Skip to main content
Science and Technology of Advanced Materials logoLink to Science and Technology of Advanced Materials
. 2008 Sep 10;9(3):033001. doi: 10.1088/1468-6996/9/3/033001

Textured silicon nitride: processing and anisotropic properties

Xinwen Zhu 1,, Yoshio Sakka 1,
PMCID: PMC5099652  PMID: 27877995

Abstract

Textured silicon nitride (Si3N4) has been intensively studied over the past 15 years because of its use for achieving its superthermal and mechanical properties. In this review we present the fundamental aspects of the processing and anisotropic properties of textured Si3N4, with emphasis on the anisotropic and abnormal grain growth of β-Si3N4, texture structure and texture analysis, processing methods and anisotropic properties. On the basis of the texturing mechanisms, the processing methods described in this article have been classified into two types: hot-working (HW) and templated grain growth (TGG). The HW method includes the hot-pressing, hot-forging and sinter-forging techniques, and the TGG method includes the cold-pressing, extrusion, tape-casting and strong magnetic field alignment techniques for β-Si3N4 seed crystals. Each processing technique is thoroughly discussed in terms of theoretical models and experimental data, including the texturing mechanisms and the factors affecting texture development. Also, methods of synthesizing the rodlike β-Si3N4 single crystals are presented. Various anisotropic properties of textured Si3N4 and their origins are thoroughly described and discussed, such as hardness, elastic modulus, bending strength, fracture toughness, fracture energy, creep behavior, tribological and wear behavior, erosion behavior, contact damage behavior and thermal conductivity. Models are analyzed to determine the thermal anisotropy by considering the intrinsic thermal anisotropy, degree of orientation and various microstructure factors. Textured porous Si3N4 with a unique microstructure composed of oriented elongated β-Si3N4 and anisotropic pores is also described for the first time, with emphasis on its unique mechanical and thermal-mechanical properties. Moreover, as an important related material, textured α-Sialon is also reviewed, because the presence of elongated α-Sialon grains allows the production of textured α-Sialon using the same methods as those used for textured β-Si3N4 and β-Sialon.

Keywords: Si3N4, Sialon, texture, seed, grain growth, hot-working, templated grain growth, strong magnetic field alignment, porosity, anisotropic properties

Introduction

Silicon nitride (Si3N4) is one of the most intensively studied ceramic materials and is suitable for a variety of structural applications such as automotive engine parts, heat exchangers, pump seal parts, ball bearings, cutting tools and ceramic armor due to its low bulk density, excellent mechanical properties at elevated temperatures, high resistance to thermal shock and chemical attack, excellent creep resistance and good tribological and wear properties [1, 2].

Si3N4 occurs in two major crystalline phases, known as α and β, both exhibiting a hexagonal structure but with different stacking sequences: ABCD in α and ABAB in β [3]. The longer stacking sequence results in the α-phase having higher hardness than the β-phase. However, the α-phase is chemically unstable compared with the β-phase. At high temperatures (HTs) when a liquid phase is present, the α-phase always transforms into the β-phase via the solution-precipitation process, during which α-grains dissolve into the liquid phase and precipitate as β-grains, known as the α–β phase transformation [46]. Therefore, β-Si3N4 is the major form used in Si3N4 ceramics. In particular, its elongated growth means that β-Si3N4 is a well-known tough ceramic because of the unique interlocking microstructure composed of elongated grains, the so-called self-reinforced microstructure. The elongated grains can act as a reinforcing phase to increase the fracture toughness by triggering various toughening effects, such as crack bridging, crack deflection and grain pull-out [7].

Self-reinforced β-Si3N4 is commonly produced from an α-Si3N4 powder that normally contains a low fraction of β-phase. Because of the strong covalent bonding, complete densification of Si3N4 generally requires sintering additives such as MgO, Al2O3 and rare-earth (RE) oxides. During sintering, the additives react with the SiO2 on the surface of Si3N4 and the Si3N4 itself to form a eutectic liquid phase, promoting densification, α–β phase transformation and β-grain growth. The α–β phase transformation facilitates the growth of large elongated grains, resulting in a typical bimodal microstructure composed of large elongated grains embedded in a small grained matrix [810], as schematically illustrated in figure 1. However, after sintering, the liquid phase invariably solidifies to an amorphous or partially crystalline secondary phase, which is located either at the grain triple junctions or the grain boundaries. The secondary phase strongly affects the thermal, physical and mechanical properties of Si3N4 ceramics, particularly at HTs.

Figure 1.

Figure 1

Scheme showing production of Si3N4 ceramics using α-Si3N4 raw powder.

The properties of Si3N4 ceramics can be improved by tailoring the microstructure, generally the grain morphology [9, 1117], secondary-phase chemistry [1838] and the grain orientation or texture [8, 3972], as summarized in table 1. Limited fracture toughness remains a major barrier to the wide structural application of Si3N4 ceramics; thus, toughening has been one of the most researched topics. It has been proven that the fracture toughness can reach 8 and even 10 MPa m1/2 by increasing the diameter and volume fraction of elongated grains [11, 12, 14, 15, 17], as predicted by the crack-bridging toughening mechanism [7]. However, the large elongated grains tend to reduce bending strength by acting as large structural flaws [11, 12, 14, 15]. For example, Kawashima et al [11] reported that the development of large elongated grains of ≈10 μm diameter produces a high fracture toughness of 11.3 MPa m1/2 but a bending strength of only 774 MPa. To resolve this contradiction, Hirao et al [13] developed a new ‘seeding’ method involving carefully tailoring the size and amount of well-dispersed large elongated grains in a fine-grained matrix, using morphologically regulated rodlike β-Si3N4 single particles. They obtained self-reinforced Si3N4 with a fracture toughness of 8.4 MPa m1/2 and a bending strength of ≈1 GPa. Using this method, Becher et al [15] showed that self-reinforced Si3N4 could achieve high fracture resistance (Inline graphic) combined with a steeply rising R-curve and high fracture strength (Inline graphic).

Table 1.

Microstructure design strategies and improved properties of Si3N4 ceramics.

Design strategy Main contents Key methods Improved propertiesa Representative examplesb
Grain morphology Diameter, aspect (i) Sintering additives Mechanical properties, 1994, Hirao et al [13], seeded Si3N4:
 ratio, grain size (ii) Sintering technique  thermal conductivity, σb=∼1 GPa and KIC=∼8.7 MPa m1/2
 distribution (bimodal (iii) Seeding  tribological and wear 2001, Hayashi et al [16],
 form)  resistance, etc. Yb2O3-MgSiN2-doped Si3N4: κ=140 W m−1 K−1.
Secondary phase Composition, (i) Sintering additives: Mechanical properties at 1999, Kleebe et al [27],
 chemistry  amount, distribution  Composition and amount  RT–HT, oxidation resistance, Sc2O3-doped Si3N4: KIC=∼10 MPa m1/2.
 and crystallization (ii) Post heating treatment  thermal conductivity, tribological 2001, Guo et al [31],
 and wear resistance, etc. Lu2O3-doped Si3N4: σb=660 MPa at 1500 °C
Grain orientation Direction of orientation (i) Templated grain growth Mechanical properties at 1999, Teshima et al [50],
 (Texture) Degree of orientation (ii) Hot-working  RT–HT, thermal conductivity, σb⊥=1.4 GPa and KIC,⊥=14 MPa m1/2.
 oxidation resistance, tribological 1999, Watari et al [52], κ=155 W m−1 K−1.
 and wear resistance, etc. 2006, Zeng et al [72],
σb⊥=739 MPa at RT and =738 MPa at 1500 °C.

aRT=room temperature, HT=high temperature.

b⊥ and ∥ mean the directions perpendicular and parallel to the c-axis of elongated β-Si3N4 grains for textured Si3N4.

However, the large elongated grains do not guarantee the occurrence of this toughening process [22]. For the crack-bridging toughening effect to occur, the reinforcing elongated grains must debond from the small matrix so that the crack tip is deflected along the grain face instead of splitting the elongated grains, while leaving intact elongated grains to bridge the crack [7]. The interfacial debonding process depends on the chemistry of the grain boundary phase, which is directly determined by the additive composition. It was reported that, in Si–Al–Y–O–N glasses, the increase in the Y/Al and O/N ratios promotes interfacial debonding, thereby improving fracture resistance [23, 26, 29]. Kleebe et al [27] reported that Sc2O3-doped Si3N4 exhibits a high fracture toughness of ≈10 MPa m1/2, because the nearly complete devitrification of the glassy phase allows the formation of a residual tensile stress field at triple pockets to trigger interfacial debonding and thus an elastic bridging effect. To specifically study the effect of the chemistry of the grain boundary phase on mechanical properties, Satet and Hoffmann [38] prepared Si3N4 ceramics with similar grain sizes using MgO and RE2O3 (RE = Sc, Lu, Yb, Y, Sm, La) as sintering additives. They showed that the lower interfacial strength that arises from the use of the larger RE3 + cation leads to higher toughness but a lower bending strength.

One of the most enticing applications of Si3N4 ceramics is their use as automotive engine components resulting in a high operation temperature, higher thermodynamic efficiency, lower engine weight and improved fuel consumption [2]. Ceramics used in this extremely severe application environment require excellent HT properties, such as strength, creep, thermal shock and oxidation resistance, which are essentially dependent on the secondary-phase chemistry. However, the glassy phase is necessary to fully sinter Si3N4 ceramics. The softening of the glassy phase is the major factor detrimental to their HT properties. There are two major strategies for improving the HT properties of Si3N4 ceramics: (i) the crystallization of the secondary amorphous phase by a postheating treatment [1821, 30] and (ii) the use of more refractory RE oxide additives that can provide crystalline secondary phases with high melting points [24, 25, 3037]. Previous studies have suggested that out of the RE oxides, Yb2O3 and Lu2O3 are the best additives for improving HT resistance up to 1400 and 1500 °C due to the formation of crystalline Yb4Si2O7N2 [24, 25, 32, 33, 36] and Lu4Si2O7N2 [31, 34, 37] grain boundary phases, respectively.

Furthermore, some superproperties in Si3N4 ceramics can be achieved by the formation of a textured microstructure in which elongated β-Si3N4 grains are oriented. Using seeding and tape casting, Hirao et al [43] developed textured β-Si3N4 with a bending strength of 1.1 GPa and a fracture toughness of 11.1 MPa m1/2 in the direction perpendicular to the grain alignment as well as a high Weibull modulus of 46. Using seeding and extrusion, Teshima et al [50] obtained highly textured Si3N4 with a bending strength of ≈1.4 GPa and a fracture toughness of ≈14 MPa m1/2 along the direction perpendicular to the grain alignment. Using the sinter-forging (SF) method, Kondo et al [51] produced textured Si3N4 with a superhigh bending strength of 2.1 GPa and a high fracture toughness of 8.3 MPa m1/2 in the direction perpendicular to the grain alignment. Moreover, texture design can lead to a further improvement in the HT properties. Using a less refractory additive composition (Y2O3–Al2O3), Park et al [68] reported the formation of textured and untextured Si3N4 exhibiting bending strengths of ≈600 and 200 MPa at 1400 °C, respectively. Using a more refractory additive composition (Lu2O3–SiO2), Zeng et al [72] reported the formation of textured Si3N4 exhibiting a bending strength of ≈740 MPa at both room temperature (RT) and 1500 °C along the direction perpendicular to the grain alignment, that is, no strength degradation occurs even at 1500 °C, indicating better HT strength retention than that of untextured materials. Moreover, textured Si3N4 exhibits unique anisotropic thermal conductivity [39, 45, 47, 52, 53, 58, 59, 62, 65, 72], tribological and wear properties [54, 55, 63, 66] and erosion behavior [67, 69]. For example, Watari et al [52] showed that thermal conductivities of 155 and 52 Wm−1 K−1 could be achieved in highly textured Si3N4 ceramic along the directions parallel and perpendicular to the grain alignment, respectively.

To date, a large number of publications are available reporting research on each microstructure design strategy. There have been several review articles reporting the progress in research on grain morphology [7375] and secondary-phase chemistry [76]. However, there is no full review article available concerning the research progress in the field of textured Si3N4 ceramics. Because texturing offers a unique opportunity to produce super-Si3N4 ceramics, a considerable amount of research has been carried out, leading to rapid progress in the processing and anisotropic properties of textured Si3N4 ceramics, which are expected to be used for a broad range of applications in extremely severe environments. Therefore, the objective of this article is to review research on textured Si3N4 ceramics, emphasizing the following areas: (i) grain growth behavior of β-Si3N4, (ii) typical texture structure and analysis, (iii) processing methods for Si3N4, (iv) anisotropic properties and (v) processing and properties of textured porous Si3N4. As an important related material, textured α-Sialon is also reviewed in this article.

Definition of textured Si3N4

There are two major forms of textured ceramics related to Si3N4 in the literature. One form comprises Si3N4 composites reinforced by secondary-phase particles or whiskers, such as Si3N4–BN [7779], SiCw/ Si3N4 [80] and SiCp/ Si3N4 [81]. In these composites, the texture results from the orientation of either elongated β-Si3N4 grains or secondary-phase particles (or whiskers) or both of them. The other form is self-reinforced Si3N4. Here, the texture results only from the preferential orientation of elongated β-Si3N4 grains. In this article, we focus only on the second form, the so-called textured Si3N4. Therefore, the term ‘textured silicon nitride (Si3N4)’ is defined as a self-reinforced silicon nitride with the preferential orientation of elongated β-Si3N4 grains. In the literature, another name, ‘anisotropic silicon nitride’, has also frequently been used for this material. In addition, it is worth emphasizing that textured Si3N4 in practice refers to textured β-Si3N4.

Grain growth of β-Si3N4

To help with understanding the texturing behavior in Si3N4 ceramics, we wish to briefly outline β-Si3N4 grain growth behavior, including anisotropic grain growth and abnormal grain growth, both of which play a critical role in the microstructural evolution and properties of Si3N4 ceramics.

Anisotropic grain growth

β-Si3N4 grains are known to grow preferentially into a typically hexagonal rodlike shape during liquid-phase sintering of either the α or β powder. This is attributed to the much faster growth rate along the [001] (or c-axis) direction than along the [210] direction, namely so-called anisotropic or elongated grain growth. The elongated grain growth allows Si3N4 ceramics to exhibit a unique self-reinforced microstructure and also a textured microstructure. Figure 2 shows (a) a typical texture microstructure containing aligned elongated β-Si3N4 grains along the c-axis direction [82] and (b) a schematic of a hexagonal rodlike β-Si3N4 grain growing from an initial nucleus or preexisting seed crystal. The white arrow indicates a typical large elongated grain growing epitaxially from an initial seed particle (dark gray color) along the c-axis direction during sintering, the so-called core-rim or shell structure [13, 15, 41, 82]. It is well-known that this elongated grain growth behavior of β-Si3N4 from the liquid phase is indeed intrinsic, although it is affected by the starting powder (α or β type) and sintering additives [83, 84]. To explain this mechanism of β-Si3N4 grain growth, several models have been proposed.

Figure 2.

Figure 2

Illustration showing typical anisotropic grain growth of β-Si3N4: (a) textured microstructure with aligned elongated β-Si3N4 grains along the c-axis and (b) schematic of an elongated hexagonal β-Si3N4 grain showing anisotropic grain growth in the [001] and [210] directions. The white arrow indicates a typical large elongated grain growing epitaxially from an initial seed particle (dark gray color) along the c-axis direction during sintering, the so-called core-rim or shell structure.

Rough–smooth plane model

On the basis of the periodic bond chain (PBC) analysis of β-Si3N4, Krämer et al [85] proposed a model that relates anisotropic grain growth to the crystal structure of β-Si3N4, in which the basal (001) plane is atomically rough but the prism (100) plane is atomically smooth. This prediction was also supported by a published transmission electron microscopy (TEM) study [86]. The rough basal plane allows a lower activation energy for nucleation and thus continuous growth in comparison with the smooth prism plane. Thus, this unique interface characteristic leads to the faster growth kinetics in the [001] direction than in the [210] direction, and thereby the observed anisotropic grain shape of β-Si3N4. On the basis of this model, Hwang et al [86] provided a reasonable explanation of the anisotropic grain growth of β-Sialon by developing new growth kinetics equations as an alternative to the classic Ostwald ripening model:

graphic file with name TSTA11660773M01.jpg

where d0 is the initial mean grain size, d is the mean grain size at time t and k is the rate constant. The Lifshitz–Slyozow–Wagner (LSW) theory predicts growth exponents of n= 2 and 3 for interfacial reaction control and diffusion control, respectively. In their work, they found that n= 3 for both length [001] and width [210] directions, indicating diffusion control. However, Lai and Tien [87] observed that n= 3 and 5 for the length and width directions of β-Si3N4 in sintered Si3N4 ceramics, respectively. The higher calculated activation energy for the width direction than for the length direction appears to support the rough–smooth plane model.

Anisotropic Ostwald ripening model

The classic Ostwald ripening model is not applicable for describing the anisotropic grain growth of β-Si3N4 because of the use of only one parameter, grain size, d(t); thus, Kitayama et al [88] developed a thermodynamic model for anisotropic Ostwald ripening using two parameters, w(t) and l(t), for the growth of β-Si3N4 in the width and length directions, respectively. In this model, the chemical potential difference between facets and the liquid phase is the driving force for mass transport. This mode reveals that the reduction of the aspect ratio after the α–β phase transformation observed experimentally is a consequence of anisotropic Ostwald ripening, and the length is a complicated function of time and the relationship between the diffusion and the interfacial reaction constants. This model predicts a growth exponent of n= 3 for a interfacial reaction controlled by kinetics, and higher values when the diffusion constants approach the interfacial reaction constants, both of which are consistent with the observed experimental results in the literature, whereas the exponent of n= 2 has never been observed for β-Si3N4, as predicted by the LSW theory. The same authors further extended this model to predict the grain growth during the α–β phase transformation [89] and the tip shape evolution of a β-Si3N4 crystal in the liquid phase [90]. The results demonstrated that the ratio of the interfacial reaction constants of the (100) and (001) planes and the α–β ratio are the key factors determining the anisotropic grain growth of β-Si3N4. The anisotropic Ostwald ripening model provides a successful explanation of why the effect of lanthanide additives on the anisotropic grain growth of β-Si3N4 is due to the change in the interfacial reaction constants between the (100) facets of β-Si3N4 crystals and the Ln–Si–O–N liquid phase [83].

Acid-base model

On the basis of the interfacial segregation mechanism of cations, Wang et al [91] proposed an acid-based model to explain the effects of the liquid-phase composition on the anisotropic grain growth of β-Si3N4. In this model, silicon and aluminum are a strong acid and weak acid, respectively, and the basicity of yttrium and RE bases increases as the ionic size increases. Again, this model treats the Si3N4 crystal as a strong acid, the (Si, Al, M)(O, N) liquid (M represents the network modifiers) as a weak acid and the modifier cations as strong bases that tend to segregate to the prismatic faces of the crystal. Thus, the anisotropic grain growth rate depends on the difference in the acidity between the crystal and the liquid. This model suggests that (i) the strong segregation inhibits the growth of prismatic planes relative to the basal plane and (ii) the anisotropic grain growth is enhanced as the ionic radius of the RE element increases because of the intensified segregation.

Differential binding energy model

The β-Si3N4 grain morphology is strongly affected by RE oxides, the most commonly used additives for sintering Si3N4 ceramics. However, the anisotropic grain growth behavior is difficult to rationalize in terms of the ionic radius of the RE cations or the cationic field strength. Painter et al [92] developed a differential-binding-energy DBE model to explain the role of RE oxides in the anisotropic grain growth of β-Si3N4. The DBE characterizes the energy contribution to the RE segregation due to the competition between the RE and Si for anion bonding. By selecting sixfold and fourfold coordinated O and N as reference hosts, respectively, the DBE is defined as:

graphic file with name TSTA11660773M02.jpg

where ΔEIJ is the binding energy of cation I (Si or RE), in an anion J (O or N) environment, which models the oxygen- or nitrogen-rich host sites in the ceramics. For an RE in an octahedral O environment, the binding energy is

graphic file with name TSTA11660773M03.jpg

where EK is the calculated total energy for the relaxed system K.

This model describes a physical mechanism in which the anisotropic grain growth originates from the site competition between REs and Si for bonding at β-Si3N4 interfaces and within the O-rich glass. This model gives a successful explanation of the observed variation of the anisotropic grain growth depending on the RE oxide [93], as illustrated in figure 3. In contrast with Lu, which has the highest DBE value, La has the lowest DBE value, leading to the strongest preferential segregation and highest binding strength to the prismatic grain surface, thereby the strongest anisotropic grain growth represented by the highest aspect ratio.

Figure 3.

Figure 3

Aspect ratio of β-Si3N4 grain correlated with DBE indicating the preference of the RE element to segregate to the Si3N4 grain surface in two representative glassy phases: Si-RE-Al and Si-RE-Mg oxynitride glasses. The DBE values of the RE elements are referenced against that of silicon (reproduced with permission from [93] ©2006 Elsevier Science Ltd). IGF=intergranular film.

Abnormal grain growth

To toughen Si3N4 ceramics, a major processing strategy is to promote the development of large elongated β-Si3N4 grains or abnormal grain growth, thereby resulting in the desired bimodal microstructure. However, the large elongated grains tend to reduce the strength of the materials, and the control of abnormal grain growth is crucial for the realization of high-performance Si3N4 ceramics. Dressler et al [94] demonstrated that the abnormal grain growth of β-Si3N4 is governed by the morphology, the grain-size distribution and the amount of β-Si3N4 nuclei present in the α-Si3N4 powder. The same authors proposed a model to explain the effect of β-Si3N4 nuclei size on abnormal grain growth, as shown in figure 4. This model predicts the dissolution of small β-grains if they are located within the diffusion gradient of a large β-grain. The authors suggest that the use of faceted and elongated β-Si3N4 nuclei crystals with a narrow size distribution is beneficial for obtaining the desired bimodal microstructure that gives Si3N4 ceramics both high strength and high fracture toughness.

Figure 4.

Figure 4

Model illustrating the dependence of the equilibrium concentrations of α- and β-phases on crystallite radius, indicating that at a given Si3N4 concentration in the liquid phase (C1 or C2) there exist radii Rα1,2 and Rβ1,2 for the α- and β-phases, respectively, above which the larger crystals grow, and below which the smaller crystals dissolve. CGα and CGβ are the equilibrium concentrations of the α- and β-phases, respectively, CGα>CGβ, because the enthalpy of the α→β transformation is negative (reproduced with permission from [94] ©1996 Elsevier Science Ltd).

Moreover, Emoto and Mitomo [95] demonstrated that the abnormal grain growth of β-Si3N4 is also governed by the morphology, the grain-size distribution and the amount of β-Si3N4 nuclei in the β-Si3N4 powders, where the α–β phase transformation is not involved. The driving force for abnormal grain growth, ΔC, is defined as

graphic file with name TSTA11660773M04.jpg

where ΔCr is the difference in solubility based on the particle size (the solubility of a smaller spherical grain is higher than that of a larger gain of the same phase, ΔCp is the crystal phase (the solubility of a spherical α-grain is higher than that of a β-grain of the same size and shape), and ΔCi is the interfacial energy (the solubility of a spherical grain is higher than that of a faceted grain of the same phase and volume). If a critical driving force exists, ΔCcrit, which causes abnormal grain growth; when ΔC> ΔCcrit abnormal grain growth occurs, but when ΔC< ΔCcrit normal grain growth occurs.

In the literature, there are several major processing methods available reported for controlling abnormal grain growth in Si3N4 ceramics, including

  • seeding with single β-Si3N4 crystal particles [1315],

  • sintering/annealing at high temperature for a long time [11, 12, 16, 17],

  • additive composition including composition and amount [16, 96, 97],

  • preheating treatment before final sintering [98, 99].

Texture structure and analysis

Texture structure

Textured Si3N4 occurs in two major microstructural forms: a, b-axis orientation and c-axis orientation. The a, b-axis oriented texture forms as a result of the orientation of the prism planes of elongated grains. This orientation is also called the planar orientation. However, the c-axis oriented texture forms as a result of the orientation of elongated grains along the long-axis (or c-axis) direction. This orientation is also called the axial orientation.

