Sunteți pe pagina 1din 10

Journal

J. Am. Ceram. Soc., 80 [8] 191928 (1997)

Evaluated Material Properties for a Sintered -Alumina


Ronald G. Munro*
Ceramics Division, National Institute of Standards and Technology, Gaithersburg, Maryland 20899

Results of a data evaluation exercise are presented for a particular specification of sintered -alumina (mass fraction of Al2O3, 0.995; relative density ( / theoretical), 0.98; and nominal grain size, 5 m). A comprehensive set of material property data is established based on published physical, mechanical, and thermal properties of alumina specimens that conform to the constraints of the material specification. The criteria imposed on the properties are that the values should be derived from independent experimental studies, that the values for physically related properties should be mutually self-consistent, and that the sets of values should be compatible with established material property relations. The properties assessed in this manner include crystallography, thermal expansion, density, sound velocity, elastic modulus, shear modulus, Poissons ratio, bulk modulus, compressive strength, flexural strength, Weibull characteristic strength, Weibull modulus, tensile strength, hardness, fracture toughness, creep rate, creep rate stress exponent, creep activation energy, friction coefficient, wear coefficient, melting point, specific heat, thermal conductivity, and thermal diffusivity. I. Introduction

LUMINA is one of the more widely used and studied advanced ceramic materials. The relative abundance and low cost of the material resource is advantageous for commercial applications, while the availability of the material in highly purified grades makes it well suited to fundamental studies in materials research. Among its desirable features, alumina shares with other advanced ceramics the characteristics of high-temperature stability and the retention of strength at high temperatures. These characteristics often are cited as the most important attributes of structural ceramics. Ironically, property

M. P. Harmercontributing editor

Manuscript No. 191842. Received May 6, 1996; approved April 28, 1997. *Member, American Ceramic Society.

data for alumina at elevated temperature often are not readily accessible in the literature, and the data that are available usually pertain to disparate material specifications. This paper addresses that situation for -alumina as part of a broader study on the issues of data evaluation and the need for temperaturedependent property data for advanced ceramics. The widespread use of alumina has prompted several previous property reviews and data compilations to provide general surveys of the properties of all grades of alumina.17 Because many of the important properties and characteristics of alumina vary considerably with composition, density, and grain size, those reviews were focused on the immense task of determining ranges of values that have been observed for the various properties. Those substantial efforts have provided an invaluable guide to the understanding and application of sintered alumina. The present work seeks to build upon those efforts in two respects while narrowing the application focus. First, there have been some rather significant results published since the previous reviews were done, and those newer results are incorporated in the present assessment. Second, the present work is directed toward deriving a single set of data from diverse independent efforts, rather than describing the distribution of the values across the numerous studies. The objective, then, is to derive a coherent, self-consistent, and comprehensive set of property values for a single specification of alumina. The constraint on the material specification is significant because many properties critical to the application of ceramics (particularly flexural strength, tensile strength, hardness, fracture toughness, and creep) are known to be sensitive to variations in the microstructure of the material.8 In the present work, that constraint is imposed by means of some relatively simple criteria pertaining to the composition, densification, and grain size of the material. These criteria are discussed in the following section. Evaluated data have an increasingly important role to play in both manufacturing and materials science applications. The current trends toward concurrent engineering practices9 in manufacturing and the growing use of the evolving Standard for the Exchange of Product Data (STEP) for communicating product specification data10 require a wide range of properties applicable to a wide range of temperature. These applications

1919

1920

Journal of the American Ceramic SocietyMunro

Vol. 80, No. 8

can benefit from the comprehensive scope of properties and the explicit interpolation formulas for temperature dependence that result from the present approach to data evaluation. Efforts to establish and understand correlations of properties and performance results also should benefit particularly from the selfconsistent relations and the coherent nature of the collection of properties, i.e., that the entire set of properties pertains to a single material specification. II. Material Specification

to be processed in the context of known material property relations. For convenience, an analytic interpolation formula is given for the temperature dependence of each property. Uncertainties are assigned to each interpolation formula according to estimates of the combined standard uncertainties appropriate to the various sets of data.16 These estimates are based on standard deviations or standard errors of the fits, as appropriate. (1) Crystallography, Density, and Thermal Expansion The crystal structure and the temperature dependences of the lattice parameters have a basic importance to the properties of polycrystalline alumina. It is appropriate, therefore, to begin this study with an examination of the lattice parameters15,1721 and their anisotropic thermal expansions.22,23 In practice, the linear coefficient of thermal expansion, CTE L1( L / T), is most commonly represented as a cumulative expansion coefficient, = L1 0 L T L0 T T0 (1)

