Вы находитесь на странице: 1из 8

Journal

J. Am. Ceram. Soc., 82 [4] 82532 (1999)

Effect of Seeding and Water Vapor on the Nucleation and Growth of -Al2O3 from -Al2O3
Roger B. Bagwell* and Gary L. Messing*
Department of Materials Science and Engineering, The Pennsylvania State University, University Park, Pennsylvania 16802

Isothermal transformation kinetics and coarsening rates were studied in unseeded and -Al2O3-seeded -Al2O3 powders heated in dry air and water vapor. Unseeded samples heated in dry air transformed to -Al2O3 with an activation energy of 567 kJ/mol. Seeding with -Al2O3 increased the transformation rates and reduced incubation times by providing low-energy sites for nucleation/growth of the -Al2O3 transformation. The activation energy for the transformation was reduced to 350 kJ/mol in seeded samples heated in dry air. Seeded samples completely transformed to -Al2O3 after 1 h at 1050C when heated in dry air compared to 1 h at 925C when heated in saturated water vapor. The combined effects of a lower nucleation barrier due to seeding and the increased diffusion due to water vapor reduced the activation energy for the transformation by 390 kJ/mol and the transformation temperature by 225C compared to the unseeded samples heated in dry air. The accelerated kinetics is believed to be due to increased surface diffusion. I. (1) Introduction

Transformation of Boehmite to -Al2O3

peratures above 1200C for complete conversion to the thermodynamically stable corundum phase.1,3,7 The -Al2O3 nucleation and growth mechanisms are directly responsible for the microstructural evolution in boehmitederived alumina. Many studies of the sintering of unseeded boehmite gels have shown that sintering temperatures of 1600C are required to approach theoretical density.811 The final grain size in these samples is >1 m, eliminating the potential microstructural benefits expected for solgel processing of alumina.7 Several factors contribute to the difficulty of sintering unseeded pseudoboehmite gels. The nucleation density affects the final grain size by dictating the volume of transition alumina matrix consumed by each nulceation event. The nucleation density of -Al2O3 colonies in unseeded boehmite-derived alumina has been reported in the range of 1081010 cm3.11,12 Nucleation of -Al2O3 occurs in a single burst with little or no subsequent nucleation.11,12 The spacing of these nucleation events contributes to the >1 m grain size because each nucleated -Al2O3 colony is able to grow extensively before impinging on another colony.10 The development of a vermicular morphology during growth, as shown in Fig. 1, makes sintering to full density even more difficult due to the formation of large pores that are often entrapped within grains. (2) Control of Nucleation through Heteroepitaxy Many authors have added a variety of potential nucleation aids such as -Al2O3, CuO, -Fe2O3, -Cr2O3, and MgO to

ONTROL of microstructural and phase evolution during calcination of alumina requires an understanding of the phase transformation sequence of the precursor.17 Boehmite ( AlOOH) is a common precursor for the transition aluminas and -Al2O3. Boehmite dehydrates at 450C to form -Al2O3, a metastable defect spinel with the oxygen atoms in cubic packing and aluminum ions in both tetrahedral and octahedral coordination.17 The transformation proceeds with further heating through the following sequence, forming several increasingly ordered transition aluminas before final rearrangement into the corundum structure:1,4

boehmite

-Al2O3

The and forms of alumina are also metastable defect spinels, making the rearrangement process from -Al2O3 through -Al2O3 topotactic and therefore of relatively low energy.14 Some authors describe the transitions between -, -, and -Al2O3 as increased ordering of the aluminum sublattice to produce an intermediate between the defect spinel structure and corundum.4 The transformation from -Al2O3 to -Al2O3 involves a significant change in the oxygen sublattice from cubic to hexagonal close packing and generally requires tem-

D. L. Johnsoncontributing editor

Manuscript No. 191235. Received February 3, 1997; approved August 10, 1998. Presented at the 98th Annual Meeting of the American Ceramic Society, Indianapolis, IN, April 16, 1996. Based in part on the Ph.D. dissertation of R. B. Bagwell. Supported by a grant from 3M Company, St. Paul, MN. *Member, American Ceramic Society.

Fig. 1. -Al2O3 colony growing into a fine-grained transition alumina matrix.

825

826

Journal of the American Ceramic SocietyBagwell and Messing

Vol. 82, No. 4

alumina gels.911,1316 The effect of these seeds is primarily dependent on the similarity of the crystal structure to -Al2O3. For example, Kumagai and Messing 9,14 added 1.5 wt% -Al2O3 seeds to pseudoboehmite gels and found the -Al2O3 crystallization peak temperature was reduced from 1215C to 1075C. Similarly, hematite ( -Fe2O3) seeds lower the crystallization peak because hematite has the corundum crystal structure and a lattice mismatch of only 5.5% with -Al2O3.10,11,13,14 Additives with a crystal structure very different from -Al2O3 tend to have less significant or no effects on the transformation to -Al2O3.11 Another important criterion for an enhanced transformation is maximum interfacial contact between the seed particles and matrix, resulting in high nucleation densities.911 Studies on the effect of seeding have demonstrated that in addition to accelerating the transformation, seeding can also be used to refine microstructures.911,15 Kumagai and Messing9,14 demonstrated that an unseeded pseudoboehmite gel requires 100 min at 1600C in order to reach full density. Seeding with 1.5 wt% -Al2O3, a concentration of 5 1013 seeds/cm3 of transition alumina, resulted in complete densification after 100 min at 1200C and a final grain size of <1 m.9 The increased number of nucleation sites resulted in early impingement of the growing -Al2O3 colonies and therefore a smaller final grain size. The early impingement due to seeding also created highangle grain boundaries that enhanced densification by providing rapid diffusion paths for the removal of porosity.10 (3) Effect of Water Vapor on Coarsening and Transformation Rates The catalytic effect of low pressure (<1 atm) water vapor on the transformation of transition aluminas to -Al2O3 has been documented in several studies.1723 Waters et al.17 found that water vapor accelerated the deactivation of alumina catalyst supports. The water vapor coarsened the transition alumina crystals and also increased the transformation rate to the lower surface area -Al2O3, which the authors attributed to the increased mobility of ions.17 Yanagida et al.18 studied the effect of water vapor on the kinetics of -Al2O3 formation by heating -Al2O3 and -Al2O3 powders in water vapor atmospheres of 0.0005, 17.5, and 760 mmHg. The 760 mmHg (1 atm) water vapor pressure accelerated the transformation of -Al2O3 to -Al2O3, at least in the initial stage of the transformation. The authors explain this effect as the hydrolysis of the AlOAl bonds in the transition alumina by interaction with the water molecule. The weakened bonds allow the breakdown of the -Al2O3 crystal structure which facilitates conversion to -Al2O3. Yanagida et al.18 also reported that the 1 atm water vapor coarsened the -Al2O3 crystals more rapidly than the dry atmosphere, possibly indicating that the water vapor enhances the transformation by allowing the -Al2O3 crystals to reach a critical size more rapidly. Peri2426 investigated the adsorption and desorption behavior of water on thin -Al2O3 plates using infrared spectroscopy and gravimetric analysis. Physisorbed water was removed from the surface by 400C, leaving a layer of hydroxyl groups. Removal of these hydroxyls is relatively easy as long as each OH has a nearby OH to condense with to form water vapor. Monte Carlo simulations indicated that once surface coverage decreases to 9.6%, the remaining hydroxyls are isolated. Further removal requires migration of a hydroxyl to a new site. Because of the difficulty of removing an isolated OH, hydroxyl groups are still present at 1000C. He also reported that a -Al2O3 surface with the hydroxyls completely removed will rapidly readsorb water in the form of hydroxyls even at temperatures as high as 1000C. Water vapor has also been shown to increase the mobility of ions in some oxides by changing defect concentrations.2729 Pope and Simkovich27 demonstrated that BaTiO3 exposed to water vapor displayed increased ionic conductivity due to species with increased mobility compared to the native defects.