Figures 5(a) and (b) show typical features of the three-dimensional (3D) a, b-axis oriented texture, in which elongated β-Si3N4 grains have planar orientation, but are randomly oriented along the c-axis (or long-axis) direction. This texture can be easily identified by standard powder x-ray diffraction (XRD) analysis, as shown in figure 5(c). On the planes parallel to the c-axis, the diffraction peaks of the (hk0) planes parallel to the c-axis of the grains are substantially stronger, typically similar to those of (200) and (210) planes, and the (002) plane perpendicular to the c-axis disappears. However, on the planes parallel to the a, b-axis, the diffraction peaks of the (hk0) planes become relatively weak, whereas the diffraction peak of the (101) plane intersecting the c-axis is the strongest, and the diffraction peak of the (002) plane is enhanced. The degree of orientation in such a texture can be intuitively evaluated from the relative peak intensity of the (101) plane in the XRD pattern of the top plane (∥ c-axis). On the whole, the degree of orientation increases as the (101) peak intensity decreases.

Figure 5.

Figure 5

Typical (a) actual and (b) schematic microstructures, and (c) XRD patterns of a, b-axis (short-axis) aligned β-Si3N4.

Figures 6(a) and (b) show typical features of the 3D c-axis oriented texture, in which elongated β-Si3N4 grains are unidirectionally oriented along the c-axis direction. This texture can also be easily identified by XRD analysis, as shown in figure 6(c). On the planes parallel to the c-axis, the diffraction peaks of the (hk0) planes exhibit stronger relative intensities, whereas those of the (101) and (002) planes are very weak. However, on the planes perpendicular to the c-axis (also parallel to the a, b-axis), the diffraction peaks of the (hk0) planes are weak, whereas the diffraction peaks of the (101) and (002) planes are the strongest. The relative intensity of the diffraction peaks between the (101) and (002) planes depends on the degree of orientation along the c-axis direction. The higher orientation allows the (002) peak to exhibit stronger intensity than the (101) peak. If the (101) peak is not detected, a perfect c-axis aligned texture is considered to have developed.

Figure 6.

Figure 6

Typical (a) actual and (b) schematic microstructures, and (c) XRD patterns of c-axis (long-axis) aligned β-Si3N4.

Texture analysis

The texture analysis methods reported for Si3N4 ceramics in the literature include the following.

X-ray diffraction method

XRD is the most widely used method for identifying and analyzing the texture development in materials. Using the XRD data, the following methods can be used to analyze the texture in Si3N4, including.

a. Relative peak intensity

Relative peak intensity is a simple and direct qualitative method for identifying the orientation of Si3N4 grains in green compact and sintered samples, particularly in green compacts with very weak orientation because of the very limited amount of β-Si3N4, the most common case for α-Si3N4 raw powder. The relative peak intensity ratios mostly used for β-Si3N4 include I(210)/I(101), I(200)/I(101), I(200)/I(101) and I(002)/I(210). For example, the a, b-axis orientation (figure 5) can be identified by the higher I(210)/I(101) value on the plane parallel to the c-axis than that on the plane parallel to the a, b-axis, as shown in figure 5(c). Nevertheless, when Y2O3 is used as a sintering additive, the texture identification of α-Si3N4 green compact, produced by strong magnetic field alignment (SMFA) using slip casting (discussed later), encounters a problem that the (400) and (411) peaks of Y2O3 overlap with the (101) and (210) peaks of β-Si3N4, respectively. As the diffraction peak of the (002) plane (2θ ≈ 64°) is too weak to be distinguished, the value of I(210)/I(101) or I(200)/I(101) is preferably used. Because of its cubic structure, the magnetic field does not lead to the crystallographic orientation of Y2O3 during slip casting. Thus, corrected values for I(210) and I(101) can be calculated by the following equations [100]:

graphic file with name TSTA11660773M05.jpg
graphic file with name TSTA11660773M06.jpg

In equations (5) and (6), the factors 0.05 and 0.23 result from the calculated values of I(411) /I(222) and I(400) /I(222) from the measured XRD peak data of the cast sample in the absence of a magnetic field, respectively. This is in agreement with the data on the standard PDF card (No. 41-1105), showing that the values of I(400) /I(222) and I(411) /I(222) are 0.24 and 0.05, respectively.

b. Orientation indices

On the basis of XRD peak intensity data, the texture in β-Si3N4 ceramics can be evaluated by the orientation index of the (hikili) plane for β-Si3N4, Nhikili, which is defined as [101]

graphic file with name TSTA11660773M07.jpg
graphic file with name TSTA11660773M08.jpg

where Ihikili is the diffraction peak intensity of the (hikili) plane, and n is the number of diffraction peaks. Fhikili0 is calculated from data on the standard PDF card (No. 33-1160) of β-Si3N4.

c. Relative facial angle

A relative facial angle measured from the c-plane, θF, has been proposed to evaluate the degree of crystalline orientation from the intensity of XRD peaks. θF is defined as [102]

graphic file with name TSTA11660773M09.jpg

where θ hkl is the facial angle between the (hkl) and (00 l) planes, Ihkl is the peak intensity of the (hkl) plane obtained from the XRD pattern. The facial angle θ F is reduced to 0° when all crystals are oriented to the c-plane and to 90° when they are oriented to the a, b-plane. Li et al [103] reported that the static magnetic field allows the β-Si3N4 grains in sintered Si3N4 to exhibit a θ F of 80°, suggesting the a, b-axis orientation of β-Si3N4 grains.

d. Lotgering orientation factor

The Lotgering orientation factor has been widely used to evaluate the degree of texture in ceramics because of its simplicity, speed and inexpensiveness. The Lotgering orientation factor, fL, can be expressed as [104]

graphic file with name TSTA11660773M10.jpg

For a, b-axis orientation,

graphic file with name TSTA11660773M11.jpg

and for c-axis orientation,

graphic file with name TSTA11660773M12.jpg

where ∑I(hk0) and ∑I(00l) are the sums of peak intensities of the (hk0) and (00l) planes parallel and perpendicular to the c-axis of β-Si3N4 crystal, respectively, and ∑I(hk0) is the sum of peak intensities of all the (hkl) planes in the range of 2θ. The values of P are obtained from the sample, and the values of P0 are obtained from the standard PDF card (No. 33-1160) of β-Si3N4. If fL= 0, no grain orientation occurs; if fL= 1, perfect orientation is considered to have developed. A larger absolute value of fL implies a higher orientation.

However, the Lotgering orientation factor does not provide quantitative information on the degree of misorientation of grains or the textured volume fraction, thus, it allows only a semiquantitative analysis. Lotgering [104] pointed out that the accuracy of this method decreases with decreasing orientation. In addition, it was reported that the value of fL is dependent on the number of reflections used in the calculation [105]. To determine this effect on the Lotgering orientation factor of textured Si3N4, we conducted calculations for various ranges of 2θ, as shown in table 2. Our results suggest that the 2θ range has no significant effect on the fL value, even for lower orientation, and the effect is less than that observed by Jones et al in bismuth titanate ceramics [105]. This calculation may be helpful for shortening the experimental time when using this method to determine the degree of texture in Si3N4 ceramics.

Table 2.

Effect of 2θ range on the calculated Lotgering orientation factor in a, b-axis and c-axis oriented Si3N4.

Lotgering orientation factor (fL)a
a, b-axis orientation
c-axis orientation
2θ(°) HD MD LD MD LD
10–70 0.97 0.51 0.18 0.49 0.17
20–55 0.97 0.52 0.11
30–70 0.50 0.18

a HD=high degree, MD=medium degree, LD=low degree of orientation of β-Si3N4 grains.

e. Pole figure

A pole figure is a common and effective quantitative method for evaluating the texture development in materials [106]. A pole figure provides information on the distribution of preferential orientation in materials. There are two major types of pole figure for determining the texture in Inline graphic ceramics: (hk0) and (00l) pole figures, corresponding to the a, b-axis and the c-axis orientations, respectively. Figure 7 shows typical (a) (200) and (b) (002) pole figures obtained from a, b-axis and c-axis oriented β-Si3N4 ceramics, which were produced from the plane-strain compression (PSC) [107] and tape-casting alignment (TCA) techniques [108], respectively. The numbers in figure 7 represents multiples of a random distribution (mrd) of the grains obtained by comparing the orientation density on the projections between the untextured and textured samples; mrd=1 and >1 for a random orientation and preferential orientation, respectively. The larger the maximum mrd value, the higher the degree of orientation. The maximum mrd values in the (200) and (002) pole figures 8 and 15, indicate the development of highly a, b-axis [107, 109, 110] and c-axis oriented β-Si3N4 ceramics [110, 111113], respectively. Details of the (200) and (002) pole figure measurement can be found in [107] and [108], respectively.

Figure 7.

Figure 7

Typical (a) (200) and (b) (002) pole figures measured from textured β-Si3N4 prepared by PSC (reproduced with permission from [107] ©2000 Blackwell Publishing Ltd) and TCA techniques (reproduced with permission from [108] ©2003 Blackwell Publishing Ltd), corresponding to highly a, b-axis and c-axis oriented β-Si3N4, respectively. The number gives the multiple of random distribution (mrd), indicating the degree of preferential orientation. Both pole figures are perpendicular to the paper, i.e., the HF direction is located at the center of the (200) pole figure, and the sheet-stacking direction is located at the center of the (002) pole figure.

Figure 8.

Figure 8

Schematic illustration of texturing mechanisms of β-Si3N4 ceramics by (a) HW and (b) TGG method.

Figure 15.

Figure 15

Geometry of a whisker (OM′) in a polar coordinate system. OM″ is the projection of whisker OM′ on the LZ plane.

Scanning electron microscopy (SEM) image analysis

On the basis of the image analysis of SEM micrographs, Imamura et al [44] developed a formula for determining the degree of orientation of elongated β-Si3N4 grains, defined as

graphic file with name TSTA11660773M13.jpg

where ψ is defined as the half bandwidth of the peak in the curve of the area fraction of elongated grains plotted against angle θ (- 90°⩽θ ⩽ 90°) between the long axis of the grains and the tape-casting direction. Moreover, the grain orientation in Si3N4 ceramics can be quantitatively evaluated using an orientation factor [114], fP, which has been used for describing the planar state of fiber orientation in a discontinuous fiber-reinforced composite, defined as [114]

graphic file with name TSTA11660773M14.jpg

where N is the number of grains, and fP= 0 for the random orientation and fP= 1 for the perfect orientation.

Electron backscattered diffraction (EBSD) method

The EBSD method has been utilized to analyze the texture in β-Si3N4 ceramics by field emission scanning electron microscopy (FE-SEM) [115, 116]. EBSD can be used to obtain a grain orientation map and pole figure map, thus, it provides a quantitative method for evaluating the texture in materials.

In addition to the above texture analysis methods, other methods, such as the Rietveld refinement and the XRD rocking-curve method [117], can also be applied to β-Si3N4 ceramics [118, 119]. Each method has its own strengths and weaknesses. The choice of method, of course, depends on the ultimate purpose of the study. However, to develop a deep and quantitative understanding of the relationship between the texture and the anisotropic properties in Si3N4 ceramics, a better and more accurate texture analysis is needed. Among the various methods discussed, the quantitative texture analysis techniques include pole figure analysis, EBSD, Rietveld refinement, and the rocking-curve method. Nevertheless, in several studies, it has been suggested that the XRD rocking-curve method is the most time-efficient and accurate texture analysis technique, as demonstrated for Al2O3 [118] and PMN-28 PT [119] ceramics.

However, the quantitative texture analysis of β-Si3N4 crystal in α-Si3N4 green compacts is still difficult because of its limited amount (<10 wt%) and weak orientation, although the orientation, for example, induced by a strong magnetic field during slurry consolidation, can indeed be identified by the XRD peak intensity ratio. Quantitative texture analysis based on a polarized-light microscopy technique has been developed to evaluate the weak orientation in alumina green compacts [120, 121]. If β-Si3N4 exhibits optical anisotropy, this method can also be applied to evaluate its grain orientation in an α-Si3N4 green compact.

Processing of textured Si3N4

In this article, the processing methods of textured Si3N4 ceramics have been classified into two types: hot-working (HW) and templated grain growth (TGG), as schematically illustrated in figure 8. In HW, the texture is primarily attributed to the rotation of elongated β-Si3N4 grains induced by a uniaxial stress imposed during sintering or forging. In TGG, the texture is attributed to the abnormal epitaxial growth of β-Si3N4 grains on initially oriented β-Si3N4 template (or seed) particles by the consumption of the fine-grained matrix during sintering. It is worth emphasizing that the term ‘templated grain growth’ was first proposed by Seabaugh et al [122] in 1997, and the authors defined this term as the use of oriented particles to obtain a texture by grain growth in the anisotropic direction of the template particles. Since then, TGG has been recognized as a term for a specific technique of fabricating textured ceramics. Both HW and TGG are thoroughly discussed in terms of their fabrication processes and the processing factors controlling the texture development.

Hot-working method

There are three different types of HW methods, including hot-pressing (HP), hot-forging (HF) and sinter-forging (SF).

Hot-pressing

Hot-pressing is a common densification method of Si3N4 ceramics that combines pressing and sintering processes while applying the uniaxial pressure. The starting material used for HP can be either loose powder or a compact form. This method enables the use of a smaller amount of sintering additives to obtain dense Si3N4 ceramics at a relatively lower temperature. At the same time, it promotes texture development in β-Si3N4. It has been reported that the texture in hot-pressed Si3N4 depends on the phase of the starting Si3N4 powder: α or β. In the case of raw α-powder, the large elongated β-Si3N4 grains, preferentially developed by the α–β phase transformation, tend to be oriented with the c-axis (long axis) in the HP direction, thereby typically resulting in the a, b-axis oriented texture (figure 5) [8, 40, 53, 62, 65, 110, 123, 124]. The texture development is due to a combination of grain rotation and preferential grain growth, as suggested by Lee and Bowman [110]. However, in several studies it has been reported that, if fine equiaxed β-powder is used, the equiaxed β-Si3N4 grains tend to be oriented with the c-axis parallel to the HP direction, or the basal plane perpendicular to the HP direction, thereby resulting in a c-axis oriented texture (figure 6) [113, 125]. Although this orientation is difficult to distinguish by SEM, it was confirmed by XRD pole figure analysis. Because of the limited grain growth, the grain orientation is responsible for the texture development during HP. When this material is submitted to annealing, the preferential grain growth contributes to the texture development [113].

In addition, Lee and Bowman [110] found that the texture development was locally nonuniform within the sample because of the simple shear form induced by the frictional effects resulting from the application of BN coating, suggesting the potential of evaluating the texture as a function of its position within the sample. Santos et al [126] reported that because of a low degree of steric hindrance, a larger amount of sintering additives leads to a lower degree of orientation of the a, b-axis of elongated β-Si3N4 grains during HP despite the increased grain size and aspect ratio, as also clearly evidenced by the tendency of anisotropy in fracture toughness.

Hot-forging

In ceramics, HF is a superplastic forming technique, in which a uniaxial compressive or tensile stress is used to forge dense Si3N4 ceramics during heating. Dense Si3N4 ceramics can be fabricated by various sintering techniques, such as HP, pressureless sintering (PLS), and gas pressure sintering (GPS). In the case of tensile stress, the c-axis oriented texture preferentially develops. Wu and Chen [112] reported that the large tensile strains at 1550 °C allow hot-pressed β-Sialon with a fine-grained microstructure at 1550 °C to develop elongated grains and a very strong c-axis oriented texture. The degree of orientation increases with increased tensile strain. At the same time, they also observed a pronounced strain-induced grain growth phenomenon at a lower temperature, probably as a result of grain coalescence and grain unimpingement during deformation. However, in the case of compressive stress, the texture microstructure depends on the operating method of HF: simple compression (SC) (or axisymmetric forging) or PSC [111], as schematically shown in figure 9. SC tends to form the a, b-axis oriented texture [110, 111], but PSC tends to form the c-axis oriented texture [48, 61, 64, 107, 110, 111, 113].

Figure 9.

Figure 9

Schematic illustration of superplastic deformation under (a) SC and (b) PSC.

Sinter-forging

SF is a novel technique that combines sintering and HF processes into a one-step process. In this technique, green compacts are produced by conventional cold-pressing and subsequent cold-isostatic pressing (CIP). Venkatachari and Raj [127] first developed this technique to obtain high-strength ceramics without large pore flaws, which were eliminated by the shear strain. They showed that, in Al2O3 ceramics, the strength began to increase at 20% strain and reached a maximum value at 60% strain. Kondo et al [51] used this technique to successfully develop textured Si3N4 with supermechanical properties. Using this technique they fabricated super-Si3N4 ceramics with a three-point bending strength of 2.1 GPa and a fracture toughness of 8.3 MPa m1/2 in the direction perpendicular to the grain orientation, which is by far the highest reported value for strength. The super-Si3N4 was prepared by the following process: powder mixture (α-Si3N4–5 wt%Y2O3–3 wt%Al2O3) → cold-pressing (steel die: 40 mm length and 20 mm width) Inline graphic Inline graphic (graphite channel die: 80 mm length and 20 mm width, pressure: 49 N, 1750 °C for 3 h in 0.1 Mpa N2) → c-axis oriented β-Si3N4. The heating profile used for SF and photographs of samples are shown in figure 10. Similar to HF, the texture also depends on the forging method: SC [128, 129] or PSC [51].

Figure 10.

Figure 10

SF of Si3N4 ceramics: (a) heating profile and (b) view of CIP green (left) and sinter-forged (right) samples. During SF, the height shrinkage is ∼73% based on the data provided by the authors (reproduced with permission from [51] ©1999 Blackwell Publishing Ltd).

Grain rotation model

Assuming a rodlike morphology for β-Si3N4 grains and the absence of any interaction between neighboring grains, grain rotation models have been developed to describe the texture development in Si3N4 ceramics by Wu and Chen [112] and Lee and Bowman [111]. For grain rotation under tension, Wu and Chen obtained the following equation

graphic file with name TSTA11660773M15.jpg

Lee and Bowman described grain rotation under compression as follows.

graphic file with name TSTA11660773M16.jpg
graphic file with name TSTA11660773M17.jpg
graphic file with name TSTA11660773M18.jpg
graphic file with name TSTA11660773M19.jpg

In equations (15)–(19), φ and θ are the Euler angles describing the orientation of the grain lone axis with respect to a fixed 3D coordinate system, the subscripts i and f, respectively, denote the initial and final states during deformation, R is the aspect ratio of the grain, ε is the tensile or compressive strain. Equations (15), (16) and (18) indicate that when R becomes sufficiently large for the term (R2-1)/(R2+ 1) to approach a constant of 1, the degree of orientation increases as the strain increases but independently of the aspect ratio. In this case, equation (15) is identical to equation (16).

Because most Si3N4 ceramics exhibit large values of R, the texture development in hot-forged Si3N4 basically depends on the strain, i.e., the texture development is essentially attributed to the grain rotation mechanism. This has been demonstrated both experimentally and theoretically by Lee and Bowman [111], as shown in figure 11. Nevertheless, the composition and amount of the additives have little effect on the degree of texture. Although the a, b-axis oriented texture can be enhanced by a larger height reduction, when the height reduction reaches a limit in the range of 57–69%, the degree of texture can no longer increase significantly because of the steric hindrance effect. Moreover, the authors demonstrated that the degree of c-axis texture increased significantly with increased height reduction during PSC. Xie et al [64, 107, 113] also demonstrated the dependence of texture development on strain during PSC within the true strain range of 0.30–0.95. The same authors further showed that when the aspect ratio was relatively small, for example, R< 2.5, it had a significant effect on the degree of texture in fine-grained β-Si3N4. As there is no difference in the mechanisms of the deformation and texture formation between HF and SF, the processing factors controlling the texture formation should be the same for both of them. Specifically, the degree of texture in sinter-forged Si3N4 should be governed by the forging strain. Table 3 lists typical examples of textured Si3N4 ceramics fabricated by HF and SF.

Figure 11.

Figure 11

Effect of height reduction during axisymmetric forging on the theoretical and experimental maximum degree of texture in the direction perpendicular to the HF. The open circles represent data from experimental pole figures, and the solid circles and curve represent the theoretical data obtained from grain rotation theory based on the Jeffery model. 5M and 10M represent 5% and 10% MgO by weight, respectively, and 10Y and 20Y represent 10% and 20% YAG by weight, respectively. The starting material was α-phase powder. The hot-forged Si3N4 was prepared from Si3N4 billets, which exhibited 100% β-phase and >85% TD after PLS at 1750 °C. Axisymmetric HF was conducted at 1750 °C for 20–60 min with a load of 18 MPa, resulting in 18–69% height reductions and >94% TD (reproduced with permission from [111] ©1994 Blackwell Publishing Ltd).

Table 3.

Examples of textured Si3N4 fabricated by HF and SF.

Sintering HF
Authors Composition Condition Phase Density (% TD) Condition Eng. strain ε (%)a Density (% TD) Texture type Degree of texture
Wu and Chen [112] α-Si3N4+AlN+Al2O3+Y2O3 HP: 1550 °C β-Sialon Tension: 1550 °C 232 c-axis oriented Pole figure analysis, Max. mrd=∼3∥ c-axis
Lee and Bowman [111] α-Si3N4+20 wt% YAG PLS: 1750 °C 100% β-Si3N4 >85 SC: 1750 °C 69 >94 a, b-axis oriented Pole figure analysis, Max. mrd=4.3⊥ c-axis
α-Si3N4+20 wt% YAG PLS: 1750 °C 100% β-Si3N4 >85 PSC: 1750 °C 65 >94 c-axis oriented Pole figure analysis, Max. mrd=6.9∥ c-axis
Xie et al [113] β-Si3N4+7 wt% cordierite HP: 1750 °C 100% β-Si3N4 >99 PSC: 1600 °C 159 c-axis oriented Pole figure analysis, Max. mrd=4.0∥ c-axis
Kondo et al [61] α-Si3N4+5 wt%Y2O3+2 wt%Al2O3 GPS 100% β-Si3N4 >99 PSC: 1750 °C 50 >99 c-axis oriented Orientation angleb
 Plane ⊥ Pressing: 33.4°
 Plane ∥ Extrusion: 24.1°
SF
Kondo et al [51] α-Si3N4+5 wt% Y2O3+2 wt%Al2O3 PSC: 1750 °C→100% β-Si3N4 73 >98 c-axis oriented Not reported
Kondo et al [128] α-Si3N4+13 wt%Yb2O3+2 wt%SiO2 SC: 1900 °C→100% β-Si3N4 60 ∼97 a, b-axis oriented Plane ⊥ Pressing:c
I(101)/I(210)=0.05
Kondo et al [129] α-Si3N4+8 wt%Lu2O3+2 wt%SiO2 SC: 2000 °C→100% β-Si3N4 60 >97 a, b-axis oriented Plane ⊥ Pressing:c
I(101)/I(210)=0.035

a Eng. strain ε=(ΔL/L0)×100% where L0 is the initial dimension of the sample, ΔL is the dimension change during deformation. If only true strain, e, is given in the literature, Eng. strain is calculated by the equation e=ln(1 +ε).

b The orientation angle is the angle between the grain long axis and the extruding direction.

c I(101)/I(210)=1.06 for the isotropic case (from PDF No. 33-1160 of β-Si3N4).

Surperplastic forming has been seen as a reliable and unique net-shape manufacturing technique of high-performance Si3N4 ceramics [130]. The constitutive equation for steady creep can generally be expressed as [131]

graphic file with name TSTA11660773M20.jpg

where Inline graphic is the strain rate, σ is the stress, d is the grain size, p and n are the grain-size and stress exponents, respectively, and A is a coefficient of proportionality. To achieve a larger superplastic deformation, a fine-grained microstructure with a tailored grain boundary phase is necessary for composition design and processing control. Therefore, the Si3N4 ceramics used for HF are normally sintered at relatively low temperatures to minimize the β-Si3N4 grain growth, as shown in table 3.

Templated grain growth method

Template synthesis

In the TGG method, the key point is the initial alignment of the single β-Si3N4 crystals, the so-called ‘seeds’ or ‘template’, during powder formation by various alignment techniques, such as cold-pressing alignment (CPA), extrusion alignment (EA), tape casting alignment (TCA) and SMFA, as summarized in table 4. In most cases, the rodlike morphology is the prerequisite for the alignment of β-Si3N4 crystals, specifically for EA and TCA. More importantly, a regulated rodlike morphology is also of particular importance to control the bimodal microstructure in self-reinforced Si3N4 ceramics. In EA and TCA, the β-Si3N4 seed particles must meet two major requirements: (i) single crystals must have a rodlike shape and (ii) the crystal size must be larger than that of the matrix of Si3N4 raw powder. In addition, an optimum aspect ratio as well as a narrow size distribution is also important. There are two major methods for synthesizing rodlike β-Si3N4 single crystals: powder sintering (PS) and combustion synthesis (CS).