The goal of the present work is not to review the properties of all aluminas, but rather to establish a definitive, comprehensive set of data for one particular material specification. The attainment of this goal is dependent on several factors, but the two most basic are the availability of the supporting data and the independence of the various experimental studies. The first factor, the availability of the data, strongly influences the specification of the material; it would be unproductive to specify a material for which there are few data. The second factor, the independence of the studies, is essential to establish or verify the reproducibility of the results. For ceramics, the insistence on independent studies invariably results in no two studies being conducted on exactly the same material. Initially, the possibility of restricting the study to one specific commercial material was considered, but that restriction in the case of alumina resulted in an insufficient supply of independent data. Although much data could be found for a few properties, a comprehensive range of properties could not be found for any one specific commercial alumina. Consequently, a more general approach to material specification was used. It was found that, if three quantities (purity, density, and mean grain size) were constrained, sufficient results for alumina could be found in the literature to establish a comprehensive range of properties, and, at the same time, the independently measured values for individual properties were reasonably consistent and reproducible. In general, stoichiometric Al2O3 has a molecular weight of 101.96. The most prevalent and stable single phase of the material is denoted as -Al2O3 and occurs in the corundum crystal structure (space group R3c) containing two formula units per unit cell.4 It is common practice, however, to represent the structure using a larger hexagonal cell containing six formula units, and that practice is followed here.11 This structure (see cover illustration) consists of planes of close-packed oxygen ions in the A-B-A-B sequence interleaved with planes of aluminum ions in an a-b-c-a-b-c sequence. In each aluminum plane, the aluminum ions occupy only two-thirds of the available octahedral sites, thereby maintaining charge neutrality (four Al3+ for every six O2). A hexagonal crystallographic cell is formed from the repeating sequence A-a-B-b-A-c-B-a-A-bB-c. The melting point1215 of this phase is 2050 4C. Polycrystalline sintered alumina ceramics are typically formed by sintering compacted powders at temperatures on the order of 1700C. These ceramic alumina materials are usually classified according to the purity of the phase composition, the mass density, and the mean grain size of the sintered material. Commonly, additives also are specified to control the rates of densification and grain growth; for example, a small amount of MgO may be added to control the grain size during sintering. Trace amounts of SiO2, CaO, Fe2O3, and Na2O also may occur in the sintered specimen. The present work is focused on high-purity sintered -alumina in which the mass fraction of the alumina phase is 99.5% or higher. The material is nearly fully densified with a mass density that is at least 98% of the theoretical density, and the material has a nominal grain size of 5 m. III. Properties and Characteristics

where L is either of the lattice parameters of the hexagonal unit cell, a or c; T is the temperature, and the subscript, 0, denotes a reference state, usually 0C or room temperature. The thermal expansion of the crystal structure can be determined either by diffraction methods, which measure the variation of the atomic-scale dimensions of the crystal structure, or by highresolution dilatometry methods that detect changes in the macroscopic dimensions of the material specimen. The two measures of expansion can have a small difference because of the presence of defects in the macroscopic specimen. In the present work, data from both types of methods are considered to determine an optimized, self-consistent representation of the lattice parameters and the coefficient of thermal expansion. To facilitate the optimization process, empirical interpolation formulas are used.
L

T = AL + BLT CL exp DLT


L

(2a) (2b)

L T = L0 1 +

TT

where AL, BL, CL, DL, and L0 are adjustable parameters, and the reference temperature is taken as T0 0C. Fitting these expressions to single-crystal data yields the results in Figs. 1 and 2 and Table I. The solid curves in Figs. 1 and 2 were produced using Eqs. (2) and the parameters in Table I. Independent thermal expansion data for polycrystalline specimens2,24,25 are shown in Fig. (2b), in which the solid curve is the mean coef-

For each property or characteristic, the experimental values obtained from published reports are viewed as raw data that are

Fig. 1.

Lattice parameters of -Al2O3 referred to the hexagonal cell.

August 1997

Evaluated Material Properties for a Sintered -Alumina

1921

(2) Elastic Properties The elastic properties of polycrystalline materials are commonly considered to be isotropic. Under such conditions, the elastic properties have the well-known relations, G= B= = E 2 1+ E 3 12 E 1 2G (4a) (4b) (4c)

Where E is the elastic modulus, G is the shear modulus, B is the bulk modulus, and is Poissons ratio. These quantities are most commonly determined by ultrasonic methods, according to which the longitudinal velocity (VL) and the shear velocity (VS) are related to the elastic properties and the bulk density ( ): G = V2 S B= = 4 V2 V2 L 3 S 1 1 1 2 VL VS 2 1 (5a) (5b) (5c)

Fig. 2. Thermal expansion of sintered polycrystalline.

-Al2O3: (a) single crystal and (b)

Experimental values of the elastic properties of alumina26,27 determined by ultrasonic methods28 are shown in Fig. 3 and can be represented by the interpolation formulas, E GPa = 417 0.0525T 7% G GPa = 169 0.0229T 6% (6a) (6b)

Table I. Parameters of Eqs. (2) Obtained for the Temperature Interval 0C T 1800C*
Value Parameter a-axis c-axis

A[106 K1] B[1010 K2] C[106 K1] D[103 K1] L0 ()

7.419 6.43 3.211 2.59 4.7602

8.026 8.17 3.279 2.91 12.9898

for 20C T 1400C. For the data sets that are currently available, the values for the elastic properties of the sintered product are indistinguishable from those obtained for hotpressed specimens, and the values derived from the singlecrystal elasticity tensor also are well within the uncertainty limits expressed by Eqs. (6).29 (3) Strength and Related Properties One of the more compelling reasons for the interest in structural ceramics is their retention of strength at high temperatures. Under compression,2,25,30 as shown in Fig. 4, the sintered

*Values obtained from Eqs. (2) using these parameters have combined standard uncertainties of 4% for values of the thermal expansion and 0.06% for values of the lattice parameters.

ficient of thermal expansion, m (2 a + c)/3, evaluated using Eq. (2 a) and the parameters in Table I. The results for the lattice parameters can be used with the molecular weight to determine the theoretical density of Al2O3 from the relation = Mz NA V (3a)

where M is the molar mass, z is the number of formula units per unit cell, NA is Avogadros number, and V is the volume of the unit cell. For -Al2O3 referred to the hexagonal crystallographic cell, M 101.96 g/mol, z 6, and V (31/2/2)a2c. With these values, the theoretical density can be represented conveniently by g cm3 = 3.9853 7.158 105T 3.035 108T 2 (3b) + 7.232 1012T 3 0.05% for 20C T 1800C, which evaluates to 3.984 0.002 g/cm3 at T 20C.
Fig. 3. Elastic and shear moduli of high-purity, high-density sintered -Al2O3.

