Availableonlineatwww.sciencedirect.com Science Direct E噩≈RS ELSEVIER Joumal of the European Ceramic Society 28(2008)2301-2308 www.elsevier.comlocate/jeurceramsoc Microstructures, crystallography of interfaces, and creep behavior of melt-growth composites L Mazerolles a,*,L. Perriere a, b,S. Lartigue-Korineka, N. Piquet a, b, M. Parlier b CECM, UPR 2801, CNRS, 15 rue Georges Urbain, F-94407 Vitry sur Seine, France b ONERADMSC 29 avenue de la Division Leclerc. F-92322 chatillon Cedex franc Available online 4 March 2008 Oxide eutectic ceramics were prepared from Al2O3 and Ln]O3-based systems by unidirectional solidification from the melt. The microstructure consists of two single-crystal phases continuously entangled in a three-dimensional interpenetrating network without grain boundaries, pores or colonies. The outstanding stability of these microstructures gives rise to a high strength and creep resistance at high temperature. Preferred growth directions, orientation relationships between phases and single-crystal homogeneity of specimens were revealed. Creep behavior at high temperatur has been studied, mechanisms of deformation by dislocation motion and twinning were revealed from Transmission Electron Microscopy (TEM) observations. Extension to ternary eutectics with a three-dimensional microstructure consisting in the addition of a toughening phase(zrO2)to the previous binary eutectics has been investigated By using this method, significant improvement of fracture toughness was obtained. C 2008 Elsevier Ltd. all rights reserved. Keywords: Oxides; Ceramic eutectics; Microstructure; Creep; Dislocations Introduction their melting point, as compared with conventional composites and monolithic ceramics. , Furthermore oxide-based materi In the field of structural materials, eutectic ceramic oxides als are very attractive because of their inherent thermochemical prepared by solidification from the melt appear as potential can- stability in oxidizing environments at high temperature. More didates in the future for thermomechanical applications at very recently, Waku et al. have developed binary eutectics, called high temperature. Indeed, challenges related to future energy melt-growth composites(MGC), with novel microstructures in equirements impose the need to develop novel ultra-high- which continuous networks of single-crystal Al2O3 phases and temperature structural materials which display good mechanical single-crystal oxide compounds interpenetrate without grain properties(tensile strength, creep resistance, fracture tough- boundaries. These composites present a flexural strength con- ness)at temperatures above 1500C. For example in aircraft stant from room temperature up to high temperatures and a ge engines, the use of Ni-based single-crystal cast superalloys creep resistance allowing to consider applications in gas turbine for turbine blades is only possible at temperatures lower than and power generation systems with non-cooled turbine blades 100-1150oC. Silicon carbide-based composites are not sta- at very high temperatures. -S In this paper, we will present ble enough in an oxidizing atmosphere when temperature is results concerning similar microstructures obtained by direc- higher than 1300C and, finally, ceramic oxides, usually pre- tional solidification in various Al2O3 and Ln2O3-based system pared by sintering, have a too high brittleness due to the grain Morphology of microstructures, crystallography of constituent boundaries and the amorphous phases observed at grain bound- phases and interfaces, and single-crystal homogeneity of grown aries. Early studies on some oxide-oxide systems(such as samples will be reported. Creep behavior at high tempera- Al2O3-ZrO2)demonstrate the outstanding mechanical proper- ture has been studied. Factors controlling the deformation ties and the thermal and microstructural stability of directionally mechanisms will be analyzed taking into account microstruc- solidified eutectic ceramic oxides up to temperatures close to tural characteristics and Transmission Electron Microscopy (TEM)observations performed on deformed specimen. Finally, first results relative to the extension to ternary systems that display a significant increase of fracture toughness will be pre- E-mail address: mazerolles@glvt-cnrs fr(L Mazerolles) sented 0955-2219/S-see front matter o 2008 Elsevier Ltd. All rights reserved. doi: 10.1016/j-jeurceramsoc 2008.01.014
Available online at www.sciencedirect.com Journal of the European Ceramic Society 28 (2008) 2301–2308 Microstructures, crystallography of interfaces, and creep behavior of melt-growth composites L. Mazerolles a,∗, L. Perriere a,b, S. Lartigue-Korinek a, N. Piquet a,b, M. Parlier b a CECM, UPR 2801, CNRS, 15 rue Georges Urbain, F-94407 Vitry sur Seine, France b ONERA/DMSC, 29 avenue de la Division Leclerc, F-92322 Ch ˆatillon Cedex, France Available online 4 March 2008 Abstract Oxide eutectic ceramics were prepared from Al2O3 and Ln2O3-based systems by unidirectional solidification from the melt. The microstructure consists of two single-crystal phases continuously entangled in a three-dimensional interpenetrating network without grain boundaries, pores or colonies. The outstanding stability of these microstructures gives rise to a high strength and creep resistance at high temperature. Preferred growth directions, orientation relationships between phases and single-crystal homogeneity of specimens were revealed. Creep behavior at high temperature has been studied, mechanisms of deformation by dislocation motion and twinning were revealed from Transmission Electron Microscopy (TEM) observations. Extension to ternary eutectics with a three-dimensional microstructure consisting in the addition of a toughening phase (ZrO2) to the previous binary eutectics has been investigated. By using this method, significant improvement of fracture toughness was obtained. © 2008 Elsevier Ltd. All rights reserved. Keywords: Oxides; Ceramic eutectics; Microstructure; Creep; Dislocations 1. Introduction In the field of structural materials, eutectic ceramic oxides prepared by solidification from the melt appear as potential candidates in the future for thermomechanical applications at very high temperature. Indeed, challenges related to future energy requirements impose the need to develop novel ultra-hightemperature structural materials which display good mechanical properties (tensile strength, creep resistance, fracture toughness) at temperatures above 1500 ◦C. For example in aircraft engines, the use of Ni-based single-crystal cast superalloys for turbine blades is only possible at temperatures lower than 1100–1150 ◦C. Silicon carbide-based composites are not stable enough in an oxidizing atmosphere when temperature is higher than 1300 ◦C and, finally, ceramic oxides, usually prepared by sintering, have a too high brittleness due to the grain boundaries and the amorphous phases observed at grain boundaries. Early studies on some oxide–oxide systems (such as Al2O3–ZrO2) demonstrate the outstanding mechanical properties and the thermal and microstructural stability of directionally solidified eutectic ceramic oxides up to temperatures close to ∗ Corresponding author. E-mail address: mazerolles@glvt-cnrs.fr (L. Mazerolles). their melting point, as compared with conventional composites and monolithic ceramics.1,2 Furthermore, oxide-based materials are very attractive because of their inherent thermochemical stability in oxidizing environments at high temperature. More recently, Waku et al. have developed binary eutectics, called melt-growth composites (MGC), with novel microstructures in which continuous networks of single-crystal Al2O3 phases and single-crystal oxide compounds interpenetrate without grain boundaries. These composites present a flexural strength constant from room temperature up to high temperatures and a good creep resistance allowing to consider applications in gas turbine and power generation systems with non-cooled turbine blades at very high temperatures.3–5 In this paper, we will present results concerning similar microstructures obtained by directional solidification in various Al2O3 and Ln2O3-based systems. Morphology of microstructures, crystallography of constituent phases and interfaces, and single-crystal homogeneity of grown samples will be reported. Creep behavior at high temperature has been studied. Factors controlling the deformation mechanisms will be analyzed taking into account microstructural characteristics and Transmission Electron Microscopy (TEM) observations performed on deformed specimen. Finally, first results relative to the extension to ternary systems that display a significant increase of fracture toughness will be presented. 0955-2219/$ – see front matter © 2008 Elsevier Ltd. All rights reserved. doi:10.1016/j.jeurceramsoc.2008.01.014