The species were determined to be either highly mobile ionic H2O species or barium vacancies formed to compensate for the additional positive charges introduced into the system.27 Water vapor also has an effect on the morphological evolution of boehmite gels.1720 Hrabe et al.20 demonstrated that water vapor resulted in increased pore size and pore volume in seeded boehmite gels. Many studies support the observation that the presence of water vapor results in the acceleration of a nondensifying transport mechanism, and correspondingly a decrease in the densification rate.17,18,20 (4) Kinetics of Nucleation and Growth Transformations The theories that are the basis for the analysis of isothermal transformation kinetics are the work of Johnson and Mehl31 and Avrami 3 2 3 4 which together comprise the JMA theory.3034. The JMA analysis is useful for nucleation and growth transformations, such as -Al2O3 to -Al2O3, because it can be applied to a variety of interface-reaction-controlled or diffusion-controlled transformations.30 The theory combines a nucleation rate equation and growth equation to develop an expression for the transformation rate of the form 1 exp(gIUmt n) (1)

where is the fraction transformed, g is a geometric shape factor, I is the nucleation rate, U is the growth rate, m is a constant dependent on the growth mechanism, t is the time, and n is a constant related to the physical characteristics of the system. Plotting ln ln (1 )1 vs ln (t) allows the determination of n and k. The reaction rate constant k can then be related to the activation energy for the transformation through the expression k A exp(Ea/RT ) (2)

where A is a constant, Ea is the activation energy, R is the gas constant, and T is the absolute temperature. In this paper we focus on the effect of water vapor on the transformation to -Al2O3 in unseeded and -Al2O3-seeded boehmite-derived -Al2O3. Obtaining kinetic data on the transformation such as activation energy and rate constants may provide insight into the mechanism by which water vapor increases diffusion and transformation rates. Information on the combined effect of water vapor and seeding will also illustrate the dependence of the transformation to -Al2O3 on both nucleation and growth. II. Experimental Procedure

(1) Sample Preparation (A) Boehmite and -Al2O3 Seed Sol Preparation: A 15 wt% boehmite ( -AlOOH) sol was prepared by adding a commercial pseudoboehmite powder (Catapal D, Vista Chemical Co.) to deionized water adjusted to pH 3 with nitric acid. The sol was stirred for 2 days with periodic sonication and pH adjustment. The major impurity in the boehmite was 0.2 wt% TiO2. These samples will be referred to as the commercial (C) boehmite. Seed dispersions were prepared by similarly dispersing -Al2O3 particles (AKP-50, Sumitomo) at pH 3. After 2 days of stabilization, the seed dispersion was centrifuged at 2300 rpm for 30 min to remove the larger particles. The remaining seed dispersion was added to the boehmite sol, followed by sonication and stirring to homogenize the seeded boehmite sol. A seeding concentration of approximately 1 1014 seed particles/cm3 -Al2O3 was obtained with a 3 wt% addition of seeds. The unseeded (U) and seeded (S) sols were gelled at >80C on a hot plate while stirring and then dried at 70C for 2 days. Gel fragments were ground to <170 mesh with an alumina mortar and pestle. The Yoldas method35 was used to prepare a high-purity boehmite for comparison with the commercial powder. Aluminum sec-butoxide was reacted at 90C with deionized, distilled

April 1999

Effect of Seeding and Water Vapor on the Nucleation and Growth of -Al2O3 from -Al2O3