1922

Journal of the American Ceramic SocietyMunro

Vol. 80, No. 8

ln ln

1 = m ln 1 PF

m ln

+ ln

V V0

(8b)

In using Eq. (8b), PF often is estimated from the rank order of a series of measurements; i.e., if the measured strengths are ordered from smallest to largest, then the failure probability of the ith value of N results is estimated as PF(i) (i 0.5)/N. In more recent years, ASTM Committee C-28 has recommended avoiding the use of this failure probability estimate and the linear regression method by using a maximum likelihood estimation method.33 In either case, the strength at any given failure probability ( (PF)) evaluated from Eq. (8) depends on the effective fracture volume of the specimen. If the flaw populations are assumed to be the same for two different, but geometrically similar, specimen sets, a and b, with volumes Va and Vb, respectively, then, from Eq. (8a), the size effect can be expressed as8,34
a b

PF Vb = PF Va

1 m

(9)

Fig. 4. Compressive strength of high-purity, high-density sintered -Al2O3.

-Al2O3 materials considered here withstand >3000 MPa at room temperature and have a temperature dependence given by CS GPa = 3.1 0.0035T + 1.1 106T 2 15% (7)

for 20C T 1400C, where CS is the compressive strength and T is the temperature. More importantly for most applications of structural ceramics is the strength of the material in flexure or tension. In these modes of stress, the strength of a brittle material is limited by the distribution of flaws in the material specimen.31 Any flaw in the material can serve as an origin of failure. Because the details of the individual flaws vary from specimen to specimen, tests of large numbers of specimens are required to characterize the strength behavior statistically.32 The most commonly used statistic is the mean value of the strength, termed the flexural strength or the tensile strength for materials in flexure or tension, respectively. When the mean value is used as the measure of strength, the standard deviation of the observed values is used as a measure of the spread of values. A more insightful description of the strength values is provided by the use of the two-parameter Weibull analysis. In the Weibull model, if PF is the probability of failure at a strength , then PF = 1 exp V V0
m 0

Consequently, results from bend tests vary with the details of the test configuration. Results also may vary with the stressing rate. The present data are intended to be appropriate for fourpoint bend tests with 14-point loading (ASTM C 1211, specimen configuration B) on a span length of 40 mm and a loading rate of 0.5 mm/min. A search of the literature for this work found no high-temperature data meeting the material and measurement conditions considered here. In lieu of such data, flexural strength results from one round-robin study8 at room temperature and two high-temperature studies35,36 on hot-pressed and hot isostatically pressed aluminas with purity, density, and grain size similar to the prescription of Section II (see Table II) are shown in Fig. 5. Also shown are handbook values over the same temperature range for a generic polycrystalline alumina.25 The agreement among these data sets, although somewhat fortuitous, suggests that these values should be a fair representation of the flexural strength in the current context. For temperatures up to 1100C, the flexural strength in Fig. 5 has a value of 380 MPa with a standard deviation of 60 MPa. Above 1100C, the flexural strength decreases linearly to 120 MPa at 1500C. The interpolation formula used in Fig. 5 is FS MPa = 380.5 1.37 105 1 + 1.76 105 exp 0.0039T (10a)

(8a)

where V is the specimen volume, V0 is a volume scale parameter, and 0 and m are characteristics of the flaw distribution known, respectively, as the characteristic strength and the Weibull modulus. Larger values of 0 imply greater strength, and larger values of m imply more reliably reproducible specimens in the sense of a narrower distribution of values. It is quite common to find Eq. (8a) rearranged into the form

for 20C T 1500C. The corresponding results for the Weibull analysis of the data36 for the hot isostatically pressed material are shown in Fig. 6. Because of the limited amount of data and the considerable scatter in the values, no trend with respect to temperature can be discerned reliably for the Weibull modulus. The Weibull parameters are best expressed as m = 11 4
0

(10b) (10c) T

MPa = 16.32 + 1.09FS 20%

where FS is given by Eq. (10a) and where 20C 1500C.

Table II.

Comparison of Material Characteristics of Specimens Used in the Estimate of Flexural Strength*


Sintered Hot pressed Hot isostatistically pressed

Processing condition

Purity (mass fraction %) Relative density (%) Grain size range ( m) Specimen dimensions (mm) Outer/inner span (mm/mm) Loading rate (mm/min) Reference
*Equation 10.