L Mazerolles et al. /Joumal of the European Ceramic Society 28(2008)2301-2308 Table I Chemical compositions and constituent phases of studied eutectics Composition( mol%) Eutectic phases Structure of phases 77Al2O3-23Gd2O3 2O3-19Er2O3 Al2O3-Yb3 AlsAlO12 Corundum(hexagonal) 77Al2O3-23Y2O3 Al2O3-Y3Als AlO12 +Garmet(cubic) 2. Experimental procedures Fabrication from the melt microstructures requires planar growth front conditions in order to keep flat solid-liquid interfaces during growth at the macro- scopic and microscopic level. These process conditions are obtained with equipments displaying high-thermal gradients in the solidification direction. Our experiments were carried out with a floating-zone translation device consisting in an arc image furnace operating with a 6-kW xenon lamp as a radiation source cation was driven in air at a constant speed ranging from 2 to 20 mmh. Cylindrical specimens of about 8 mm in diame ter,and 50 mm in length were grown by this method. hism elt- 10 m powders(99.99%)were used for starting materials. Before melt ing, these powders were mixed and molded in green cylinder Fig. 2. Al2O3-Eu2O3 eutectic: SEM micrograph of longitudinal section cold isostatic pressure then, consolidated by sintering at 1300°Cfor12h. major axis being parallel to the solidification direction Microstructural observations and chemical analyses of eutec- Constant-load compressive creep tests were conducted in air tic phases were performed by Scanning Electron Microscopy environment at 1450 and 1600C within the stress range from using a Leo 1530 ( Leo, Germany)equipped with a Princeton 70 to 200 MPa. Compression rams were made from sintered Gamma Tech(USA) EDX spectrometer. They were carried out alumina with sapphire rods for applying loads on the specimen on sections of rods parallel and perpendicular to the growth direction. The crystalline homogeneity of specimen was con- 3. Results and discussion trolled at the macroscopic scale(1-2 mm) by using the Electron ack Scattering Diffraction(EBSD)technique on a Zeiss DSM 3.1. Microstructure 950 microscope equipped with a TSL detector. Maps of the crys- tallographic orientations were obtained by this method. Growth The studied eutectic composites were prepared in directions, orientation relationships between eutectic phases and Al2O3-Ln2O3 systems. Phase diagrams of these systems tron Microscopy. These observations were performed on thinned at temperatures close to 1800C. Eutectic phases conss e structure of interfaces were investigated by Transmission Elec- all display an eutectic composition on the rich alumina si foils of transverse sections using either a conventional micro- an a-Al2O phase(corundum structure)associated to either scope (EOL 2000EX)or a high-resolution electron microscope a perovskite-type phase LnAIO3(Ln=Sm, Eu, and Gd)or (Topcon 002B) both operating at 200k a garnet-type phase Ln AlsO12 for other elements of the From cylinder bars, compressive creep specimens were lanthanides series Data about the studied eutectic compositions machined with dimensions of 3 mm x mm x 5.5 mm, their are reported in Table 1 10 um μm Fig. 1. SEM micrographs of directionally solidified eutectic cross-sections showing the interlocking microstructure(a)Al2O3-Gd2O3: (b)Al2O3-Eu2O3;(c) Al2O3-Er2O3; (d)Al2O3-Y2O3). The white regions are perovskite (a and b)or garnet(c and d) and the black regions are alumina
2302 L. Mazerolles et al. / Journal of the European Ceramic Society 28 (2008) 2301–2308 Table 1 Chemical compositions and constituent phases of studied eutectics Composition (mol%) Eutectic phases Structure of phases 77Al2O3–23Gd2O3 Al2O3–GdAlO3 Corundum (hexagonal) 76Al2O3–24Eu2O3 Al2O3–EuAlO3 +Perovskite (orthorhombic) 81Al2O3–19Er2O3 Al2O3–Yb3Al5AlO12 Corundum (hexagonal) 77Al2O3–23Y2O3 Al2O3–Y3Al5AlO12 +Garnet (cubic) 2. Experimental procedures Fabrication from the melt of homogeneous eutectic microstructures requires planar growth front conditions in order to keep flat solid–liquid interfaces during growth at the macroscopic and microscopic level. These process conditions are obtained with equipments displaying high-thermal gradients in the solidification direction. Our experiments were carried out with a floating-zone translation device consisting in an arc image furnace operating with a 6-kW xenon lamp as a radiation source. Solidification was driven in air at a constant speed ranging from 2 to 20 mm h−1. Cylindrical specimens of about 8 mm in diameter, and 50 mm in length were grown by this method. High-purity powders (99.99%) were used for starting materials. Before melting, these powders were mixed and molded in green cylinder bars by cold isostatic pressure then, consolidated by sintering at 1300 ◦C for 12 h. Microstructural observations and chemical analyses of eutectic phases were performed by Scanning Electron Microscopy using a Leo 1530 (Leo, Germany) equipped with a Princeton Gamma Tech (USA) EDX spectrometer. They were carried out on sections of rods parallel and perpendicular to the growth direction. The crystalline homogeneity of specimen was controlled at the macroscopic scale (1–2 mm2) by using the Electron Back Scattering Diffraction (EBSD) technique on a Zeiss DSM 950 microscope equipped with a TSL detector. Maps of the crystallographic orientations were obtained by this method. Growth directions, orientation relationships between eutectic phases and structure of interfaces were investigated by Transmission Electron Microscopy. These observations were performed on thinned foils of transverse sections using either a conventional microscope (JEOL 2000EX) or a high-resolution electron microscope (Topcon 002B) both operating at 200 kV. From cylinder bars, compressive creep specimens were machined with dimensions of 3 mm × 3 mm × 5.5 mm, their Fig. 2. Al2O3–Eu2O3 eutectic: SEM micrograph of longitudinal section. major axis being parallel to the solidification direction. Constant-load compressive creep tests were conducted in air environment at 1450 and 1600 ◦C within the stress range from 70 to 200 MPa. Compression rams were made from sintered alumina with sapphire rods for applying loads on the specimen. 3. Results and discussion 3.1. Microstructure The studied eutectic composites were prepared in Al2O3–Ln2O3 systems. Phase diagrams of these systems all display an eutectic composition on the rich alumina side at temperatures close to 1800 ◦C. Eutectic phases consist in an -Al2O3 phase (corundum structure) associated to either a perovskite-type phase LnAlO3 (Ln = Sm, Eu, and Gd) or a garnet-type phase Ln3Al5O12 for other elements of the lanthanides series. Data about the studied eutectic compositions are reported in Table 1. Fig. 1. SEM micrographs of directionally solidified eutectic cross-sections showing the interlocking microstructure ((a) Al2O3–Gd2O3; (b) Al2O3–Eu2O3; (c) Al2O3–Er2O3; (d) Al2O3–Y2O3). The white regions are perovskite (a and b) or garnet (c and d) and the black regions are alumina.
