Sie sind auf Seite 1von 6

journal

J. Am. Ceram. Soc., 85 [7] 171318 (2002)

Bismuth Oxide Nanoparticles by Flame Spray Pyrolysis


Lutz Ma dler and Sotiris E. Pratsinis
Institute of Process Engineering, ETH Zu rich, CH-8092 Zu rich, Switzerland Bismuth oxide nanostructured particles were made via the flame spray pyrolysis (FSP) of bismuth nitrate that had been dissolved in a solution of ethanol/nitric acid or in acetic acid. These self-sustaining spray flames produced tetragonal -Bi2O3. The use of ethanol/nitric acid solutions resulted in a mixture of hollow, shell-like, and solid nanograined particles. The particle homogeneity was improved as the content of acetic acid in the precursor solution increased. Solid bismuth oxide nanoparticles were prepared consistent with percolation theory, accounting for the specific volume of the product and the precursor. Using pure acetic acid as the solvent, the effect of FSP variables on spray flame and product powder characteristics was investigated. The specific surface area of the Bi2O3 particles could be controlled over a range of 20 80 m2/g by the liquid feed and oxygen gas flow rates for powder production rates of 6 46 g/h. I. Introduction volatility of the starting materials represents the most-severe limitation. Bi2O3 nanoparticles 100 nm in diameter were made by Kaito et al.,13 via the direct oxidation of bismuth metal that was burned on a filament support. Suzuki et al.14 prepared spherical Bi2O3 particles 26 nm in diameter by atomizing a 1M bismuth nitrate solution into fine droplets and reacting them in an argon atmosphere (high-temperature, radio-frequency (rf)) inductively coupled plasma (ICP) reactor. The spraying of an aqueous bismuth nitrate solution into a hot-wall reactor resulted in Bi2O3 particles 100 nm in diameter.15 El-Shall et al.16 prepared pure Bi2O3 nanoparticles with well-defined morphology and crystallinity via laser vaporization/condensation in a diffusion cloud chamber. Here, the synthesis of Bi2O3 nanoparticles is investigated using flame spray pyrolysis (FSP), because flame technology is used for the manufacture of oxide commodities and, as such, FSP has great potential for the manufacture of oxide nanoparticles.17 The FSP process, in particular, can synthesize metal oxides and mixedmetal oxides of high purity at high production rates and under controlled conditions.18 20 Here, bismuth nitrate is used as a raw material, because it is widely used in the wet chemical processing of Bi2O3 and it is one of the least-expensive precursors available. However, in conventional spray pyrolysis, the use of this precursor has resulted in hollow particles.2123 Therefore, special emphasis is placed on control of the morphology and the primary particle size of the Bi2O3 nanoparticles that are produced.

ISMUTH OXIDE (Bi2O3) is used as an additive in paints and in cataphoresis; it also is used as a substitute for lead oxide in glass or porcelain.1 More specifically, Bi2O3-based compounds are much better solid electrolytes than the well-known stabilized zirconia, because the face-centered cubic (fcc) -Bi2O3 exhibits the highest ion conductivity of all oxide ion conductors.2 However, this phase is only stable within a temperature range of 10031097 K.3,4 The presence of a small amount of Bi2O3 in calcined zinc oxide (ZnO) ceramics promotes nonlinear currentvoltage characteristics, which are used in varistor production.5 When 1 mol% of Bi2O3 is added to niobium oxide (Nb2O5) varistor-type materials before calcination, it improves the hydrogen gas sensing.6 Also, tin oxide-based gas sensors (for carbon monoxide) exhibit improved sensitivity and selectivity when doped with Bi2O3.7 Furthermore, the total oxidation of isobutene is enhanced on bismuth-containing tin oxide catalysts.8 Bismuth oxide itself also serves as a catalyst for the conversion of propylene to 1,5-hexadiene9 and can be regenerated when supported on -Al2O3.10 In electrolyte, varistor, sensor, and catalyst applications, high purity and small Bi2O3 particle sizes with controlled morphology are required. The available surface area of Bi2O3 presently limits the applications, especially in catalysis.8 Typically, Bi2O3 is prepared via the oxidation of bismuth metal at 750 800C or via the thermal decomposition of carbonates or the addition of alkali-metal hydroxides to a bismuth salt solution.11 In the latter case, hydrated bismuth oxides precipitate, followed by calcination that yields Bi2O3. The high-temperature synthesis of Bi2O3 from precursor gases was proposed by Liu and Kleinschmit,12 using the established Aerosil route of reacting chloride precursors in a hydrogen/air (or oxygen) flame; however, the

II.