Table 4.

Various orientation techniques of β-Si3N4 template for producing textured Si3N4 by the TGG method as reported in the literature.

Product
Orientation technique Requirements for template Sintering method Texture Form Shape Key factors affecting texture Comments
CPA Preexisting or externally added HP a, b-axis Bulk Simple Size and amount of template Poor texture control and
 Shape: rodlike PLS  aligned Pressing pressure  few studies
 Size: sufficiently large GPS Sintering conditions
 Aspect ratio: sufficiently high HIP
SPS
EA Externally added c-axis Rod Simple Size and amount of template Suitable for producing
 Shape: rodlike  aligned tube EA conditions c-axis aligned Si3N4
 Size: sufficiently large bulk Sintering conditions  parts with limited
 Aspect ratio: ≽4  shapes
TCA Externally added c-axis Sheet Simple Size and amount of template Suitable for large-scale
 Shape: rodlike  aligned bulk Slurry dispersion  production of c-axis
 Size: sufficiently large TCA conditions  aligned Si3N4 parts with
 Aspect ratio: ≽4 Sintering condition  limited shapes
SMFA Preexisting or externally added a, b-axis Film Complex Size and amount of template Suitable for producing
 Shape: free  or c-axis bulk Slurry dispersion a, b-axis or c-axis
 Size: ∼ nanosize  aligned SMFA conditions  aligned Si3N4 parts with
Sintering conditions  complex shapes

a. PS method [132134]

This method involves the heat treatment of an α-Si3N4 powder normally mixed with sintering additives at high temperatures, followed by grinding, sieving and a series of acid rinse treatments. During heating, the sintering additives provide a liquid phase to allow the α-phase to transform into the β-phase and grow anisotropically into rodlike β-Si3N4 crystals. The crystal morphology can be tailored by the type of α-Si3N4 powder, additives and heating conditions. Hirao et al [132] developed a systematic method for synthesizing high-quality rodlike β-Si3N4 single crystals from a powder system of α-Si3N4, Y2O3 and SiO2, in which crystals with average aspect ratios of 4 and 10 were obtained using molar ratios of Y2O3:SiO2= 1:2 and 2:1, respectively. The mechanism underlying the effect of the Y2O3: SiO2 ratio on the β-Si3N4 crystal morphology was studied by Kitayama et al [135], and they proposed that this effect is due to the difference in the solubility of Si3N4 caused by a change in the Y2O3: SiO2 ratio in the liquid phase. To obtain purer single β-Si3N4 crystals, Hirao et al [132] developed an efficient four-step acid rinse treatment method: (i) a mixed solution of concentrated hydrofluoric acid (HF) and nitric acid (HNO3) at 60 °C to dissolve the residual glassy phase, (ii) concentrated sulfuric acid (H2SO4) solution at 160 °C to eliminate yttrium compounds possibly formed during the previous treatment, (iii) dilute HF solution to dissolve the SiO2 film formed on the particle surface and (iv) concentrated ammonium solution at 60 °C to remove the HF adsorbed at the surface. Each rinse treatment was followed by washing the powders in distilled water at room temperature. Figure 12 shows SEM micrographs of commercially available β-Si3N4 whiskers (SN-WB) provided by UBE Industries Co., Ltd., and rodlike β-Si3N4 single crystals with an average aspect ratio of 4 prepared by PS. These two typical types of β-Si3N4 crystals have been widely used as the seeds or template for investigating the texture development and anisotropic properties of Si3N4 ceramics via TGG.

Figure 12.

Figure 12

SEM micrographs of (a) commercially available β-Si3N4 whiskers (SN-WB, UBE industries Co., Ltd) and (b) rodlike β-Si3N4 single crystals with an average aspect ratio of 4 prepared by PS (reproduced with permission from [56] ©2000 Blackwell Publishing Ltd).

b. CS method [136140]

CS, so-called self-propagating high-temperature synthesis (SHS), is a very powerful technique for manufacturing various ceramic materials such as powders and bulk materials that utilizes the heat generated by an exothermic reaction to sustain itself in the form of a combustion wave after external ignition. This method has been applied to the synthesis of rodlike β-Si3N4 single crystals using Si raw powder with α-Si3N4 powder as a diluent and an additive (e.g. Y2O3) powder. During combustion, rodlike β-Si3N4 crystals are formed by the nitridation of Si and the α–β phase transformation. The crystal morphology can be controlled by the type and amount of additives [139]. To obtain pure single crystals, acid rinse treatments are required, similar to those used in PS. Table 5 gives the processes for producing rodlike β-Si3N4 seed crystals and their characteristics, as reported in the literature.

Table 5.

Processes for obtaining rodlike β-Si3N4 seed crystals reported in the literature.

Seed characteristicsb
Authors Methoda Powder composition Heating conditions w (μm) R
Hirao et al [132] PS α-Si3N4(SN-E5)+5 mol%Y2O3+2.5 mol%SiO2 1850 °C, 2 h, 0.5 MPa N2 1–2 10
Hirao et al [13] PS α-Si3N4(SN-E5)+5 mol%Y2O3+10 mol%SiO2 1850 °C, 2 h, 0.5 MPa N2 0.96 4
Imamura et al [56] PS α-Si3N4(SN-E10)+5 mol%Y2O3+10 mol%SiO2 1850 °C, 2 h 0.5 MPa N2 0.47 4.2
α-Si3N4(SN-E2)+5 mol%Y2O3+10 mol%SiO2 1850 °C, 2 h 0.5 MPa N2 1.29 4.0
Ramesh et al [133] PS α-Si3N4(SN-E10)+0.5 wt%Y2O3 1800 °C, 0.35 atm N2 ∼4
Chen et al [138] CS Si+48.8 wt% α-Si3N4+4.8 wt%Y2O3 5 MPa N2 ∼10
Peng et al [140] CS Si+48.8 wt% α-Si3N4+2.4 wt%MgSiN2 3 MPa N2 0.5–2 10–15

a PS=powder sintering, CS=combustion synthesis.

b w=width, R is the aspect ratio, given by R=l/w (l=length).

Template alignment technique

5.2.2.1. CPA technique

Cold-pressing is most widely used for forming ceramic powders, usually using a metal die, particularly a stainless-steel die. During pressing, the shear strain tends to orient the anisotropic particles (e.g. plateletes, ellipsoids, whiskers) by grain rotation, thereby resulting in texture development via TGG during sintering. Goto et al [141] reported the formation of textured Si3N4 by CPA of an α-Si3N4 powder containing needle-like β-Si3N4 particles, followed by PLS. In their work, the needlelike β-Si3N4 particles were obtained by the heat treatment of α-Si3N4 powder with some Y2O3 and Al2O3 at 1700 °C. The texture of Si3N4 is characterized by the c-axis of elongated β-Si3N4 grains being perpendicular to the cold-pressing direction, i.e. the a, b-axis oriented texture. Although this method has seldom been studied or used because of its poor texture control, it implies that the cold-pressing operation should be conducted carefully for Si3N4 powders containing rodlike β-Si3N4 particles to produce randomly oriented Si3N4 ceramics. Particular care is required in controlling the pressure parameter, because the orientation of rodlike β-Si3N4 particles is enhanced at a higher pressure.

5.2.2.2. EA technique

In this technique, the unidirectional alignment of rodlike β-Si3N4 seed particles can be achieved by extrusion through a die of the desired shape (e.g. square, circle) in an extrusion machine [50, 63, 142144], as schematically illustrated in figure 13(a). The alignment is due to the grain rotation induced by shear strain. The grain rotation mechanism can be described by the model of Wu and Chen, i.e. the degree of orientation directly depends on the shear strain and aspect ratio of the seeds. Prior to the preparation of a ceramic paste for extrusion, a two-step mixing process is usually performed to achieve better homogenization between the seed particles, sintering additive and Si3N4 powder. In both steps an organic medium is preferably used, such as methanol or isopropyl alcohol. In the first step, intense milling, such as planetary milling, is used to mix the sintering additives and Si3N4 powder. The addition of seed particles follows the second step, and gentle milling should be used to avoid damage to the seed crystals, such as ball milling using nylon balls. To prepare the ceramic paste, water is normally used as a mixing medium, and other organic additives are added to improve the formability of ceramic powders. Table 6 lists the EA processes used for producing textured Si3N4 ceramics as reported in the literature. Two shapes are generally used for the extruded body: rods and rectangles. The stacking condition depends on the sintering method used. To achieve complete densification, PLS and GPS require CIP to increase the green density in comparison with HP.

Figure 13.

Figure 13

Schematic illustration of producing Si3N4 green body with oriented rodlike β-Si3N4 seeds by (a) EA and (b) TCA techniques.

Table 6.

Extrusion alignment processes for producing textured Si3N4 as reported in the literature.

β-Si3N4 seed
Authors Morphology Amount Additives for aqueous ceramic paste Extruded body Stacking methoda GDb (% TD) Sintering SDb (% TD)
Muscat et al [142] Whisker (SN-WB) 15 wt% Plasticizer: Rectangle CIP: 340 MPa ∼53 PLS, 95%
w=0.1–0.3 μm, Hydroxy-propyl cellulose (HPC) 60 mm×8 mm×8 mm  1800 °C, 1 h
R=15–50 (Mw=300 000)
Teshima et al [50] Rodlike shape 5 vol% Binder: YB-113G and YB-113C Rod CIP: 500 MPa ∼54 GPS, ∼96%
w=0.5 μm, R=4 8 mm×50 mm  1850 °C, 2 h
 0.9 MPa N2
Nakamura et al [63] Rodlike shape 5 wt% Binder: YB-113G and YB-113C Rod CIP: 500 MPa ∼54 GPS, >98%
w=0.44 μm, R=8 8 mm×50 mm  1850 °C, 2 h
 0.9 MPa N2
Zou et al [143] Whisker (SN-WB) 3 wt% Binder: PVA Wire 0.2 mm Cold pressing: 40 MPa HP,
Plasticizer: glycerine  1820 °C, 1.5 h,
Lubricant: paraffin  0.1 MPa N2
Belmonte et al [144] Rodlike shape 5 wt% Binder: PEG 1000 Rectangle Cold pressing: 1 MPa HP, ∼99%
w=1.6 μm, R=4 Plasticizer: cellulose derivative 50 mm×20 mm×2 mm  1750 °C, 2 h,
Lubricant: PEG 200  0.1 MPa N2

aCIP=cold isostatic pressing, CP=cold pressing.

b GD=green density, SD=sintered density and TD=theoretical density.

5.2.2.3. TCA technique

Tape casting is a low-cost technique for the large-scale fabrication of ceramic substrates and bulk ceramics, in which a slurry is cast on a substrate through the action of a doctor blade that levels the slurry. It is also a powerful technique for producing a green sheet with oriented anisotropic ceramic particles, thereby developing textured ceramics, as schematically illustrated in figure 13(b). Tape casting can produce a green body with unidirectionally oriented rodlike β-Si3N4 seed particles. Figure 14 shows a typical tape-casting procedure for producing textured Si3N4 ceramics, in which a nonaqueous slurry is prepared by a three-step milling process. The initial planetary milling is used to homogenize the Si3N4 powder and sintering additives, which is followed by a two-step ball milling process. Note that various additives are needed for the preparation of the tape-casting slurry. The role of the dispersant is to ensure the stability of the slurry, and the roles of the binder and plasticizer are to ensure adequate strength and flexibility of the tape. Details of the tape-casting technology for producing ceramics can be found in several review articles [145147]. Two major nonaqueous tape-casting processes have been developed to produce textured Si3N4, as listed in table 7.

Figure 14.

Figure 14

Processing route for TGG using TCA and GPS.

Table 7.

Nonaqueous TCA processing of textured Si3N4 by GPS as reported in the literature.

Stacking condition Green body
Slurry additives Temp. Pressure CIP GD GPS SD
Authors Solvent Dispersant Binder Plasticizer (°C) (MPa) (MPa) (% TD) conditions (% TD)
Hirao et al [43] Toluene/ n-butanol=4/1 Diamine PRT Polyvinyl butyral (PVB) Dioctyl adipate 130 70 500 54 1850 °C, 6 h, 1 MPa N2 >99
Park and Kim [153] Methylisobutyl ketone (MIBK) Hypermer KD1 Polyvinyl butyral (PVB) Dibutyl phthalate 80 50 250 1875 °C, 4 h, 2 MPa N2 >99

Recently, because of the safety, economic and environmental considerations, the aqueous tape-casting of Si3N4 ceramics has received attention [148150], as listed in table 8. Although these studies do not involve the use of β-Si3N4 seeds, they will undoubtedly be of great help in producing textured Si3N4 by the aqueous TCA technique. Similarly to EA, the slurry incorporated with β-Si3N4 seeds should be gently mixed, typically by ball milling using nylon balls. Also, CIP treatment is required for PLS and GPS to achieve complete densification.

Table 8.

Aqueous tape casting of Si3N4 ceramics as reported in the literature.

Slurry additives
Authors Sintering additives Dispersant Binder Plasticizer Others Slurry pH Stacking Sintering SD (% TD)
Bitterlich et al [148, 149] Y2O3+Al2O3 Dolapix A88: Copolymer Defoamer: 10 No stacking, GPS: 1800 °C, 1 h, ∼98
 PC 33=90:1  (Mowilith DM  Fatty alcohol  Single tape for  5 MPa N2 [148]
 (in weight)  765)  sintering
 Pressure: GPS: 1800 °C, 1 h, ∼98
 14 MPa 5 MPa N2 [149]
Zhang et al [150] Y2O3+Al2O3 PEI+Citric acid PVA Glycerol 9.3 Simple stacking HP: 1800 °C—30 min >95
ammonium salt  in a graphite die SPS: 1600 °C—3 min

a. Whisker orientation model

A model for whisker orientation during tape casting has been proposed by Wu and Messing [151], described as follows

graphic file with name TSTA11660773M21.jpg
graphic file with name TSTA11660773M22.jpg

where R is the aspect ratio of a whisker, or more generally, any shaped particle (e.g. platelet), G is the shear rate during tape casting, and α and β are the angles determining the position of the whisker in a polar coordinate system, as shown in figure 15. In addition, if φ is defined as the angle between the tape-casting direction and the whisker orientation, the relationship between φ, α and β is

graphic file with name TSTA11660773M23.jpg

The perfect orientation requires φ = 0 or α = β = 90. Equations (21) and (22) predict that the whisker orientation is governed by the following factors

  • Shear rate

    For a Newtonian suspension, the shear rate G is directly proportional to the tape-casting rate and inversely proportional to the tape thickness. Therefore, the rates of whisker orientation can be increased by increasing the shear rate (e.g. casting rate) or by decreasing the blade opening during casting.

  • Aspect ratio

    As R becomes large, the term (R2-1)/(R2+ 1) approaches 1. Thus, longer whiskers undergo a faster rate of orientation. Moreover, if R is greater than 10, these longer whiskers should behave in a similar manner assuming that the whisker content in the slurry is below the percolation threshold and that whiskers do not interact with one another or with the particulate phase during tape casting.

  • Initial position

    The initial position of a whisker during tape casting also affects the orientation. There are three different initial positions as follows.

  • Pure rotation: if the whisker lies in the LZ plane, it only rotates about the T-axis, thereby it is aligned in the L-axis or casting direction.

  • Pure spinning: if the whisker lies in the T-axis, it only spins about its own axis and cannot be aligned in the L-axis for a 1D velocity gradient.

  • Mixed mode: during tape casting, the whisker alignment occurs as a result of both rotation and spinning. As β approaches 90°, the net torque exerted by the fluid on the whisker approaches zero.

However, the factors controlling the shear rate G are not considered in Wu and Messing's model. During tape casting, the flow behavior of a Newtonian fluid is due to a combination of pressure and viscous forces. To describe the flow characteristics of the slurry during tape casting, Kim et al [152] proposed a parameter Π to define the ratio of pressure force to viscous force as follows:

graphic file with name TSTA11660773M24.jpg
graphic file with name TSTA11660773M25.jpg

where ΔP is the pressure exerted by the slurry head, H1 is the blade gap used, η is the viscosity of the slurry, L is the length of the doctor blade, U is the casting velocity, ρ is the density of the slurry and H is the height of the slurry in the reservoir. The viscous force determines the shear behavior necessary for particle rotation and alignment. Equation (24) indicates that the degree of whisker orientation can be increased by increasing the casting velocity (U), slurry viscosity (η) and whisker fraction.

b. Modified TCA technique

As discussed above, the whisker orientation strongly depends on the casting velocity, i.e. the velocity gradient along a single whisker. As reported previously, the length of rodlike β-Si3N4 seed crystals is usually in the order of ∼10 μm, which is smaller than 1/104 of the width of the flow during tape casting, thereby resulting in a lower velocity gradient in the local region. In this case, using the method of only adjusting tape-casting velocity, it is very difficult to achieve a highly unidirectional alignment of seed particles by conventional tape casting, particularly for a low content of seeds, as shown in figure 13(b). This is the reason why the TCA technique does not normally produce as highly c-axis aligned β-Si3N4 as the EA technique. To solve this problem, Park and Kim [153] developed a modified TCA, in which an array of sharp guides (e.g. pins) are set at the exit of the reservoir to divide the flow into n narrow flows, thereby increasing the torque for aligning the whiskers by a factor of n2, as schematically illustrated in figure 16. Using this method, they produced highly c-axis oriented β-Si3N4 with 3 wt% β-Si3N4 whiskers (UBE SN-WB), using an array of sharpened pins 0.7 mm apart from each other, a casting speed of 10 mm s−1 and a blade gap of 0.45 mm.

Figure 16.

Figure 16

Schematic illustration of modified TCA of the whiskers using an array of sharpened pins set at the exit of the reservoir.

5.2.2.4. SMFA technique

Recently, SMFA has received more attention for the fabrication of textured nonmagnetic ceramics, such as Al2O3, TiO2, AlN, ZnO, hydroxyapatite (HAp). When a strong magnetic field (typically Inline graphic ) is imposed, the nonmagnetic ceramic particles are oriented during slurry consolidation, thereby resulting in the formation of textured ceramics during sintering. Important developments have been discussed in a review article by Sakka and Suzuki [154]. Compared with the other three techniques, SMFA has outstanding advantages of the independence of grain morphology and its applicability to even very fine nanoscale grains. Another attraction is the versatility and simplicity of the fabrication process, because various colloidal forming approaches are well developed, such as slip casting [155], electrophoretic deposition [156] and gel casting [157].

a. Magnetic field alignment theory

  • Rotating condition When a single crystal with magnetic anisotropy is placed in a magnetic field, it will rotate if the anisotropic magnetic energy is higher than the energy of the thermal motion. The condition for rotation is expressed as [158]
    graphic file with name TSTA11660773M26.jpg
    where Δ χ = χ a, b- χ c, is the anisotropy of the magnetic susceptibilities, V is the volume of the material, B is the applied magnetic field, μ 0 is the permeability in a vacuum, kB is the Boltzmann constant and T is the absolute temperature.
  • Static magnetic field alignment model If a crystal is dispersed in a viscous liquid of viscosity η, the crystal rotation in a static magnetic field (SMF) can be described as [158]
    graphic file with name TSTA11660773M27.jpg
    where ξ is the angle between the magnetic field direction and the crystal axis of the largest magnetic susceptibility. The subscripts ‘i’ and ‘f’, respectively, denote the initial and final states during alignment, t is the time and τ −1 is defined as the alignment rate
    graphic file with name TSTA11660773M28.jpg
    where L is a hydrodynamic term related to the particle shape. For a spherical particle of radius d, equations (26) and (28) can be rewritten as
    graphic file with name TSTA11660773M29.jpg
    graphic file with name TSTA11660773M30.jpg
    For an ellipsoidal (or prolate) crystal with short-axis radius d and aspect ratio R, equations (26) and (28) can rewritten as
    graphic file with name TSTA11660773M31.jpg
    graphic file with name TSTA11660773M32.jpg
    graphic file with name TSTA11660773M33.jpg
    If Inline graphic , then F(R) = 1 and equation (30) is identical to equation (32); if Inline graphic , then F(R) = - 3/2 and equation (32) describes the case of a whisker (or fiber).
  • Rotating magnetic field alignment model In an SMF, the crystal is aligned with the largest magnetic susceptibility axis parallel to the magnetic field. If χ c> χ a, b, the crystal is aligned with the c-axis parallel to the magnetic field, namely, the uniaxial or unidirectional alignment; If χ c< χ a, b, the crystal is aligned with the a, or b-axis parallel to the magnetic field, namely, the planar alignment. In the case of χ c< χ a, b, a rotating magnetic field (RMF) can be used to achieve the alignment of the c-axis. As schematically shown in figure 17, the crystal alignment model in an RMF (B) can be described as follows [159]
    graphic file with name TSTA11660773M34.jpg
    graphic file with name TSTA11660773M35.jpg
    where ω is the angular velocity of the rotation of the magnetic field on the XY plane. The alignment rate τ −1 is given by equations (30) and (32) for a spherical and ellipsoidal crystal, respectively.
Figure 17.

Figure 17

Polar coordinate system describing the whisker orientation in an RMF on the XY plane with an angular velocity of ω. ξ is the angle between the magnetic field direction and the whisker long axis.

Equations (29)–(32), (34) and (35) indicate the conditions necessary for the magnetic alignment of crystals during slurry consolidation, including the following.

  • The particle should be anisotropically magnetic, normally, with a noncubic crystal structure (e.g. hexagonal, tetragonal).

  • The size of the particle is sufficiently large to overcome the energy of thermal motion.

  • The suspension should be well deagglomerated or dispersed and exhibit low viscosity, because only single crystals can be aligned in the magnetic field.

  • The rotating velocity of the magnetic field should be sufficiently high to align along the c-axis. Note that the RMF is achieved by rotating the sample in the SMF, where the magnetic field is parallel to the rotating plane [159]. In this case, the rotating velocity corresponds to the rotating velocity of the sample.

Figure 18 shows the dependence of the minimum particle size necessary for the rotation of a spherical particle on the magnetic flux density, according to equation (29). This figure implies that a magnetic field of 10 T allows the minimum size of a crystal with |Δ χ | = 10− 6 (a typical feeble magnetic material) to be reduced to 30 nm, i.e. the nanoparticles of this material may be aligned by a strong magnetic field of 10 T. The alignment of spherical TiO2 [160] and ZnO [161] nanoparticles as well as TiO2 [162] whiskers has been demonstrated by slip casting in a magnetic field of 10 T. In these works, it was demonstrated that the magnetic alignment is independent of the particle shape. This is the most attractive feature of SMFA in comparison with the other three alignment techniques.

Figure 18.

Figure 18

Dependence of minimum particle size necessary for the rotation of a spherical particle on the magnetic flux density (T), calculated by equation (5).

b. SMFA of Inline graphic

Because of their hexagonal structure, a strong magnetic field of Inline graphic leads to the alignment of both α- and β-Si3N4 crystals during the slip casting of α-Si3N4 powders. However, the SMF leads only to the alignment of both α- and β-Si3N4 crystals with the a or b-axis parallel to the magnetic field, or the planar orientation because of the magnetic susceptibility of χ a,b χ c [163], as schematically illustrated in figure 19(a). Hence, the a, b-axis oriented β-Si3N4 is preferentially developed by sintering, during which the α-phase completely transforms into the β-phase [103, 163]. To align the c-axis of β-Si3N4 crystals, an RMF has been used for the slip casting of α-Si3N4 powders that contain a small fraction of β-Si3N4 crystals [82, 164, 165], as shown in figure 19(b). In the RMF, slip casting is perpendicular to the magnetic field direction, and the rotating speed of the sample should be optimized to achieve the highest orientation. It was found that a rotating speed of 10 rpm is sufficient to achieve greater orientation of the axis of β-Si3N4 single crystals [82].

Figure 19.

Figure 19

Schematic drawings of orientation mechanisms of β-Si3N4 grains during slip casting in an (a) SMF and (b) RMF.