L Mazerolles et al. Journal of the European Ceramic Society 28(2008)2301-2308 S mm/h 30mm熙8 S20 Fig. 3. Al2O3-Y2O3 eutectic: Morphology of microstructure vs. the solidification rate The eutectic microstructures grown in various composites are by using an arc image furnace). When this rate increases the shown in Fig. 1. These SEM micrographs of sections perpen- eutectic growth undergoes a transition from the planar to the dicular to the growth direction, reveal, in every case, continuous cellular regime that does not correspond anymore to a couple networks of an alumina phase(dark contrast)and lanthanide growth. Fig. 3 reveals that the three-dimensional microstruc and aluminum oxide compounds ((a) GAP=GdAlO3;(b) ture of the Al2O3-YAG eutectic, is not immediately modified AP=EuAIO3;(c)EAG= Er3AlsO12: (d) YAG=Y3Al5O12). when the growth rate increases and persists up to rates close to The observations performed on sections parallel (Fig. 2)to 30 mmh- with similar solidification conditions(thermal gra- the growth direction show similar morphologies indicating the dient, diameter of specimen) before entering into the cellular three-dimensional configuration of the microstructure. The two growth regime(Fig. 3). Similar results were obtained with the phases are continuously entangled in a three-dimensional inter- other eutectic compositions penetrating network without grain boundaries, pores or colonies The average size of each phase does not vary with the added 3. 2. Crystallography and interfaces rare-earth oxide except for the Al2O3-YAG composite( Fig. ld) ndeed, at this eutectic composition, prepared at a growth rate Aligned eutectic microstructures(lamellae, fibers or disper (5 mmh- )similar to other composites, a microstructure with soids), grown by unidirectional solidification, usually consist of harp angle facets and dimensions about five times larger than single-crystal phases growing preferentially along well-defined that of other eutectics has been observed. This difference could crystallographic directions. These directions are not necessarily be related to the diffusion coefficient values of yttrium higher the directions of easy-growth of the components but often cor- than that of lanthanide elements. However, for the Al2O3-YAG respond to minimum interfacial energy configurations between eutectic, a microstructure displaying dimensions very similar phases. These perfectly aligned lattices are related by orienta much higher solidification ales o d ut modifying the Chinese tion relationships which are unique in most systems and produce to other AlO3-Ln2 O3 eutectics with script morphology has been prod for growth conditions with well-defined interface planes corresponding to dense atomic arrangements in the component phases. 8- Contrarily to these In most directionally solidified oxide eutectics, coupled aligned microstructures, interconnected microstructures, shown growth is mainly controlled by the growth rate. For example, in in the previous figures, display a very isotropic morphology the case of the Al2O3-ZrO2(Y2O3)eutectic, the planar growth However, electron diffraction studies performed on thin plates matrix,only exists at very low solidification rates(<5mmh-I cut perpendicularly to the rod axes reveal growth directions also regime, leading to zirconia fibers embedded in an alumina corresponding to well-defined crystallographic directions 121 Fig 4. Electron diffraction patterns performed at the Al2O3-YAG (a)and Al2O3-EuAlO(b)interfaces
L. Mazerolles et al. / Journal of the European Ceramic Society 28 (2008) 2301–2308 2303 Fig. 3. Al2O3–Y2O3 eutectic: Morphology of microstructure vs. the solidification rate. The eutectic microstructures grown in various composites are shown in Fig. 1. These SEM micrographs of sections perpendicular to the growth direction, reveal, in every case, continuous networks of an alumina phase (dark contrast) and lanthanide and aluminum oxide compounds ((a) GAP = GdAlO3; (b) EAP = EuAlO3; (c) EAG = Er3Al5O12; (d) YAG = Y3Al5O12). The observations performed on sections parallel (Fig. 2) to the growth direction show similar morphologies indicating the three-dimensional configuration of the microstructure. The two phases are continuously entangled in a three-dimensional interpenetrating network without grain boundaries, pores or colonies. The average size of each phase does not vary with the added rare-earth oxide except for the Al2O3–YAG composite (Fig. 1d). Indeed, at this eutectic composition, prepared at a growth rate (5 mm h−1) similar to other composites, a microstructure with sharp angle facets and dimensions about five times larger than that of other eutectics has been observed. This difference could be related to the diffusion coefficient values of yttrium higher than that of lanthanide elements. However, for the Al2O3–YAG eutectic, a microstructure displaying dimensions very similar to other Al2O3–Ln2O3 eutectics without modifying the Chinese script morphology has been produced for growth conditions with much higher solidification rates.6 In most directionally solidified oxide eutectics, coupled growth is mainly controlled by the growth rate. For example, in the case of the Al2O3–ZrO2(Y2O3) eutectic, the planar growth regime, leading to zirconia fibers embedded in an alumina matrix, only exists at very low solidification rates (<5 mm h−1 by using an arc image furnace). When this rate increases the eutectic growth undergoes a transition from the planar to the cellular regime that does not correspond anymore to a coupled growth.7 Fig. 3 reveals that the three-dimensional microstructure of the Al2O3–YAG eutectic, is not immediately modified when the growth rate increases and persists up to rates close to 30 mm h−1 with similar solidification conditions (thermal gradient, diameter of specimen) before entering into the cellular growth regime (Fig. 3). Similar results were obtained with the other eutectic compositions. 3.2. Crystallography and interfaces Aligned eutectic microstructures (lamellae, fibers or dispersoids), grown by unidirectional solidification, usually consist of single-crystal phases growing preferentially along well-defined crystallographic directions. These directions are not necessarily the directions of easy-growth of the components but often correspond to minimum interfacial energy configurations between phases. These perfectly aligned lattices are related by orientation relationships which are unique in most systems and produce well-defined interface planes corresponding to dense atomic arrangements in the component phases.8–11 Contrarily to these aligned microstructures, interconnected microstructures, shown in the previous figures, display a very isotropic morphology. However, electron diffraction studies performed on thin plates cut perpendicularly to the rod axes reveal growth directions also corresponding to well-defined crystallographic directions. Fig. 4. Electron diffraction patterns performed at the Al2O3–YAG (a) and Al2O3–EuAlO3 (b) interfaces