Experimental Procedure

T. M. Besmanncontributing editor

Manuscript No. 187440. Received September 28, 2001; approved March 25, 2002. The research was supported by the Kommission fu r Technologie und Innovation (KTI) TOP NANO 21, through Grant No. 5351.1, Switzerland. Author to whom correspondence should be addressed.

A flame-spray apparatus that consisted of an external mixinggas-assisted nozzle surrounded by a flamelet ring was used.20 Oxygen (99.95% purity, Pan Gas, Luzern, Switzerland) was used as the oxidant and dispersion gas, atomizing the liquid feed that was supplied by a syringe pump (0.515 mL/min) into a fine spray with a droplet-mass median diameter of 10 m, as measured with Fraunhofer laser diffraction spectrometry (SympaTec, Basel, Switzerland). The flow rate of the oxygen dispersion gas was controlled by a mass flow controller (Bronkhorst, Ruurlo, The Netherlands), which maintained a constant pressure decrease (1.5 bar) at the nozzle tip, regardless of oxygen flow rate. Evaporation and ignition of the spray was initiated by eighteen supporting flamelets that were uniformly distributed at a radius of 6 mm from the center of the nozzle. Premixed methane/oxygen gas flowed through all eighteen supporting flamelets, at a total flow rate of 2.96 L/min (the CH4 flow rate was 1.58 L/min, and the O2 flow rate was 1.38 L/min). An oxygen sheath flow (5 L/min) was supplied through a sintered metal plate ring that was 8 mm wide and had an inner radius of 9 mm. Bismuth trinitrate pentahydrate (Bi(NO3)35H2O, with a purity of 98%; Aldrich Chemical Co., Milwaukee, WI) was used as a precursor; it is soluble in strong inorganic acids and in glacial acetic acid (C2H4O2).11 Thus, two 0.41M Bi(NO3)35H2O solutions were prepared by dissolving the precursor either in a solution of 85 vol% ethanol (Merck, Darmstadt, Germany) and 15 vol% concentrated nitric acid (65% in water, Merck) or in pure glacial acetic acid (reagent-grade, Scharlau Chemie S.A., Barcelona, Spain). The standard Bi2O3 production rate was 5.76 g/h, using either precursor feed solution. The highest production rate achieved was 46.08 g/h. 1713

1714

Journal of the American Ceramic SocietyMa dler and Pratsinis III. Results and Discussion

Vol. 85, No. 7

The particles were collected using a vacuum pump on a glass fiber filter (GF/A, Whatmann, Springfield Mill, Maidstone, Kent, U.K.) that had a diameter of 15 cm. The powder specific surface area (As) was measured via nitrogen adsorption at 77 K and BrunauerEmmettTeller (BET) theory (Model Gemini 2375, Micromeritics, Norcross, GA, with a five-point isotherm) after degassing the sample for at least 1 h at 150C in nitrogen. Assuming equally sized spheres (monodisperse particles), the average BET-equivalent particle diameter (dBET) is d BET 6 Asp (1)

where p is the density of Bi2O3 (8.9 g/cm3).11 Samples of the product powder were analyzed by transmission electron microscopy (TEM) (Model H600, operated at 100 kV; Hitachi, Tokyo, Japan). For a homogeneous deposition on the TEM grid, a 1 mg sample of the product powder was dispersed in 10 cm3 of toluene (99.0% purity, Fluka, Ronkonkoma, NY) in an ultrasonic bath for 15 min. Afterward, a droplet of the suspension was dried on a carbon-coated copper TEM grid (Plano GmbH, Wetzlar, Germany). Scanning electron microscopy (SEM) (Model S900, operated at 30 kV; Hitachi) was used to obtain secondary-electron (SE) and back-scattered electron (BS) SEM images. The XRD spectra were recorded with an advance diffractometer (Model D8, Bruker, Karlsruhe, Germany) over a 2 range from 20 to 70. Crystallite characteristics from the XRD spectra were obtained using the Topas 2.0 software (Model AXS 2000, Bruker).