99.9 >99 16 3 4 50 40/20 0.5 8

99.9 >98 12 3.81 6.35 44.45 38.1/12.7 0.15 33

99.9 >99 26 3 4 45 40/20 0.5, 1.5 34

August 1997

Evaluated Material Properties for a Sintered -Alumina

1923

search of the literature found no paper in which material and measurement details were reported with the tensile strength. One paper2 gave the tensile strength for an alumina with the Al2O3 mass fraction of 99.9%, relative density of >99%, and a grain size range of 16 m, but no measurement details were given. Two other papers25,40 gave results with neither material nor measurement details. The data from these three sources are shown in Fig. 7. Based on these data, the tensile strength can be expected to have a value of 270 MPa for temperatures up to 1100C, whereafter the tensile strength decreases more sharply than the flexural strength, decreasing to a value of <50 MPa at 1300C. The interpolation formula for Fig. 7 is TS MPa =267 256 1 + 5.8 109 exp 0.018T
1 2

10% (11)

Fig. 5. Flexural strength of high-purity, high-density sintered -Al2O3.

for 20C T 1500C. The results expressed by Eqs. (10) and (11) were obtained from independent sets of data. Given also the scarcity of the data, it is desirable to verify the consistency of these two important measures of the strength of the material. If it is assumed that the flaw systems are the same for the two sets of specimens and that the specimen volumes are the same, then the Weibull analysis can be used to establish a relation between the mean flexural strength and the mean tensile strength. It is shown in Panel A that the ratio of these strengths should be [4(m + 1)2/(m + 2)]1/m 1.4 when m 11. Using 380 MPa for the mean flexural strength below 1000C in Fig. 5 and 270 MPa for the mean tensile strength in Fig. 7, the ratio is 380/270 1.4, in agreement with the Weibull analysis. Two other characteristics, hardness and fracture toughness, are commonly discussed in the context of the strength of structural ceramics. Hardness41 is intended to be a measure of the resistance to plastic deformation, which may include effects such as material displacement and fracture. Fracture toughness42 is intended to be a measure of the resistance to the extension of cracks. Quantitative results for these characteristics depend on the procedural details of how the deformation is produced or the cracks are caused to propagate. In the present work, attention is restricted to the Vickers indentation method for hardness measurements and the indentation strength method for fracture toughness measurements. The hardness obtained by the Vickers method43 is HV = 1.8544 P d2 (12)

Fig. 6. Weibull parameters of high-purity, high-density sintered -Al2O3: (a) modulus (m) and (b) characteristic strength ( 0).

The flexural strength exhibits steady or monotonically decreasing values with increasing temperature. When a significant volume fraction of a polycrystalline ceramic is occupied by a glassy grain-boundary phase, the material is susceptible to the onset of a brittle-to-ductile transition at elevated temperature.37,38 The transition appears to be a softening of the glassy phase, which subsequently permits a degree of crack blunting. As a result, it is characteristic of this transition that there is a sudden increase in the measured strength of the material and its fracture toughness.39 The absence of such an increase in either property in the present case may be attributable to the high mass fraction (99.5%) of the alumina grains in the material, which may preclude the possibility of a substantial glassy phase. Data on the tensile strength of alumina are very scarce. The

Fig. 7. Tensile strength of high-purity, high-density sintered -Al2O3.

1924

Journal of the American Ceramic SocietyMunro

Vol. 80, No. 8

Panel A.

Flexural versus Tensile Strength


t

If a single Weibull distribution is assumed to characterize the flaw system in a test specimen, then a relation between two different test configurations can be developed.64 For this purpose, it is preferable to consider Eq. (8a) in a form that explicitly accounts for the nonuniform stress that occurs in bend tests. Using the substitution v V/ V0 for convenience, PF = 1 exp
m 0

=1 m+2 4 m+1
2

(A4)

For a four-point bend test with quarter-point loading,


4-point,1 4

(A5)

dv

(A1)

With these relations, the mean fracture strength can be calculated using the observation that the number of specimens that fail at stress is the difference between the number of specimens surviving at stress /2 and + /2. Thus,
mean

Within the nonuniform stress distribution, the maximum stress ( max) that occurs in the specimen is the stress of primary relevance to the test result. Dividing and 0 by max, Eq. (A1) can be rewritten as PF = 1 exp v where = 1 v
m max m max 0

dPF d

(A6)

from which it can be found that


4-point,1 4,mean t,mean

(A2)

4 m+1 m+2

1 m

(A7)

dv

(A3)

The value of depends on the specific test configuration. For a tensile test with uniform tension, max, and, hence,

assuming that the specimen volumes are the same. Using m 11 from Eq. (10b), we would expect 4-point,1/4,mean/ 1.4 for the present material. For temperat,mean tures <1000C, we find in Figs. 5 and 7, respectively, 4-point,1/4,mean 380 MPa and t,mean 270 MPa, which results in a ratio of 1.4, in agreement with Eq. (A7).

where P is the applied load and d is the mean diagonal length of the irreversible impression produced by the indentor. For most ceramics, the value of HV varies with the load, at least for lower loads. Empirically, the size of the indentation impression is often related to the load by the expression P d (sometimes called the Meyer law),44,45 where is a constant and is a parameter (sometimes called the indentation size effect index for brittle materials).46 Substituting this expression into Eq. (12) shows that HV is independent of load only for the exceptional case that 2. For most structural ceramics, < 2 under low loads. Consequently, it is prudent to consider the value of the hardness with due reference to the load used to measure it. Two applicable papers47,48 were found according to which the hardness can be represented as in Fig. 8 when a load of 1

kg is used. Although there is a significant amount of scatter in the data, the general trends can be represented by the interpolation formula HV GPa = 15.5 exp 0.0012T 15% (13)

where HV is hardness and 20C T 1000C. Fracture toughness is a more complicated characteristic that depends quite significantly on the microstructure of the specimen. Alumina appears to be one of the materials known to exhibit R-curve behavior; i.e., the toughness increases with increasing crack length. Only one paper discussing both the temperature and crack-length dependences was found.49 In that work, a Vickers indentor was used with loads from 12 to 300 N on hot-pressed specimens that measured 2.5 mm 4 mm 25 mm. The strength tests were conducted in four-point bending with inner and outer spans of 10 mm and 20 mm, respectively, and with a constant crosshead speed of 0.25 mm/min. The fracture toughness, KI, was evaluated as KI = Y f 1
2