L Mazerolles et al. /Joumal of the European Ceramic Society 28(2008)2301-2308 th directions and orientation relationships of eutectic phases in the AlzO3-Ln2O3 eutectics Eutectic phases Growth directions Orientation relationships [1010]A1203//[1 10]perovskite (1 120)Al203//(00 1)perovskit (Ln= Gd, Eu) or[11 20Al203//100 1]perovskite (0001)Al2O3/(1 00)perovskit Al2O3-Ln3 Als O12 [10I0Al2O3 (0001)Al2O3//(12 1)garnet _n=Er, Dy, Yb, Y) (2 10)or(1 10)gamet In these eutectic composites, the most frequently observed The minimal energy configuration at the interfaces is well growth direction for Al2O3 was [1010]. Sometimes, the illustrated by the HRTEM image(Fig. 5a)of the inter- [1 120] direction was also observed but, in all cases, the basal face between the corundum and perovskite structures for the plane of the corundum structure is always parallel to the solid- Al2O3-EuAlO3 eutectic. This image corresponds to a eutectic ification axis whatever the considered eutectic system. The structure grown along the [1 120]alumina and [00 l]perovskite electron diffraction patterns shown in Fig. 4 were performed directions. No intermediary phase is detected at the interface and on platelets cut perpendicularly to the growth directions of the transition on both sides of the interface operates on one or A12O3-YAG and Al2O3-GAP eutectics. The selected area aper- two atomic planes. Bragg filtering in the reciprocal space, from ture is centered on the interface, and consequently diffraction the numerical Fourier Transform of the digitized image, reveals spots of both phases are superimposed on the same pattern Crys- that accommodation between the two structures is restored by a llographic principal directions are strictly aligned according to periodic array of dislocations along the interface(Fig 5b). Sim- the following relations ilar results were also obtained with eutectics consisting of garnet and alumina phases. These TEM observations did not reveal any (0001)A2O3//(121)Y3Al5O12 stress fields at the interfaces, and are in good agreement with low and:(1120)Al2O3//(001GdAO3 values of residual stresses measured at room temperature from X-ray diffraction ex riments or from spectroscopic studies through the shift of fluorescence lines of Cr in sapphire. 4 Other growth directions of the garnet phase were also observed((11 0)in major cases) but the orientation relationship 3.3. Compressive creep deformation between the two phases persisted In the case of eutectics associ- ating perovskite and corundum structures, two sets of orientation Fig. 6 shows a typical strain vs. time relationship for the relations were determined. These results are summarized in Al2O3-YAG eutectic in a compressive test at 1450C with Table 2 stresses ranging from 70 to 200 MPa. The creep deformation EBSD studies have shown that these preferred growth direc- curves show a primary creep regime where the deformation rate tions and orientation relationships are retained on a large central decreases continuously. After this short primary stage(. g strain part of the specimenand confirmed the single-crystal quality of about 0.5%), when the secondary creep rate is reached, the applied load is modified and primary and secondary creep rates 102104 b 5.(a)HRTEM image of the AlzO3-EuAlO3 interface(b)Inverse Fourier Transform from the digitized image of the (a) interface built with 0003A10, and
2304 L. Mazerolles et al. / Journal of the European Ceramic Society 28 (2008) 2301–2308 Table 2 Growth directions and orientation relationships of eutectic phases in the Al2O3–Ln2O3 eutectics Eutectic phases Growth directions Orientation relationships Al2O3–LnAlO3 [1 0 1 0]Al ¯ 2O3//[1 1 0]perovskite (1 1 2 0) Al ¯ 2O3//(0 0 1) perovskite (Ln = Gd, Eu) or [1 1 2 0]Al ¯ 2O3//[0 0 1]perovskite (0 0 0 1) Al2O3//(1 0 0) perovskite Al2O3–Ln3Al5O12 [1 0 1 0]Al ¯ 2O3 (0 0 0 1)Al2O3//(1 2 1)garnet (Ln = Er, Dy, Yb, Y)) 2 1 0¯ or110garnet In these eutectic composites, the most frequently observed growth direction for Al2O3 was [1 0 1 0]. Sometimes, the ¯ [1 1 2 0] direction was also observed but, in all cases, the basal ¯ plane of the corundum structure is always parallel to the solidification axis whatever the considered eutectic system. The electron diffraction patterns shown in Fig. 4 were performed on platelets cut perpendicularly to the growth directions of Al2O3–YAG and Al2O3–GAP eutectics. The selected area aperture is centered on the interface, and consequently diffraction spots of both phases are superimposed on the same pattern. Crystallographic principal directions are strictly aligned according to the following relations: (0 0 0 1)Al2O3//(1 2 1)Y3Al5O12 and : (1 1 2 0)Al ¯ 2O3//(0 0 1)GdAlO3 Other growth directions of the garnet phase were also observed (110 in major cases) but the orientation relationship between the two phases persisted. In the case of eutectics associating perovskite and corundum structures, two sets of orientation relations were determined. These results are summarized in Table 2. EBSD studies have shown that these preferred growth directions and orientation relationships are retained on a large central part of the specimen12 and confirmed the single-crystal quality of grown samples. The minimal energy configuration at the interfaces is well illustrated by the HRTEM image (Fig. 5a) of the interface between the corundum and perovskite structures for the Al2O3–EuAlO3 eutectic. This image corresponds to a eutectic structure grown along the [1 1 2 0] alumina and [0 0 1] perovskite ¯ directions. No intermediary phase is detected at the interface and the transition on both sides of the interface operates on one or two atomic planes. Bragg filtering in the reciprocal space, from the numerical Fourier Transform of the digitized image, reveals that accommodation between the two structures is restored by a periodic array of dislocations along the interface (Fig. 5b). Similar results were also obtained with eutectics consisting of garnet and alumina phases. These TEM observations did not reveal any stress fields at the interfaces, and are in good agreement with low values of residual stresses measured at room temperature from X-ray diffraction experiments13 or from spectroscopic studies through the shift of fluorescence lines of Cr in sapphire.14 3.3. Compressive creep deformation Fig. 6 shows a typical strain vs. time relationship for the Al2O3–YAG eutectic in a compressive test at 1450 ◦C with stresses ranging from 70 to 200 MPa. The creep deformation curves show a primary creep regime where the deformation rate decreases continuously. After this short primary stage (e.g. strain of about 0.5%), when the secondary creep rate is reached, the applied load is modified and primary and secondary creep rates Fig. 5. (a) HRTEM image of the Al2O3–EuAlO3 interface. (b) Inverse Fourier Transform from the digitized image of the (a) interface built with 0003Al2O3 and 200EuAlO3 Bragg spots
L Mazerolles et al. Journal of the European Ceramic Society 28(2008)2301-2308 100 MPa 8E 06 △Al2O3 A2O3·GAF 0 104 Applied Stress(MPa) Time(s) Fig. 8. Plot of the creep rates vs. the applied stress for various eutectics. Fig. 6. Al2O3-YAG eutectic: Creep curve at 1450C with load increments and decrements(70, 100, 140, 200 then 140 and 100 MPa) low stresses(<140 MPa) and even higher at high stresses. Thi behavior is especially noteworthy as the Al2O3-ZrO2 eutec- e is very sensitive to the applied stress o following a power-law Al2O3-Ln2O3 eutectics (1710<Tm<1827C. Consequently, relationship: for the same T/Tm, these former display a better creep resistance. The incremental application of the load during one single E=Ao exp RT periment allows to determine readily the n exponent at each stress step by extrapolation of the minimum creep rate values where A is a material constant, n is the stress exponent, Q is for a given strain value. Eventual changes of the n value for var- the activation energy for creep, R is the gas constant and T ious applied stresses reveal different creep mechanisms. From the absolute temperature. The quasi-steady-state regime in thi the results summarized in Table 3 we can see that