The magnetic field alignment requires a suspension with well-agglomerated β-Si3N4 particles. Slurry consolidation also requires a well-stabilized suspension for developing a high-quality green body, better sinterability and high-quality sintered products. In SMFA, the dispersion and stability of the Si3N4 suspension (aqueous or nonaqueous) are crucial for controlling the texture development and properties of β-Si3N4 ceramics. Because of the need for safety, aqueous processing is preferable for producing Si3N4 ceramics. The most common polyelectrolytes used for stabilizing aqueous Si3N4 suspensions are acrylic-based polymers, such as poly(acrylic acid) (PAA) or its ammonium salt (PAA-NH4), known as anionic polyelectrolytes [166168]. Recent studies have shown that polyethylenimine (PEI) is an effective cationic dispersant for aqueous Si3N4 suspension, because it can provide better stabilization of the suspension by maintaining the suspension pH and producing a steric or electrosteric effect [169171].

For the SMFA of α-Si3N4 powders, previous studies have revealed one important fact: the β-Si3N4-phase crystal exhibits stronger orientation ability than the α-Si3N4 phase crystal despite the minor phase, for example, ∼1 and ∼5 wt% for SN-E5 and SN-E10 powders, respectively [163]. One reason for this is that, compared with α-Si3N4 particles with an equiaxed shape, β-Si3N4 particles are less agglomerated and have a predominantly rodlike shape. The other reason is probably because the β-Si3N4 crystal exhibits a larger value of the anisotropic magnetic susceptibility than the α-Si3N4 crystal according to equation (26). Furthermore, as the Inline graphic phase transformation occurs via the solution-reprecipitation process, the solution probably disturbs the crystallographic orientation of the α-phase. However, due to the greater stability, the oriented β-Si3N4 grains act as seeds to promote the development of oriented large elongated β-Si3N4 grains via anisotropic grain growth after and during the phase transformation, namely, the TGG mechanism. Although highly a, b-axis oriented β-Si3N4 can be produced via SMFA by the slip casting of α-Si3N4 powder (SN-E10) without additional β-Si3N4 seed particles, the use of β-Si3N4 seeds with tailored morphology is more efficient for not only increasing the degree of orientation, particularly the c-axis orientation, but also improving the bimodal microstructure.

The authors studied the effect of Si3N4 powder type (α or β-Si3N4) on texturing β-Sialon by an SMF and reaction PLS [172] and clearly demonstrated that β-Si3N4 crystals exhibit substantially stronger orientation than α-Si3N4 crystals. The β-raw powder produced a highly a, b-axis oriented β-Si3N4 green body with a Lotgering orientation factor of up to 0.97. During sintering, the β-raw powder allowed the a, b-axis oriented β-Sialon to retain the Lotgering orientation factor similar to and even higher than that of β-Si3N4 in the green body. In contrast, the α-raw powder led to a faster transformation rate of α /β-Si3N4 to β-Sialon but a substantially lower texture in β-Sialon. Using the slip casting of Si powder containing β-Si3N4 particles in an SMF, the authors also obtained a, b-axis aligned sintered reaction-bonded silicon nitride (SRBSN) [100]. Because of the cheap Si raw powder, the 60% weight gains by silicon nitridation, the machinability of RBSN body and the smaller sintering shrinkage, SRBSN is a well-known cost-effective Si3N4 ceramic material, thus, this work suggested that the combination of SMFA and SRBSN is a promising method for fabricating cost-effective textured Si3N4 as well as Sialon.

Anisotropic sintering shrinkage

Shrinkage anisotropy is an important phenomenon that occurs during the sintering of Si3N4 compact-containing oriented β-Si3N4 seed particles. The anisotropy is primarily associated with the texture development via the preferential grain growth of elongated β-Si3N4 seeds. The shrinkage anisotropy behavior in 3D space depends on the processing route and texture structure, as schematically illustrated in figure 20. The shrinkage is defined as

graphic file with name TSTA11660773M36.jpg

where L0 and L are the sample dimensions before and after sintering, respectively. Furthermore, the shrinkage anisotropy factor, fSA, is defined as

graphic file with name TSTA11660773M37.jpg

where Inline graphic and Inline graphic are the shrinkages in the directions parallel and perpendicular to the c-axis of elongated β-Si3N4 grains, corresponding to the Inline graphic and Inline graphic or Inline graphic , Inline graphic and Inline graphic or Inline graphic , Inline graphic or Inline graphic and Inline graphic , and Inline graphic and Inline graphic or Inline graphic in figures 20(a)–(d), respectively. The largest and smallest shrinkages occur in the directions perpendicular and parallel to the c-axis of elongated β-Si3N4, respectively. Muscat et al [142] reported an fSA value of 75% for extruded Si3N4 with β-Si3N4 whiskers. Teshima et al [50] reported an fSA value of 67% for extruded Si3N4 with rodlike β-Si3N4 seeds. Using a modified TCA technique, Bae et al [108] reported an fSA value of 70% for textured Si3N4 with fine β-Si3N4 whiskers, which is higher than that (∼52%) for tape-cast Si3N4 with coarser β-Si3N4 whiskers.

Figure 20.

Figure 20

Schematic illustrations of sintering shrinkage behavior of textured Si3N4 by (a) EA, (b) TCA, (c) slip casting in an SMF and (d) slip casting in an RMF. The left figure is the schematic texture and the right figure is the shrinkage curve as a function of time (t). t0 is the time when the shrinkage begins.

Generally, the fSA value increases with the increased degree of orientation by increasing the degree of orientation of seed particles and the seed content. Figure 21 illustrates the increased shrinkage anisotropy with increased β-Si3N4 whisker content obtained by modified TCA. On the basis of sintering shrinkage anisotropy, Park and Kim [173] proposed a model to quantify the degree of orientation of the large elongated grains in c-axis aligned Si3N4 developed from β-Si3N4 whisker templates. Figure 22 gives the shrinkage anisotropy in textured Si3N4 prepared by SMFA using slip casting [82, 163]. It is clear that texture development results in substantially higher shrinkage anisotropy in comparison with an untextured sample. Nevertheless, the addition of β-Si3N4 seeds does not seem to contribute to the shrinkage anisotropy, which is different from the report of Part and Kim [173]. This is most likely to be due to the morphology of the β-Si3N4 seeds used. If the β-Si3N4 seeds exhibit a whiskerlike shape and a larger size than the matrix grains, the shrinkage anisotropy should be enhanced [174]. Despite this, the shrinkage anisotropy can be seen as a direct indicator of texture development in Si3N4 ceramics.

Figure 21.

Figure 21

Effect of β-Si3N4 whisker content on the shrinkage anisotropy of Si3N4 compacts prepared by modified TCA during GPS. The authors only present the data of Inline graphic and Inline graphic in their work. If Inline graphic were given, the fSA values would be larger.

Figure 22.

Figure 22

Shrinkage anisotropy in textured Si3N4 prepared by SMFA using slip casting. Sample TE10S contains 5 wt% equiaxed β-Si3N4 seeds relative to sample TE10, and E10 means the raw UBE SN-E10 powder. Sample UTE10 was prepared in the absence of a magnetic field. For the RMF of 12 T, the rotating speed of the sample was fixed at 10 rpm.

Factors affecting texture development

a. β-Si3N4 seed morphology and content

Texture development via TGG is determined by the alignment of seed particles, which is controlled by the β-Si3N4 seed morphology and content. The dependence of texture development on β-Si3N4 seed morphology is related to the seed alignment technique. For EA and TCA, the β-Si3N4 seeds must be externally added because of the requirements of a rodlike morphology and a larger size than the matrix grains. In both EA and TCA, the texture development can be enhanced by increasing the aspect ratio and the β-Si3N4 seed content. Imamura et al [56] reported that texture development in β-Si3N4 is enhanced by increasing the aspect ratio and decreasing the diameter. The larger aspect ratio results in a higher degree of orientation, as the Wu and Messing model predicts [151]. When the aspect ratio remains constant, the seed density in the matrix decreases as the diameter increases, i.e., the number of oriented seed particles decreases as the diameter increases, thereby increasing the degree of orientation of large elongated grains during sintering. Bae et al [108] reported that for tape casting, fine whiskers with a diameter of 1.1 μm and an aspect ratio of 11.1 result in a higher degree of texture in sintered Si3N4 than coarse whiskers with a diameter of 2.8 μm and an aspect ratio of 6.7. The enhanced texture development is due to a combination of the larger aspect ratio and the larger seed density in the finer whiskers. Furthermore, the degree of texture increases with increased seed density regardless of the seed size. Also, for a given β-Si3N4 seed, the degree of texture increases as the seed content increases [75, 175].

Although SMFA is independent of the rodlike morphology, it strongly depends on the β-Si3N4 seed content. Again, β-Si3N4 seeds with size larger than the α-Si3N4 matrix should be beneficial for enhancing the texture via TGG during sintering. According to the Dressler model [94], if the size of β -nuclei is below the critical size, they will dissolve and reprecipitate on the surrounding large β-Si3N4 grains. This means that, to guarantee the stability of oriented β-nuclei during sintering, the β-nuclei should be sufficiently large. Nevertheless, because of the high solubility, α-grains should dissolve more rapidly than β-grains with the same size. This suggests that the β-nuclei may have the same order of grain size as the matrix α-Si3N4 particles used for the SMFA technique. This has been clearly demonstrated in the literature [82, 103, 163165]. Although the texture can develop from the preexisting β-Si3N4 nuclei in the α-Si3N4 raw powder, the number of the preexisting β-Si3N4 nuclei is usually insufficient to achieve a high degree of texture in β-Si3N4, particularly for the c-axis texture generated by an RMF. Table 9 gives the effects of α-Si3N4 powder and β-Si3N4 seeds on the degree of orientation of β-Si3N4 grains in green bodies by slip casting using an SMF and RMF. It is evident that the addition of β-seeds markedly increases the degree of orientation, as indicated by the larger XRD intensity ratios on the plane perpendicular to the magnetic field for SMF and on the plane parallel to the magnetic field for RMF. Figure 23 illustrates that the β-Si3N4 nuclei control the texture development in Si3N4 via SMFA, and the addition of β-Si3N4 seeds is more efficient for increasing the degree of texture than prolonged sintering.

Table 9.

Degree of orientation of β-Si3N4 grains in green bodies by slip casting using an SMF and RMF [82,163].

XRD intensity ratio
SMF (I(200)/I(101)) RMF (I(002)/I(200))
Sample TS(⊥B) SS(∥B) TS(∥B) SS(⊥B)
TE10 2.90 0.75 ≈0.5 ≈0
TE10S 12.4 0.89 5.07 ≈0
TE5 1.67 1.43
TE5S 13.43 1.08

TS=top surface, perpendicular to the magnetic field for SMF but parallel to the magnetic field for RMF; SS=side surface, parallel to the magnetic field for SMF but perpendicular to the magnetic field for RMF, see figure 19.

Figure 23.

Figure 23

Lotgering orientation factor fL of textured Si3N4 samples TE10, TE10S, TE5 and TE5S prepared by (a) SMF (reproduced with permission from [163] ©2006. The Ceramic Society of Japan) and (b) RMF using slip casting. The sintering was carried out at 1800 °C in both cases. Samples TE10S and TE5S contain 5 wt% equiaxed β-Si3N4 seeds relative to samples TE10 and TE5, and E10 and E5 denote raw UBE SN-E10 and SN-E5 powders, which contain 4.5 and 1.3 wt% β-Si3N4 nuclei, respectively.

b. Sintering conditions

Sintering is responsible for most of the texture development in ceramics via TGG. During sintering, the oriented β-Si3N4 seeds grow both anisotropically and abnormally by consuming the surrounding small matrix grains by the α–β phase transformation and subsequent grain growth by Ostwald ripening, eventually resulting in the enhanced texture. For EA and TCA, the addition of large β-Si3N4 seeds tends to hinder the densification; thus, the GPS method at a higher temperature (Inline graphic ) is preferable, particularly for the case of low amounts of sintering additives, as shown in tables 6 and 7. The higher temperature and prolonged sintering favors the grain growth and thus the texture development. However, the degree of texture eventually reaches a limit as the sintering time increases, because the grain growth is hindered by the impingement of neighboring β-Si3N4 grains [45, 75, 163]. As shown in figure 23, the use of β-seeds is more efficient for developing highly textured Si3N4 via SMFA, particularly c-axis textured Si3N4 by an RMF, compared with prolonged sintering. As shown in figure 24, the addition of β-seeds increases the number of initially oriented β-Si3N4 nuclei and promotes the growth of large elongated grains, thereby increasing the degree of texture. The core-rim structure (indicated by white arrows) implies that the oriented large elongated grains grow epitaxially from the initially oriented β-Si3N4 particles during slip casting. Although some of the β-Si3N4 seeds do not have a rodlike morphology, the elongated grain growth enables them to grow into large elongated grains. In addition, Park et al [176] also observed some large elongated grains with more than one core grown using a Y2O3–Al2O3 additive composition, probably as a result of the coalescence of grains growing from separate β-Si3N4 seeds.

Figure 24.

Figure 24

SEM micrographs of polished and plasma-etched surfaces of samples (a) TE10 and (b) TE10S sintered at 1800 °C for 3 h using an RMF of 12 T. The surface is perpendicular to the rotating plane and the vertical direction in the page is the preferential c-axis orientation of elongated β-Si3N4 grains. White arrows indicate the core-rim structure grown from the seed particles initially aligned by epitaxial grain growth.

c. Seed alignment technique and processing parameters

Because the c-axis oriented β-Si3N4 can have the maximum anisotropy in its properties, the alignment of the c-axis of β-seeds is of great interest. Compared with TCA and SMFA, EA is more effective for achieving this goal. However, the high orientation of the c-axis can also be achieved by modified TCA (figure 16) through tailoring the slurry properties and casting parameters. Furthermore, the orientation of the c-axis can also be achieved by an RMF. Although no data is available concerning the effect of slurry properties on the orientation of the β-seed nuclei via SMFA, it should be emphasized that a well-dispersed slurry is a prerequisite for this alignment technique, as discussed by Sakka and Suzuki in their review article [154].

Therefore, the key points for developing highly textured β-Si3N4 are as follows:

  • High orientation of the β-Si3N4 seed crystals,

  • Optimum content of β-Si3N4 seed,

  • High-temperature sintering/annealing.

From the industrial point of view, TCA is a low-cost process for producing high-quality textured Si3N4 components. Nevertheless, SMFA is a promising process for producing textured Si3N4 components with complex shapes.

Anisotropic properties of textured Si3N4

Hardness

In addition to the porosity, grain boundary phase and grain size [177], the orientation of elongated β-Si3N4 grains also affects the hardness of Si3N4 ceramics [50, 63, 126, 141, 144], as clearly demonstrated by the experimental data listed in table 10. The plane parallel to the grain alignment (or c-axis) exhibits greater hardness than the plane perpendicular to the grain alignment. Thus, the texture allows the textured material to possess greater hardness in the plane parallel to the grain alignment than the untextured material prepared under the same conditions. The hardness anisotropy depends on the degree of orientation. Higher orientation of the c-axis of elongated grains results in higher hardness anisotropy. The anisotropy in textured Si3N4 is essentially due to the intrinsic anisotropy in the hardness of β-Si3N4 single crystals [178180], as shown in table 11. The intrinsic hardness anisotropy reaches ∼46% for the β-Si3N4 single crystal. The intrinsic hardness anisotropy of the single crystal is due to the difference in the number of bonds per unit area on different crystallographic planes. As Chakraborty and Mukerji [178] proposed, the intrinsic hardness anisotropy of a single crystal, fIHA, can be expressed in theory as

graphic file with name TSTA11660773M38.jpg

where a and c are the unit cell parameters. For β-Si3N4, a= 0.7608 and c= 0.2911 nm, and fIHA is determined to be 39% by equation (38). This value is in agreement with the experimental data of Vickers hardness measurements on β-Si3N4 single crystals either grown from Si melt or present in sintered Si3N4 ceramics. The dissolution of Al and O leads to a decrease in the hardness of the β-Si3N4 lattice by the formation of β-Sialon (Si6-zAlzOzN8-z) [181]. Tanaka et al [182] found that the decrease in hardness with the z value can be expressed as

graphic file with name TSTA11660773M39.jpg

Therefore, the c-axis alignment of β-Si3N4 grains combined with the use of a small amount of an Al-free sintering additive is expected to result in the formation of high-hardness β-Si3N4 ceramics in the plane perpendicular to the grain alignment.

Table 10.

Vickers hardness (HV) of untextured and textured Si3N4 as reported in the literature.

HV (GPa) fCHA Schematic of
Author Material c-axis c-axis (%)a hardness test
Goto et al [141] a, b-axis aligned 15.7 14 11 Inline graphic
Santos et al [126] Weakly a, b-axis aligned 16.3 15.9 3
Belmonte et al [144] c-axis aligned 15.2 14.3 6
Weakly a, b-axis aligned 15.7 15.1 4
Teshima et al [50] Highly c-axis aligned 15.8 13.4 15
Nakamura et al [63] Highly c-axis aligned 17 12.6 26
Randomly aligned 15.6

a Hardness anisotropy factor, fHA=(1−HV/HV)×100%.

Table 11.

Vickers hardness (HV) of β-Si3N4 single crystals as reported in the literature.

HV (GPa) fCHA Schematic of
Author Material Test technique c-axis c-axis (%)a hardness test
Chakraborty Grown from Microindentation 36.2 20 45 Inline graphic
 and Mukerji [178]  Si melt Load=100 g
Dusza et al [179] GPS Si3N4 Microindentation 20.8 13 38
Load=25 g
Hay et al [180] GPS Si3N4 Nanoindentation 35 19 46
Load=4 mN

a Hardness anisotropy factor, fCHA=(1−HV/HV)×100%.

Elastic modulus

The elastic modulus is an important physical parameter for understanding the mechanical properties and structural applications of Si3N4 ceramics. The Young's moduli of Si3N4 ceramics were reported to be in the range of 300–330 GPa, depending on porosity, grain boundary phases, texture and the relative contents of α and β-phases [180]. Lee and Bowman [110] first reported the elastic modulus anisotropy in hot-pressed Si3N4 ceramics, which showed Young's moduli of 332 and 315 GPa parallel and perpendicular to the HP direction, respectively. The authors supposed that the small anisotropy in the Young's modulus is due to the low anisotropy in the Young's modulus of the β-Si3N4 single crystal. This assumption has been confirmed by Hay et al [180] using nanoindentation measurements on individual large elongated hexagonal β-Si3N4 grains, as shown in figure 25. In their work, the measured elastic moduli are specifically defined as ‘indentation moduli’ rather than ‘Young's moduli’ because the data analysis procedures used to obtain indentation moduli from the nanoindentation load-displacement data are based on contact mechanics solutions for elastically isotropic media and therefore only formally apply to isotropic materials. For anisotropic materials, the indentation modulus is generally different from the Young's modulus, and its value depends on all the elastic constants of the materials in a complex way, as shown in figure 25. In the stiffest direction (α = 0, i.e. the c-axis of the crystal) the indentation modulus underestimates the Young's modulus of the single crystal by about 20%, and in the most compliant direction it overestimates Young's modulus by about 10%.

Figure 25.

Figure 25

Young's modulus calculated from the single-crystal elastic constants compared with the indentation modulus for a range of crystallographic directions (reproduced with permission from [180] ©1998 Blackwell Publishing Ltd). The solid curve of the indentation modulus was calculated using Vlassak–Nix analysis and the single-crystal elastic constants. The dotted lines represent possible deviations due to uncertainties in the Poisson ratios.

Figure 25 provides a measure of the magnitude of elastic anisotropy in β-Si3N4. The estimated Young's moduli for β-Si3N4 are 280 and 540 GPa in the directions perpendicular and parallel to the c-axis, respectively, indicating that the elastic anisotropy of an elongated hexagonal β-Si3N4 crystal is highly intrinsic. If the intrinsic elastic anisotropy factor for a β-Si3N4 crystal, fIEA, is defined as

graphic file with name TSTA11660773M40.jpg

where Ec and Ec are the Young's moduli in the directions perpendicular and parallel to the c-axis of the β-Si3N4 crystal, respectively. Thus, fIEA is determined to be 48%. The elastic anisotropy was also reported by Kondo et al [183] in textured porous Si3N4 ceramics prepared by partial SF, in which the Young's modulus is lower in the direction perpendicular to the c-axis of elongated grains than in the direction parallel to the c-axis. Therefore, the c-axis alignment of elongated β-Si3N4 grains should reduce the Young's modulus in the direction perpendicular to the grain alignment, thereby giving Si3N4 ceramics excellent strain tolerance [184]. In addition to the intrinsic elastic anisotropy, the β-Si3N4 grain morphology, such as the aspect ratio, should also affect the elastic anisotropy of textured Si3N4 ceramics [185]. To obtain a full understanding of the elastic anisotropy behavior, further studies are needed, including experimental and model simulation studies.

Strength and toughness

RT strength and toughness

It has been well documented that texturing leads to anisotropy in the strength and toughness of Si3N4 ceramics. This offers a unique opportunity to maximize the strength and toughness in the direction perpendicular to the grain alignment. Figure 26 shows the important progress achieved in improving the strength and toughness of untextured and textured Si3N4 ceramics. Compared with untextured Si3N4, textured Si3N4 has the capability of reaching a bending strength of up to 1.4 GPa and a fracture toughness of up to 14 MPa m1/2 in the direction perpendicular to the grain alignment. The grain orientation provides more opportunities for crack deflection and bridging in the crack propagation plane perpendicular to the grain alignment, thereby higher fracture toughness [15, 4143, 186]. On the one hand, the high bending strength is due to the more uniform 2-D distribution of larger elongated grains, thereby preventing the formation of large defects such as clusters. On the other hand, the high fracture toughness in the same direction also contributes to the high strength, as indicated by the relationship between bending strength (σb) and fracture toughness (KIC) [187]

graphic file with name TSTA11660773M41.jpg

where a is half the crack length. The higher toughness allows textured Si3N4 to have higher tolerance to flaws generated during machining, thereby improving the mechanical reliability. Hirao et al [43] developed c-axis aligned Si3N4 that exhibited a Weibull modulus of 46. Moreover, Ohji et al [188] observed a steep R-curve behavior for textured Si3N4, as shown in figure 27, which is believed to be advantageous for avoiding catastrophic fractures and for increasing the Weibull modulus by narrowing the strength distribution. Pezzotti et al [189] used Raman microprobe spectroscopy to reveal that crack-face bridging is the most effective mechanism for toughening Si3N4 ceramics. The macroscopic fracture behavior of untextured and textured Si3N4 can be modeled as a function of the microscopic maps of bridging tractions; the higher the bridging stress, the steeper the R-curve.

Figure 26.

Figure 26

Bending strength versus fracture toughness for untextured and highly c-axis aligned Si3N4.

Figure 27.

Figure 27

R-curve behavior of untextured (open symbols) and textured Si3N4 (TSN) ceramics during transverse crack propagation (closed symbols), determined by as-indented crack lengths (circles) and instability crack lengths (squares). The closed, triangular symbols indicate the textured Si3N4 during longitudinal crack propagation determined by as-indented crack lengths (reproduced with permission from [188] ©1995 Blackwell Publishing Ltd).

The bending strength anisotropy increases with increased degree of orientation. The bending strength anisotropy factor, fBSA, is defined as

graphic file with name TSTA11660773M42.jpg

where σb⊥ and σb∥ are the bending strengths in the directions perpendicular and parallel to the grain alignment, respectively. Figure 28 shows the bending strength anisotropy versus the bending strength σb⊥ of c-axis aligned Si3N4. The bending strength anisotropy increases with increased strength σ b⊥.

Figure 28.

Figure 28

Bending strength anisotropy versus bending strength (σb⊥) in the direction perpendicular to the grain alignment, corresponding to the direction of the tensile stress plane parallel to the grain alignment.