827

water and then peptized with nitric acid to form a 4 wt% sol. The sol was refluxed at 90C for 10 days and became translucent. The total metallic impurity content in the alkoxide was less than 10 ppm. After forming a dispersed sol, the alkoxidederived boehmite was seeded, gelled, and ground to powder in the same manner as the commercial boehmite. These samples will be referred to as the alkoxide (A) boehmite. (B) Heat Treatment: The batches of boehmite powder were all precalcined at 600C for 1 h in dry air to obtain -Al2O3, thereby eliminating any effects on the transformation kinetics due to the water liberated during dehydration. Powder samples of the -Al2O3 weighing 0.25 g were placed in a 3 cm 1 cm 1 cm platinum crucible which produced an 3 mm deep powder bed. The crucible was inserted to a point just outside the hot zone, approximately 200C below the soak temperature, and allowed to equilibrate with the furnace atmosphere for 1 min. The crucible was then inserted into the hot zone of the furnace. Temperature control was within 3C during the soak. The time between insertion of the crucible and equilibration with the furnace temperature was 10 to 20 s based on previous experience. After the desired calcination time the crucible was removed from the hot zone and cooled to at least 200C below the soak temperature within 20 s. The shortest soak time was 2 min in order to prevent the equilibration or cooling time from becoming a significant fraction of the transformation time. Water vapor of 1 atm, designated saturated (W) water vapor, was produced by connecting the furnace tube to a 1-L flask of boiling distilled, deionized water. An 0.5 atm of water vapor atmosphere, designated moderate (M) water vapor, was obtained by bubbling air at 60 cm3/min through a 1-L flask of dionized water maintained at 80C. The water vapor pressures were determined from a vapor pressure table.36 A compressed air (<0.01 atm water vapor) flow rate of 60 cm3/ min was used to produce the dry atmosphere. The ends of the furnace tube were maintained at 175C using heating tape to prevent condensation. The different heating conditions used are summarized in Table I. (2) Sample Characterization The -Al2O3 to -Al2O3 transformation temperature was evaluated by heating 30 mg samples in platinum cups in a differential thermal analysis system (DTA) at 10C/min in dry flowing air. DTA measurements in water vapor were not possible because of condensation of the water vapor in the thermal analysis system. Quantitative X-ray diffraction (XRD) was performed on samples containing CaF2 as an internal standard from 24 to 30 2 at 2/min using CuK radiation. The integrated intensities of the 25.6 2 peak of -Al2O3 and the 28.3 2 peak of CaF2 were determined and the ratio compared with a standard curve to determine percent conversion to -Al2O3. Error in the percent transformation was estimated as 3 wt% based on repeated measurements under the same conditions. Each data point in the transformation versus time plot represents one measurement. Incubation times for the onset of the transformation were determined by extrapolating the linear portion of the transformation curve to the time axis. The curves on the transformation as a function of time plots are not the result of a curve-fitting algorithm, but are drawn to represent an Avrami-type transformation. Transformation data were entered directly into the formula ln ln (1 )1 for plotting as a func-

tion of ln (t) to determine the kinetics data and corresponding activation energies. Data points below 5% and for 100% transformation were not included in the data analysis to avoid effects due to determination of incubation time and final transformation time on the calculated values. XRD line broadening measurements were performed to determine crystallite size. Scans were run from 30 to 50 2 at 4 2 /min using CuK radiation. The 43.3 2 (113) reflection was used to determine the size of the -Al2O3 crystals. The 32.9 2 (202) reflection was used for the -Al2O3 crystals since it was the only distinct, isolated peak in the pattern. The full width at half-maximum was measured for each peak and the crystallite size, t, was calculated from the Scherrer equation:37 t 0.9 /B cos (3)

where is the wavelength of CuK radiation, B is the broadening of the peak in radians, and cos is determined for the peak position in . Line broadening due to the equipment was subtracted from the peak width before calculating the crystallite size using the formula B2 B2 B2 meas equip (4)

The specific surface area of the -Al2O3 powders was determined by multiple point BET in nitrogen after outgassing at 200C for 30 min. III. Results and Discussion

The unseeded commercial and alkoxide-derived boehmite samples transformed to -Al2O3 at 1198 and 1181C, respectively, in the dry atmosphere, as measured by the peak maximum in the DTA. The transformation temperatures were reduced in the seeded samples to 1121C for both the commercial and alkoxide powders. (1) -Al2O3 Transformation Kinetics (A) Transformation in Dry Air: Isothermal transformation curves for the unseeded and seeded commercial samples heated in the dry atmosphere are shown in Fig. 2. The transformation curves for the unseeded and seeded alkoxide samples in the dry atmosphere are not presented. The transformation to -Al2O3 follows Avrami-type transformation kinetics. The formation of -Al2O3 is preceded by an incubation time which decreases steadily with increasing temperature. The slopes of the curves are shallow initially, which is generally attributed to a slow initial nucleation rate.30 This interpretation seems questionable for the transformation to -Al2O3 since many authors report an initial burst of -Al2O3 nucleation with no further nucleation events.11,12 The shallow slopes may be partially due to difficulties in determining the exact time of the initial nucleation event. The slope of the curves increases at 10% transformation as the -Al2O3 colonies reach their maximum growth rate. The curves begin to plateau at 90% transformation presumably because of extensive impingement of the growing -Al2O3 colonies. Some impingement may occur at lower volume fractions of -Al2O3, based on TEM observations, but does not appear to significantly affect the calculated growth rate. Seeding the transformation accelerates the transformation by providing low-energy nucleation/growth sites, allowing transformation rates similar to the unseeded system at 100C lower temperatures. The incubation time is also substantially reduced in the seeded samples. The unseeded commercial sample heated at 1100C had an incubation time of 126 min while the seeded sample heated at 1050C had an incubation time of only 27 min. The longer incubation times for the unseeded samples are typically interpreted as the time required to form -Al2O3 nuclei. It is reasonable to postulate that in the presence of

Table I.

Heating Conditions and Sample Designations for Isothermally Transformed Powders


Dry 1 atm water vapor 0.5 atm water vapor

Sample

Unseeded commercial Unseeded alkoxide Seeded commercial Seeded alkoxide

UCD UAD SCD SAD

UCW UAW SCW SAW

UCM SCM

828

Journal of the American Ceramic SocietyBagwell and Messing

Vol. 82, No. 4

Fig. 2. Isothermal transformation curves for -Al2O3 heated in dry air: (a) unseeded, commercial, and (b) seeded, commercial.