the work was determined by plots of the strain rate as a function of is higher than 2 from a 100-140 MPa step suggesting a defor- the true strain(Fig. 7). Fig. 8 shows the quasi-steady-state rates mation mechanism controlled by a dislocation motion. Values for the eutectics and regime stress tested. These plots reveal the close to 4 have also been measured for creep tests performed at high creep resistance of these interlocked microstructures with 1600 C. These high stress components are incompatible with strain rates very similar to that of Al2O3-ZrO2 eutectics5, 16 the interpretation of plasticity controlled by pure diffusion. In the case of deformation due to pure lattice diffusion or grain boundary diffusion the resulting creep rate is linearly prop tional to the stress and n=1. Moreover, hrtEM images of interfaces( Fig. 5)indicate a considerable coherence and strong bonding between phases at the interfaces and consequently boundary sliding mechanisms are impossible. TEM studies were performed on these specimens deformed at 1600C. Images presented in Fig. 9 corresponding to the AlzO3-GAP eutectic, clearly show dislocations in the two eutectic phases. In alumina, the dislocations observed in Fig. 9a are basal type dislocations(b= 1 /3(21)aligned in parallel basal slip planes(0001). The basal plane is seen edge-on in this orienta tion. The basal slip system has the lowest critical resolved shear Table 3 Values of the stress exponent as a function of the stress increment for vari AlO3-Lng O3 eutectics deformed at 1450oC 02040608 70→100MPa 100→140MPa 140→200MPa True strain(%) Al2O3-EAG AlO3-YAG 1.13 2.06 -state regime In various eutectics
L. Mazerolles et al. / Journal of the European Ceramic Society 28 (2008) 2301–2308 2305 Fig. 6. Al2O3–YAG eutectic: Creep curve at 1450 ◦C with load increments and decrements (70, 100, 140, 200 then 140 and 100 MPa). are again measured at a higher stress. The secondary creep rate ε˙ is very sensitive to the applied stress σ following a power–law relationship: ε˙ = Aσn exp − Q RT where A is a material constant, n is the stress exponent, Q is the activation energy for creep, R is the gas constant and T is the absolute temperature. The quasi-steady-state regime in this work was determined by plots of the strain rate as a function of the true strain (Fig. 7). Fig. 8 shows the quasi-steady-state rates for the eutectics and regime stress tested. These plots reveal the high creep resistance of these interlocked microstructures with strain rates very similar to that of Al2O3–ZrO2 eutectics15,16 at Fig. 7. Strain rates as a function of true strain allowing to determine the minimum creep rates corresponding to the quasi-steady-state regime in various eutectics. Fig. 8. Plot of the creep rates vs. the applied stress for various eutectics. low stresses (<140 MPa) and even higher at high stresses. This behavior is especially noteworthy as the Al2O3–ZrO2 eutectic has melting temperature (Tm = 1910 ◦C) higher than that of Al2O3–Ln2O3 eutectics (1710 < Tm < 1827 ◦C). Consequently, for the same T/Tm, these former display a better creep resistance. The incremental application of the load during one single experiment allows to determine readily the n exponent at each stress step by extrapolation of the minimum creep rate values for a given strain value. Eventual changes of the n value for various applied stresses reveal different creep mechanisms. From the results summarized in Table 3 we can see that the n value is higher than 2 from a 100–140 MPa step suggesting a deformation mechanism controlled by a dislocation motion. Values close to 4 have also been measured for creep tests performed at 1600 ◦C. These high stress components are incompatible with the interpretation of plasticity controlled by pure diffusion. In the case of deformation due to pure lattice diffusion or grain boundary diffusion the resulting creep rate is linearly proportional to the stress and n = 1.17 Moreover, HRTEM images of interfaces (Fig. 5) indicate a considerable coherence and strong bonding between phases at the interfaces and consequently boundary sliding mechanisms are impossible. TEM studies were performed on these specimens deformed at 1600 ◦C. Images presented in Fig. 9 corresponding to the Al2O3–GAP eutectic, clearly show dislocations in the two eutectic phases. In alumina, the dislocations observed in Fig. 9a are basaltype dislocations (b = 1/3[2 1¯ 1 0]) aligned in parallel basal slip ¯ planes (0 0 0 1). The basal plane is seen edge-on in this orientation. The basal slip system has the lowest critical resolved shear Table 3 Values of the stress exponent as a function of the stress increment for various Al2O3–Ln2O3 eutectics deformed at 1450 ◦C 70→100 MPa 100→140 MPa 140→200 MPa Al2O3–GAP 1.10 2.08 2.60 Al2O3–EAG 1.20 2.10 2.72 Al2O3–YAG 1.13 2.06 2.99
306 L Mazerolles et al. /Joumal of the European Ceramic Society 28(2008)2301-2308 g=(1213) AL,O3 galo 100nm GdAIO,AL,O (022) gM=(1102) d) 200nm Fig 9. TEM images of microstructural features in AlzO3-GAP sample after compressive deformation at 1600C. (a)2-Beam dark field image showing dislocations in basal planes in the alumina phase.(b) Bright field image showing a dislocation pile-up into the gaP phase and a basal twin occuring in alumina at the interaction etween the pile-up and the interface. (c)Dark field image(with a twin diffraction vector)revealing dislocations in the twin boundaries. (d)Bright field image: Two isolated dislocations occur inside the GAP phase Interfaces display dislocations arranged by pairs(indicated by arrows). stress at high temperature. 8 Fig 9b shows a dislocation pile-up case of AlO3-EAG and Al2O3-GAP eutectics) the glide of in the GAP phase. The pile-up interacts with the GAP-alumina dislocations require higher stresses to be activated. interface at the bottom of the image, and a basal twin is seen in umina close to the interaction. The twin has probably occurred 3.4. Extension to ternary systems by accommodation of stresses resulting from the pile-up. Dis- locations are present in the twin boundaries(Fig. 9c). It is Despite a high flexural strength reported in the literature,a corth noting that usually large stresses are necessary to acti- high temperature as we have seen previously, the A1203-Ln203 observed in alumina submitted to plastic deformation. 9,20These observations are in agreement with the high value of the stress display a low fracture toughness at room temperature. We tried to increase the Kic values by the addition of a toughen- Finally, interfacial dislocations arranged by pairs are present ing phase(ZrO2)to these binary eutectics without modifying in most interfaces(Fig. 9d). Work is underway to understan the three-dimensional interpenetrating network. This singular the origin of these defects, either from misfit dislocations or microstructure observed in binary systems results from eutectic resulting from the interaction with lattice dislocation coupled-growth conditions Consequently, we looked for prepar- We can note from results presented in Table 3 when the stress ing ternary eutectic compositions in the Al2O3-Ln2O3-Zr02 is higher than 140MPa, that the n value corresponding to the systems with the same growth conditions. The phase diagrams of Al2O3-YAG eutect slightly higher than that of the two other these systems are not well known. One eutectic composition has eutectics indicating a higher creep rate. If we consider that the been reported and prepared in the AlO3/ZrO2/Y2 O3 system" stress required to drive a dislocation across the interfaces is at but for other rare-earth oxides very few experimental data are least two orders of magnitude higher than the resolved stresses available except phase diagrams calculated with CALPHAD methods. We experimentally determined the eutectic composi- applied in the tests, 2 both phases must deform independently tion 58Al203-19Gd2O3-23ZrO2(mol%)from microstructural in order to creep at the same rate. Consequently, the dislocations have to move in each component and must bow within the phase observations. Similarly to binary eutectics, the as prepared spacing(A). The stress necessary for this process is microstructures consist of an interpenetrated network of two major eutectic phases and the zirconia phase is located at the interfaces between the alumina and garnet(or perovskite) type phases. ZrO2 is fully stabilized by the rare-earth ions in the fluorite structure(Fig. 10). where G is the shear modulus and b the Burgers vector.Con In the case of the Al2O3-Y203-ZrO2 eutectic, the compari- sequently, when the size of microstructure decreases (it is the son with the binary eutectic, reveals a significant decrease of the