(1) Effect of Precursor Solution on Particle Morphology Spraying under standard conditions (1 mL/min of the Bi(NO3)35H2O/ethanol/HNO3 solution at an oxygen dispersion gas flow rate of 3 L/min) resulted in a spray flame 63 mm long. Figure 1 shows a SE-SEM image (Fig. 1(a)) and a BS-SEM image (Fig. 1(b)) of the bright-yellow Bi2O3 powder that was made; this powder consisted of hollow, almost-transparent spheres 1 m in diameter (Fig. 1(b)). However, smaller (50 100 nm) dense spheres (high contrast in the BS-SEM image; see Fig. 1(b)) also can be observed. Figure 1(c) shows a TEM image of a hollow sphere, where an opening is visible (noted by the arrow), which indicates a shell-like structure. Furthermore, Fig. 1(d) shows a TEM image of the dense spheres (50 100 nm in diameter) and nanograined particles (10 nm in size) at higher magnification. Clearly, this powder had an inhomogeneous morphology, and the specific surface area was 51 m2/g (which corresponded to a dBET value of 13.3 nm (see Eq. (1))). The XRD spectrum of the powder indicated metastable tetragonal -Bi2O3, which has been observed in the rapid quenching of molten Bi2O3.11 Calculation of the average crystal size from the Scherrer equation,24 using the (201) peak, gives a value of 46 nm, which is different from that of the BET analysis. A more-detailed analysis of this sample (Fig. 2), the so-called fundamental parameter approach,2527 was used to convolute the XRD spectrum in two superimposed modes of -Bi2O3 crystal size28 without changing their lattice constants and considering microstrain: a

Fig. 1. Electron microscope images of hollow, large, and fine solid Bi2O3 particles made by FSP of Bi(NO3)35H2O/HNO3/ethanol solution under standard conditions. In the secondary-electron (SE) image (Fig. 1(a)), the majority of the particles are on the order of 1 m, whereas in the backscattered-electron (BS) image (Fig. 1(b)), hollow structures (low contrast) and solid particles (50 100 nm) (high contrast) can be distinguished. Figure 1(c) shows a TEM image of the shell-like structure of the hollow particles, whereas at higher magnification (Fig. 1(d)), solid large and fine particles can be identified.

July 2002

Bismuth Oxide Nanoparticles by Flame Spray Pyrolysis

1715

Fig. 2. XRD spectrum of Bi2O3 powder prepared using ethanol and nitric acid as the bismuth nitrate solvent. Spectrum was deconvoluted with two superimposed -Bi2O3 crystal-size modes with a crystal size of 3 nm (65% by mass) and 104 nm (35% by mass), using the fundamental parameter approach.

mode with an average crystal size of 104 nm (35% of the total mass), which corresponds to the dense spheres, and another mode with a size of 3 nm (65% of the total mass), which corresponds to the nanograined powder and polycrystallinity of the hollow spheres. These results are consistent qualitatively with the SEM/ TEM images (Fig. 1). However, for this powder, there is no clear evidence supporting the idea that the larger particles are single crystals. In conventional spray pyrolysis, the formation of hollow particles is observed frequently.29 In these processes, the final product particles are directly related to the initial solute concentration and droplet size. Jain et al.23 correlated the morphology of singlecomponent particles to the specific volume of nitrate, acetate, and oxalate precursors, based on the percolation theory.30 Assuming that the precursor melts before its decomposition, they proposed a transition point from hollow to solid particles when the specific volume of the product (metal oxide) is larger than 16% of the specific volume of the molten precursor,30 based on their data and that of other investigators. Here, the melting point of Bi(NO3)3 is 76C, and its specific volume is 231 cm3/mol (at the molten state); its decomposition starts at 250C.1 The specific volume of Bi2O3 is 52 cm3/mol, which is 11% of the specific volume of the molten bismuth nitrate, given the stoichiometry. Therefore, hollow particles are expected, according to percolation theory.30 Some of the