-Al2O3 seeds that the transformation process involves growth only. Plots of ln ln (1 )1 vs ln (time) for both the commercial and alkoxide samples heated in dry air were used to determine the kinetic data presented in Table II. The rate constants for the transformation increased with increasing temperature, as expected for a thermally activated, or Arrhenius, process. The value of the time exponent, n, varies significantly even for one

Table II. Kinetic Data Obtained from Isothermal Transformation Analysis of Powders Heated in Dry Air
Sample Temp (C) r n k (s1) Inc time (min)

UCD

1100 1125 1150 1175 975 1000 1025 1050 1100 1125 1150 1175 975 1000 1025 1050

0.99 0.99 0.99 0.94 0.99 0.98 0.97 0.95 0.97 0.99 0.95 0.97 0.99 0.97 0.99 0.98

3.47 3.40 3.39 3.31 1.30 1.47 1.37 1.31 2.26 3.13 3.27 2.87 1.08 1.33 1.46 1.18

7.22 1.75 3.82 9.70 3.09 6.23 1.04 2.16 1.44 3.94 8.41 1.75 4.23 9.78 1.65 4.31

105 104 104 104 105 105 104 104 104 104 104 103 105 105 104 104

126 47 21 9 230 132 59 27 42 21 8 6 132 66 30 12

SCD

UAD

SAD

Correlation coefficient.

type of powder. The change in the time exponent is reported to imply a change in the nucleation conditions.3034 However, the JMA model does not include effects such as the presence of porosity and a volume change during the transformation, which makes interpretation of the time exponent in terms of a physical process in the powders difficult.3034 The low correlation coefficients for some of the data sets (for example, 0.94 for UCD 1175C) may also be indicative of problems with using the JMA method to analyze the -Al2O3 to -Al2O3 transformation. The alkoxide-derived samples exhibited slightly faster transformation kinetics and shorter incubation times than the commercial boehmite-derived samples. McArdle38 reported that the relatively small differences in transformation rates between alkoxide-derived and commercial boehmites were not due to impurities or surface area. One possible explanation for the difference in transformation rates in this study is the crystallite packing in the powders. The alkoxide-derived boehmite was refluxed for 10 days at 90C and contains uniform 5 nm crystals while the commercial boehmite contains 6 nm anisometric crystals and significant quantities of heterogeneities such as agglomerates, based on TEM observations. The alkoxide synthesis procedure may have resulted in a more densely packed transition alumina matrix which would favor more rapid growth. Since the transformation to -Al2O3 followed an Arrhenius relation, plots of ln k vs 1/T can be used to determine the activation energy for the transformation. The correlation coefficients were 0.99 for all of the activation energy plots. Seeding with -Al2O3 reduced the activation energy for the transformation from 567 kJ/mol to 350 kJ/mol for the commercial samples and from 547 kJ/mol to 411 kJ/mol for the alkoxide samples. The error in the activation energies was estimated to be 5% based on the scatter in the isothermal transformation data. The activation energies for the unseeded samples are not significantly different within the estimated error, which is also true for the seeded samples. The activation energies are within the range reported in other studies. McArdle and Messing10 calculated an activation energy of 578 kJ/mol for the transformation to -Al2O3 in an unseeded powder and 476 kJ/mol when seeded with -Fe2O3. Other authors have reported activation energies for the transformation to -Al2O3 in the range of 400600 kJ/mol.39,40 Direct comparison of activation energies from different studies is problematic because of variations in experimental conditions and starting materials. The 30% reduction in the activation energy for the transformation in the seeded system may be due to the removal of the energy barrier for nucleation. However, interpreting the decrease in activation energy in terms of only one component of the transformation process is questionable. Clearly, increasing the number of potential growth sites from 1010 to >1014 sites/cm3 (each seed can provide multiple growth sites) also biases the activation energy for growth. The vermicular growth morphology, shown in Fig. 1, results in the -Al2O3 colonies losing contact with the -Al2O3 matrix. The loss of interfacial area for the transfer of atoms from -Al2O3 crystals to -Al2O3 significantly limits the growth rate in the unseeded system. The effect of the loss of contact between the -Al2O3 and -Al2O3 on the growth rate is not as significant in the seeded gels because the scale of the vermicular morphology is significantly reduced. The effects of seeding on the transformation rate in boehmite-derived alumina must be interpreted as both a removal of the barrier to nucleation by epitaxy and also as an alteration of the overall growth process by a change in the scale of the developing microstructure. (B) Transformation in Water Vapor: The transformation curves for the commercial boehmite-derived samples heated in saturated water vapor are shown in Fig. 3. The water vapor accelerates the transformation, especially in the seeded systems. For example, the transformation to -Al2O3 was complete after 1 h at 925C in the seeded powder heated in water vapor and after 1.5 h at 1100C in the unseeded sample. Table

April 1999

Effect of Seeding and Water Vapor on the Nucleation and Growth of -Al2O3 from -Al2O3

829

Fig. 3. Isothermal transformation curves for -Al2O3 heated in saturated water vapor: (a) unseeded, commercial, and (b) seeded, commercial.

III summarizes the kinetic data obtained for the samples heated in saturated water vapor. The incubation time for the unseeded commercial boehmitederived -Al2O3 heated at 1100C decreased from 126 min to 38 min when heated in water vapor. The decrease in incubation
Table III. Kinetic Data Obtained from Isothermal Transformation Analysis of Powders Heated in Saturated Water Vapor
Sample Temp (C) r n k (s1) Inc time (min)

time may not be entirely due to a change in the energy barrier for nucleation. Since nucleation consumes <<1% of the total matrix, nucleation could be identical in dry air and saturated water vapor. The water vapor may also accelerate the rate at which a sufficient fraction (2%) of -Al2O3 is produced for detection by X-ray diffraction. Water vapor reduces the activation energy for the transformation from 567 to 519 kJ/mol for the unseeded commercial boehmite-derived -Al2O3 and from 547 to 356 kJ/mol for the unseeded alkoxide-derived -Al2O3. The larger reduction in activation energy for the unseeded alkoxide sample may be related to surface area. The surface areas of the powders after precalcination at 600C for 1 h in dry air were 167 and 218 m2/g for the commercial and alkoxide-derived -Al2O3, respectively. Since the accelerated diffusion due to water vapor involves surface adsorption, as discussed below, a larger surface area favors this mechanism. Water vapor reduced the activation energy for the transformation in the seeded samples from 350 to 177 kJ/mol for the commercial samples and from 411 to 191 kJ/mol for the alkoxide samples. The regulation of nucleation through seeding and growth by water vapor reduced the activation energies by 69% from the unseeded samples heated in dry air. The transformation in the seeded samples heated in water vapor exhibited rate constants similar to seeded samples heated at 175C higher temperatures in the dry atmosphere. The transformation to -Al2O3 in moderate water vapor occurs at a rate intermediate between that of the dry atmosphere and saturated water vapor, as illustrated in Fig. 4. The rate constants for the transformation are also intermediate between the saturated water vapor and dry atmospheres, as shown in Table IV. The activation energies are 504 and 215 kJ/mol for the unseeded and seeded samples, respectively. Table V summarizes the activation energies for the transformation to -Al2O3 for all the conditions studied. The largest decreases in activation energy are for seeded samples heated in saturated water vapor, emphasizing the importance of both nucleation and growth in the transformation to -Al2O3. The relative effect of water vapor on the rate constants depends on the transformation temperature. Table VI shows the ratio of the rate constants for samples heated in water vapor (saturated and moderate) to that in the dry atmosphere at various temperatures. Since seeded samples were not heated at the same temperature in water vapor and dry air, the k values were extrapolated, assuming Arrhenius kinetics, from the measured data in order to compare reaction rate constants at the same temperature. The ratios of the reaction rate constants for saturated water vapor compared to the dry atmosphere approach a value of 3 at the higher temperatures but increase steadily with decreasing temperature and are >100 at 800C. The moderate water vapor exhibited a similar trend with ratios lower