2306 L. Mazerolles et al. / Journal of the European Ceramic Society 28 (2008) 2301–2308 Fig. 9. TEM images of microstructural features in Al2O3–GAP sample after compressive deformation at 1600 ◦C. (a) 2-Beam dark field image showing dislocations in basal planes in the alumina phase. (b) Bright field image showing a dislocation pile-up into the GAP phase and a basal twin occuring in alumina at the interaction between the pile-up and the interface. (c) Dark field image (with a twin diffraction vector) revealing dislocations in the twin boundaries. (d) Bright field image: Two isolated dislocations occur inside the GAP phase. Interfaces display dislocations arranged by pairs (indicated by arrows). stress at high temperature.18 Fig. 9b shows a dislocation pile-up in the GAP phase. The pile-up interacts with the GAP–alumina interface at the bottom of the image, and a basal twin is seen in alumina close to the interaction. The twin has probably occurred by accommodation of stresses resulting from the pile-up. Dislocations are present in the twin boundaries (Fig. 9c). It is worth noting that usually large stresses are necessary to activate basal twins. Only in a few cases, basal twinning has been observed in alumina submitted to plastic deformation.19,20 These observations are in agreement with the high value of the stress component. Finally, interfacial dislocations arranged by pairs are present in most interfaces (Fig. 9d). Work is underway to understand the origin of these defects, either from misfit dislocations or resulting from the interaction with lattice dislocations. We can note from results presented in Table 3 when the stress is higher than 140 MPa, that the n value corresponding to the Al2O3–YAG eutectic is slightly higher than that of the two other eutectics indicating a higher creep rate. If we consider that the stress required to drive a dislocation across the interfaces is at least two orders of magnitude higher than the resolved stresses applied in the tests,21 both phases must deform independently in order to creep at the same rate. Consequently, the dislocations have to move in each component and must bow within the phase spacing (λ). The stress necessary for this process is: τ = Gb λ where G is the shear modulus and b the Burgers vector.22 Consequently, when the size of microstructure decreases (it is the case of Al2O3–EAG and Al2O3–GAP eutectics) the glide of dislocations require higher stresses to be activated. 3.4. Extension to ternary systems Despite a high flexural strength reported in the literature, a good thermal stability and an outstanding creep resistance at high temperature as we have seen previously, the Al2O3–Ln2O3 eutectics (particularly for systems containing a garnet phase) display a low fracture toughness at room temperature. We tried to increase the KIC values by the addition of a toughening phase (ZrO2) to these binary eutectics without modifying the three-dimensional interpenetrating network. This singular microstructure observed in binary systems results from eutectic coupled-growth conditions. Consequently, we looked for preparing ternary eutectic compositions in the Al2O3–Ln2O3–ZrO2 systems with the same growth conditions. The phase diagrams of these systems are not well known. One eutectic composition has been reported and prepared in the Al2O3/ZrO2/Y2O3 system23 but for other rare-earth oxides very few experimental data are available except phase diagrams calculated with CALPHAD methods.24 We experimentally determined the eutectic composition 58Al2O3–19Gd2O3–23ZrO2 (mol%) from microstructural observations. Similarly to binary eutectics, the as prepared microstructures consist of an interpenetrated network of two major eutectic phases and the zirconia phase is located at the interfaces between the alumina and garnet (or perovskite) type phases. ZrO2 is fully stabilized by the rare-earth ions in the fluorite structure (Fig. 10). In the case of the Al2O3–Y2O3–ZrO2 eutectic, the comparison with the binary eutectic, reveals a significant decrease of the
L Mazerolles et al. Journal of the European Ceramic Society 28(2008)2301-2308 2307 Fig. 10. SEM micrographs of ternary eutectics: (a)Al2O3-Y2O3-ZrO2;(b)AlO3-Gd2O3-ZrO2 Alumina phase corresponds to dark contrast, YAG and GAP phases to the grey contrast and zirconia phase is indicated by arrows mean size of microstructure and changes of morphology(growth relative to the Al2O3-Y203-ZrO2 eutectic shown in Fig. 11 facets have disappeared) likely due to growth mechanisms modi- reveal single growth directions for the three constituent phases fied by the third eutectic phase. The single-crystal homogeneity in the analyzed area(the highest density of the 10 10A1203 is not modified by the addition of a third phase. The electron 00 lYAG and 001zro, orientations is located in the center of diffraction studies revealed the following growth directions: the stereographic projections corresponding to the normal to [1010A2O3//001Y3AlsO12//[001zrO2 the surface). Localization of poles on the great circle is also in agreement with electron diffraction results and the epitaxial relationships between constituent phases The addition of ZrO2 phase to the binary systems gives rise to significant increase of fracture toughness. The most important (0001)Al203//100)ZrO2(000 1)AlO3//(00 1)Y3Al5O12 effect was observed in the case of eutectics associating alumina The same epitaxial relations between alumina and zirco- and garnet phase. Mechanism of crack deflection by the zirconia nia phases were already observed for the AlO3-Zroz binary phase coupled with the decrease of the size of microstructure eutecticbut the preferred growth direction of the garnet struc- resulted into an increasing of the Kic values of about 70%.25 ture observed in the binary systems is modified by the addition This effect is less significant for the eutectics containing a per of zirconia and occurs along the(100) direction in the ternary ovskite phase but a systematic increase of about 15% was also system. Similarly to the results obtained on binary eutectics, measured. Study of creep behavior of these temary eutectics the EBSD analysis confirmed the single-crystal homogeneity on Is in progress large areas(some millimeters). The pole figures(density maps) 4. Conclusion 0001 Our study on the oxide eutectic composites grown from the melt has shown that the outstanding mechanical properties at low or high temperature could result from the combination