(14)

where the geometric factor Y has been determined to be 1.02, is the crack length. f is the fracture stress, and The results for the fracture toughness (Fig. 9) show a linear dependence on the square root of the crack length and a nonlinear dependence on temperature. Treating these dependences simultaneously, a suitable interpolation formula is found to be KI MPa m1
2

= 2775

exp 0.0000476T + 0.084 T + 1323

1 2

15% (15)

where 25C T 1300C and 1/2 is in the range 721 m1/2. The solid curves in Fig. 9 are produced from Eq. (15). (4) Creep Characteristics The lifetime of a material component operating at high temperature may be limited by the long-term creep deformation characteristics of the material.50,51 Ceramics generally exhibit a temperature and stress-dependent steady-state creep rate of the form

Fig. 8. Vickers hardness of high-purity, high-density sintered -Al2O3.

August 1997

Evaluated Material Properties for a Sintered -Alumina

1925

FIG. 9. Fracture toughness of high-purity, high-density sintered -Al2O3. Fig. 11. Friction and wear transition diagram for high-purity, highdensity sintered -Al2O3 (COF is coefficient of friction and Kw is dimensionless wear coefficient).

d =A dt

exp Eact RT

(16)

where d /dt is the creep rate, is the applied stress, n is a parameter known as the creep stress exponent, Eact is the activation energy, R is the fundamental molar gas constant, and A is a scale parameter. The parameters A, n, and Eact are determined from least-squares fits to the experimental data. In the present case (see Fig. 10) the creep rate under flexural stress conditions5254 spans 6 orders of magnitude on the temperature range from 1200 to 1800C for applied stresses in the interval from 100 to 200 MPa. For Fig. 10, the solid interpolation curves are given by Eq. (16) using the parameter values A 3.6 1011 s1, n 1.08, and Eact 323 kJ/mol. These curves have a combined standard uncertainty for log10(d /dt) of 8%. (5) Tribological Characteristics In many applications of advanced ceramics, the durability of the material under sliding conditions is an important design consideration, and the sliding friction and wear characteristics

are required for estimates of lifetime, heat generation, and energy consumption. These tribological properties are difficult to assess because they depend on both microscopic and macroscopic contact conditions and are dependent on the specific material pairs in contact, rather than being intrinsic properties of one material.55 The measurements of friction and wear properties vary with the configurations of the apparatus and the specimens as well as varying with the specifications of the material pairs. Consequently, studies of these properties often present results within the context of a fixed measurement protocol using a single measurement apparatus. Further, to interpret the measurement results succinctly, attempts have been made to provide maps56 or transition diagrams57 that identify stress and temperature domains of distinguishable friction and wear behaviors. A transition diagram for alumina57 is shown in Fig. 11 that shows the coefficient of friction (COF) and a dimensionless measure of wear called the wear coefficient: Kw = Vw H Fn Ds (17)

where Vw is the wear volume, H is the hardness, Fn is the normal force, and Ds is the total sliding distance. Regions I, II, and III in Fig. 11 were considered to be regions of relatively mild wear that were distinguished from each other by the mechanisms involved in the wear process. In region I, reactions between alumina and water vapor may have formed a protective surface that limited the amount of wear.58 At the higher temperatures of region II, plastic flow made significant contributions to the wear process. In region III, protective surface formations were found containing silicon, apparently from a silicon impurity in the bulk material. Region IV was considered to be a region of severe wear in which the wear coefficients were at least 2 orders of magnitude greater than those of regions I, II, and III. Wear in region IV was characterized by intergranular fracture. (6) Thermal Properties Temperature gradients, thermally induced strains, and the rate of transport of thermal energy are especially significant concerns in applications of materials at high temperature. The

Fig. 10. Steady-state creep rate of high-purity, high-density sintered -Al2O3.

1926

Journal of the American Ceramic SocietyMunro

Vol. 80, No. 8

Fig. 12.

Specific heat of high-purity, high-density sintered -Al2O3.

primary properties of interest in this context are specific heat, thermal conductivity, and thermal diffusivity. The specific heat describes the increase of temperature when a quantity of heat is added to a material; the thermal conductivity characterizes the steady-state heat flow in the material; and the thermal diffusivity is related to the transient response of the material to a heat pulse. For isotropic materials, the thermal transport properties are correlated by the relation = CpD (18)
Fig. 13. Thermal transport properties of high-purity, high-density sintered -Al2O3: (a) conductivity and (b) diffusivity.

Table III.
Property 20

Selected Property Values for Sintered -Al2O3*


Temperature (C) 500 1000 1200 1400 1500

Bulk modulus (GPa) Compressive strength (GPa) Creep rate at 150 MPa (109 s1) Density (g/cm3) Elastic modulus (GPa) Flexural strength (MPa) Fracture toughness for crack length of 300 m (MPa m1/2) Friction coefficient at 2 GPa Hardness Vickers, 1 kg (GPa) Lattice parameter, a () Lattice parameter, c () Poissons ratio Shear modulus (GPa) Sound velocity, longitudinal (km/s) Sound velocity, shear (km/s) Specific heat (J kg1 K1) Tensile strength (MPa) Thermal conductivity (W m1 K1) Thermal diffusivity (cm2/s) Thermal expansion from 0C (106 K1) Wear coefficient at 2 GPa (log10) Weibull modulus Weibull characteristic strength (MPa)