hollow structures may further collapse and form larger dense spheres that have similar mass, according to the SEM observations (Fig. 1). Such a restructuring mechanism also was reported by Ortega et al.31 in the spray pyrolysis of barium calcium titanate in a hot-wall flow reactor. They observed hollow spheres at 1200C and larger dense particles at 1500C. Bi2O3 has a low melting point (825C); therefore, restructuring and sintering is very likely within the spray flame. However, no unique temperature exists for when this occurs, because this microstructural process is a function of both temperature and time; therefore, the temperature history of each individual particle determines its likelihood to restructure, and, even at temperatures of 1500C, particles may sustain their hollow shapes if their residence time is short enough. In spray pyrolysis at furnace temperatures of 300C, the formation of hollow particles has been related to the solute concentration distribution inside the droplet and the critical supersaturation and precipitation.29,32 However, the critical solution concentration and its temperature-dependent nucleation or crystal growth of the related materials are unknown and difficult to obtain.29,32 Here, in addition to the hollow and dense particles, a fraction of the Bi2O3 product had the form of aggregates of primary particles 10 nm in diameter. The latter are similar to fumed silica,17 so they may be formed in the gas phase from evaporated precursor after the burning of the solvent. However, this process might be hindered by the endothermic oxidation of bismuth nitrate.1 To avoid the formation of an inhomogeneous mixture of hollow and dense spherical particles, Bi(NO3)35H2O was dissolved in glacial acetic acid. Spraying this solution under standard conditions also resulted in a stable spray flame 63 mm long. The total enthalpy of the spray flame was reduced by 20%, using the acetic acid (16.4 kJ/mL) instead of the ethanol/nitric acid solution (19.6 kJ/mL). Figures 3(a) and (b) respectively show the SE-SEM and TEM images of homogeneous, nanograined product powder; hollow spheres and dense spherical particles were not observed. The grain size in the TEM image (Fig. 3(b)) is consistent with the average dBET value, 12.2 nm, which corresponds to a specific surface area of 55 m2/g. Figure 4 shows the XRD spectrum of the powder in Fig. 3(a); this spectrum corresponds to tetragonal -Bi2O3. Here, the average crystal size is 9.4 nm, by applying the Scherrer formula to the (201) peak (2 28). Using the fundamental parameter approach,2527 the average crystal size is 9.0 nm; this value is smaller than the dBET value, which results possibly from the different BET and XRD averages or from multiple crystals in a particle, as well as particle necking. The change in product morphology by replacing the solvent of the bismuth precursor may be attributed to the formation of intermediate organic bismuth species. It is likely that metastable bismuth oxide acetate is formed during FSP, as in the oxidation of

Fig. 3. Homogeneous, nanograined Bi2O3 powder made by FSP of Bi(NO3)35H2O/acetic acid solution under standard conditions ((a) SEM secondary-electron (SE) image and (b) TEM image).

1716

Journal of the American Ceramic SocietyMa dler and Pratsinis

Vol. 85, No. 7

Fig. 4. XRD spectra of Bi2O3 powder prepared by FSP of (a) bismuth nitrate and (b) bismuth acetate, each dissolved in acetic acid with an equivalent addition of nitric acid.

bismuth citrate.33 The specific volume of the product (Bi2O3) is 32% of that of bismuth oxide acetate (81 cm3/mol, at the molten state); therefore, no hollow particles are formed, according to the percolation theory. The bismuth oxide acetate reacts further exothermally to form bismuth oxide33 and, therefore, contributes to the thermal process, which results in the observed nanograined powder. The change of product powder morphology, from an inhomogeneous mixture of hollow spheres (1 m) and 50 100 nm particles to a homogeneous nanograined product, was studied systematically by preparing 0.41M Bi(NO3)35H2O solutions in solvents of increasing acetic acid concentration. A mixture of concentrated nitric acid, ethanol, and acetic acid with a constant HNO3 content of 15 vol% was prepared. The volume fractions of ethanol and acetic acid were adjusted to obtain various ratios (0.0, 0.2, . . . , 1.0) of acetic acid to ethanol. These solutions were processed via FSP under standard conditions. The flame height did not change, and the color of the product powders was yellow under all conditions. The specific surface area of the product powder slightly increased (up to 20%) as the acetic acid concentration increased, which was consistent with the reduction of flame enthalpy and the change from hollow and large particles to nanograined powder. Increasing the acetic acid concentration substantially increased the homogeneity of the powder crystallinity. Figure 5 shows a clear transition in the XRD spectra from the convoluted peaks (e.g., the (201) peak at 2 28) of small (3 nm) and large (104 nm) crystals in the absence acetic acid (R C2H4O2/(C2H4O2 EtOH) 0.0), as in Fig. 2, to the homogeneous XRD spectrum of nanograined powder (crystal diameter of 9 nm at R 1.0), as Fig. 4(a). This finding was confirmed via TEM, where no hollow spheres or fragmented shells were observed for R 0.6. To explore further the importance of the precursor solvent during oxide formation, a 0.41M solution was prepared by dissolving bismuth acetate (Bi(ac), 99.99% pure; Aldrich Chemical Co.) in acetic acid, with the addition of an equivalent amount of nitric ions (8.8 vol% of concentrated nitric acid), to obtain an identical ion concentration, as with the Bi(NO3)35H2O in acetic acid. The Bi(ac) solution was sprayed under standard conditions, and the product powder exhibited an XRD spectrum that was identical to that of powder obtained using Bi(NO3)35H2O as a precursor (Fig. 4(b)). The specific surface area also was almost identical: 56 and 55 m2/g for powders made using Bi(ac) and Bi(NO3)35H2O precursors, respectively. In addition, no mixtures of hollow or dense particles were present, as observed via TEM. These observations may indicate the importance of the ion composition within the precursor solution rather than the original bismuth precursor (nitrate or acetate) that is used. In this case, the same reaction probably occurs, which results in similar product