of differ- ent factors. Firstly, eutectic composition gives rise to a very good hemical and thermal stability of the microstructure up to tem peratures close to the melting temperature. The very isotropic morphology of that microstructure in the Al2O3-Ln2O3 sys- tems, induces a mechanical behavior that will not depend on the direction of the applied mechanical stress. Growth of these 001 001 eutectics. whatever the rare-earth oxides associated to alumina and solidification conditions, follows a small set of crystallo- graphic directions often corresponding to dense atomic rows of structures. These directions do not change all over the speci men, and consequently a very low amount of grain boundaries, often at the origin of brittleness phenomena, is observed. The absence of intermediary phase at the interfaces and the crystal lographic orientation relationships between phases give rise to a strong cohesion between components. The outstanding creep resistance of these materials is likely related to the quality of YAG ZrO2(Y2O3) these interfaces. tEM studies revealed deformation mechanisms Fig. 11. A1203-YAG-ZrO2 eutectic: EBSD pole figures of a by dislocation motion in the two phases with pile-up phenomena 1.75mm x 1.75 mm area corresponding to the 0001Ah03, 1010A1 03, 0 at the interfaces. Other mechanisms, such as basal twinning in 0 lyAG and 00 1zo, onentations. alumina usually requiring large stresses, were also observed
L. Mazerolles et al. / Journal of the European Ceramic Society 28 (2008) 2301–2308 2307 Fig. 10. SEM micrographs of ternary eutectics: (a) Al2O3–Y2O3–ZrO2; (b) Al2O3–Gd2O3–ZrO2. Alumina phase corresponds to dark contrast, YAG and GAP phases to the grey contrast and zirconia phase is indicated by arrows. mean size of microstructure and changes of morphology (growth facets have disappeared) likely due to growth mechanisms modi- fied by the third eutectic phase. The single-crystal homogeneity is not modified by the addition of a third phase. The electron diffraction studies revealed the following growth directions: [1 0 1 0]Al ¯ 2O3//[0 0 1]Y3Al5O12//[0 0 1]ZrO2 and the epitaxial relationships between constituent phases: (0 0 0 1)Al2O3//(1 0 0)ZrO2(0 0 0 1)Al2O3//(0 0 1)Y3Al5O12 The same epitaxial relations between alumina and zirconia phases were already observed for the Al2O3–ZrO2 binary eutectic9 but the preferred growth direction of the garnet structure observed in the binary systems is modified by the addition of zirconia and occurs along the 100 direction in the ternary system. Similarly to the results obtained on binary eutectics, the EBSD analysis confirmed the single-crystal homogeneity on large areas (some millimeters). The pole figures (density maps) Fig. 11. Al2O3–YAG–ZrO2 eutectic: EBSD pole figures of a 1.75 mm × 1.75 mm area corresponding to the 0001Al2O3 , 1 0 1 0¯ Al2O3 , 0 0 1YAG and 0 0 1ZrO2 orientations. relative to the Al2O3–Y2O3–ZrO2 eutectic shown in Fig. 11, reveal single growth directions for the three constituent phases in the analyzed area (the highest density of the 1 0 1 0¯ Al2O3 , 001YAG and 0 0 1ZrO2 orientations is located in the center of the stereographic projections corresponding to the normal to the surface). Localization of poles on the great circle is also in agreement with electron diffraction results. The addition of ZrO2 phase to the binary systems gives rise to a significant increase of fracture toughness. The most important effect was observed in the case of eutectics associating alumina and garnet phase. Mechanism of crack deflection by the zirconia phase coupled with the decrease of the size of microstructure resulted into an increasing of the KIC values of about 70%.25 This effect is less significant for the eutectics containing a perovskite phase but a systematic increase of about 15% was also measured.12 Study of creep behavior of these ternary eutectics is in progress. 4. Conclusion Our study on the oxide eutectic composites grown from the melt has shown that the outstanding mechanical properties at low or high temperature could result from the combination of different factors. Firstly, eutectic composition gives rise to a very good chemical and thermal stability of the microstructure up to temperatures close to the melting temperature. The very isotropic morphology of that microstructure in the Al2O3–Ln2O3 systems, induces a mechanical behavior that will not depend on the direction of the applied mechanical stress. Growth of these eutectics, whatever the rare-earth oxides associated to alumina and solidification conditions, follows a small set of crystallographic directions often corresponding to dense atomic rows of structures. These directions do not change all over the specimen, and consequently a very low amount of grain boundaries, often at the origin of brittleness phenomena, is observed. The absence of intermediary phase at the interfaces and the crystallographic orientation relationships between phases give rise to a strong cohesion between components. The outstanding creep resistance of these materials is likely related to the quality of these interfaces. TEM studies revealed deformation mechanisms by dislocation motion in the two phases with pile-up phenomena at the interfaces. Other mechanisms, such as basal twinning in alumina usually requiring large stresses, were also observed
L Mazerolles et al. /Joumal of the European Ceramic Society 28(2008)2301-2308 In order to increase fracture toughness of binary eutec- 8. Stubican, V.S. and Bradt, R C, Eutectic solidification in ceramic systems tics, a third phase(zrO2)was added. These composites were Ann. Rev. Mater Sci. 1981.11. 267-297 prepared at a ternary eutectic composition. A new eutectic com- 9. Mazerolles, L, Michel, D. and Portier, R, Interfaces in oriented position was determined for the Al2O3-Gd2O3-ZrO2 system (10). Revcoleyschi, A. Dallhenne, G. and Michel. D. Extermal and internal inter- ces of metal oxides. Mater Sci. Form. Trans. Tech Publications. 1988 The single-crystal homogeneity of composites was preserved 173-197. with preferred growth directions and orientation relationship 11. Mazerolles, L, Michel, D and Hytch, M.J., Microstructure and interfaces between phases. Fracture toughness was improved by a fac in directionally solidified oxide-oxide eutectics J. Eur. Ceram. Soc, 2005 25,1389-1395 tor ranging from 1.5 to 1.7 compared to binary systems. Creep 12Mazerolles, L.Piquet,N, Trichet, M.F. and Parlier,M,Microstructures behavior of these ternary eutectics compared to the binary sys- and interfaces in melt growth Al2O3-Ln2O3 based eutectic compo tems and the investigation of deformation mechanisms from Sci. technol,2006,45,1377-1384 TEM observations is in progress 13. Dickey, E. C, Frazer, C. S, Watkins, T R. and Hubbard, C.R., Residual resses in high temperature ceramic eutectics. J. Eur. Ceram. Soc., 1999, Acknowledgements 19,2503-2509 14. Gouadec, G, Colomban, Ph, Piquet, N, Trichet, M. F and Mazerolles L,Raman/Cr+ fluorescence mapping of a melt-grown Al203/GdAl The authors would like to thank D. Boivin (ONERA/DMMP) eutectic. Eur Ceram Soc. 2005. 25. 1447-1453 for the ebsd studies. L. Perrier would also like to thank the Argon, A.S., Yi,J and Sayir, A, Creep ce of directionally solidified French Defence Research Organization(DGA)for a Doctorate eutectics of Al2O3/cZrO2 with sub-microncolumnar morphologies Mater. Studentship. Sci.Eng,2001,A319,838-842. 