The Vickers indentation fracture (VIF) method has been widely used for determining the fracture toughness of untextured and textured Si3N4 ceramics. For textured Si3N4, the toughness anisotropy can be determined by the anisotropic crack propagation between the directions perpendicular and parallel to the grain alignment, as typically shown in figure 29. In the direction perpendicular to the grain alignment (⊥ c-axis), the cracks are shorter and toughness is higher because more energy is released per unit crack length, compared with the direction parallel to the grain alignment (∥ c-axis). This means that in the VIF method, one indentation point can produce one toughness anisotropy value. In contrast, the single-edge-precracked-beam (SEPB) and the single-edge-V-notched-beam (SEVNB) methods require two samples to produce one toughness anisotropy value, because only one precracked plane is operated on for each sample. The fracture toughness anisotropy factor, fFTA, is defined as

graphic file with name TSTA11660773M43.jpg

where KIC⊥ and KIC∥ are the fracture toughnesses perpendicular and parallel to the grain alignment, respectively. Table 12 lists the fracture toughness anisotropy in textured Si3N4 as reported in the literature. Generally, fFTA increases as the degree of orientation increases, which is governed by the processing method. The c-axis aligned texture results in a higher fFTA than the a, b-axis aligned texture. Because of the limited a, b-axis orientation, HP produces the lowest fFTA, whereas because of the high c-axis orientation, EA produces the highest fFTA; the TCA and HF methods produce intermediate values of fFTA. Therefore, EA is the most efficient method for developing textured Si3N4 with the highest anisotropic mechanical properties.

Figure 29.

Figure 29

Schematic illustration of typical anisotropic crack propagation measured by Vickers indentation method. For a, b-axis aligned Si3N4, this indentation pattern is obtained from the surface parallel to the a, b-axis of elongated β-Si3N4 grains, which is perpendicular to the HP or HF direction under SC. For c-axis aligned Si3N4, this indentation pattern is obtained from the surface parallel to the c-axis of elongated grains, which is parallel to the tape casting or extruding direction.

Table 12.

Fracture toughness anisotropy in textured Si3N4 as reported in the literature.

Processing KIC (MPa m1/2)c
Authors method Key factor DOTa Methodb fFTA
Lee and Bowman [111] HW HP VIF 5.72 4.63 19
 HF_SP 8.41 5.88 30
 HF_PSC 8.07 3.90 52
Xie et al [64] HF-PSC True strain:
 0.3 VIF 3.32 2.67 19
 0.5 3.90 2.67 31
 0.7 4.66 2.51 46
 0.95 5.22 2.54 51
Santos et al [126] HP Additive (vol%)
 5 VIF 6.6 5.2 21
 10 6.5 5.6 14
 15 6.4 5.8 9
 20 6.2 5.9 5
Imamura et al [56] TCA Seed type
w=0.44 μm, R=3.1 SEPB 10.4 7.1 32
w=1.29 μm, R=4 11.6 7.3 37
w=0.47 μm, R=4.2 12.4 7.2 42
Muscat et al [142] EA β-Si3N4 whisker seed VIF 13.5 4.6 66
Nakamura et al [63] EA β-Si3N4 rod seed VIF 9.7 3.9 60
Park et al [57] TCA β-Si3N4 whisker seed VIF 8.52 5.13 40

a DOT=degree of texture, ‘↓’ means increased DOT upon varying parameters from up to down and vice versa.

b VIF=Vickers indentation fracture, SEPB=single-edge-precracked-beam method.

c ⊥ and ∥ denote the directions perpendicular and parallel to the alignment of elongated grains, respectively.

HT strength

It is well-known that the HT strength of Si3N4 ceramics is mainly determined by the softening temperature of the grain boundary phase. Therefore, two common methods are used for improving HT strength: choosing a more refractory additive composition and crystallizing grain boundary amorphous phases by a postheating treatment. Table 13 lists the important progress achieved in improving the HT strength of Si3N4 ceramics. In addition to the two common methods, texturing the microstructure is also an efficient method for improving HT strength. The highly c-axis aligned texture allows Si3N4 with less refractory Y2O3–Al2O3 additives to obtain a strength retention as high as 63% at 1400 °C [68], and allows Si3N4 with more refractory Lu2O3–SiO2 additives to obtain a strength retention as high as ∼100% at 1500 °C [72], which is the highest reported value to date. Again, the c-axis aligned texture results in strength anisotropy at both RT and high temperatures. However, the strength retention is lower in the direction parallel to grain alignment than in the direction perpendicular to the grain alignment.

Table 13.

HT strength of untextured and textured Si3N4 as reported in the literature.

Bending strength σb (MPa)
HT (°C)
RT 1400 1500 Strength retention (%)a
Authors Additive Microstructure Atmosphere
Lu and Huang [30] b Y2O3−Al2O3 Randomly aligned c 853 375 N2 44
Yb2O3−Al2O3 926 530 N2 57
Park et al [68] d Y2O3−Al2O3 Randomly aligned 735 203 Air 28
Highly c-axis aligned 930 583 Air 63
Nishimura et al [24] b Yb2O3−SiO2 Weakly a, b-axis aligned 977 484 N2 50
Park et al [25] d Yb2O3 Weakly a, b-axis aligned 1092 870 N2 80
Kondo et al [128] d Yb2O3−SiO2 Highly a, b-axis aligned 1258 441 Air 35
Guo et al [31] b Lu2O3 Weakly a, b-axis aligned 982 660 N2 67
Kondo et al [129] d Lu2O3−SiO2 Highly a, b-axis aligned 940 696 Air 74
Zeng et al [72] d Lu2O3−SiO2 Highly c-axis aligned 739 556 738 463 Air 100 83

a Strength retention=(σb, HTb, RT)×100%.

b Four-point bending test.

c Post heating treatment was conducted at 1400 °C to crystallize the amorphous grain boundary phase. Other reports did not use post heating treatment.

d Three-point bending test.

Park et al [68] found that using an Y2O3–Al2O3 additive, texturing led to an increase in both RT and HT strengths with increased amount of additive. However, the untextured sample exhibited a large decrease in the HT strength with increased amount of additive, which was opposite the effect on the RT strength. The load-displacement curves at 1400 °C indicated that the untextured sample exhibited significant plastic deformation due to the softening of the grain boundary phase, whereas the textured sample exhibited linear elastic behavior, indicating that the fracture behavior is not governed by deformation of the glassy phase. This should be due to the unidirectional orientation of elongated β-Si3N4 grains, allowing the neighboring grains to be sufficiently close to replace the glassy boundary with low-angle boundaries. At HTs, the fracture of untextured samples is due to a combination of the failure along the grain boundary and the fracture of the grains, whereas the fracture of textured samples is dominated by the brittle fracture of large aligned grains. Therefore, an effective processing strategy is to combine the unidirectionally aligned texture with a more refractory additive composition (the Lu2O3–SiO2 system is the most suitable that has been verified to date).

Fracture energy

Fracture energy is the most significant factor determining the strength of Si3N4 ceramics. The relationship between fracture energy (γ) and strength (σ) is expressed by Griffith's fracture equation [187]

graphic file with name TSTA11660773M44.jpg

where E is the Young's modulus, C is the length of the crack that initiates the fracture and A is a dimensionless number that depends on the mode of stressing, specimen configuration and dimensions, and the type of crack under consideration. The fracture energy can be understood as being the rate at which energy is absorbed by the growth of the crack. Equation (44) indicates that a higher fracture energy can result in a material having a higher strength. Si3N4 ceramics are well-known for high strength in comparison with other structural ceramics such as Al2O3, AlN and SiC. The high strength is closely related to the higher fracture energy, because the interlocking microstructure of elongated β-Si3N4 grains increases the energy dissipation by the frictional sliding of one surface against another during crack propagation.

Lange [8] revealed that the fracture energy of Si3N4 ceramics could be increased not only by tailoring the grain morphology, but also by developing the texture or grain orientation. He first reported that the fracture energy is about 16% higher in the direction perpendicular to the c-axis of elongated β-Si3N4 than in the direction parallel to the c-axis in hot-pressed Si3N4. Table 14 lists the fracture energy of untextured and textured Si3N4 ceramics as reported in the literature. At RT, the texture design is more efficient for increasing the fracture energy than the additive composition. The unidirectional aligned texture results in a higher fracture energy than the a, b-axis aligned texture in the direction perpendicular to the c-axis of the elongated grains at both RT and HTs, corresponding to the higher anisotropy. The testing atmosphere has no significant effect on the fracture energy at HTs. However, the anisotropy of fracture energy becomes small at HTs. The fracture energy anisotropy factor, fFEA, is defined as

graphic file with name TSTA11660773M45.jpg

where γ and γ are the fracture energies in perpendicular and parallel to the c-axis of elongated β-Si3N4 grain, respectively. Thus, fFEA decreases significantly from 61% at RT to 7% at 1500 °C, according to the data of Zeng et al [70]. The large increase in fracture energy by grain orientation is attributed to the effects of grain bridging and grain pullout. The fracture energy anisotropy is responsible for the anisotropy of the RT strength of textured Si3N4. At HTs, the higher fracture energy is due to the softening of the secondary glassy phases and the melting of secondary crystalline phases, resulting in enhanced grain pullout.

Table 14.

Fracture energy of untextured and textured Si3N4 as reported in the literature.

Fracture energy (J m−2)
HT (°C)
RT 1300 1500
Authors Additive Microstructure Atmosphere
Lange [8] a MgO Weakly a, b-axis aligned 82 69
Kondo et al [61] b Y2O3−Al2O3 Randomly aligned 48 75 Air
Weakly c-axis aligned 177 620 Air
720 N2
Kondo et al [128] c Yb2O3−SiO2 Weakly a, b-axis aligned 80 522 Air
570 N2
Highly a, b-axis aligned 187 734 Air
753 N2
Zeng et al [70] c Lu2O3−SiO2 Randomly aligned 109 454 Air
c-axis aligned 301 118 781 727 Air

a Double-cantilever (DC) method with a crosshead speed of 0.5 mm min−1.

b Chevron-notched-beam (CNB) method with a displacement rate of 0.005 mm min−1.

c CNB method with a displacement of 0.01 mm min−1.

Creep behavior

The creep behavior of Si3N4 ceramics has been extensively studied, and a review article of the processes involved in this area has been published recently [190]. Generally, the creep deformation mechanisms of Si3N4 ceramics are attributed to viscous flow, solution precipitation, cavitation and shear thickening, depending on the various factors involved in determining the characteristics of Si3N4 ceramics and the conditions of creep tests [190]. However, studies concerning the role of texture in creep deformation behavior are very limited. Yoon et al [48] studied the effect of the grain orientation on the compressive creep deformation of hot-pressed Si3N4 in air at temperatures of 1300–1400 °C and pressures of 30–100 MPa in air. They observed that the creep rate in the loading direction parallel to the HP was higher than that in the loading direction perpendicular to the HP. The stress exponent of the creep rate was determined to be about 1 and the apparent activation energy in both parallel and perpendicular directions at the temperature range of 1300–1400 °C is about 500 kJ mol−1. This suggests that the creep deformation is mainly determined by diffusion-controlled solution precipitation. Combined with TEM observation, the authors attributed the observed anisotropic creep behavior to the difference in the rate of the solution-precipitation step. Compared with the loading direction perpendicular to the HP, the direction parallel to the HP exhibits a higher solution-precipitation rate because of the unstrained basal (002) plane. Santos et al [71] also reported similar anisotropic creep behavior in hot-pressed Si3N4 between the directions parallel and perpendicular to the HP.

Kondo et al [60] studied the tensile creep behavior of untextured and textured Si3N4 ceramics prepared by GPS and SF, respectively, at temperatures of 1200–1250 °C in air. They found that the creep rate along the grain alignment direction in the textured sample (c-axis) was lower about one order of magnitude than that of the untextured sample. The stress exponents of the untextured and textured samples at 1200 °C were 2.1 and 2.6, respectively, and the value for the textured sample at 1250 °C was 3.6. This reveals that the grain alignment can improve the tensile creep resistance of Si3N4 ceramics in the grain alignment direction. The authors proposed several possible mechanisms for achieving significantly improved tensile creep resistance by the c-axis alignment of elongated grains: (i) increasing the sliding resistance to the grain boundaries, (ii) suppressing the formation of cavities and (iii) inhibiting viscous flow by providing a large number of film-free boundaries between well-aligned elongated grains. Hirao [75] also reported that c-axis aligned Si3N4 exhibits higher flexural creep resistance in the direction perpendicular to the grain alignment in comparison with untextured samples. Furthermore, the creep resistance increases with increasing amount of aligned elongated grains, i.e., degree of orientation, as shown in figure 30. The textured material can possess flexural creep resistance that is 1–2 magnitudes higher than that of the untextured material. The results suggest that texture design is an effective method for improving the creep resistance of Si3N4 ceramics. To produce highly creep-resistant Si3N4, an effective processing strategy should involve a combination of texture design, effective additive composition and optimized processing parameters.

Figure 30.

Figure 30

Flexural creep behavior at 1200 °C in air for untextured and textured Si3N4 ceramics prepared by tape casting without and with rodlike β-Si3N4 seed crystals. The bending tests were conducted with the tensile stress parallel to the grain alignment (reproduced with permission from [75] ©2006 The Ceramic Society of Japan).

Tribological and wear resistance

Miller and Bowman [46] first observed through sliding-ball tests that textured Si3N4 exhibited less damage in the plane perpendicular to the grain alignment than in the direction parallel to the grain alignment and that cracks tend to propagate in the sliding direction. Liang et al [54, 55] investigated the effects of the degree of orientation and sintering temperature on the friction and wear behavior of textured Si3N4 ceramics using ball-on-disk reciprocating sliding tests with Si3N4 balls under dry friction. They showed that a highly c-axis aligned sample exhibited much lower wear resistance in the direction parallel to the grain alignment than in the direction perpendicular to the grain alignment, and the higher degree of orientation resulted in lower wear resistance parallel to the sliding direction but higher wear resistance perpendicular to the sliding direction. An increased sintering temperature tends to degrade the wear resistance both parallel and perpendicular to the grain alignment. Furthermore, they also showed that the cross-alignment design of elongated grains allowed the wear resistance of this material to be between those of the highly c-axis aligned material in the directions parallel and perpendicular to the grain alignment. The effect of grain orientation on the friction coefficient depends on the load, but the dependence is not sufficiently strong to affect the wear resistance.

Nakamura et al [63, 66] conducted comparative studies of tribological and wear behavior between textured and untextured Si3N4 under dry friction using block-on-ring tests. They observed that for both Si3N4 and stainless steel rings, the unidirectionally aligned Si3N4 exhibited higher wear resistance in the plane perpendicular to the elongated grain alignment than in the direction both parallel and perpendicular to the grain alignment, and also higher wear resistance than the untextured sample for all sliding conditions. In the case of the Si3N4 ring, the wear resistance is greater perpendicular to the grain alignment than parallel to the grain alignment [63], whereas in the case of the stainless-steel ring, the wear resistance depends on the sliding conditions [66]. Table 15 lists the data of the friction coefficient and wear rate of unidirectionally aligned and randomly aligned Si3N4 ceramics for a Si3N4 ring, as reported by the authors.

Table 15.

Tribological and wear properties of unidirectionally aligned Si3N4 (UASN) and randomly aligned Si3N4 (RASN) as reported by Nakamura et al [63].

Test plane or Sliding Friction Wear rate Schematic of test
Material directiona speed (m s−1) coefficient (μ) (×10−8 mm2/N) (plane or direction)
UASN S-para 0.15 0.7–0.8 1.3 Inline graphic
1.5 –0.7 11.5
S-perp 0.15 0.7–0.8 0.8
1.5 ∼0.6 7
N-plane 0.15 0.7–0.8 0.6
1.5 –0.3 0.3
RASN 0.15 0.7–0.8 1.5
1.5 –0.9 3.3

a The wear test was conducted using a block-on-ring tester under a dry condition with a normal load of 5 N, and a commercial Si3N4 ceramic was used for the ring test piece.

The wear behavior of ceramics is dominated by the microfracture mechanism [191]. The sliding process generates compressive and tensile stresses near the contact surface, resulting in the debonding of the elongated β-Si3N4 grains from the small matrix by the initiation and propagation of microcracks. Thus, different debonding strengths lead to different wear behaviors. For highly c-axis aligned Si3N4, during the sliding process a lower number of elongated grains are pulled out in the direction perpendicular to the grain alignment than in the direction parallel to the grain alignment because of the higher crack propagation resistance, thereby resulting in a lower wear rate. Comparing the directions parallel and perpendicular to the grain alignment, the plane perpendicular to the grain alignment exhibits the lowest probability of grain pullout; therefore, it has the lowest wear rate. The anisotropic friction and wear mechanisms in textured Si3N4 are schematically illustrated in figure 31. The worn surface in the sliding direction parallel to the grain alignment (figure 31(a)) is typically covered by a thick layer of wear debris, and the worn surface in the sliding direction perpendicular to the grain alignment (figure 31(b)) is covered by a small amount of wear debris, whereas the worn surface in the plane perpendicular to the grain alignment (figure 31(c)) is reasonably smooth.

Figure 31.

Figure 31

Schematic illustrations of mechanisms for anisotropic friction and wear behavior in unidirectionally aligned Si3N4.

Erosion resistance

Lim et al [67] studied the erosion behavior of unidirectionally aligned Si3N4 by a gas-blast-type erosion tester using SiC grit at impact angles from 45 to 90°, as schematically shown in figure 32. They found that the direction parallel to the grain alignment exhibits higher erosion resistance (or lower erosion rate) than the direction perpendicular to the grain alignment for all tested impact angles. They attributed this erosion anisotropy to the higher probability of grain pull-out because of the ease of lateral crack formation in the direction perpendicular to the elongated grain alignment, thereby resulting in the greater amount of material removal.

Figure 32.

Figure 32

Schematic illustrations of impact direction of erosive SiC particles with respect to the alignment of elongated grains: (a) impact perpendicular to the grain alignment and (b) impact parallel to the grain alignment. α is the impact angle.

Zhang et al [69] conducted a comparative study of the erosion behavior of textured and untextured Si3N4 by the same testing technique using SiC grit with two impact angles of 30 and 90°. They found that the textured Si3N4 exhibits lower erosion resistance than the untextured Si3N4 at both impact angles, as listed in table 16. Generally, the erosion rate of ceramics should have a strong inverse dependence on the fracture toughness but a much weaker dependence on the hardness. According to theoretical predictions, highly unidirectionally aligned Si3N4 should have exhibited better erosion resistance in the direction perpendicular to the grain alignment than the randomly aligned Si3N4 because of the higher fracture resistance. The authors attributed this to the microstructure effect. The unidirectionally aligned microstructure promotes grain pullout, whereas the randomly aligned microstructure provides an interlocking effect that hinders it. In addition, because of the higher incidence of grain pullout, the direction parallel to the grain alignment exhibited a higher erosion rate than the direction perpendicular to the grain alignment, which is contrary to the findings of Lim et al [67]. This contradiction concerning the effect of grain orientation on erosion resistance may be due to the testing conditions. To obtain a full understanding of the effect of texture on the erosion behavior, more erosion conditions should be considered, such as the characteristics of erosive particles, the velocity and dosage of erosive particles and the distance between the nozzle and sample.

Table 16.

Erosion data of UASN and RASN as reported by Zhang et al [69].

Impact Erosion rate
Material angle (°) Test direction (m3 kg−1)×10−3
UASN 30 c-axis 26
c-axis 24
90 66
RASN 30 13
90 37

Hertzian contact damage

The Hertzian indentation technique has been used to characterize the contact damage of Si3N4 ceramics, which is crucial for some structure applications, e.g., frictional or erosive environments where the surfaces undergo impact due to a load. Lee et al [192, 193] studied the effect of the microstructure on the Hertzian contact damage of Si3N4 ceramics. They showed that there is competition between brittle (ring or cone fracture) and quasi-plastic damage (distributed shear-activated microfaulting) modes. The microstructure with relatively equiaxed grains exhibits damage in the form of classical cone cracks, whereas the microstructure with large elongated grains exhibits damage characterized by grain-localized microfailures within a subsurface, indicating higher contact damage resistance. Furthermore, in several studies it has been revealed that the grain orientation has a strong effect on the contact damage behavior of Si3N4 ceramics [46, 144]. Figure 33 shows typical Hertzian contact damage patterns of weakly a, b-axis and highly c-axis aligned Si3N4 ceramics, which show the random and unidirectional alignment of the c-axis in testing plane A, respectively [144]. Compared with the a, b-axis aligned sample, the c-axis aligned sample exhibits fewer and shallower surface ring cracks, which have an elliptical shape. This elliptical shape is due to the deflection of crack trajectories perpendicular to the interface caused by unidirectionally aligned elongated grains quasi-parallel to the bonding interface (plane B). It is clear that beneath the contact area, the unidirectional alignment of elongated grains causes the contact damage to change from cone cracks to a combination of cone cracks and a subsurface damage zone. This subsurface damage zone increases in size with the indentation load. Moreover, in the subsurface zone, multiple twins/slips are observed at the large Si3N4 seeds, together with debonding at seed interfaces. The twins are found to be oriented at about 45° to the contact axis, which is the direction of maximum shear stress [46]. The results indicate that the grain alignment affects the contact damage behavior in two ways: (i) conferring an elliptical shape to the radial surface cracks and (ii) promoting the multiple twin/slip formation at the large elongated grains within the high-shear-strain region.

Figure 33.

Figure 33

Optical micrographs of half-surface (top) and section (bottom) views of Hertzian contact damage at load P=2000 N for (a) weakly a, b-axis aligned Si3N4 and (b) highly c-axis aligned Si3N4, prepared by only HP without seeds and HP of extruded compacts with seeds. The right inset shows the subsurface damage of the highly c-axis aligned sample under Nomarski illumination. (c) schematic illustration of the Hertzian contact damage test and corresponding planes for the highly c-axis aligned sample. Plane A is parallel to the extruding direction and perpendicular to the HP direction, and Plane B is parallel to both the extruding and HP directions (reproduced with permission from [144] ©2007 Blackwell Publishing Ltd).

Thermal conductivity

In 1995, Haggerty and Lightfoot [194] first pointed out that Si3N4 has an intrinsic thermal conductivity of over 200 W m−1 K−1 at RT. Si3N4 has since been listed as a new high-thermal-conductivity ceramic material, similar to AlN and SiC. A large amount of experimental work has resulted in the establishment of a state-of-the-art method of producing β-Si3N4 ceramics with thermal conductivities of over 100 W m−1 K−1, including the utilization of effective Al-free sintering additives, HT (and/or long-time) sintering/annealing and grain orientation techniques. The improvement in thermal conductivity is primarily attributed to the purification of β-Si3N4 grains, such as by removing the dissolved oxygen in the β-Si3N4 lattice, because lattice defects cause the scattering of phonons that dominate heat conduction. There are several review articles dealing with the progress achieved in this field [195197]. Because of the combination of high thermal conductivity, superior mechanical properties and an excellent insulating property, β-Si3N4 has been regarded as a promising substrate material for semiconductor devices. Enhanced thermal conductivity is also helpful for expanding its application to structural materials such as engine components and heat exchangers, where high thermal-shock resistance is required to improve the long-term reliability.

Textured β-Si3N4 ceramics exhibit an interesting thermal anisotropy: the thermal conductivity is higher in the direction of the elongated grain alignment than in the direction perpendicular to the grain alignment [39, 45, 47, 52, 53, 58, 59, 62, 65, 72]. Thus, texturing has become another important method for improving the thermal conductivity of β-Si3N4 ceramics. Below we will discuss the thermal anisotropy of textured β-Si3N4 both theoretically and experimentally.

Theoretical analysis

To analyze the thermal anisotropy of β-Si3N4 ceramics, it is necessary to understand the thermal anisotropy of a single β-Si3N4 crystal. Several studies have been carried out to predict the intrinsic thermal conductivity of a single β-Si3N4 crystal [194, 196, 198, 199], as summarized in table 17. However, the thermal anisotropy was only taken into account in the work of Hirosaki et al [198]. Using the molecular dynamics method, they predicted that a single β-Si3N4 crystal has intrinsic thermal conductivities of 170 and 450 W m−1 K−1 along the a- and c-axis directions, respectively. To describe the thermal anisotropy of a single β-Si3N4 crystal, the factor, fITA, is defined as

graphic file with name TSTA11660773M46.jpg

where κ c and κ a are the thermal conductivities along the c- and a-axis directions, respectively. Thus, for an ideal single crystal, fITA is 62%. Moreover, such strong thermal anisotropy has also been demonstrated experimentally in actual rodlike β-Si3N4 crystals to be independent of the diameter by Li et al [200], as shown in table 17. The fact that the theoretical value agrees well with the experimental value suggests that the thermal anisotropy of β-Si3N4 crystals is highly intrinsic.