Fig. 5. XRD spectra of Bi2O3 powder prepared by FSP of bismuth nitrate dissolved in a mixture of nitric acid, ethanol, and acetic acid (C2H4O2); the nitric acid concentration was kept constant (15 vol%), whereas various fractions of acetic acid in ethanol (0%100%) were prepared. Bismuth oxide made in the absence of acetic acid is inhomogeneous and contains hollow, large, and fine solid particles with low crystallinity. Increasing the acetic acid concentration (R 0.6) facilitates the formation of homogeneous, nanograined -Bi2O3 particles 10 nm in diameter, as shown by TEM (Fig. 3(b)).

characteristics. Here, nanoparticles are formed most likely in the gas phase by rapid evaporation of the precursor, followed by the particle-forming reactions. (2) Control of Particle Size The previous section showed that homogeneous, nanograined bismuth oxide powders can be produced via the FSP of Bi(NO3)35H2O in acetic acid solutions. As a result, a detailed study of the effect of FSP process parameters on particle crystallinity and specific surface area was performed using this solution. Specifically, the effect of the precursor solution feed rate (or powder production rate), and the oxygen dispersion gas flow rate, on the characteristics of product particles were investigated. First, the feed rate of the precursor solution was varied over a range of 1 8 mL/min, while maintaining an oxygen dispersion gas flow rate of 6 L/min. Figure 6 shows the visually measured height of the spray flame at different liquid feed rates, as a function of the

Fig. 6. Flame height as a function of the equivalence ratio (Eq. (2)) (() constant oxygen flow rate of 6 L/min and a precursor feed rate of 1 8 mL/min; (E) precursor feed rate of 1 mL/min and an oxygen flow rate of 3 8 L/min; and () precursor feed rate of 8 mL/min and an oxygen flow rate of 3 8 L/min).

July 2002

Bismuth Oxide Nanoparticles by Flame Spray Pyrolysis

1717

resulting stoichiometric ratio ( 1) of the liquid spray and the oxygen dispersion gas:


n oxidant n fuel n oxidant n fuel

stoich.

(2)

real

The open triangles in Fig. 6 correspond to the condition where the oxygen dispersion gas flow was maintained at 6 L/min. The error bars indicate the variation of the flame height for different runs. Here, an almost-linear relationship was observed, which was consistent with the results from Karpetis and Gomez34 and Ma dler et al.20 The flame height increased from 48 mm at a feed rate of 1 mL/min to 97 mm at a feed rate of 8 mL/min. The specific surface area of the product powder steadily decreased from 76 m2/g to 28 m2/g by increasing the liquid feed rate from 1 mL/min to 8 mL/min. This decrease in the specific surface area corresponds to an increase in the equivalent average primary particle size from 9 nm to 24 nm (Fig. 7). The increase in primary particle size is attributed to the greater concentration of the product powder and the increased residence time at high temperature for higher feed rates, as has been observed in the vapor flame synthesis of TiO2.35 Greater concentrations lead to larger particles, because of enhanced rates of coagulation and/or surface growth.36 At a feed rate of 1 mL/min, the residence time of product particles within the hot temperature zone is shorter than that at 8 mL/min. Shorter residence times at high temperatures lead to slower sintering and smaller primary particle sizes.36 Figure 7 shows a comparison of the dBET value and the average XRD crystal size, using the fundamental parameter approach. Both averaging methods are in good agreement, which indicates that the primary particles are single crystals. The small deviation in particle size between the BET and XRD results possibly results from particle necking37 and the difference in obtaining an average value of the total particle population. This difference increases as the particle size decreases (Fig. 7): necking becomes more important as the particle size decreases, because sintering occurs faster in smaller particles than larger particles. The effect of oxidant/dispersion gas flow rate was investigated by varying the oxygen dispersion gas flow rate in the range of 3 8 L/min at two precursor feed rates: 1 and 8 mL/min. At a precursor feed rate of 1 mL/min, the flame height decreased from 63 mm to 48 mm, as shown in Fig. 6 (circle symbols). At a precursor feed rate of 8 mL/min, the flame height decreased from 141 mm to 100