257(50) 3.0(5) 0 3.984(2) 416(30) 380(50) 3.5(5) 0.40(5) 15(2) 4.761(3) 12.991(7) 0.231(1) 169(1) 11.00(30) 6.51(20) 755(15) 267(30) 33(2) 0.111(20) 4.6(2) 6(1) 11(4) 395(25)

247 1.6 0 3.943 390 375 3.0 0.8 8.5 4.777 13.040 0.237 158 10.77 6.33 1165 267 11.4 0.0251 7.1 4 11 390

237 0.7 4 3.891 364 345 2.7 0.4 4.6 4.797 13.102 0.244 146 10.54 6.14 1255 243 7.22 0.0150 8.1 6 11 360

233 0.4 280 3.868 354 300 2.6 3.7 4.806 13.129 0.247 142 10.44 6.06 1285 140 6.67 0.0136 8.3 11 310

229 0.3 6600 3.845 343 210 2.5 2.9 4.815 13.156 0.250 137 10.35 5.97 1315 22 6.34 0.0127 8.5 11 210

227 0.28 24600 3.834 338 130 2.5 2.5 4.820 13.169 0.252 135 10.30 5.93 1330 13 6.23 0.0124 8.6 11 125

*Alumina mass fraction, 99.5%; mass density, 98% of the theoretical density; and nominal grain size, 5 m. Numbers in parentheses are representative combined standard uncertainties of the final digits; i.e., the quantity 3.0(5) is equivalent to 3.0 0.5. Units are given in parentheses.

August 1997

Evaluated Material Properties for a Sintered -Alumina

1927

where is the thermal conductivity, is the mass density, Cp is the specific heat at constant pressure, and D is the thermal diffusivity. Specific-heat data for alumina25,59 are shown in Fig. 12. The corresponding interpolation formula is Cp J kg1 K1 = 1117 + 0.14T 411 exp 0.006T 2% (19) for 20C t 1800C. Results from independent studies of the thermal conductivity36,60,61 and the thermal diffusivity40,62,63 are shown in Fig. 13. For these data, the interpolation formulas are optimized simultaneously giving W m1 K1 = 5.85 + for 20C T 15360 exp 0.002T 6% T + 516 (20)

1800C, and 18.9 exp 0.0014T 18% T + 164 (21)

D cm2 s1 = 0.011 + for 20C T 1400C. IV.

Conclusion

Property data for a constrained specification of sintered -alumina have been reviewed to produce a comprehensive set of values for the major material properties as a function of temperature. The mass fraction of the alumina phase of the material is designated to be at least 99.5%, which allows a small amount of MgO to be added to control the grain size. The mass density is constrained to be at least 98% of the theoretical density, and the nominal grain size is 5 m. Interpolation formulas, contained in Eqs. (1) through (21), have been given whenever feasible. Based on these relations, representative values of all the properties on the temperature range from 20 to 1500C are given in Table III. References
1 W. H. Gitzen, Alumina as a Ceramic Material; pp. 1253. American Ceramic Society, Columbus, OH, 1970. 2 R. N. Kleiner, Handbook of Materials Science, Vol. II, Metals, Composites, and Refractory Materials; pp. 35585. Edited by C. T. Lynch. CRC Press, Cleveland, OH, 1975. 3 Engineering Property Data on Selected Ceramics, Volume III, Single Oxides, Rept. No. MCIC-HB-07, Metals and Ceramics Information Center, Battelle Columbus Laboratories, Columbus, OH, 1981. 4 E. D. H. Hubner (Ed.), Alumina Processing, Properties, and Applications; pp. 1317. Springer-Verlag, New York, 1984. 5 R. Morrell, Handbook of Properties of Technical and Engineering Ceramics, Part 2: Data Reviews; Section I: High-Alumina Ceramics; pp. 1255. Her Majestys Stationery Office, London, U.K., 1987. 6 L. D. Hart (Ed.), Alumina Chemicals: Science and Technology Handbook; pp. 1617. American Ceramic Society, Westerville, OH, 1990. 7 K. Komeya and M. Matsui, Materials Science and Technology, Vol. 11, Structure and Properties of Ceramics; pp. 51765. Edited by M. Swain. VCH, Weinheim, Germany, 1994. 8 G. D. Quinn, Flexure Strength of Advanced Ceramics: A Round Robin Exercise, Rept. No. MTL TR 89-62, U.S. Army Materials Technology Laboratory, Watertown, MA, 1989. 9 Enabling Technologies for Unified Life-Cycle Engineering of Structural Components, Publication NMAB-455, National Materials Advisory Board, National Academy Press, Washington, DC, 1991. 10 J. Rumble Jr. and J. Carpenter Jr., Materials STEP into the Future, Adv. Mater. Processes, 142, 2327 (1992). 11 B. G. Hyde, J. G. Thompson, and R. L. Withers, Crystal Structures of Principal Ceramic Materials, Mater. Sci. Technol., 11, 145 (1994). 12 W. H. Gitzen; see Ref. 1, p. 64. 13 H. Holleck, Material Selection for Hard Coatings, J. Vacuum Sci. Technol., A4, 266169 (1986). 14 M. E. Schlesinger, Melting Points, Crystallographic Transformation, and Thermodynamic Values; pp. 88391 in Engineered Materials Handbook, Vol. 4, Ceramics and Glasses. Edited by S. J. Schneider Jr. ASM International, Metals Park, OH, 1991. 15 P. Aldebert and J. P. Traverse, -Al2O3: A High-Temperature Thermal Expansion Standard, High Temp.-High Pressures, 16, 12735 (1994). 16 B. N. Taylor and C. E. Kuyatt, Guidelines for Evaluating and Expressing the Uncertainty of NIST Measurement Results, NIST Tech. Note 1297, U.S. Government Printing Office, Washington, DC, 1993. 17 H. E. Swanson and R. K. Fuyat, Standard X-ray Diffraction Powder Pat-