Table 17.

Intrinsic and experimental thermal conductivity (κ) of a single β-Si3N4 crystal at RT.

Intrinsic κ(Wm−1 K−1)
Year Authors Method a-axis c-axis fITA (%)
1995 Haggerty and Lightfoot [194] Slack equation 200(n=14) or 300(n=7)
2001 Watari et al [196] Slack equation 320
2003 Hirosaki et al [198] Molecular dynamics 170 450 62
2005 Bruls et al [199] Temperature-dependent thermal 105
 diffusivity method
Experimental κ(Wm−1 K−1)
a-axis c-axis fITA (%)
1999 Li et al [200] Thermoreflectance microscopy 69 180 62
 measurements on large rodlike
 β-Si3N4 grains in a ceramic

Kitayama et al [53] developed a modified Wiener model to describe the thermal conductivity of β-Si3N4 ceramics that contain oriented rodlike grains surrounded by a continuous grain boundary phase. They simply treated the textured microstructure as an idealized 2D microstructure composed of perfectly oriented β-Si3N4 grains along the c-axis direction. Using this new model, they predicted thermal conductivities as a function of grain boundary film thickness and grain size, which were in good agreement with experimental results. At the same time, they also showed that the thermal anisotropy is affected by the grain boundary film thickness and grain size. However, in their work, the intrinsic thermal anisotropy of β-Si3N4 grains was not considered. In this article, we use the Kitayama model to analyze the thermal anisotropy of textured β-Si3N4, as expressed by the following equations

graphic file with name TSTA11660773M47.jpg
graphic file with name TSTA11660773M48.jpg

where κ and κ are defined as the bulk thermal conductivities parallel and perpendicular to the c-axis (long-axis) of β-Si3N4 grains, respectively; κ c and κ g are the thermal conductivities of the β-Si3N4 crystal and the grain boundary glassy phase, respectively; w and l are the width and length of elongated β-Si3N4 grains, respectively; δ is the grain boundary film thickness; and fp is the volume fraction of the glass pockets, given by

graphic file with name TSTA11660773M49.jpg

where x is the volume fraction of the glassy phase. To describe the thermal anisotropy of textured β-Si3N4 ceramics, the factor, fTA, is defined as

graphic file with name TSTA11660773M50.jpg

A larger value of fTA implies stronger thermal anisotropy.

First, to study the effect of only the aspect ratio of β-Si3N4 grains on the thermal anisotropy of c-axis oriented β-Si3N4 ceramics, only the isotropic thermal conductivity is considered for the β-Si3N4 single crystal. The values used in the calculations are w= 1 μm, κ c= 180 W m−1 K−1 [45, 201], κ g= 1 W m−1 K−1 [202] and δ = 1 nm. Figure 34(a) shows that the elongated shape leads to thermal anisotropy. The thermal anisotropy initially increases rapidly with increasing aspect ratio, then remains almost constant above a critical value (≈4). Obviously, the isotropic grain shape (R= 1) does not lead to thermal anisotropy. In addition, the thermal anisotropy decreases slightly with increasing volume fraction of the glassy phase.

Figure 34.

Figure 34

Kitayama model prediction for the effects of (a) aspect ratio (R) of β-Si3N4 grains and (b) intrinsic thermal anisotropy of β-Si3N4 grains on the thermal anisotropy of ideal c-axis aligned β-Si3N4 ceramics. In (a), only the isotropic thermal conductivity is considered for the β-Si3N4 crystal, where κc=180 W m−1 K−1 and κg=1 W m−1 K−1. In (b), the intrinsic thermal anisotropy (fITA) is considered for the β-Si3N4 crystal, where w=1 μm, κg=1 W m−1 K−1, κc∥=180 W m−1 K−1 and κc⊥=69–180 W m−1 K−1, which is given by κc∥(1−fITA/100), where fITA=0–62%.

Second, to study the effect of the degree of orientation on the thermal anisotropy, we use the varied intrinsic thermal anisotropy in the range of fITA= 0–62%. When fCA= 0, no orientation occur; when fITA= 62%, the orientation is perfect. The values used in the calculations are w= 1 μm, κ c∥= 180 W m−1 K−1 and κ c⊥= κ c∥(1 - fITA/100) W m−1 K−1. Figure 34(b) shows the effect of the intrinsic thermal anisotropy on the thermal anisotropy of textured Si3N4 under various simulated conditions. Clearly, the intrinsic thermal anisotropy or the degree of orientation is the dominant factor governing the thermal anisotropy of textured β-Si3N4, and the aspect ratio has a very limited effect. In the perfectly oriented β-Si3N4, varying the microstructural factors such as the aspect ratio, volume fraction of the glassy phase and the grain boundary film thickness, results in a slight change in the thermal anisotropy from 58 to 66%.

Experimental results

In 1978, thermal anisotropy in β-Si3N4 ceramics was first reported for hot-pressed Si3N4 by Hirai et al [39]. They showed that the thermal conductivity along the direction parallel to the HP is 36% lower than that along the direction perpendicular to the HP at 20 °C. Moreover, this value was found to decrease to 15% at 1300 °C. Kitayama et al conducted a series of studies on the thermal conductivity of hot-pressed Si3N4, including the roles of the starting powders [53], lattice oxygen [58] and RE oxide additives [62], as summarized in figure 35. Figure 35(d) shows the two cases for measuring thermal conductivity by the laser flash method. Figure 35(a) shows that fine SN-E10 powder leads to a larger grain size but a lower thermal anisotropy than coarse SN-E05 powder, primarily as a result of the larger aspect ratio, which is believed to increase the degree of orientation during HP [53]. However, the annealing time has little effect on the thermal anisotropy, indicating that the grain growth plays a less important role in the degree of orientation during annealing. The textured β-Si3N4, which was hot-pressed from β-Si3N4 powder in a shorter time without annealing, exhibits a substantially smaller grain size than the sample hot-pressed from SN-E05 powder followed by annealing, but both samples exhibit almost the same thermal anisotropy. Figure 35(b) shows the effect of the ionic radius of RE cations (RE3 +) on the thermal anisotropy. It was found that during HP, the La2O3-doped sample exhibited the lowest thermal anisotropy, corresponding to the largest ionic radius of RE3 +. This is attributed to the incomplete α–β phase transformation, thereby resulting in the lowest degree of orientation. However, prolonged annealing allows samples to exhibit an increase in the thermal anisotropy with increased ionic radius, indicating that the grain growth contributes slightly to the texture development. Annealing has no effect on the thermal anisotropy in the Sc2O3-doped sample. This may be due to the smallest anisotropic grain growth rate of β-Si3N4 in this sample because of the smallest ionic radius of RE3 +, as predicted by DBE theory [93]. Nevertheless, no direct relationship can be observed between the thermal anisotropy and the grain size. Figure 35(c) shows that the thermal anisotropy increases gradually with increased Y2O3: SiO2 molar ratio, which seems to correlate with the increased grain size. Actually, this should be attributed to the increased aspect ratio with increased Y2O3: SiO2 molar ratio, thereby resulting in a higher degree of orientation during HP and subsequent annealing [62, 135]. However, it is found that thermal anisotropy is limited to below 30% in hot-pressed Si3N4 regardless of the processing conditions. This is attributed to the a, b-axis aligned texture, in which the β-Si3N4 elongated grains are randomly oriented along the c-axis.

Figure 35.

Figure 35

Effects of (a) Si3N4 raw powder [53], (b) ionic radius of RE cation [62] and (c) Y2O3/SiO2 ratio [58] on the grain size and thermal anisotropy of textured β-Si3N4 produced by HP at 1800 °C for 2 h and subsequent annealing at 1850 °C. Both SN-E10 and SN-E05 are α-Si3N4 type, and SN-21FC is β-Si3N4 type. SN-E05 powder was used in both (b) and (c). The grain size was determined from SEM micrographs of the plane perpendicular to the HP direction by the linear intercept method. (d) shows the two cases for the thermal conductivity measurements by the laser flash method.

Table 18 shows a summary of the reported thermal conductivities of highly c-axis aligned β-Si3N4 ceramics produced by the TCA technique. The c-axis aligned β-Si3N4 exhibits thermal anisotropy values of Inline graphic , which are substantially higher than those of a, b-axis aligned β-Si3N4. The highest thermal anisotropy reaches 66%, almost the same as the intrinsic thermal anisotropy of single β-Si3N4 grains, suggesting that a nearly perfect orientation is developed in such a material. Evidently, the thermal anisotropy is insensitive to the composition and amount of additives, but is sensitive to the degree of orientation, as theoretically predicted. It is interesting to see that by using the improved TCA method to increase the degree of orientation of seed particles, GPS also can produce textured β-Si3N4 that exhibits thermal anisotropy as strong as that produced by HP and hot-isostatic pressing (HIP). However, GPS is a relatively inexpensive process. As a result, to maximize the thermal anisotropy in β-Si3N4 ceramics, a highly c-axis aligned texture is of particular importance. To obtain higher thermal conductivity, the other processing conditions should be controlled at the same time, such as the use of effective Al-free additives and sintering techniques.

Table 18.

Thermal conductivity (κ) of highly textured Si3N4 produced by the TCA technique.

Si3N4 Sintering/annealing κ(Wm−1 K−1) Schematic of three thermal
Year Authors powder β-Si3N4 seed Sintering additives conditions x y z fTA (%) conduction directions
1996 Hirao et al [45] α-Si3N4 5 vol% 5 wt% Y2O3 HP (1800 °C, 2 h, 121 75 60 50 Inline graphic
 (SN-E10) w=1 μm  flow N2)→
l=10 μm Annealing (1850 °C,
 66 h, 0.9 MPa N2)
1996 Hirosaki et al [47] β-Si3N4 5 vol% 0.5 mol% Y2O3+ HP (1940 °C, 0.5 h, 137 97 72 47
 (SN- w=1 μm  0.5 mol% Nd2O3  1 MPa N2)→
 P21FC) l=10 μm Annealing (2200 °C,
 4 h, 30 MPa N2)
1999 Watari et al [52] α-Si3N4 5 vol% 5 wt% Y2O3 HP (1800 °C, 2 h, 70 30 57
 (SN-E10) w=1.3 μm  flow N2)
l=5.4 μm
HIP (2500 °C, 2 h, 155 67 52 66
 200 MPa N2) Inline graphic
2000 Lee et al [59] α-Si3N4 10 wt% 6 wt% Y2O3 + GPS (2000 °C, 4 h, 78 28 64
 (SN-E10) w=0.1–1.5 μm  2 wt% Al2O3  3 MPa N2)
l=10–15 μm

As discussed above, textured Si3N4 ceramics exhibit anisotropy in their thermal, physical and mechanical properties, such as the hardness, elastic modulus, bending strength, fracture toughness, fracture energy, creep resistance, tribological and wear resistance, erosion resistance, contact damage resistance and thermal conductivity. The anisotropic properties result either from the intrinsic anisotropy or from the grain orientation or from both, as summarized in table 19. On the basis of previous experimental evidence, figure 36 shows a scheme of the directions and planes giving the maximum properties in the c-axis oriented β-Si3N4 ceramics, the directions giving the lowest and highest elastic modulus are also shown, because the required elastic modulus depends on the structural applications. Although it was reported that, corresponding to the directions perpendicular and parallel to the grain orientation, c-axis oriented β-Si3N4 exhibits higher bend and tensile creep resistance, respectively, than untextured β-Si3N4, the effect of the grain orientation on the bend and tensile creep resistance of textured β-Si3N4 is still unknown. In addition, the effect of the grain orientation on the erosion and contact damage resistance is also still unknown. Therefore, their maximum properties cannot be shown in figure 36. Further studies are needed to clarify the effect of the grain orientation on the bend and creep resistance, erosion and contact damage resistance.

Table 19.

Anisotropic properties of textured Si3N4 and their origins.

Anisotropic properties Origins
Hardness Intrinsic anisotropy
 and grain orientation
Elastic modulus Intrinsic anisotropy
 and grain orientation
Bend strength Grain orientation
Fracture toughness Grain orientation
Fracture energy Grain orientation
Creep resistance Grain orientation
Tribological and wear resistance Grain orientation
Erosion resistance Grain orientation
Contact damage resistance Grain orientation
Thermal conductivity Intrinsic anisotropy
 and grain orientation
Figure 36.

Figure 36

Scheme of the directions and planes giving the maximum properties in the c-axis oriented β-Si3N4, which have been clearly demonstrated experimentally in the literature. Note that the directions giving the highest and lowest elastic modulus are shown, because the required elastic modulus depends on the structural applications.

Textured porous Si3N4

Textured porous Si3N4 is a type of novel porous ceramic composed of oriented elongated grains and anisotropic pores. This unique porous microstructure gives this material excellent mechanical properties [183, 184, 203206] and thermal-mechanical properties [207], including strength, fracture toughness, fracture energy, crack growth resistance and thermal shock resistance. Figure 37 schematically illustrates the microstructure of textured porous Si3N4 with unidirectionally oriented elongated grains and elongated pores developed by Shigegaki et al [184] by tape casting using a β-Si3N4 whisker raw powder. They showed that this unidirectionally textured porous Si3N4 with a porosity of 14.4% exhibits superior strain tolerance to untextured dense Si3N4 because of the lower Young's modulus and the same high strength. Moreover, this material exhibited a fracture energy of 490 J m− 2, about 7 times that of untextured dense Si3N4 [203].

Figure 37.

Figure 37

Schematic microstructure of a textured porous Si3N4 ceramic with unidirectionally aligned elongated β-Si3N4 grains and ellipsoidal pores.

Processing

The methods of processing textured porous Si3N4 are similar to those used for producing textured dense Si3N4, such as HW and TGG. The only difference is that the processing of textured porous Si3N4 requires incomplete densification by tailoring the Si3N4 raw powder, additive composition and sintering conditions. Nevertheless, two major methods have been commonly used in previous studies: tape casting and SF. In tape casting, β-Si3N4 whiskers or rodlike seed crystals are used as raw powders to produce unidirectionally textured porous Si3N4, because they tend to retard densification during sintering because of the smaller specific surface area. Inagaki et al [204] showed that a similar porous texture develops when β-Si3N4 whiskers and rodlike seed crystals are used as raw materials. In partial SF, α-Si3N4 raw powders are used to preferentially develop elongated β-Si3N4 grains during sintering, thereby resulting in a porous texture due to grain rotation during forging. Nevertheless, it is difficult to produce unidirectionally textured porous Si3N4 by partial SF compared with by tape casting. From the industrial point of view, tape casting is an efficient processing method for producing high-quality textured porous Si3N4. β-Si3N4 whisker powders are no longer commercially available because of safety considerations. However, rodlike seed crystals are very expensive as a result of the limited manufacturing capacity. Therefore, alternative raw powders should be considered, particularly commercially available α-Si3N4 raw powder in conjunction with a seeding method, as used for the processing of textured dense Si3N4. Actually, the addition of large rodlike β-Si3N4 seed crystals tends to retard the densification of fine α-Si3N4 compacts [13, 44, 56]. Thus, the porous texture can be tailored by optimizing the morphology and amount of seed crystals, the additive composition and processing conditions. The use of fugitive fibers should be an efficient method for tailoring an elongated pore morphology and alignment; they can undoubtedly be easily aligned together with rodlike β-Si3N4 seeds during tape casting.

Properties

It has been shown that pores can provide more bridging sites by detaching the small matrices from the large elongated grains, thereby resulting in greater crack growth resistance than in untextured dense Si3N4, as shown in figure 38. This is also the major reason why textured porous Si3N4 possesses a high fracture energy than other types of Si3N4 ceramics, such as textured dense Si3N4 and untextured dense and porous Si3N4, as illustrated in figure 39. The fracture energy is affected by porosity and texture. The fracture energy is higher in the direction perpendicular to the grain alignment than in the direction parallel to the grain alignment. The effect of porosity on the fracture energy appears to depend on the texture. In the direction parallel to the grain alignment, the porosity always tends to reduce the fracture energy. However, in the direction perpendicular to the grain alignment, up to a porosity of ∼\!15%, the fracture energy increases with increasing porosity, but above this value the fracture energy decreases. This suggests that 15% may be the critical porosity for the fracture energy of textured porous Si3N4. However, no linear relationship in the semi-log plots of fracture energy against porosity can be found.

Figure 38.

Figure 38

R-curves of textured porous Si3N4 and untextured dense Si3N4 based on the chevron-notched-beam (CNB) method. The square dots represent the experimental results and solid lines represent the fitting curves. The porous textured Si3N4 exhibits the c-axis alignment of elongated β-Si3N4 grains with 84.3% of TD (reproduced with permission from [206] ©2005 Blackwell Publishing Ltd).

Figure 39.

Figure 39

Fracture energy versus porosity for textured porous Si3N4 in the directions perpendicular and parallel to the alignment of elongated grains. The fracture energy data was measured by the CNB method using the chevron-notch planes perpendicular and parallel to the grain alignment. For comparison, some data for untextured Si3N4 are also shown in the figure.

In contrast to the fracture energy, the reported data for the strength, toughness and Young's modulus are found to have a linear relationship with porosity in semi-log plots, as illustrated in figures 4042, respectively, i.e., the dependences of these mechanical properties on porosity can be described by the general formula [208]

graphic file with name TSTA11660773M51.jpg

where Pr and Pr0 are the properties of the materials with and without pores, respectively, b is the parameter determined by the pore characteristics and Φ is the volume fraction of the pores instead of the volume percentage. Table 20 shows a summary of the fitting results of bending strength, fracture toughness and Young's modulus against porosity using equation (51). It is found that there is a similar value of b= ∼5 in the bending strength for the directions perpendicular and parallel to the grain alignment. Because the bending strength data in figure 40 was obtained from unidirectionally textured porous Si3N4 materials, it appears to suggest that the parameter b is independent of the grain alignment in the materials. However, there is a large difference in the b value for the Young's modulus in the directions perpendicular and parallel to the grain alignment. This is very likely to be due to the a, b-axis instead of the c-axis aligned texture, because the Young's modulus data in the direction parallel to the grain alignment were obtained only from a, b-axis aligned porous Si3N4 prepared by partial SF using SC, in which the alignment of the c-axis is random. Therefore, the Young's modulus in the direction perpendicular to the grain alignment should be equal to the Young's modulus in the direction perpendicular to the c-axis of β-Si3N4, but the Young's modulus in the direction parallel to the grain alignment should be equal to the Young's modulus of the untextured porous Si3N4 materials. At the same time, the results suggest that, compared with untextured porous Si3N4, the excellent mechanical properties of textured porous Si3N4 should be primarily attributed to the alignment of elongated β-Si3N4 grains.

Figure 40.

Figure 40

Porosity dependence of bending strength for textured porous Si3N4 ceramics in the tensile stress plane parallel and perpendicular to the alignment of elongated grains. All data were obtained from three-point bending tests. For comparison, some data for untextured Si3N4 are also shown in the figure.

Figure 42.

Figure 42

Porosity dependence of Young's modulus for textured porous Si3N4 in the directions perpendicular and parallel to the alignment of elongated grains. Young's modulus data were measured by the pulse echo method. For comparison, some data for untextured Si3N4 are also shown in the figure.

Table 20.

Fitting results of bending strength, fracture toughness and Young's modulus by equation (51).

ba
Property Equation Property without porosity
Bending strength σb σbb0ebf3 4.6 5.1 σb0⊥=1863, σb0∥=1096 MPa
Fracture toughness KIC KIC=KIC0ebf3 2.8 KIC0⊥=20 MPa m1/2
Young's modulus E E=E0ebf3 3.8 2.3 E0⊥=359, E0∥=326 GPa

a⊥ and ∥ denote the directions perpendicular and parallel to the alignment of elongated grains, corresponding to the directions of the tensile stress plane parallel and perpendicular to the grain alignment for the bending tests, see figures 4042.

Figure 41.

Figure 41

Porosity dependence of fracture toughness for textured porous Si3N4 ceramics perpendicular to the alignment of elongated grains [204]. Fracture toughness, KIC, was calculated using the equation KIC=(2E⊥γ)1/2, where E is the Young's modulus of each porous sample in the plane-strain condition given by E=E/(1−ν2) with Poisson ratio ν=0.25.

Figure 43 shows the significantly improved thermal shock resistance of textured porous Si3N4 in comparison with untextured porous Si3N4 using the water-quenching method. The critical temperature differences are ∼900 and ∼1400 C for untextured and textured porous Si3N4, respectively. The critical temperature difference of both untextured and textured porous Si3N4, ΔTc, can be determined by [209]

graphic file with name TSTA11660773M52.jpg

where σ b is the flexural strength, E is the Young's modulus, α is the thermal expansion coefficient, ν is the Poisson's ratio, Q is a shape factor dependent on the sample geometry (Q= 3.25 for a plate and Q= 4.3 for a rod), κ is the thermal conductivity, a is the characteristic heat-transfer length and h is the surface heat-transfer coefficient. According to equation (52), the ΔTc values of untextured and textured porous Si3N4 are 725 and 1347 °C, respectively, essentially consistent with the experimental values. The slight discrepancy between the calculated and experimental values may arise from the uncertainty of the values of h.

Figure 43.

Figure 43

Strength damage factor as a function of quenching-temperature difference for textured and untextured porous Si3N4 by the water-quenching method [207]. Strength damage factor, Ds, is defined as Ds=(1−σbb0)×100%, where σb is the bending strength of as-received samples and σb0 is the bending strength after water-quenching.

Textured α-Sialon

α and β-Sialon are two major solution ceramics of Si3N4, corresponding to α-Si3N4 and β-Si3N4, respectively. However, compared with the elongated growth of β-Sialon grains, α-Sialon grains normally grow into an equiaxed shape, which results in α-Sialon ceramics having low fracture toughness and it being difficult to form a textured microstructure. However, the discovery of the preferential orientation of elongated Ca-α-Sialon grains during HP indicates the possibility of developing textured α-Sialon by facilitating elongated grain growth, similar to the cases of β-Sialon and β-Si3N4 [210]. The texturing behavior has also been observed in hot-pressed Ca-α-Sialon [211], Li-α-Sialon [212], and Nd-(α + β)-Sialon [213], as clearly evidenced by the stronger relative intensities of all (hk0) diffractions on the surface perpendicular to the HP than on the surface parallel to the HP.

Carman et al [214] conducted a detailed study of texturing behavior in Sm-α-Sialon ceramics during HP and subsequent HF, as schematically illustrated in figure 44. The texture development during HP and HF was characterized by pole figure analysis, as shown in figure 45. The maximum mrd values at the edge of the pole figures (figure 45(a)) indicate that the unit cell basal plane is perpendicular to the HP direction (Direction 1), i.e. the elongated α-Sialon crystals are preferentially oriented with the c-axis perpendicular to the HP direction. During HF, the texture development is found to depend on the forging time. When the forging time is 15 min, a unidirectional or c-axis oriented texture along Direction 3 is formed, as indicated by the higher maximum mrd value in Direction 3 than in Direction 1 at the edge of the pole figure (figure 45(b)). When the forging time increases to 30 min, the c-axis oriented texture is enhanced, as indicated by the increased maximum mrd value in Direction 3 at the edge of the pole figure (figure 45(c)) compared with the corresponding value in figure 45(b). However, with further forging (up to 60 min), the c-axis oriented texture along Direction 3 tends to change into a 2D oriented texture with the c-axis preferentially oriented perpendicular to the HF direction (Direction 2), as indicated by figure 45(d). Despite this, the c-axis of the α- Sialon crystals is still preferentially oriented along Direction 3. In contrast, the 2D oriented texture in the hot-pressed sample features the random alignment of the c-axis of α-Sialon crystals, as indicated by the symmetry of the pole figure (figure 45(a)). Furthermore, the authors carried out a multiple-forging experiment to further enhance the orientation of the c-axis as follows: a hot-pressed sample along Direction 1 was then hot-forged three times at 1800 °C for 30 min—first along Direction 2, then along Direction 1, and then along Direction 2 again. Thus, a higher degree of c-axis orientation along Direction 3 was eventually obtained. The texture development in α-Sialon ceramics is further indicated by the measured toughness anisotropy on the different faces using the Vickers indentation technique, as shown in figure 46. Clearly, as the c-axis orientation is enhanced, corresponding to 15 and 30 min forging and multiple forging, the toughness anisotropy of both Faces 1 and 2 (figure 45(b)) exhibits an increase, whereas the toughness anisotropy of Face 3 remains almost constant. Specifically, multiple forging results in a toughness anisotropy of as high as 50%, similar to that of c-axis oriented β-Si3N4 prepared by tape casting (table 12). Although the α- Si3N4 (and α-Sialon) crystal exhibits inherent anisotropic hardness, similar to the β-Si3N4 (and β-Sialon) crystal [178, 215], no significant anisotropy is observed in the hardness of textured α-Sialon ceramics [214].