mm, as also shown in Fig. 6 (square symbols). Here, the proposed linear relation between the flame height and the stoichiometric parameter ( 1) can be determined by considering all flame heights.34 For high oxygen dispersion gas flow rates at a feed rate of 8 mL/min (small ( 1) values; square symbols in Fig. 6), the flame height decreases to a lesser extent, which implies that the additional oxygen cannot promote the evaporation and burning of the fuel any further. However, increasing the oxygen dispersion gas flow rate reduces the total particle concentration and residence time at high temperature. The measured maximum flame temperature is reduced substantially as the oxygen dispersion flow rate increases.20 This leads to faster quenching of particle formation and, therefore, smaller particles, as observed in Fig. 8, where the BET-equivalent particle diameter decreased from dBET 12.1 nm to dBET 8.5 nm at a liquid feed rate of 1 mL/min and from dBET 34.6 nm to dBET 23.8 nm at a liquid feed rate of 8 mL/min by increasing the oxygen dispersion gas flow rate from 3 L/min to 8 L/min. For oxygen flow rates of 5 L/min (smaller ), the reduction in flame height diminishes for both liquid flow rates; the dBET value decreases in a similar manner. The corresponding XRD sizes were similar to the dBET values, as shown in Fig. 7. IV. Summary

Pure, crystalline, and homogeneous bismuth oxide (Bi2O3) nanoparticles have been prepared via the flame spray pyrolysis (FSP) of bismuth nitrate solutions. The decomposition behavior and subsequent reaction path of the precursor/solvent systems affect the morphology and homogeneity of the products. The use of ethanol and nitric acid as the solvent of the nitrate precursor results in the formation of hollow (1 m in diameter) and sintered dense spheres (50 100 nm in diameter), as well as aggregates of primary particles, each of which are 10 nm in diameter. The formation of hollow structures is consistent with percolation theory; therefore, a portion of the product formation occurs within the precursor spray droplet. The mass fraction of two crystalline sizes has been identified using the fundamental parameter approach in X-ray diffractometry (XRD). In contrast, the dissolution of the bismuth nitrate in acetic acid leads to the synthesis of solid nanoparticles. Acetic acid solvent may have favored the formation of an acetobismuth complex, thus facilitating precursor evaporation and gas phase formation of the oxide particles.

Fig. 7. Average BET-equivalent primary particle diameter and XRD crystal size of Bi2O3 made via the FSP of Bi(NO3)3 in acetic acid solution, as a function of the precursor feed rate, at an oxygen dispersion gas flow rate of 6 L/min; the good agreement between the two size lengths indicates single primary particle crystallites. BET size is slightly bigger than the XRD size, probably because of necking, which is an effect that diminishes with increasing size and, subsequently, decreasing sintering rate.

Fig. 8. Average BET-equivalent primary particle diameter of Bi2O3 made via the FSP of Bi(NO3)3 in acetic acid solution, as a function of the oxygen dispersion gas flow rate at two constant precursor feed rates (1 and 8 mL/min). Particle size decreases as the increasing oxygen flow rate decreases the particle concentration and, therefore, the particle growth rate by coagulation. Furthermore, a shorter flame height reduces the particle residence time at high temperatures, reducing the particle sintering rate and coalescence.

1718

Journal of the American Ceramic SocietyMa dler and Pratsinis


12

Vol. 85, No. 7

The role of the precursor solvent was systematically investigated via the FSP of solvents with different ratios of acetic acid to ethanol, where an increase of particle homogeneity and crystallinity was observed for an increasing acetic acid fraction. For an acetic acid fraction of 0.6, only nanograined product powder was observed using electron microscopy and XRD. The role of acetic acid was elucidated further using a solution that contained the same ions, obtained from bismuth acetate instead of bismuth nitrate, which resulted in the same product powder. This result indicates the importance of the ion composition within the precursor, rather than the precursor itself. The specific surface area of the powders was controlled by the precursor feed rate and oxygen dispersion flow rate, using the bismuth nitrate pentahydrate in acetic acid solution. The specific surface area decreased from 80 m2/g (which corresponded to an average primary particle diameter of 8.4 nm) to 20 m2/g (which corresponded to an average primary particle diameter of 33.7 nm). Increasing the liquid precursor feed rate decreased the Bi2O3 specific surface area, whereas increasing the oxygen dispersion gas flow rate increased the Bi2O3 specific surface area. This observation correlated well with precursor concentrations and the measured flame height and, therefore, the particle residence time at high temperatures. Thus, the production of homogeneous Bi2O3 nanoparticles with controlled size and crystallinity was achieved at production rates of 5.76 46.08 g/h. Acknowledgments
The authors gratefully acknowledge the technical support of Dr. M. Mu ller (for providing the EM facilities) and the stimulating discussions with Mr. T. Tani (visiting scientist from Toyota Central R&D Labs, Inc., Aichi, Japan).