UCW

1075 1100 1125 1150 800 850 900 925 1075 1100 1125 1150 800 850 900 925

0.99 0.99 0.93 0.98 0.97 0.97 0.99 0.98 0.99 0.99 0.99 0.92 0.99 0.99 0.96 0.99

2.92 2.44 4.05 3.59 1.20 1.42 1.24 1.27 2.18 2.23 2.45 2.60 1.07 0.95 1.36 1.48

1.09 2.13 6.02 1.16 4.39 1.10 2.21 3.67 3.45 7.14 1.21 1.85 4.28 9.02 2.81 3.61

104 104 104 103 105 104 104 104 104 104 103 103 105 105 104 104

87 38 17 8 179 56 22 13 21 11 6 3 132 56 32 21

SCW

UAW

SAW

Correlation coefficient.

Fig. 4. Isothermal transformation curves for unseeded commercial samples heated at 1100C in the three atmospheres studied.

830

Journal of the American Ceramic SocietyBagwell and Messing

Vol. 82, No. 4

Table IV. Kinetic Data Obtained from Isothermal Transformation Analysis of Powders Heated in Moderate Water Vapor
Sample Temp (C) r n k (s1) Inc time (min)

UCM

1075 1100 1125 1150 900 950 1000 1025

0.99 0.95 0.99 0.97 0.99 0.99 0.99 0.99

4.07 3.63 3.50 2.13 1.54 1.04 1.84 1.13

6.31 1.53 3.59 6.59 6.30 1.42 4.27 4.59

105 104 104 104 105 104 104 104

115 47 18 7 105 30 12 7

SCM

Correlation coefficient.

Table V. Activation Energies for the Transformation of -Al2O3 to -Al2O3


Sample Activation energy (kJ/mol)

UCD SCD UAD SAD UCW SCW UAW SAW UCM SCM

567 350 547 411 519 177 356 191 504 215

Fig. 5. vapor.

Coarsening of -Al2O3 crystals in dry air and saturated water

Table VI.
Samples

Ratio of Rate Constants in Water Vapor to Rate Constants in Dry Air


Temperature (C) kwater /kdry (kwater /kdry)1/3

UCW/UCD

1100 1125 1150 1100 1125 1150 800 850 925 800 850 925 1100 1125 1150 925 1000 1025

2.95 3.44 3.04 4.96 3.07 2.20 113 84.0 40.0 159 85.1 33.4 2.12 2.05 1.73 11.2 6.85 4.41

1.43 1.51 1.45 1.71 1.45 1.30 4.83 4.38 3.42 5.42 4.40 3.22 1.28 1.27 1.20 2.24 1.90 1.64

where r is the radius of the -Al2O3 crystal after heating, r0 is the original radius of the crystal, and t is the heating time. The slope of the plot is k , which is proportional to the diffusion coefficient, D, and the surface energy, . The initial slopes of the curves are used to compare the relative effect of the water vapor on coarsening. The ratio of the slopes in Fig. 5 for samples heated in water vapor to that of the dry atmosphere is 50 and 11 at 950 and 1050C, respectively. The relative effect of the water vapor is greater at the lower temperatures. The water vapor results in an increase in the diffusion coefficient of at least a factor of 50 at 950C. The effect of water vapor on D is slightly larger if the surface energy is assumed to be 10% lower as a result of the adsorbed water.42 (3) Comparison to Other Studies The transformation data in this study can be compared to other studies involving alternate transport paths and accelerated diffusion. Yanagida et al.18 reported an increase in the transformation rate in the initial stage of the transformation to -Al2O3 when heated in 1 atm of water vapor followed by a plateau. Shelleman and Messing16 studied the effect of enhancing diffusion in the -Al2O3 to -Al2O3 transformation by adding V2O5 to both seeded and unseeded -Al2O3. The solubility of alumina in V2O5, which melts at 673C, allowed the rapid transport of aluminum and oxygen ions. The V2O5 lowered the -Al2O3 transformation temperature, as measured by DTA, of the unseeded gel by 25C. Seeding reduced the transformation temperature by 83C. The combined effect of enhanced nucleation and diffusion lowered the transformation temperature by 205C. The activation energy for the transformation was reduced from 442 kJ/mol in the untreated system to 116 kJ/mol in the seeded, V2O5 system. The transformation to 100% -Al2O3 occurred in 30 min at 875C. Shaklee and Messing43 used HF as a mineralizer for the transformation of -Al2O3 to -Al2O3. They proposed that vapor transport by AlF2O increased the transformation rates. The transformation of -Al2O3 to -Al2O3 was complete after 30 min at 900C when optimum reactant concentrations were used. The transformation rates in the seeded powders heated in water vapor in this study are slightly slower, but comparable to both the liquid- and vapor-assisted transformations. (4) Water-Vapor-Assisted Diffusion Several possible mechanisms could contribute to enhanced diffusion by water vapor, including vapor transport, changes in surface energies, and increased defect concentration. Previous studies43,44 have shown that the presence of a gaseous species