Figure 44.

Figure 44

Schematic illustration of HF of hot-pressed Sm-α-Sialon: (a) HP (Face 1 ⊥ uniaxial HP load) and (b) HF (Face 2 ⊥ uniaxial HF load).

Figure 45.

Figure 45

Calculated (00l) basal pole figures of hot-pressed sample (a) (the HP direction, or Direction 1, is located at the center of the pole figure), and samples hot-forged for (b) 15 min, (c) 30 min and (d) 60 min (the HF direction, or Direction 2, is located at the center of the pole figures) (reproduced with permission from [214] ©2006 Blackwell Publishing Ltd).

Figure 46.

Figure 46

Toughness anisotropy of the three faces in textured α-Sialon obtained by HP and subsequent HF. The toughness was determined from the three faces shown in figure 44 by the Vickers indentation technique, as schematically illustrated in figure 29. The toughness anisotropy factor, fFTA, is determined by equation (43) [214].

It has been reported that rodlike α-Sialon single crystals can be prepared by various methods, such as the sintering of powder pellets [216] and combustion synthesis [217]. Therefore, in addition to the HW method, the TGG method is undoubtedly applicable to the production of textured α-Sialon ceramics, although there have apparently been no reports on this method. Moreover, SMFA is a novel processing technique for producing textured α-Sialon using α-Si3N4 as the starting powder, as has been recently demonstrated by Zhu et al [218]. They obtained a, b-axis aligned Ca-α-Sialon using slip casting of α-Si3N4, CaCO3, Al2O3 and AlN powders in an SMF of 12 T, followed by reaction sintering. The texture is characterized by the long axis (c-axis) of elongated Ca-α-Sialon grains perpendicular to the magnetic field, and the texturing mechanism is attributed to heterogeneous nucleation and epitaxial grain growth on α-Si3N4 seeds initially oriented.

Summary and outlook

A considerable amount of work over the past 15 years has led to a better understanding of the fundamental aspects of the processing and anisotropic properties of textured Si3N4 ceramics. Several theoretical models, such as the rough–smooth plane model, the anisotropic Ostwald ripening model, the acid-base model and the differential-binding-energy model, have been proposed to interpret elongated grain growth of β-Si3N4, which is directly responsible for the development of the unique bimodal microstructure as well as textured microstructure. In addition, the texture development and bimodal microstructure are also related to the abnormal grain growth, which is governed by the morphology, grain-size distribution and the amount of β-Si3N4 nuclei.

The processing techniques for obtaining textured Si3N4 described in this review have been classified into two types: HW and TGG. The techniques involved in HW are HP, HF and SF, and the techniques involved in TGG are CPA, EA, TCA and SMFA based on the alignment of the β-Si3N4 template (or seeds). Textured β-Si3N4 occurs in two major structural forms: a, b-axis (or planar) orientation and c-axis orientation, depending on the processing technique. Highly c-axis oriented β-Si3N4 can be obtained by HF and SF using a large tensile or PSC deformation. The texture development is primarily due to the grain rotation, and the strain is the dominant factor controlling the degree of texture. In TGG, both EA and TCA are efficient for producing highly c-axis oriented β-Si3N4 by optimizing the aspect ratio and content of β-Si3N4 seeds and using HT sintering/annealing. However, the prerequisite for both orientations is a rodlike morphology and a grain size larger than the matrix grains. In contrast, the most attractive characteristic of SMFA is the independence of the particle shape of ceramics powders. Owing to the magnetic susceptibility of χ a, b> χc for β-Si3N4, a, b-axis oriented β-Si3N4 is preferentially developed in an SMF. However, an RMF can be used to obtain the c-axis oriented β-Si3N4. It is clearly demonstrated that β-Si3N4 nuclei play a key role in texture development in Si3N4 via SMFA, and the addition of β-Si3N4 seeds is more efficient for increasing the degree of texture than varying sintering conditions.

The anisotropy in the hardness, elastic modulus, bending strength, fracture toughness, fracture energy, tribological and wear behavior and thermal conductivity has been clearly demonstrated and is well understood. It has been revealed both theoretically and experimentally that the thermal anisotropy is mainly determined by the intrinsic anisotropy and degree of orientation, and the effect of the rodlike grain morphology becomes less important. The anisotropic properties are maximized for the perfect c-axis orientation. Although it was reported that c-axis oriented β-Si3N4 exhibits higher bend and tensile creep resistance than untextured β-Si3N4, corresponding to the directions perpendicular and parallel to the grain orientation, respectively, the effect of the grain orientation on the bend and tensile creep resistance of textured β-Si3N4 is unknown. In addition, the effect of grain orientation on the erosion and contact damage resistance is also still unknown. Therefore, further studies are needed to clarify the effect of the grain orientation on the creep, erosion, contact damage and chemical corrosion behavior.

Moreover, the existence of oriented anisotropic pores allows textured porous Si3N4 to possess comparable and even superior mechanical and thermal-mechanical properties to textured dense Si3N4 and untextured dense and porous Si3N4. The presence of elongated α-Sialon grains allows the production of textured α-Sialon using the same methods as for producing textured β-Si3N4 and β-Sialon. The toughness anisotropy has been demonstrated in textured α-Sialon prepared by HF. Because transient liquid-phase sintering enables the production of glass-free α-Sialon ceramics, it is expected that highly c-axis oriented texture should lead to the further improvement in the HT properties of α-Sialon ceramics. To establish a better understanding of the processing and properties of textured α-Sialon, further studies are needed.

Cost and mechanical reliability are always major barriers preventing the widespread applications of Si3N4 ceramics. Texturing offers a unique opportunity to improve the mechanical reliability, which will help expand the applications of Si3N4 ceramics, particularly in extremely severe environments. From the industrial point of view, TCA is a low-cost process for producing textured Si3N4. However, TCA and EA have limitations for producing textured β-Si3N4 components with complex shapes. In contrast, SMFA is a promising process for producing textured Si3N4 components with complex shapes, because various colloidal forming approaches are well developed, such as slip casting, electrophoretic deposition and gel casting. However, the properties of textured β-Si3N4 produced by this technique are still not known. Therefore, intensive studies should be carried out on the processing and properties of textured β-Si3N4 by SMFA.