16. Yi, J, Argon, A.S. and Sayir, A, Creep resistance of the directionally solid- fied ceramic eutectic of Al2 O3/cZrO2(Y203): experiments and models. J. References Eur Ceram Soc,2005,25,1201-1214 17. Poirier, J, Creep of Crystals. Cambridge University Press, Cambridge, 1985 1.Hulse, C.O. and Batt, J. A. The effect of eutectic microstructures on the 18. Lagerloff, K P D, Heuer, A. H Castaing, J, Riviere, J.P. and Mitchell,T. echanical properties of ceramic oxides. Final. Rept UARL-N910803-10 E, Slip and twinning in sapphire(a-Al2O3).J. Am. Ceram Soc., 1994, 77, NTIS AD-781995/6GA,1974. 385-397. 2. Minford, w J, Bradt, R.C. and Stubican, V S, Crystallography and 19. Heuer, A H, Lagerlof, K P D and Castaing, J, Slip and twinning dislo- microstructure of directionally solidified oxide eutectics. J. Am. Ceram. ations in sapphire(a-AlO3) Phil Mag. A 747-763 20. Lartigue-Korinek, S and Dupau, F, Grain bour havior in superplastic 3. Waku, Y, Nakagawa, N, Wakamoto, T, Ohtsubo, H, Shimizu, K. and Mg-doped alumina with yttria codoping. Act Mater,1994,42, Kohtoku, Y, A ductile ceramic eutectic composite with high strength at 293-302. 873K. Nature,1997,389,49-52 21. Martinez Femandez, J. and Sayir, A, Creep of directionally solidified 4. Ochiai, S, Ueda, T, Sato, K, Hojo, M, Waku, Y, Nakagawa, N. er ai Al2O3/Er Als O12 fibers with hypo-eutectic composition. Ceram Eng. Sci Deformation and fracture behavior of an Al2 O3/YAG composite from room Pmc.2001,22,421-428 emperature to 2023. Comp. Sci. Technol., 2001, 61, 2117-2128 22. Hirth, J. P and Lothe, J. Theory of Dislocations. John Wiley Sons, 1982 5. Nakagawa, N. Ohtsubo, H. Mitani, A, Shimizu, K and Waku, Y. High 23. Waku, Y, Sakata, S. Mitani, A and Shimizu, K Temperature dependence temperature strength and thermal stability for melt growth composite. J. Eur f flexural strength and microstructure of Al2O3-Y3 AlsO12-ZrO2 ternary Ceram.Soc.,2005,25,1251-1257 melt grown composites. J Mater. Sci, 2002, 37, 2975-2982. 6. Pastor. J.Y. LIe ar. A. Olite. P B. de fransisco I and Pena. 24 Akiza, S. M, Fabrichnaya, O, Wang, Ch, Zinchevich, M. and Aldinger, L, Mechanical p F, Phase diagram of the ZrO2-Al2O3-Gd2O3 system J. Eur Ceram Soc. eutectics up to 1 Am. Ceran.Soc,2005,88,1488-1495 2006.26.233-246 7. Echigoya, J, Takabayashi, Y, Sasaki, K Hayashi, S and Suto, H J, Solid- 25. Perriere, L Valle, R, Mazerolles, L and Parlier, M, Crack propagation ification microstructure of Y2O3- added Al2O3-ZrO2 eutectic. Trans Jp in directionally solidified eutectic ceramics. J. Eur. Ceram. Soc., 2008, 28 lnst.Met.,1986,27,102-1
2308 L. Mazerolles et al. / Journal of the European Ceramic Society 28 (2008) 2301–2308 In order to increase fracture toughness of binary eutectics, a third phase (ZrO2) was added. These composites were prepared at a ternary eutectic composition. A new eutectic composition was determined for the Al2O3–Gd2O3–ZrO2 system and interconnected microstructures were successfully grown. The single-crystal homogeneity of composites was preserved with preferred growth directions and orientation relationships between phases. Fracture toughness was improved by a factor ranging from 1.5 to 1.7 compared to binary systems. Creep behavior of these ternary eutectics compared to the binary systems and the investigation of deformation mechanisms from TEM observations is in progress. Acknowledgements The authors would like to thank D. Boivin (ONERA/DMMP) for the EBSD studies. L. Perriere would also like to thank the ` French Defence Research Organization (DGA) for a Doctorate Studentship. References 1. Hulse, C. O. and Batt, J. A. The effect of eutectic microstructures on the mechanical properties of ceramic oxides. Final. Rept UARL-N910803-10, NTIS AD-781995/6GA, 1974. 2. Minford, W. J., Bradt, R. C. and Stubican, V. S., Crystallography and microstructure of directionally solidified oxide eutectics. J. Am. Ceram. Soc., 1979, 62, 154–157. 3. Waku, Y., Nakagawa, N., Wakamoto, T., Ohtsubo, H., Shimizu, K. and Kohtoku, Y., A ductile ceramic eutectic composite with high strength at 1873 K. Nature, 1997, 389, 49–52. 4. Ochiai, S., Ueda, T., Sato, K., Hojo, M., Waku, Y., Nakagawa, N. et al., Deformation and fracture behavior of an Al2O3/YAG composite from room temperature to 2023. Comp. Sci. Technol., 2001, 61, 2117–2128. 5. Nakagawa, N., Ohtsubo, H., Mitani, A., Shimizu, K. and Waku, Y., High temperature strength and thermal stability for melt growth composite. J. Eur. Ceram. Soc., 2005, 25, 1251–1257. 6. Pastor, J. Y., Llorca, J., Salazar, A., Oliete, P. B., de Fransisco, I. and Pena, J. I., Mechanical properties of melt-grown alumina–yttrium aluminium garnet eutectics up to 1900 K. J. Am. Ceram. Soc., 2005, 88, 1488–1495. 7. Echigoya, J., Takabayashi, Y., Sasaki, K., Hayashi, S. and Suto, H. J., Solidification microstructure of Y2O3-added Al2O3–ZrO2 eutectic. Trans Jpn. Inst. Met., 1986, 27, 102–107. 8. Stubican, V. S. and Bradt, R. C., Eutectic solidification in ceramic systems. Ann. Rev. Mater. Sci., 1981, 11, 267–297. 9. Mazerolles, L., Michel, D. and Portier, R., Interfaces in oriented Al2O3–ZrO2(Y2O3) eutectics. J. Am. Ceram. Soc., 1986, 69, 252–255. [10].Revcolevschi, A., Dalhenne, G. and Michel, D. External and internal interfaces of metal oxides. Mater. Sci. Forum, Trans. Tech. Publications, 1988, 173–197. 11. Mazerolles, L., Michel, D. and Hytch, M. J., Microstructure and interfaces ¨ in directionally solidified oxide–oxide eutectics. J. Eur. Ceram. Soc, 2005, 25, 1389–1395. 12. Mazerolles, L., Piquet, N., Trichet, M. F. and Parlier, M., Microstructures and interfaces in melt growth Al2O3–Ln2O3 based eutectic composites. Adv. Sci. Technol., 2006, 45, 1377–1384. 13. Dickey, E. C., Frazer, C. S., Watkins, T. R. and Hubbard, C. R., Residual stresses in high temperature ceramic eutectics. J. Eur. Ceram. Soc., 1999, 19, 2503–2509. 14. Gouadec, G., Colomban, Ph., Piquet, N., Trichet, M. F. and Mazerolles, L., Raman/Cr3+ fluorescence mapping of a melt-grown Al2O3/GdAlO3 eutectic. J. Eur. Ceram. Soc., 2005, 25, 1447–1453. 15. Argon, A. S., Yi, J. and Sayir, A., Creep resistance of directionally solidified eutectics of Al2O3/cZrO2 with sub-microncolumnar morphologies. Mater. Sci. Eng., 2001, A319, 838–842. 16. Yi, J., Argon, A. S. and Sayir, A., Creep resistance of the directionally solidified ceramic eutectic of Al2O3/cZrO2(Y2O3): experiments and models. J. Eur. Ceram. Soc., 2005, 25, 1201–1214. 17. Poirier, J.,Creep of Crystals. Cambridge University Press, Cambridge, 1985. 18. Lagerloff, K. P. D., Heuer, A. H., Castaing, J., Riviere, J. P. and Mitchell, T. ` E., Slip and twinning in sapphire (-Al2O3). J. Am. Ceram. Soc., 1994, 77, 385–397. 19. Heuer, A. H., Lagerlof, K. P. D. and Castaing, J., Slip and twinning dislocations in sapphire (-Al2O3). Phil. Mag. A, 1998, 78, 747–763. 20. Lartigue-Korinek, S. and Dupau, F., Grain boundary behavior in superplastic Mg-doped alumina with yttria codoping. Acta Metall. Mater., 1994, 42, 293–302. 21. Martinez Fernandez, J. and Sayir, A., Creep of directionally solidified Al2O3/Er3Al5O12 fibers with hypo-eutectic composition. Ceram Eng. Sci. Proc., 2001, 22, 421–428. 22. Hirth, J. P. and Lothe, J., Theory of Dislocations. John Wiley & Sons, 1982. 23. Waku, Y., Sakata, S., Mitani, A. and Shimizu, K., Temperature dependence of flexural strength and microstructure of Al2O3–Y3Al5O12–ZrO2 ternary melt grown composites. J. Mater. Sci., 2002, 37, 2975–2982. 24. Lakiza, S. M., Fabrichnaya, O., Wang, Ch., Zinchevich, M. and Aldinger, F., Phase diagram of the ZrO2–Al2O3–Gd2O3 system. J. Eur. Ceram. Soc., 2006, 26, 233–246. 25. Perriere, L., Valle, R., Mazerolles, L. and Parlier, M., Crack propagation in directionally solidified eutectic ceramics. J. Eur. Ceram. Soc., 2008, 28, 2337–2343.