UAW/UAD

SCW/SCD

SAW/SAD

UCM/UCD

SCM/SCD

Indicates some rate constant data were extrapolated to the temperature listed.

than observed in the saturated water vapor. The last column in Table VI is the cube root of the rate constant ratios, which facilitates later comparison of the data to the change in the diffusion coefficient. (2) Effect of Water Vapor on Coarsening Figure 5 shows how coarsening of the -Al2O3 crystals changes when heated in the water vapor. The data for the coarsening of the crystals is plotted according to Eq. (5), which describes grain growth dominated by surface diffusion within a porous body,41 r4 r4 0 kt (5)

April 1999

Effect of Seeding and Water Vapor on the Nucleation and Growth of -Al2O3 from -Al2O3

831

capable of transporting aluminum and oxygen, as in the case of HF mineralization, drastically increases the coarsening and transformation rates. Vapor transport of aluminum is generally disregarded as a mechanism for the enhanced diffusion observed in water vapor due to the low vapor pressures (<1015 atm) of aluminum-containing species.23 However, a recent study on the oxidation rates of silicon carbide in wet oxygen45 showed that at 1200C aluminum could be transported through the vapor phase from the walls of the alumina tube furnace. Increased aluminum transport rates would have a significant effect if aluminum is the rate-limiting species. The above discussion emphasizes the importance of understanding the rate-limiting species for diffusion in alumina. The rate-limiting species for the coarsening of both -Al2O3 and -Al2O3 and also during the -Al2O3 to -Al2O3 transformation is not well-defined. Aluminum lattice diffusion is faster than oxygen lattice diffusion because of the smaller ionic radius of the aluminum.46 Diffusion along boundaries is much more complicated. LeGall et al.47 reported that aluminum subboundary diffusion is faster than oxygen diffusion in -Al2O3 single crystals. The general consensus is that oxygen grain boundary diffusion is faster than aluminum grain boundary diffusion in polycrystalline alumina.46 The rate-limiting species for surface diffusion has not been determined, although considerable data exist for the overall surface diffusion coefficient.46 Several other factors contribute to the difficulty in assigning a rate-limiting species to coarsening of transition aluminas or the transformation to -Al2O3. Defect concentration and therefore diffusion in alumina are controlled by aliovalent impurity contents as low as 10 ppm due to the low concentration of intrinsic defects.46,48 This dependence of diffusion on low impurity concentrations is one cause of the variations in diffusion data between different investigations.46,48 Another complication in determining a rate-limiting species is that most of the available diffusion data are for -Al2O3 rather than the transition aluminas. A change in surface energies is another possibility proposed for the effect of water vapor on the coarsening and transformation rates in alumina.21 Since adsorbed water lowers the surface energy of alumina,42 it can alter coarsening, nucleation, and growth processes. The energy barrier to nucleation of -Al2O3 would be lowered, thereby accelerating the transformation to -Al2O3, if water lowers the interfacial energy between -Al2O3 and -Al2O3. On the other hand, adsorbed water would result in a slower coarsening rate since the driving force for coarsening is the reduction of surface free energy. Thus, the effect of the adsorption of hydroxyls on the transformation rate would depend on the relative contributions to nucleation and growth. Another potential mechanism for the enhanced diffusion is a change in the defect concentration due to the adsorbed water molecules. Peri2426 showed that transition alumina and -Al2O3 crystals adsorb a high concentration of water molecules which form hydroxyls in a variety of configurations and are highly mobile due to a proton hopping mechanism. Pope and Simkovich27 showed that water vapor increases the ionic conductivity in BaTiO3 by increasing the defect concentration, possibly in the form of barium vacancies. Similarly, the diffusion of adsorbed hydroxyls into the alumina would result in a local positive charge that must be compensated by the formation of a negatively charged defect. The formation of aluminum vacancies would balance the additional charge while increasing the diffusion coefficient of aluminum. One final point that should be addressed is the critical -Al2O3 crystal size for the nucleation of -Al2O3. Many authors proposed that -Al2O3 crystals must coarsen a certain degree before the nucleation of -Al2O3. The rapid coarsening and diffusion due to water vapor could accelerate the transformation to -Al2O3 by increasing the rate at which the -Al2O3 crystals reach this critical size. This explanation is unlikely based on evidence for the critical -Al2O3 crystal size in the

nucleation of -Al2O3, which will be discussed in a future paper. Based on the number of potential mechanisms to explain the effect of water vapor, it is important to discuss the relative increase in the diffusion coefficient necessary to explain the increased transformation rates. The growth rate, u, in a transformation is related to the diffusion coefficient by Eq. (6), u D b/ j (6)

where Db is the boundary diffusion coefficient and j is the jump distance. The rate constant is proportional to the growth rate cubed. The diffusion coefficient is therefore proportional to the cube root of the rate constant. The data in the last column of Table VI therefore indicate a fivefold increase in D at temperatures <850C and 1.5-fold at temperatures >1100C. Comparing the rate constant data with the coarsening data in Fig. 5 indicates a relative increase in D in the range of 550 at temperatures less than 900C and 1.53 at temperatures greater than 1100C. (A) Potential for Vapor Transport: The potential for vapor transport must be considered, because of the similarities in the transformation rates between the HF-mineralized vapor transport system43 and the water-vapor-treated samples in this study. As discussed earlier, vapor-phase transport of a species that contains aluminum is generally disregarded in the literature because of the thermodynamic calculations of low vapor pressures, and correspondingly low fluxes.23 We calculated using SOLGAS mix that the most concentrated aluminum oxygen vapor species, AlHO2, has a vapor pressure of 1019 atm at 900C and 1016 atm at 1100C. The net flux due to vapor transport can be calculated from J
c(