References
P. Sulcova and M. Trojan, The Thermal Synthesis of the ZnOBi2O3 Pigments, J. Therm. Anal., 60 [1] 20913 (2000). 2 P. Suk, H. D. Wiemhofer, U. Guth, W. Gopel, and M. Greenblatt, Oxide Ion Conducting Solid Electrolytes Based on Bi2O3, Solid State Ionics, 89 [3 4] 179 96 (1996). 3 E. Oniyama and P. G. Wahlbeck, Phase Equilibria in the BismuthOxygen System, J. Phys. Chem. B, 102 [22] 4418 25 (1998). 4 D. Risold, B. Hallstedt, L. J. Gauckler, H. L. Lukas, and S. G. Fries, The BismuthOxygen System, J. Phase Equilib., 16 [3] 22334 (1995). 5 M. Matsuoka, Nonohmic Properties of Zinc Oxide Ceramics, Jpn. J. Soc. Appl. Phys., 10 [6] 736 46 (1971). 6 T. Hyodo, E. Kanazawa, Y. Takao, Y. Shimizu, and M. Egashira, H2 Sensing Properties and Mechanism of Nb2O5Bi2O3 Varistor-Type Gas Sensors, Electrochemistry, 68 [1] 24 31 (2000). 7 G. S. Devi, S. V. Manorama, and V. J. Rao, SnO2:Bi2O3 Based CO Sensor: Laser-Raman, Temperature Programmed Desorption and X-ray Photoelectron Spectroscopic Studies, Sens. Actuators B, 56 [12] 98 105 (1999). 8 L. Moens, P. Ruiz, B. Delmon, and M. Devillers, Enhancement of Total Oxidation of Isobutene on Bismuth-Promoted Tin Oxide Catalysts, Catal. Lett., 46 [12] 9399 (1997). 9 H. E. Swift, J. E. Bozik, and J. A. Ondrey, Dehydrodimerization of Propylene Using Bismuth Oxide as the Oxidant, J. Catal., 21, 21224 (1971). 10 F. E. Massoth and D. A. Scarpiello, Kinetics of Bismuth Oxide Reduction with Propylene, J. Catal., 21, 22538 (1971). 11 J. Kru ger, P. Winkler, E. Lu deritz, M. Lu ck, and H. U. Wolf, Bismuth, Bismuth Alloys, and Bismuth Compounds; in Ullmanns Encyclopedia of Industrial Chemistry. WileyVCH Verlag GmbH, Weinheim, Germany, 2000.
1