References

  1. Ziegler G, Heinrich J. and Wötting W. J. Mater. Sci. 1987;22:3041. doi: 10.1007/BF01161167. [DOI] [Google Scholar]
  2. Riley F L. J. Am. Ceram. Soc. 2000;83:245. [Google Scholar]
  3. Hardie D. and Jack K H. Nature. 1957;180:332. doi: 10.1038/180332a0. [DOI] [Google Scholar]
  4. Bowen L J, Weston R J, Carruthers T G. and Brook R J. J. Mater. Sci. 1978;13:341. doi: 10.1007/BF00647779. [DOI] [Google Scholar]
  5. Natansohn S. and Sarin V K. Ceramic Powder Processing Science. In: Hausner H, , Messing G L, Hirano S, , editors. Berchtesgaden: Deutsche Keramische Gesellschaft; 1988. p. p 433. [Google Scholar]
  6. Suematsu H, Mitomo M, Mitchell T E, Petrovic J J, Fukunaga O. and Ohashi N. J. Am. Ceram. Soc. 1997;80:615. [Google Scholar]
  7. Becher P F. J. Am. Ceram. Soc. 1991;74:255. doi: 10.1111/j.1151-2916.1991.tb06872.x. [DOI] [Google Scholar]
  8. Lange F F. J. Am. Ceram. Soc. 1973;56:518. doi: 10.1111/j.1151-2916.1973.tb12401.x. [DOI] [Google Scholar]
  9. Lange F F. J. Am. Ceram. Soc. 1979;62:428. doi: 10.1111/j.1151-2916.1979.tb19096.x. [DOI] [Google Scholar]
  10. Mitomo M. and Uenosono S. J. Am. Ceram. Soc. 1992;75:103. doi: 10.1111/j.1151-2916.1992.tb05449.x. [DOI] [Google Scholar]
  11. Kawashima T, Okamoto H, Yamamoto H. and Kitamura A. J. Ceram. Soc. Japan. 1991;99:320. [Google Scholar]
  12. Hirosaki N, Akimune Y. and Mitomo M. J. Am. Ceram. Soc. 1993;76:1892. doi: 10.1111/j.1151-2916.1993.tb06670.x. [DOI] [Google Scholar]
  13. Hirao K, Nagaoka T, Brito M E. and Kanzaki S. J. Am. Ceram. Soc. 1994;77:1857. doi: 10.1111/j.1151-2916.1994.tb07062.x. [DOI] [Google Scholar]
  14. Björklund H, Falk L K L. J. Eur. Ceram. Soc. 1997;17:13. doi: 10.1016/S0955-2219(96)00075-1. [DOI] [Google Scholar]
  15. Becher P F, et al J. Am. Ceram. Soc. 1998;81:2821. [Google Scholar]
  16. Hayashi H, Hirao K, Toriyama M, Kanzaki S. and Itatani K. J. Am. Ceram. Soc. 2001;84:3060. [Google Scholar]
  17. Peillon F C. and Thevenot F. J. Eur. Ceram. Soc. 2002;22:271. doi: 10.1016/S0955-2219(01)00290-4. [DOI] [Google Scholar]
  18. Tsuge A, Nishida K. and Komatsu M. J. Am. Ceram. Soc. 1975;58:323. doi: 10.1111/j.1151-2916.1975.tb11488.x. [DOI] [Google Scholar]
  19. Cinibulk M K, Thomas G. and Johnson S M. J. Am. Ceram. Soc. 1992;75:2037. doi: 10.1111/j.1151-2916.1992.tb04462.x. [DOI] [Google Scholar]
  20. Cinibulk M K, Thomas G. and Johnson S M. J. Am. Ceram. Soc. 1992;75:2050. doi: 10.1111/j.1151-2916.1992.tb04464.x. [DOI] [Google Scholar]
  21. Cinibulk M K, Thomas G. and Johnson S M. J. Am. Ceram. Soc. 1992;75:2044. doi: 10.1111/j.1151-2916.1992.tb04463.x. [DOI] [Google Scholar]
  22. Pyzik A J. and Beaman D R. J. Am. Ceram. Soc. 1993;76:2737. doi: 10.1111/j.1151-2916.1993.tb04010.x. [DOI] [Google Scholar]
  23. Becher P F, Sun E Y, Hsueh C H, Alexander K B, Hwang S L, Waters S B. and Westmoreland C G. Acta Mater. 1996;44:3881. doi: 10.1016/S1359-6454(96)00069-9. [DOI] [Google Scholar]
  24. Nishimura T, Mitomo M. and Suematsu H. J. Mater. Res. 1997;12:203. doi: 10.1557/JMR.1997.0027. [DOI] [Google Scholar]
  25. Park H J, Kim H E. and Niihara K. J. Am. Ceram. Soc. 1997;80:750. doi: 10.1111/j.1151-2916.1997.tb02797.x. [DOI] [Google Scholar]
  26. Sun E Y, Becher P F, Plucknett K P, Hsueh C H, Alexander K B, Waters S B, Hirao K. and Brito M E. J. Am. Ceram. Soc. 1998;81:2831. doi: 10.1111/j.1151-2916.1998.tb02774.x. [DOI] [Google Scholar]
  27. Kleebe H J, Pezzotti G. and Ziegler G. J. Am. Ceram. Soc. 1999;82:1857. [Google Scholar]
  28. Pezzotti G. and Kleebe H J. J. Eur. Ceram. Soc. 1999;19:451. doi: 10.1016/S0955-2219(98)00216-7. [DOI] [Google Scholar]
  29. Becher P F, Painter G S, Sun E Y, Hsueh C H. and Lance M J. Acta Mater. 2000;48:4493. doi: 10.1016/S1359-6454(00)00236-6. [DOI] [Google Scholar]
  30. Lu H H. and Huang J L. Ceram. Int. 2001;27:621. doi: 10.1016/S0272-8842(01)00008-6. [DOI] [Google Scholar]
  31. Guo S Q, Hirosaki N, Yamamoto Y, Nishimura T. and Mitomo M. Scr. Mater. 2001;45:867. doi: 10.1016/S1359-6462(01)01111-3. [DOI] [Google Scholar]
  32. Guo S Q, Hirosaki N, Yamamoto Y, Nishimura T. and Mitomo M. J. Mater. Res. 2001;16:3254. doi: 10.1557/JMR.2001.0448. [DOI] [Google Scholar]
  33. Guo S Q, Hirosaki N, Nishimura T, Yamamoto Y. and Mitomo M. Philos. Mag. 2002;A 82:3027. [Google Scholar]
  34. Guo S Q, Hirosaki N, Yamamoto Y, Nishimura T. and Mitomo M. J. Am. Ceram. Soc. 2002;85:1607. [Google Scholar]
  35. Guo S Q, Hirosaki N, Nishimura T, Yamamoto Y. and Mitomo M. J. Am. Ceram. Soc. 2003;86:1900. doi: 10.1111/j.1151-2916.2003.tb03579.x. [DOI] [Google Scholar]
  36. Guo S Q, Hirosaki N, Yamamoto Y, Nishimura T. and Tanaka H. Mater. Lett. 2003;57:3257. doi: 10.1016/S0167-577X(03)00044-2. [DOI] [Google Scholar]
  37. Guo S Q, Hirosaki N, Yamamoto Y, Nishimura T. and Kagawa Y. Mater. Sci. Eng. 2005;A 408:9. doi: 10.1016/j.msea.2005.05.025. [DOI] [Google Scholar]
  38. Satet R L. and Hoffmann M J. J. Am. Ceram. Soc. 2005;88:2485. doi: 10.1111/j.1551-2916.2005.00421.x. [DOI] [Google Scholar]
  39. Hirai T, Hayashi S. and Niihara K. Am. Ceram. Soc. Bull. 1978;57:1126. [Google Scholar]
  40. Weston J E. J. Mater. Sci. 1980;15:1568. doi: 10.1007/BF00752139. [DOI] [Google Scholar]
  41. Muscat D, Pugh M D, Drew R A L, Pickup H. and Steele D. J. Am. Ceram. Soc. 1992;75:2713. doi: 10.1111/j.1151-2916.1992.tb05494.x. [DOI] [Google Scholar]
  42. Lee F J, Sandlin M S. and Bowman K J. J. Am. Ceram. Soc. 1993;76:1793. doi: 10.1111/j.1151-2916.1993.tb06649.x. [DOI] [Google Scholar]
  43. Hirao K, Ohashi M, Brito M E. and Kanzaki S. J. Am. Ceram. Soc. 1995;78:1687. doi: 10.1111/j.1151-2916.1995.tb08871.x. [DOI] [Google Scholar]
  44. Imamura H, Hirao K, Brito M E, Toriyama M. and Kanzaki S. J. Ceram. Soc. Japan. 1996;104:748. [Google Scholar]
  45. Hirao K, Watari K, Brito M E, Toriyama M. and Kanzaki S. J. Am. Ceram. Soc. 1996;79:2485. doi: 10.1111/j.1151-2916.1996.tb09002.x. [DOI] [Google Scholar]
  46. Miller P D. and Bowman K J. Acta Mater. 1996;44:3025. doi: 10.1016/1359-6454(95)00443-2. [DOI] [Google Scholar]
  47. Hirosaki N, Ando M, Okamoto Y, Munakata F, Akimune Y, Hirao K, Watari K, Brito M E, Toriyama M. and Kanzaki S. J. Ceram. Soc. Japan. 1996;104:1171. [Google Scholar]
  48. Yoon S Y, Akatsu T. and Yasuda E. J. Mater. Sci. 1997;32:3813. doi: 10.1023/A:1018631924934. [DOI] [Google Scholar]
  49. Kondo N, Ohji T. and Wakai F. J. Am. Ceram. Soc. 1998;81:713. doi: 10.1111/j.1151-2916.1998.tb02759.x. [DOI] [Google Scholar]
  50. Teshima H, Hirao K, Toriyama M. and Kanzaki S. J. Ceram. Soc. Japan. 1999;107:1216. [Google Scholar]
  51. Kondo N, Suzuki Y. and Ohji T. J. Am. Ceram. Soc. 1999;82:1067. [Google Scholar]
  52. Watari K, Hirao K, Brito M E, Toriyama M. and Kanzaki S. J. Mater. Res. 1999;14:1538. doi: 10.1557/JMR.1999.0206. [DOI] [Google Scholar]
  53. Kitayama M, Hirao K, Toriyama M. and Kanzaki S. J. Am. Ceram. Soc. 1999;82:3105. [Google Scholar]
  54. Liang Y N, Lee S W. and Park D S. Wear. 1999;224:202. doi: 10.1016/S0043-1648(98)00284-1. [DOI] [Google Scholar]
  55. Liang Y N, Lee S W. and Park D S. Wear. 1999;225–229:1327. doi: 10.1016/S0043-1648(99)00058-7. [DOI] [Google Scholar]
  56. Imamura H, Hirao K, Brito M E, Toriyama M. and Kanzaki S. J. Am. Ceram. Soc. 2000;83:495. [Google Scholar]
  57. Park D S, Choi M J, Roh T W, Kim H D. and Han B D. J. Mater. Res. 2000;15:130. doi: 10.1557/JMR.2000.0022. [DOI] [Google Scholar]
  58. Kitayama M, Hirao K, Tsuge A, Watari K, Toriyama M. and Kanzaki S. J. Am. Ceram. Soc. 2000;83:1985. [Google Scholar]
  59. Lee S W, Chae H B, Park D S, Choa Y H, Niihara K. and Hockey B J. J. Mater. Sci. 2000;35:4487. doi: 10.1023/A:1004807516791. [DOI] [Google Scholar]
  60. Kondo N, Suzuki Y, Brito M E. and Ohji T. J. Mater. Res. 2001;16:2182. doi: 10.1557/JMR.2001.0297. [DOI] [Google Scholar]
  61. Kondo N, Inagaki Y, Suzuki Y. and Ohji T. J. Am. Ceram. Soc. 2001;84:1791. [Google Scholar]
  62. Kitayama M, Hirao K, Watari K, Toriyama M. and Kanzaki S. J. Am. Ceram. Soc. 2001;84:353. [Google Scholar]
  63. Nakamura M, Hirao K, Yamauchi Y. and Kanzaki S. J. Am. Ceram. Soc. 2001;84:2579. [Google Scholar]
  64. Xie R J, Mitomo M, Xu F F, Zhan G D, Bando Y. and Akimune Y. J. Eur. Ceram. Soc. 2002;22:963. doi: 10.1016/S0955-2219(01)00401-0. [DOI] [Google Scholar]
  65. De Pablos A, Osendi M I. and Miranzo P. J. Am. Ceram. Soc. 2002;85:200. [Google Scholar]
  66. Nakamura M, Hirao K, Yamauchi Y. and Kanzaki S. Wear. 2000;252:484. doi: 10.1016/S0043-1648(02)00005-4. [DOI] [Google Scholar]
  67. Lim D S, Cho C H. and Park D S. Wear. 2003;255:110. doi: 10.1016/S0043-1648(03)00217-5. [DOI] [Google Scholar]
  68. Park D S, Hahn B D, Bae B C. and Park C. J. Am. Ceram. Soc. 2005;88:383.. doi: 10.1111/j.1551-2916.2005.00074.x. [DOI] [Google Scholar]
  69. Zhang Y, Cheng Y B, Lathabai S. and Hirao K. J. Am. Ceram. Soc. 2005;88:114. [Google Scholar]
  70. Zeng Y P, Yang J F, Kondo N, Ohji T, Kita H. and Kanzaki S. J. Am. Ceram. Soc. 2005;88:1622. doi: 10.1111/j.1551-2916.2005.00242.x. [DOI] [Google Scholar]
  71. Santos C, Strecker K, Neto F P, Baldacim S A, Silva O M M, Silva C R M. Cerâmica. 2005;51:96. [Google Scholar]
  72. Zeng Y P, Kondo N, Hirao K, Kita H, Ohji T. and Kanzaki S. Key Eng. Mater. 2006;317–318:593. [Google Scholar]
  73. Hoffmann M J. and Petzow G. Pure Appl. Chem. 1994;66:1807. doi: 10.1351/pac199466091807. [DOI] [Google Scholar]
  74. Mitomo M, Hirosaki N, Nishimura T. and Xie R J. J. Ceram. Soc. Japan. 2006;114:867. doi: 10.2109/jcersj.114.867. [DOI] [Google Scholar]
  75. Hirao K. J. Ceram. Soc. Japan. 2006;114:665. doi: 10.2109/jcersj.114.665. [DOI] [Google Scholar]
  76. Nishimura T, Guo S Q, Hirosaki N. and Mitomo M. J. Ceram. Soc. Japan. 2006;114:880. doi: 10.2109/jcersj.114.880. [DOI] [Google Scholar]
  77. Niihara K, Bentsen L D, Haselman D P H. and Mazdiyasni K S. J. Am. Ceram. Soc. 1981;64:C117. doi: 10.1111/j.1151-2916.1981.tb10328.x. [DOI] [Google Scholar]
  78. Lienard S Y, Kovar D, Moon R J, Bowman K J. and Halloran J W. J. Mater. Soc. 2000;35:3365. doi: 10.1023/A:1004880901978. [DOI] [Google Scholar]
  79. Wang C A, Huang Y, Zan Q F, Zou L H. and Cai S Y. J. Am. Ceram. Soc. 2002;85:2457. doi: 10.1111/j.1151-2916.2002.tb00480.x. [DOI] [Google Scholar]
  80. Goto Y. and Tsuge A. J. Am. Ceram. Soc. 1993;76:1420. doi: 10.1111/j.1151-2916.1993.tb03920.x. [DOI] [Google Scholar]
  81. Kondo N, Suzuki Y. and Ohji T. J. Am. Ceram. Soc. 2000;83:1816. [Google Scholar]
  82. Zhu X W, Suzuki T S, Uchikoshi T. and Sakka Y. Key Eng. Mater. 2008;368–372:871. [Google Scholar]
  83. Kitayama M, Hirao K, Toriyama M. and Kanzaki S. J. Ceram. Soc. Japan. 1999;107:995. [Google Scholar]
  84. Satet R L. and Hoffmann M J. J. Eur. Ceram. Soc. 2004;24:3437. doi: 10.1016/j.jeurceramsoc.2003.10.034. [DOI] [Google Scholar]
  85. Krämer M, Wittmüss D, Küppers H, Hoffmann M J. and Petzow G. J. Cryst. Growth. 1994;140:157. doi: 10.1016/0022-0248(94)90509-6. [DOI] [Google Scholar]
  86. Hwang C M, Tien T Y. and Chen I W. Sintering' 87. In: Somiya S, , Shimada M, , Yoshimura M, , Watanabe R, , editors. New York: Elsevier; 1988. p. p 1034. [Google Scholar]
  87. Lai K R. and Tien T Y. J. Am. Ceram. Soc. 1993;76:91. doi: 10.1111/j.1151-2916.1993.tb03693.x. [DOI] [Google Scholar]
  88. Kitayama M, Hirao K, Toriyama M. and Kanzaki S. Acta Mater. 1998;46:6541. doi: 10.1016/S1359-6454(98)00290-0. [DOI] [Google Scholar]
  89. Kitayama M, Hirao K, Toriyama M. and Kanzaki S. Acta Mater. 1998;46:6551. doi: 10.1016/S1359-6454(98)00291-2. [DOI] [Google Scholar]
  90. Kitayama M, Hirao K, Toriyama M. and Kanzaki S. Acta Mater. 2000;48:4635. doi: 10.1016/S1359-6454(00)00250-0. [DOI] [Google Scholar]
  91. Wang L L, Tien T Y. and Chen I W. J. Am. Ceram. Soc. 2003;86:1578. [Google Scholar]
  92. Painter G S, Becher P F, Shelton W A, Satet R L. and Hoffmann M J. Phys. Rev. 2004;B 70:144108. doi: 10.1103/PhysRevB.70.144108. [DOI] [Google Scholar]
  93. Becher P F, Painter G S, Shibata N, Satet R L, Hoffmann M J. and Pennycook S J. Mater. Sci. Eng. 2006;A 422:85. doi: 10.1016/j.msea.2006.01.006. [DOI] [Google Scholar]
  94. Dressler W, Kleebe H J, Hoffmann M J, Rühle M. and Petzow G. J. Eur. Ceram. Soc. 1996;16:3. doi: 10.1016/0955-2219(95)00175-1. [DOI] [Google Scholar]
  95. Emoto H. and Mitomo M. J. Eur. Ceram. Soc. 1997;17:797. doi: 10.1016/S0955-2219(96)00139-2. [DOI] [Google Scholar]
  96. Park D S, Lee S Y. and Kim H D. J. Am. Ceram. Soc. 1998;81:1876. [Google Scholar]
  97. Zhu X W, Hayashi H, Zhou Y. and Hirao K. J. Mater. Res. 2004;19:3270. doi: 10.1557/JMR.2004.0416. [DOI] [Google Scholar]
  98. Rhee S H, Lee J D. and Kim D Y. Mater. Lett. 1997;32:115. doi: 10.1016/S0167-577X(97)00018-9. [DOI] [Google Scholar]
  99. Lehner W, Kleebe H J. and Ziegler G. J. Eur. Ceram. Soc. 2006;26:201. doi: 10.1016/j.jeurceramsoc.2004.10.015. [DOI] [Google Scholar]
  100. Zhu X W, Suzuki T S, Uchikoshi T. and Sakka Y. J. Eur. Ceram. Soc. 2008;28:929. doi: 10.1016/j.jeurceramsoc.2007.09.019. [DOI] [Google Scholar]
  101. Willson K S. and Rogers J A. Tech. Proc. Am. Electroplaters Soc. 1964;51:92. [Google Scholar]
  102. Tahashi M, Ishihara M, Sassa K. and Asai S. Mater. Trans. 2003;44:285. doi: 10.2320/matertrans.44.285. [DOI] [Google Scholar]
  103. Li S Q, Sassa K. and Asai S. J. Am. Ceram. Soc. 2004;87:1384. doi: 10.1111/j.1151-2916.2004.tb07743.x. [DOI] [Google Scholar]
  104. Lotgering F K. J. Inorg. Nucl. Chem. 1959;9:113. doi: 10.1016/0022-1902(59)80070-1. [DOI] [Google Scholar]
  105. Jones J L, Slamovich E B. and Bowman K J. J. Mater. Res. 2004;19:3414. doi: 10.1557/JMR.2004.0440. [DOI] [Google Scholar]
  106. Dahms M. and Bunge H J. J. Appl. Crystallogr. 1989;22:439. doi: 10.1107/S0021889889005261. [DOI] [Google Scholar]
  107. Xie R J, Mitomo M, Kim W J. and Kim Y W. J. Am. Ceram. Soc. 2000;84:3147. [Google Scholar]
  108. Bae B C, Park D S, Kim Y W, Kim W J, Han B D, Kim H D. and Park C. J. Am. Ceram. Soc. 2003;86:1008. [Google Scholar]
  109. Kossowsky R. J. Mater. Sci. 1973;8:1603. doi: 10.1007/BF00754896. [DOI] [Google Scholar]
  110. Lee F J. and Bowman K J. J. Am. Ceram. Soc. 1992;75:1748. doi: 10.1111/j.1151-2916.1992.tb07192.x. [DOI] [Google Scholar]
  111. Lee F J. and Bowman K J. J. Am. Ceram. Soc. 1994;77:947. doi: 10.1111/j.1151-2916.1994.tb07251.x. [DOI] [Google Scholar]
  112. Wu X. and Chen I W. J. Am. Ceram. Soc. 1992;75:2733. doi: 10.1111/j.1151-2916.1992.tb05497.x. [DOI] [Google Scholar]
  113. Xie R J, Mitomo M, Kim W J, Kim Y W. and Zhan G D. J. Mater. Res. 2001;16:590. doi: 10.1557/JMR.2001.0085. [DOI] [Google Scholar]
  114. Pipes R B, Mccullough R L. and Taggart D G. Polym. Compos. 1982;3:34. doi: 10.1002/pc.750030107. [DOI] [Google Scholar]
  115. Yasutomi Y, Sakaida Y, Hirosaki N. and Ikuhara Y. J. Ceram. Soc. Japan. 1998;106:980. [Google Scholar]
  116. Nakano H, Nakano H. and Watari K. Adv. Tech. Mat. Mat. Proc. J. 2006;8:67. [Google Scholar]
  117. Vaudin M D, Rupich M W, Jowett M, Riley G N. and Bingert J F. J. Mater. Res. 1998;13:2910. doi: 10.1557/JMR.1998.0398. [DOI] [Google Scholar]
  118. Seabaugh M M, Vaudin M D, Cline J P. and Messing G L. J. Am. Ceram. Soc. 2000;83:2049. [Google Scholar]
  119. Brosnan K H, Messing G L, Meyer R J. and Vaudin M D. J. Am. Ceram. Soc. 2006;89:1965. doi: 10.1111/j.1551-2916.2006.01049.x. [DOI] [Google Scholar]
  120. Tanaka S, Makiya A, Shoji D, Watanabe S, Kato Z, Uchida N. and Uematsu K. J. Ceram. Soc. Japan. 2004;112:276. doi: 10.2109/jcersj.112.276. [DOI] [Google Scholar]
  121. Makiya A, Tanaka S, Shoji D, Ishikawa T, Uchida N. and Uematsu K. J. Eur. Ceram. Soc. 2007;27:3399. doi: 10.1016/j.jeurceramsoc.2006.11.041. [DOI] [Google Scholar]
  122. Seabaugh M M, Kerscht I H. and Messing G L. J. Am. Ceram. Soc. 1997;80:1181. [Google Scholar]
  123. Nuttall K. and Thompson D P. J. Mater. Sci. 1974;9:850. doi: 10.1007/BF00761806. [DOI] [Google Scholar]
  124. Westron J E, Pratt P L, Steele B C H. J. Mater. Sci. 1978;13:2137. doi: 10.1007/BF00541667. [DOI] [Google Scholar]
  125. Zhan G D, Mitomo M, Ikuhara Y. and Sakuma T. J. Am. Ceram. Soc. 2000;83:3179. [Google Scholar]
  126. Santos C, Strecker K, Baldacim S A, De Silva O M M, Da Silva C R M. Ceram. Int. 2004;30:653. doi: 10.1016/j.ceramint.2003.07.011. [DOI] [Google Scholar]
  127. Venkatachari K R. and Raj R. J. Am. Ceram. Soc. 1987;70:514. doi: 10.1111/j.1151-2916.1987.tb05686.x. [DOI] [Google Scholar]
  128. Kondo N, Suzuki Y, Miyajima T. and Ohji T. J. Eur. Ceram. Soc. 2003;23:809. doi: 10.1016/S0955-2219(02)00189-9. [DOI] [Google Scholar]
  129. Kondo N, Asayama M, Suzuki Y. and Ohji T. J. Am. Ceram. Soc. 2003;86:1430. [Google Scholar]
  130. Chen I W. and Xue L A. J. Am. Ceram. Soc. 1990;73:2585. doi: 10.1111/j.1151-2916.1990.tb06734.x. [DOI] [Google Scholar]
  131. Wakai F, Kodama Y, Sakaguchi S, Murayama N, Izaki K. and Niihara K. Nature. 1990;344:421. doi: 10.1038/344421a0. [DOI] [Google Scholar]
  132. Hirao K, Tsuge A, Brito M E. and Kanzaki S. J. Ceram. Soc. Japan. 1993;101:1078. [Google Scholar]
  133. Ramesh P D, Oberacker R. and Hoffmann M J. J. Am. Ceram. Soc. 1999;82:1608. [Google Scholar]
  134. Hirata T, Akiyama K. and Morimoto T. J. Eur. Ceram. Soc. 2000;20:1191. doi: 10.1016/S0955-2219(99)00266-6. [DOI] [Google Scholar]
  135. Kitayama M, Hirao K, Toriyama M. and Kanzaki S. J. Ceram. Soc. Japan. 1999;107:930. [Google Scholar]
  136. Rodriguez M A, Makhonin N S, Escrina J A, Borovinskaya I P, Osendi M I, Barba M F, Iglesias J E. and Moya J S. Adv. Mater. 1995;7:745. doi: 10.1002/adma.19950070815. [DOI] [Google Scholar]
  137. Chen D Y, Zhang B L, Zhuang H R, Li W L. and Xu S Y. Mater. Res. Bull. 2002;37:1481. doi: 10.1016/S0025-5408(02)00775-4. [DOI] [Google Scholar]
  138. Chen D Y, Zhang B L, Zhuang H R, Li W L. and Xu S Y. Mater. Lett. 2002;57:399. doi: 10.1016/S0167-577X(02)00799-1. [DOI] [Google Scholar]
  139. Li W K, Chen D Y, Zhang B L, Zhuang H R. and Li W L. Mater. Lett. 2004;58:2322. doi: 10.1016/S0167-577X(04)00154-5. [DOI] [Google Scholar]
  140. Peng G H, Jiang G J, Zhuang H R, Li W L. and Xu S Y. Mater. Res. Bull. 2005;40:2139. doi: 10.1016/j.materresbull.2005.07.002. [DOI] [Google Scholar]
  141. Goto Y, Ohta H, Komatsu M. and Komeya K. Yogyo-Kyokai-Shi. 1986;94:167. [Google Scholar]
  142. Muscat D, Pugh M D, Drew R A L. Ceramic Transactions vol 19 Advanced Composite Materials. In: Sacks M D, editor. Westerville, OH: The American Ceramic Society; 1991. p. p 137. [Google Scholar]
  143. Zou L H, Park D S, Cho B U, Huang Y. and Kim H D. Mater. Lett. 2004;58:1587. doi: 10.1016/j.matlet.2003.10.031. [DOI] [Google Scholar]
  144. Belmonte M, Miranzo P. and Osendi M I. J. Am. Ceram. Soc. 2007;90:1157. doi: 10.1111/j.1551-2916.2007.01620.x. [DOI] [Google Scholar]
  145. Moreno R. Am. Ceram. Soc. Bull. 1992;71:1521. [Google Scholar]
  146. Moreno R. Am. Ceram. Soc. Bull. 1992;71:1647. [Google Scholar]
  147. Hotza D. and Greil P. Mater. Sci. Eng. 1995;A 202:206. doi: 10.1016/0921-5093(95)09785-6. [DOI] [Google Scholar]
  148. Bitterlich B. and Heinrich J G. J. Eur. Ceram. Soc. 2002;22:2427. doi: 10.1016/S0955-2219(02)00029-8. [DOI] [Google Scholar]
  149. Bitterlich B. and Heinrich J G. J. Am. Ceram. Soc. 2005;88:2713. doi: 10.1111/j.1551-2916.2005.00512.x. [DOI] [Google Scholar]
  150. Zhang J X, Ye F, Jiang D L. and Iwasa M. Ceram. Int. 2006;32:277. doi: 10.1016/j.ceramint.2005.03.003. [DOI] [Google Scholar]
  151. Wu M X. and Messing G L. J. Am. Ceram. Soc. 1994;77:2586. doi: 10.1111/j.1151-2916.1994.tb04646.x. [DOI] [Google Scholar]
  152. Kim H J, Krane M J M, Trumble K P. and Bowman K J. J. Am. Ceram. Soc. 2006;89:2769. [Google Scholar]
  153. Park D S. and Kim C W. J. Mater. Sci. 1999;34:5827. doi: 10.1023/A:1004770520830. [DOI] [Google Scholar]
  154. Sakka Y. and Suzuki T S. J. Ceram. Soc. Japan. 2005;113:26. doi: 10.2109/jcersj.113.26. [DOI] [Google Scholar]
  155. Suzuki T S, Sakka Y. and Kitazawa K. Adv. Eng. Mater. 2001;3:490. doi: 10.1002/1527-2648(200107)3:7&#x0003c;490::AID-ADEM490&#x0003e;3.0.CO;2-O. [DOI] [Google Scholar]
  156. Uchikoshi T, Suzuki T S, Okuyama H, Sakka Y. and Nicholson P S. J. Eur. Ceram. Soc. 2004;24:225. doi: 10.1016/S0955-2219(03)00242-5. [DOI] [Google Scholar]
  157. Zimmerman M H, Faber K T. and Fuller E R. J. Am. Ceram. Soc. 1997;80:2725. [Google Scholar]
  158. Kimura T. Polym. J. 2003;35:823. doi: 10.1295/polymj.35.823. [DOI] [Google Scholar]
  159. Kimura T, Yoshino M, Yamane T, Yamato M. and Tobita M. Langmuir. 2004;20:5669. doi: 10.1021/la049347w. [DOI] [PubMed] [Google Scholar]
  160. Suzuki T S. and Sakka Y. Japan J. Appl. Phys. 2002;41:L1272. doi: 10.1143/JJAP.41.L1272. [DOI] [Google Scholar]
  161. Suzuki T S. and Sakka Y. Chem. Lett. 2002;31:1204. doi: 10.1246/cl.2002.1204. [DOI] [Google Scholar]
  162. Sakka Y, Suzuki T S, Tanabe N, Asai S. and Kitazawa K. Japan. J. Appl. Phys. 2002;41:L1416. doi: 10.1143/JJAP.41.L1416. [DOI] [Google Scholar]
  163. Zhu X W, Suzuki T S, Uchikoshi T, Nishimura T. and Sakka Y. J. Ceram. Soc. Japan. 2006;114:979. doi: 10.2109/jcersj.114.979. [DOI] [Google Scholar]
  164. Li S Q, Sassa K, Iwai K. and Asai S. Mater. Trans. 2004;45:3124. doi: 10.2320/matertrans.45.3124. [DOI] [Google Scholar]
  165. Li S Q, Sassa K. and Asai S. Mater. Lett. 2005;59:153. doi: 10.1016/j.matlet.2004.07.043. [DOI] [Google Scholar]
  166. Hackley V A. J. Am. Ceram. Soc. 1997;80:2315. [Google Scholar]
  167. Hackley V A. J. Am. Ceram. Soc. 1998;81:2421. [Google Scholar]
  168. Laarz E. and Bergström L. J. Eur. Ceram. Soc. 2000;20:431. doi: 10.1016/S0955-2219(99)00187-9. [DOI] [Google Scholar]
  169. Zhang J X, Ye F, Jiang D L. and Iwasa M. Colloids Surf. 2005;A 259:117. doi: 10.1016/j.colsurfa.2005.02.006. [DOI] [Google Scholar]
  170. Zhu X W, Uchikoshi T, Suzuki T S. and Sakka Y. J. Am. Ceram. Soc. 2007;90:797. doi: 10.1111/j.1551-2916.2007.01491.x. [DOI] [Google Scholar]
  171. Zhu X W, Uchikoshi T. and Sakka Y. Mater. Sci. Forum. 2007;534–536:1009. [Google Scholar]
  172. Zhu X W, Suzuki T S, Uchikoshi T. and Sakka Y. J. Am. Ceram. Soc. 2008;91:620. doi: 10.1111/j.1551-2916.2007.02125.x. [DOI] [Google Scholar]
  173. Park D S. and Kim C W. J. Am. Ceram. Soc. 1999;82:780. [Google Scholar]
  174. Ozer I O, Suvaci E, Karademir B, Missiaen J M, Carry C P. and Bouvard D. J. Am. Ceram. Soc. 2006;89:1972. doi: 10.1111/j.1551-2916.2006.01039.x. [DOI] [Google Scholar]
  175. Hirao K, Imamura H, Teshima H, Brito M E, Toriyama M. and Kanzaki S. Sintering Science and Technology. In: German R M, , Messing G L, Cornwall R G, , editors. Pennsylvania: The Pennsylvania State University; 2000. p. p 343. [Google Scholar]
  176. Park D S, Roh T W, Hockey B J, Takigawa Y, Han B D, Kim H D. and Yasutomi Y. J. Eur. Ceram. Soc. 2003;23:555. doi: 10.1016/S0955-2219(02)00145-0. [DOI] [Google Scholar]
  177. Chakraborty D. and Mukerji J. J. Mater. Sci. 1980;15:3051. doi: 10.1007/BF00550375. [DOI] [Google Scholar]
  178. Chakraborty D. and Mukerji J. Mater. Res. Bull. 1982;17:843. doi: 10.1016/0025-5408(82)90003-4. [DOI] [Google Scholar]
  179. Dusza J, Eschner T. and Rundgren K. J. Mater. Sci. Lett. 1997;16:1664. doi: 10.1023/A:1018521930114. [DOI] [Google Scholar]
  180. Hay J C, Sun E Y, Pharr G M, Becher P F. and Alexander K B. J. Am. Ceram. Soc. 1998;81:2661. [Google Scholar]
  181. Ekström T, Käll P O, Nygren M. and Olsson P O. J. Mater. Sci. 1989;24:1853. doi: 10.1007/BF01105715. [DOI] [Google Scholar]
  182. Tanaka I, Nasu S, Adachi H, Miyamoto Y. and Niihara K. Acta Metall. Mater. 1992;40:1995. doi: 10.1016/0956-7151(92)90185-H. [DOI] [Google Scholar]
  183. Kondo N, Inagaki Y, Suzuki Y. and Ohji T. Mater. Sci. Eng. 2002;A 335:26. doi: 10.1016/S0921-5093(01)01907-4. [DOI] [Google Scholar]
  184. Shigegaki Y, Brito M E, Hirao K, Toriyama M. and Kanzaki S. J. Am. Ceram. Soc. 1997;80:495. [Google Scholar]
  185. Ganesh V V. and Chawla N. Mater. Sci. Eng. 2005;A 391:342. doi: 10.1016/j.msea.2004.09.017. [DOI] [Google Scholar]
  186. Ohji T, Kondo N, Suzuki Y. and Hirao K. Mater. Lett. 1999;40:5. doi: 10.1016/S0167-577X(99)00039-7. [DOI] [Google Scholar]
  187. Griffith A A. Phil. Trans. R. Soc. Lond. 1920;A 221:163. doi: 10.1098/rsta.1921.0006. [DOI] [Google Scholar]
  188. Ohji T, Hirao K. and Kanzaki S. J. Am. Ceram. Soc. 1995;78:3125. doi: 10.1111/j.1151-2916.1995.tb09095.x. [DOI] [Google Scholar]
  189. Pezzotti G, Ichimaru H, Ferroni L P, Hirao K. and Sbaizero O. J. Am. Ceram. Soc. 2001;84:1785. [Google Scholar]
  190. Meléndez-Martínezl J J, Domínguez-Rodríguez A. Prog. Mater. Sci. 2004;49:19. doi: 10.1016/S0079-6425(03)00020-3. [DOI] [Google Scholar]
  191. Kong H. and Ashby M F. Acta Metall. Mater. 1992;40:2907. doi: 10.1016/0956-7151(92)90455-N. [DOI] [Google Scholar]
  192. Lee S K, Wuttiphan S. and Lawn B R. J. Am. Ceram. Soc. 1997;80:2367. [Google Scholar]
  193. Lee S K, Lee K S, Lawn B R. and Kim D K. J. Am. Ceram. Soc. 1998;81:2061. [Google Scholar]
  194. Haggerty J S. and Lightfoot A. Ceram. Eng. Sci. Proc. 1995;16:475. doi: 10.1002/9780470314715.ch52. [DOI] [Google Scholar]
  195. Hirao K, Watari K, Hayashi H. and Kitayama M. MRS Bull. 2001;26:451. [Google Scholar]
  196. Watari K. J. Ceram. Soc. Japan. 2001;109:S7. [Google Scholar]
  197. Watari K, Hirao K, Brito M E, Toriyama M. and Ishizaki K. Adv. Tech. Mat. Mat. Proc. J. 2005;7:191. [Google Scholar]
  198. Hirosaki N, Ogata S, Kocer C, Kitagawa H. and Nakamura Y. Phys. Rev. 2002;B 65:134110. doi: 10.1103/PhysRevB.65.134110. [DOI] [Google Scholar]
  199. Bruls R J, Hintzen H T. and Metselaar R. J. Eur. Ceram. Soc. 2005;25:767. doi: 10.1016/j.jeurceramsoc.2004.05.003. [DOI] [Google Scholar]
  200. Li B C, Pottier L, Roger J P, Fournier D, Watari K. and Hirao K. J. Eur. Ceram. Soc. 1999;19:1631. doi: 10.1016/S0955-2219(98)00258-1. [DOI] [Google Scholar]
  201. Yokota H. and Ibukiyama M. J. Eur. Ceram. Soc. 2003;23:1183. doi: 10.1016/S0955-2219(02)00292-3. [DOI] [Google Scholar]
  202. Kingery W D, Bowen H K. and Uhlmann D R. 2nd edn. New York: Wiley; 1976. Introduction to Ceramics; p. p 626. [Google Scholar]
  203. Inagaki Y, Ohji T, Kanzaki S. and Shigegaki Y. J. Am. Ceram. Soc. 2000;83:1807. [Google Scholar]
  204. Inagaki Y, Shigegaki Y, Ando M. and Ohji T. J. Eur. Ceram. Soc. 2004;24:197. doi: 10.1016/S0955-2219(03)00603-4. [DOI] [Google Scholar]
  205. Kondo N, Inagaki Y, Suzuki Y. and Ohji T. J. Ceram. Soc. Japan. 2004;112:316. doi: 10.2109/jcersj.112.316. [DOI] [Google Scholar]
  206. Deng Z Y, Inagaki Y, She J H, Tanaka Y, Liu Y F, Sakamoto M. and Ohji T. J. Am. Ceram. Soc. 2005;88:462. doi: 10.1111/j.1551-2916.2005.00067.x. [DOI] [Google Scholar]
  207. She J H, Yang J F, Jayaseelan D D, Kondo N, Ohji T, Kanzaki S. and Shigegaki Y. J. Am. Ceram. Soc. 2003;86:738. [Google Scholar]
  208. Rice R W. J. Mater. Sci. 1996;31:102. doi: 10.1007/BF00355133. [DOI] [Google Scholar]
  209. Becher P F, Lewis D III, Carman K R. and Gonzalez A C. Am. Ceram. Soc. Bull. 1980;59:542. [Google Scholar]
  210. Wang H, Cheng Y B, Muddle B C, Gao L. and Yen T S. J. Mater. Sci. Lett. 1996;15:1447. doi: 10.1007/BF00275302. [DOI] [Google Scholar]
  211. Wang P L, Zhang C, Sun W Y. and Yan D Y. J. Eur. Ceram. Soc. 1999;19:553. doi: 10.1016/S0955-2219(98)00249-0. [DOI] [Google Scholar]
  212. Yu Z B, Thompson D P. and Bhatti A R. J. Eur. Ceram. Soc. 2001;21:2423. doi: 10.1016/S0955-2219(01)00197-2. [DOI] [Google Scholar]
  213. Carman A, Pereloma E. and Cheng Y B. J. Eur. Ceram. Soc. 2006;26:1337. doi: 10.1016/j.jeurceramsoc.2005.02.011. [DOI] [Google Scholar]
  214. Carman A, Pereloma E. and Cheng Y B. J. Am. Ceram. Soc. 2006;89:478. doi: 10.1111/j.1551-2916.2005.00768.x. [DOI] [Google Scholar]
  215. Reimanis I E, Suematsu H, Petrovic J J. and Mitchell T E. J. Am. Ceram. Soc. 1996;79:2065. doi: 10.1111/j.1151-2916.1996.tb08938.x. [DOI] [Google Scholar]
  216. Zenotchkine M, Shuba R, Kim J S. and Chen I W. J. Am. Ceram. Soc. 2001;84:1651. doi: 10.1111/j.1151-2916.2001.tb00760.x. [DOI] [Google Scholar]
  217. Fu R L, Chen K X, Xu X, Ferreira J M F. Mater. Lett. 2004;58:1956. doi: 10.1016/j.matlet.2003.12.014. [DOI] [Google Scholar]
  218. Zhu X W, Suzuki T S, Uchikoshi T. and Sakka Y. J. Ceram. Soc. Japan. 2007;115:701. doi: 10.2109/jcersj2.115.701. [DOI] [Google Scholar]

Articles from Science and Technology of Advanced Materials are provided here courtesy of National Institute for Materials Science and Taylor & Francis

RESOURCES