, PALMERO Paola, REVERON Helen