A. T. Liu and P. Kleinschmit, Production of Fumed Oxides by Flame Hydrolysis; pp. 110 in Novel Ceramic Fabrication Processes and Applications, Vol. 38. Edited by R. W. Davidge. Institute of Ceramics, Staffs, U.K., 1986. 13 C. Kaito, K. Fujita, H. Shibahara, and M. Shiojiri, Electron Microscopic Study of Metal Oxide Smoke Particles Prepared by Burning in Ar-O2 Gas, Jpn. J. Soc. Appl. Phys., 16 [5] 697704 (1977). 14 M. Suzuki, M. Kagawa, Y. Syono, and T. Hirai, Synthesis of Ultrafine Single-Component Oxide Particles by the Spray-ICP Technique, J. Mater. Sci., 27 [3] 679 84 (1992). 15 Y. Xiong, S. W. Lyons, T. T. Kodas, and S. E. Pratsinis, Volatile Metal-Oxide Evaporation during Aerosol Decomposition, J. Am. Ceram. Soc., 78 [9] 2490 96 (1995). 16 M. S. El-Shall, W. Slack, W. Vann, D. Kane, and D. Hanley, Synthesis of Nanoscale Metal-Oxide Particles Using Laser Vaporization Condensation in a Diffusion Cloud Chamber, J. Phys. Chem., 98 [12] 306770 (1994). 17 S. E. Pratsinis, Flame Aerosol Synthesis of Ceramic Powders, Prog. Energy Combust. Sci., 24 [3] 197219 (1998). 18 R. M. Laine, R. Baranwal, T. Hinklin, D. Treadwell, A. Sutorik, C. Bickmore, K. Waldner, and S. S. Neo, Making Nanosized Oxide Powders from Precursors by Flame Spray Pyrolysis, Key Eng. Mater., 159 [1] 1724 (1999). 19 A. Kilian and T. F. Morse, A Novel Aerosol Combustion Process for the High Rate Formation of Nanoscale Oxide Particles, Aerosol Sci. Technol., 34 [2] 22735 (2001). 20 L. Ma dler, H. K. Kammler, R. Mueller, and S. E. Pratsinis, Controlled Synthesis of Nanostructured Particles by Flame Spray Pyrolysis, J. Aerosol Sci., 33 [2] 369 89 (2002). 21 T. J. Gardner and G. L. Messing, Preparation of MgO Powder by Evaporative Decomposition of Solutions, Am. Ceram. Soc. Bull., 63 [12] 1498 501 (1984). 22 T. J. Gardner, D. W. Sproson, and G. L. Messing, Precursor Chemistry Effects on Development of Particulate Morphology during Evaporative Decomposition of Solutions; pp. 22732 in Better Ceramics through Chemistry, Materials Research Society Symposium Proceedings, Vol. 32. Edited by C. J. Brinker, D. E. Clark, and D. R. Ulrich. NorthHolland, New York, Amsterdam, The Netherlands, and Oxford, U.K., 1984. 23 S. Jain, D. J. Skamser, and T. T. Kodas, Morphology of Single-Component Particles Produced by Spray Pyrolysis, Aerosol Sci. Technol., 27 [5] 57590 (1997). 24 H. P. Klug and L. E. Alexander, X-ray Diffraction Procedures: For Polycrystalline and Amorphous Materials, 2nd Ed. Wiley, New York, 1974. 25 R. W. Cheary and A. Coelho, A Fundamental Parameters Approach to X-ray Line-Profile Fitting, J. Appl. Crystallogr., 25, 109 21 (1992). 26 R. W. Cheary and A. A. Coelho, Axial Divergence in a Conventional X-ray Powder Diffractometer. I, Theoretical Predictions, J. Appl. Crystallogr., 31, 851 61 (1998). 27 R. W. Cheary and A. A. Coelho, Axial Divergence in a Conventional X-ray Powder Diffractometer. II, Realization and Evaluation in a Fundamental-Parameter Profile Fitting Procedure, J. Appl. Crystallogr., 31, 862 68 (1998). 28 S. K. Blower and C. Greaves, The Structure of beta-Bi2O3 from Powder Neutron-Diffraction Data, Acta Crystallogr., Sect. C: Cryst. Struct. Commun., 44, 587 89 (1988). 29 G. L. Messing, S.-C. Zhang, and G. V. Jayanthi, Ceramic Powder Synthesis by Spray Pyrolysis, J. Am. Ceram. Soc., 76 [11] 270726 (1993). 30 R. Zallen, The Physics of Amorphous Solids, Wiley Classics Library Edition. Wiley, New York, 1998. 31 J. Ortega, T. T. Kodas, S. Chadda, D. M. Smith, M. Ciftcioglu, and J. E. Brennan, Formation of Dense Ba0.86Ca0.14TiO3 Particles by Aerosol Decomposition, Chem. Mater., 3 [4] 746 51 (1991). 32 I. W. Lenggoro, T. Hata, F. Iskandar, M. M. Lunden, and K. Okuyama, An Experimental and Modeling Investigation of Particle Production by Spray Pyrolysis Using a Laminar Flow Aerosol Reactor, J. Mater. Res., 15 [3] 733 43 (2000). 33 S. A. A. Mansour, Thermal-Decomposition of Anhydrous Bismuth Citrate, Thermochim. Acta, 233 [2] 257 68 (1994). 34 A. N. Karpetis and A. Gomez, An Experimental Study of Well-Defined Turbulent Non-premixed Spray Flames, Combust. Flame, 121 [12] 123 (2000). 35 S. E. Pratsinis, W. H. Zhu, and S. Vemury, The Role of Gas Mixing in Flame Synthesis of Titania Powders, Powder Technol., 86 [1] 8793 (1996). 36 S. K. Freidlander, Smoke, Dust, and Haze; 2nd Ed. Oxford University Press, New York, 2000. 37 S. Vemury and S. E. Pratsinis, Dopants in Flame Synthesis of Titania, J. Am. Ceram. Soc., 78 [11] 2984 92 (1995).

Das könnte Ihnen auch gefallen