P)/(2 MRT )0.5

(7)

where c is the condensation coefficient which is assumed to be 1, P is the difference in partial pressure of the AlHO2 over the -Al2O3 and -Al2O3, M is the molecular weight of the transporting species, R is the gas constant, and T is the absolute temperature. The difference in pressure was estimated, as an upper limit of the potential flux, to be of the same order of magnitude as the calculated vapor pressure. The maximum flux at 1100C was estimated as 9.3 1013 g/(m2 s), which is far too small to account for the increased coarsening and transformation rates observed. A 0.25 g sample would take several years to transform if this flux were the only material transport mechanism. The flux at temperatures <900C would be several orders of magnitude lower. Vapor transport of oxygen by water molecules also cannot account for the 100-fold increase in the diffusion coefficient at low temperatures since aluminum boundary diffusion is ratelimiting. However, Rahaman41 demonstrates that in an ionic crystal the effective diffusion coefficient can be increased by accelerating diffusion of the faster species. The diffusion of the faster species sets up a large chemical potential gradient that increases the diffusion of the slower species. In the case of alumina, the effective diffusion coefficient can be expressed as41 Deff 5DAlDO /(3DAl + 2DO) (8)

For rapid oxygen diffusion, Eq. (8) reduces to41 Deff 5DAl/2 (9)

The effective diffusion coefficient would be increased by a factor of 2.5 due to rapid oxygen diffusion through the vapor, which may account for the high-temperature (>1100C) behavior in this study. However, a different mechanism must be responsible for the rapid diffusion observed at the lower temperatures. (B) Changes in Solid-State Diffusion Coefficient: Some

832

Journal of the American Ceramic SocietyBagwell and Messing

Vol. 82, No. 4

of the increase in the diffusion coefficient must be due to a solid-state process since the increase at low temperatures cannot be attributed to vapor-phase transport. A change in the defect concentrations explains the enhanced coarsening and transformation rates and the increased relative effect of the water vapor at lower temperatures. The extrinsic vacancy concentration due to the adsorbed water would be more significant than at higher temperatures where more intrinsic vacancies exist. Surface adsorption would also be favored at lower temperatures. Unfortunately, little literature exists to explain how water vapor affects the already defective structure of -Al2O3. Also, complicating the analysis is the difficulty of the experiments and the absence of model single crystals for detailed surface studies. Without further detailed experiments we can only speculate about the true mechanism for enhanced transport in high-temperature water vapor. The change in defect concentration due to adsorbed water molecules is a reasonable explanation for the increased diffusion and transformation rates, but reliable data for calculating an intrinsic defect concentration is needed to confirm this hypothesis. IV. Conclusions The transformation to -Al2O3 in boehmite-derived -Al2O3 was accelerated by altering both the nucleation and growth processes. Seeding the transformation with -Al2O3 effectively eliminated the barrier to nucleation. The presence of water vapor accelerated diffusion in the system. The combined effects of seeding and water vapor increased the transformation rate by several orders of magnitude compared to an unseeded sample heated in dry air. The changes in the transformation rates and activation energy illustrate the strong dependence of the formation of -Al2O3 on both nucleation and growth. The above discussion leads to several important conclusions regarding the presence of favorable nucleation sites, such as seed particles or other heterogeneities, or water vapor during the processing of alumina and alumina precursors. The elimination of low-energy -Al2O3 nucleation sites is beneficial to the production of thermally stable transition aluminas since any defect may act to nucleate the transformation to -Al2O3. Water vapor should be minimized during the sintering of -Al2O3 due to the nondensifying transport mechanism that is accelerated at temperatures ranging from 800 to 1200C. Water vapor should also be eliminated, if possible, when the stability of high surface area, and thus the retention of a transition alumina, are critical. References
1 K. Wefers and C. Misra, Oxides and Hydroxides of Aluminum, Alcoa Technical Paper No. 19, Aluminum Company of America, Pittsburgh, PA, 1987. 2 G. Ervin, Structural Interpretation of the DiasporeCorundum and Boehmite -Al2O3 Transitions, Acta Crrystallogr., 5, 103108 (1952). 3 R. K. Iler, Fibrillar Colloidal Boehmite; Progressive Conversion to Gamma, Theta, and Alpha Aluminas, J. Am. Ceram. Soc., 44 [12] 61824 (1961). 4 R.-S. Zhou and R. L. Snyder, Structures and Transformation Mechanisms of the , , and Transition Aluminas, Acta Crystallogr., B47, 61730 (1991). 5 B. C. Lippens and J. J. Steggenda, Active Alumina; pp. 171211 in Physical and Chemical Aspects of Adsorbents and Catalysts. Edited by B. G. Linsen. Academic Press, London, U.K., 1970. 6 B. E. Yoldas, Effect of Variations in Polymerized Oxides on Sintering and Crystalline Transformations, J. Am. Ceram. Soc., 65 [8] 38793 (1982). 7 P. A. Badkar and J. E. Bailey, The Mechanism of Simultaneous Sintering and Phase Transformation in Alumina, J. Mater. Sci., 11, 1794806 (1976). 8 P. A. Badkar, J. E. Bailey, and H. A. Barker, Sintering Behaviour of Boehmite Gel; pp. 31121 in Sintering and Related Phenomena. Plenum Press, 1972. 9 M. Kumagai and G. L. Messing, Enhanced Densification of Boehmite Sol Gels by Alpha-Alumina Seeding, J. Am. Ceram. Soc., 67 [11] C-230C-231 (1984). 10 J. L. McArdle and G. L. Messing, Transformation, Microstructure Development, and Densification in -Fe2O3 Seeded Boehmite-Derived Alumina, J. Am. Ceram. Soc., 76 [1] 21422 (1993). 11 W. A. Yarbrough and R. Roy, Microstructural Evolution in Sintering of AlOOH Gels, J. Mater. Res., 2 [4] 494515 (1987).