terns, National Bureau of Standards Circular 539, Vol. II, pp. 2023, National Bureau of Standards, Gaithersburg, MD, 1953. 18 R. E. Newnham and Y. M. de Haan, Refinement of the -Al2O3, Ti2O3, V2O3, and Cr2O3 Structures, Z. Kristallogr., 117, 23537 (1962). 19 H. E. Steinwehr, Gitterkonstanten im System -(Al,Fe,Cr,)2O3 und ihr Abweichen von der Vegardregel, Z. Kristallogr., 125, 377403 (1967). 20 S. V. Grum-Grzhimailo and M. V. Klassen-Neklyudova, Essential Properties of Ruby and Leucosapphire Single Crystals Based on Data in the Literature pp. 211 in Ruby and Sapphire. Edited by L. M. Belyaev. Amerind, New Delhi, India, 1980. 21 N. Ishizawa, T. Miyata, I. Minato, F. Marumo, and S. Iwai, A Structural Investigation of -Al2O3 at 2170 K, Acta Crystallogr., Sect. B. Struct. Sci., 36, 22830 (1980). 22 J. B. Wachtman Jr., T. G. Scuderi, and G. W. Cleek, Linear Thermal Expansion of Aluminum Oxide and Thorium Oxide from 100 to 1100 K, J. Am. Ceram. Soc., 45, 31923 (1962). 23 A. N. Amatuni, T. I. Malyutina, V. Y. Chekhovskoi, and V. A. Petukhov, Standard Samples for Dilatometry, High Temp.-High Pressures, 8, 56570 (1976). 24 R. Morrell; see Ref. 5, p. 19. 25 M. Miyayama, K. Koumoto, and H. Yanagida, Engineering Properties of Single Oxides; see Ref. 14, pp. 74857. 26 N. Soga and O. L. Anderson, High-Temperature Elastic Properties of Polycrystalline MgO and Al2O3, J. Am. Ceram. Soc., 49, 35559 (1966). 27 D. H. Chung and G. Simmons, Pressure and Temperature Dependences of the Isotropic Elastic Moduli of Polycrystalline Alumina, J. Appl. Phys., 39, 531626 (1968). 28 Standard Test Method for Dynamic Youngs Modulus, Shear Modulus, and Poissons Ratio for Advanced Ceramics by Sonic Resonance, ASTM Designation No. C 1198-91. 1995 Annual Book of ASTM Standards, Vol. 15.01. American Society for Testing and Materials, Philadelphia, PA. 29 T. Goto and O. L. Anderson, Elastic Constants of Corundum up to 1825 K, J. Geophys. Res. B, 94 [6] 7588602 (1989). 30 J. Lankford, Compressive Strength and Microplasticity in Polycrystalline Alumina, J. Mater. Sci., 12, 79196 (1977). 31 S. W. Freiman, Brittle Fracture Behavior of Ceramics, Am. Ceram. Soc. Bull., 67, 392402 (1988). 32 G. D. Quinn and R. Morrell, Design Data for Engineering Ceramics: A Review of the Flexure Test, J. Am. Ceram. Soc., 74, 203766 (1991). 33 Standard Practice for Reporting Uniaxial Strength Data and Estimating Weibull Distribution Parameters for Advanced Ceramics, Designation C 123994a. 1995 Annual Book of ASTM Standards, Vol. 15.01, pp. 35673. American Society for Testing and Materials, Philadelphia, PA, 1995. 34 D. G. S. Davies, The Statistical Approach to Engineering Design in Ceramics, Proc. Br. Ceram. Soc., 22, 42952 (1973). 35 R. M. Spriggs, J. B. Mitchell, and T. Vasilos, Mechanical Properties of Pure, Dense Aluminum Oxide as a Function of Temperature and Grain Size, J. Am. Ceram. Soc., 47, 32327 (1964). 36 J. Kubler, Mechanische Charakterisierung von Hochleistungskeramik Festigkeitsuntersuchung, Rept. No. EMPA-Nr. 129 747, pp. 7788, Swiss Federal Laboratories for Materials Testing and Research, Duebendorf, Switzerland, 1993. 37 M. Khantha, The Brittle-to-Ductile Transition, II: Dislocation Dynamics and the Strain-Rate Dependence of the Transition Temperature, Scr. Metall. Mater., 31, 135560 (1994). 38 H. Kim and S. Roberts, BrittleDuctile Transition and Dislocation Mobility in Sapphire, J. Am. Ceram. Soc., 77, 3099104 (1994). 39 C. R. Cheeseman and G. W. Groves, The Mechanism of the Peak in Strength and Toughness at Elevated Temperatures in Alumina Containing a Glass Phase, J. Mater. Sci., 20, 261422 (1985). 40 K. Wefers, Nomenclature, Preparation, and Properties of Aluminum Oxides, Oxide Hydroxides, and Trihydroxides; pp. 1322 in Alumina Chemicals: Science and Technology Handbook. Edited by L. D. Hart. American Ceramic Society, Westerville, OH, 1990. 41 I. J. McColm, Ceramic Hardness; pp. 110. Plenum Press, New York, 1990. 42 R. M. Anderson, Testing Advanced Ceramics, Adv. Mater. Processes, 135, 3136 (1989). 43 Standard Test Method for Microhardness of Materials, Designation No. E 384. 1995 Annual Book of ASTM Standards, Vol. 03.01. American Society for Testing and Materials, Philadelphia, PA, 1995. 44 E. Meyer, Unterschugen uber Prufung and Harte, Z. Ver. Dtsch. Ing., 52, 64554 (1908). 45 B. W. Mott, Micro-Identation Hardness Testing; pp. 101107. Butterworths Scientific Publications, London, U.K., 1956. 46 P. M. Sargent, Use of the Indentation Size Effect on Microhardness for Materials Characterization; pp. 16074 in Microindentation Techniques in Materials Science and Engineering, ASTM STP 889. Edited by P. J. Blau and B. R. Lawn. American Society for Testing and Materials, Philadelphia, PA, 1986. 47 H. Matsui, M. Kawai, T. Takae, K. Hashimoto, S. Takato, and H. Jinno, Microstructure and Mechanical Properties of Al2O3, Si3N4, SiC, and ZrO2 (PSZ) Ceramics for Slurry Valves, J. Soc. Mater. Sci., Jpn., 35, 4147 (1986). 48 Y. S. Wang, S. M. Hsu, and R. G. Munro, A Model for Ceramic Sliding Wear; pp. 122532 in Proceedings of the Japan International Tribology Conference, Vol. II (Nagoya, Japan). Japanese Society of Tribologists, Tokyo, Japan, 1990. 49 H. H. K. Xu, C. P. Ostertag, and R. F. Krause Jr., Effect of Temperature on Toughness Curves in Alumina, J. Am. Ceram. Soc., 78, 26062 (1995).