12 F. W. Dynys and J. W. Halloran, Alpha Alumina Formation in AlumDerived Gamma Alumina, J. Am. Ceram. Soc., 65, 44248 (1982). 13 G. C. Bye and G. T. Simpkin, Influence of Cr and Fe on Formation of -Al2O3 from -Al2O3, J. Am. Ceram. Soc., 57 [8] 36771 (1974). 14 M. Kumagai and G. L. Messing, Controlled Transformation of a Boehmite SolGel by -Alumina Seeding, J. Am. Ceram. Soc., 68 [9] 500505 (1985). 15 J. L. McArdle and G. L. Messing, Transformation and Microstructure Control in Boehmite-Derived Alumina by Ferric Oxide Seeding, Adv. Ceram. Mater., 3 [4] 38792 (1988). 16 R. A. Shelleman and G. L. Messing, Liquid-Phase-Assisted Transformation of Seeded -Alumina, J. Am. Ceram. Soc., 71 [5] 31722 (1988). 17 R. F. Waters, J. B. Peri, G. S. John, and H. S. Seelig, Alumina Structure and Activity of PlatinumAlumina Catalysts, Ind. Eng. Chem., 52 [5] 41516 (1960). 18 H. Yanagida, G. Yamaguchi, and J. Kubota, The Role of Water Vapor in Formation of Alpha Alumina from Transient Alumina, J. Ceram. Assoc. Jpn., 74 [12] 37177 (1966). 19 C. H. Chang, R. Gopalan, and Y. S. Lin, A Comparative Study on Thermal and Hydrothermal Stability of Alumina, Titania, and Zirconia Membranes, J. Membr. Sci., 91, 2745 (1994). 20 Z. Hrabe, S. Komarneni, L. Pach, and R. Roy, The Influence of Water Vapor on Thermal Transformations of Boehmite, J. Mater. Res., 7 [2] 44449 (1992). 21 Z. Hrabe, O. M. Spaldon, L. Pach, and J. Kozankova, Thermal Transformations of Boehmite Gel in Controlled Furnace Atmospheres, Mater. Res. Bull., 27 [4] 397404 (1992). 22 M. Pijolat, M. Dauzat, and M. Soustelle, Influence of Additives and Water Vapor on the Transformation of Transition Aluminas into Alpha Alumina, Thermochim. Acta, 122, 7177 (1987). 23 F. W. Dynys, Alpha Alumina Formation in Alumina SolGels; Ph.D. Thesis. Case Western Reserve University, Cleveland, OH, 1985. 24 J. B. Peri, Surface Hydroxyl Groups on -Alumina, J. Phys. Chem., 64 [10] 152630 (1960). 25 J. B. Peri, Infrared and Gravimetric Study of the Surface Hydration of -Alumina, J. Phys. Chem., 69 [1] 21119 (1965). 26 J. B. Peri, A Model for the Surface of -Alumina, J. Phys. Chem., 69 [1] 22030 (1965). 27 J. M. Pope and G. Simkovich, The Use of BaTiO3 as a Solid-Electrolyte to Determine Water Vapor Effects Upon Electrical Transport Mechanisms, Mater. Res. Bull., 9 [9] 111118 (1974). 28 R. Waser, Solubility of Hydrogen Defects in Doped and Undoped BaTiO3, J. Am. Ceram. Soc., 71 [1] 5863 (1988). 29 S. K. Mohapatra, S. K. Tiku, and F. A. Kroger, The Defect Structure of Unintentionally Doped -Al2O3 Crystals, J. Am. Ceram. Soc., 62 [12] 5057 (1979). 30 J. W. Christian, The Theory of Transformation in Metals and Alloys, 2nd ed. Pergamon Press, Oxford, U.K., 1975. 31 W. A. Johnson and R. F. Mehl, Reaction Kinetics in Processes of Nucleation and Growth, Trans. AIME, 135, 41633 (1939). 32 M. Avrami, Kinetics of Phase ChangeI, J. Chem. Phys., 20, 110312 (1939). 33 M. Avrami, Kinetics of Phase ChangeII, J. Chem. Phys., 8, 21224 (1940). 34 M. Avrami, Kinetics of Phase ChangeIII, J. Chem. Phys., 9, 17784 (1941). 35 B. E. Yoldas, Alumina Gels that Form Porous Transparent Al2O3, J. Mater. Sci., 10, 185660 (1975). 36 CRC Handbook of Chemistry and Physics, 73rd ed. CRC Press, Cleveland, OH, 1992. 37 B. D. Cullity, Elements of X-ray Diffraction, 2nd ed. Addison-Wesley, Reading, MA, 1978. 38 J. L. McArdle, Seeding with Ferric Oxide for Enhanced Transformation and Microstructure Development in Boehmite-Derived Alumina; Ph.D. Thesis. Pennsylvania State University, University Park, PA, 1989. 39 R. A. Shelleman, G. L. Messing, and M. Kumagai, Alpha-Alumina Transformation in Seeded Boehmite Gels, J. Non-Cryst. Solids, 82, 27785 (1986). 40 H. Schaper and L. L. Van Reijen, A Quantitative Investigation of the Phase Transformation of Gamma to Alpha Alumina with High Temperature DTA, Thermochim. Acta, 77, 38393 (1984). 41 M. N. Rahaman, Ceramic Processing and Sintering. Marcel Dekker, New York, 1995. 42 D. Hardie and N. J. Petch, Lowering of Surface Energy by Adsorption on Alumina, Proc. Br. Ceram. Soc., 5, 8596 (1965). 43 C. A. Shaklee and G. L. Messing, Growth of -Al2O3 Platelets in the HF -Al2O3 System, J. Am. Ceram. Soc., 77, 297784 (1994). 44 V. N. Kuklina, E. A. Levitskii, L. M. Plyasova, and V. I. Zharkov, The Role of a Mineralizer in the Polymorphic Transition of Aluminum Oxide, Kinet. Katal., 13 [5] 126974 (1972). 45 E. J. Opila, Oxidation Kinetics of Chemically Vapor Deposited Silicon Carbide in Wet Oxygen, J. Am. Ceram. Soc., 77 [3] 73036 (1994). 46 E. Doore and H. Hubner, Alumina: Processing, Properties, and Applications. Springer-Verlag, Berlin, Germany, 1984. 47 M. LeGall, A. M. Huntz, B. Lesage, C. Monty, and J. Bernardini, SelfDiffusion in -Al2O3 and Growth Rate of Alumina Scales Formed by Oxidation: Effect of Y2O3 Doping, J. Mater. Sci., 30, 20111 (1995). 48 R. E. Mistler and R. L. Coble, Rate Determining Species in DiffusionControlled Processes in Al2O3, J. Am. Ceram. Soc., 54 [1] 6061 (1971).

Вам также может понравиться