1928

Journal of the American Ceramic SocietyMunro

Vol. 80, No. 8

50 R. W. Evans and B. Wilshire, Creep Property Characterization of Ceramics, Met. Mater. (Inst. Met.), 7, 36366 (1991). 51 S. M. Wiederhorn, B. J. Hockey, and T. J. Chuang, Creep and Creep Rupture of Structural Ceramics; pp. 55576 in Toughening Mechanisms in Quasi-Brittle Materials. Edited by S. P. Shah. Kluwer, Dordrecht, The Netherlands, 1991. 52 R. C. Folweiler, Creep Behavior of Pore-Free Polycrystalline Aluminum Oxide, J. Appl. Phys., 32, 77378 (1961). 53 A. G. Robertson, D. S. Wilkinson, and C. H. Caceres, Creep and Creep Fracture in Hot-Pressed Alumina, J. Am. Ceram. Soc., 74, 91521 (1991). 54 T. Ohji, A. Nakahira, T. Hirano, and K. Niihara, Tensile Creep Behavior of Alumina/Silicon Carbide Nanocomposite, J. Am. Ceram. Soc., 77, 325962 (1994). 55 F. P. Bowden and D. Tabor, The Friction and Lubrication of Solids; pp. 35064. Clarendon, London, U.K., 1964. 56 S. M. Hsu, D. S. Lim, Y. S. Wang, and R. G. Munro, Ceramics Wear Maps: Concept and Method Development, Lubr. Eng., 47, 4954 (1991).

57 X. Dong, S. Jahanmir, and S. M. Hsu, Tribological Characteristics of -Alumina at Elevated Temperatures, J. Am. Ceram. Soc., 74, 103644 (1991). 58 R. S. Gates, S. M. Hsu, and E. E. Klaus, Tribochemical Mechanism of Alumina with Water, Tribol. Trans., 32, 35763 (1989). 59 R. Morrell; see Ref. 5, p. 17. 60 D. W. Lee and W. D. Kingery, Radiation Energy Transfer and Thermal Conductivity of Ceramic Oxides, J. Am. Ceram., Soc., 43, 594607 (1960). 61 T. Nishijima, T. Kawada, and A. Ishihata, Thermal Conductivity of Sintered UO2 and Al2O3 at High Temperatures, J. Am. Ceram. Soc., 48, 3134 (1965). 62 R. Morrell; see Ref. 5, p. 20. 63 D. P. H. Hasselman, R. Syed, and T. Y. Tien, The Thermal Diffusivity and Conductivity of Transformation-Toughened Solid Solutions of Alumina and Chromia, J. Mater. Sci., 20, 25492556 (1985). 64 J. B. Wachtman, Mechanical Properties of Ceramics; pp. 89115. Wiley, New York, 1996.

Ronald G. Munro is a member of the Ceramics Division of the National Institute of Standards and Technology. Dr. Munro was awarded a National Research Council postdoctoral appointment at NIST in 1976 (when NIST was called the National Bureau of Standards) and subsequently became a staff member in 1978. His work has included many-body theory, molecular dynamic simulations, physics of materials at high pressure, multivariate statistics, tribology, and advanced data evaluation methodologies. Currently, Dr. Munro is a member of the Data Technologies Group, where his research activities include the modeling of materials and their property relations, the development of data evaluation methodologies, and the development of evaluated materials property databases for structural ceramics and high-temperature superconductors. Dr. Munro is a member of the NIST Editorial Review Board, the Editoral Board of the Journal of Physical and Chemical Referencce Data, the American Ceramic Society, the American Physical Society, and the American Society for Testing and Materials.

S-ar putea să vă placă și