Sie sind auf Seite 1von 13

This article was downloaded by: [203.250.80.

20] On: 22 December 2011, At: 18:04 Publisher: Taylor & Francis Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Aerosol Science and Technology


Publication details, including instructions for authors and subscription information: http://www.tandfonline.com/loi/uast20

Chemical Aerosol Engineering as a Novel Tool for Material Science: From Oxides to Salt and Metal Nanoparticles
Evagelos K. Athanassiou , Robert N. Grass & Wendelin J. Stark
a a a a

Institute for Chemical and Bioengineering, ETH Zurich, Zurich, Switzerland

Available online: 24 Oct 2011

To cite this article: Evagelos K. Athanassiou, Robert N. Grass & Wendelin J. Stark (2010): Chemical Aerosol Engineering as a Novel Tool for Material Science: From Oxides to Salt and Metal Nanoparticles, Aerosol Science and Technology, 44:2, 161-172 To link to this article: http://dx.doi.org/10.1080/02786820903449665

PLEASE SCROLL DOWN FOR ARTICLE Full terms and conditions of use: http://www.tandfonline.com/page/terms-and-conditions This article may be used for research, teaching, and private study purposes. Any substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any form to anyone is expressly forbidden. The publisher does not give any warranty express or implied or make any representation that the contents will be complete or accurate or up to date. The accuracy of any instructions, formulae, and drug doses should be independently verified with primary sources. The publisher shall not be liable for any loss, actions, claims, proceedings, demand, or costs or damages whatsoever or howsoever caused arising directly or indirectly in connection with or arising out of the use of this material.

Aerosol Science and Technology, 44:161172, 2010 Copyright American Association for Aerosol Research ISSN: 0278-6826 print / 1521-7388 online DOI: 10.1080/02786820903449665

Chemical Aerosol Engineering as a Novel Tool for Material Science: From Oxides to Salt and Metal Nanoparticles
Evagelos K. Athanassiou, Robert N. Grass, and Wendelin J. Stark
Institute for Chemical and Bioengineering, ETH Zurich, Zurich, Switzerland

Downloaded by [203.250.80.20] at 18:04 22 December 2011

Aerosol nanotechnology has rapidly evolved in the past years. This fascinating technology has resulted in the development of functional nanomaterials providing novel solutions in industrial applications. The extensive research on the physical understanding of gas phase processes has strongly contributed to the present industrial use of single and mixed oxides and the design of industrial aerosol reactors. Recent advances have shown that chemical aerosol engineering can be established on the interface between classical aerosol science and chemical engineering. The emerging new methods give access to a much broader class of functional materials including salt and metal nanoparticles. The latter implies that aerosol production units can now be considered as chemical reactors. The incorporation of thermodynamic considerations and chemical kinetics in the modelling of gas phase processes will further boost the development of aerosol engineering and will provide deeper understanding of the fundamentals of particle formation mechanisms. This will ultimately enable access to new multicomponent materials with various structures or morphologies and the development of more sustainable, energy efcient gas phase processes.

INTRODUCTION Aerosol particle technology has gone through a tremendous development in the past thirty years (Kruis et al. 1998a; Pratsinis 1998; Rosner 2005; Roth 2007; Swihart 2003). Fundamental studies on particle dynamics, combustion science and innovative engineering now enable the scalable synthesis of nanoparticles through a broad variety of processes. The increasing amount of research on nanoparticle synthesis has been triggered by very distinct properties of such particles, leading to novel materials with possible new and improved applications (Gleiter 1989). Some of the most advantageous properties were

Received 20 August 2009; accepted 26 October 2009. The authors acknowledge nancial support by ETH Zurich in the form of a TH grant (TH-02 07-3) and the Swiss National Science Foundation (SNF 200021-116123). Address correspondence to Wendelin J. Stark or Evagelos K. Athanassiou, Institute for Chemical and Bioengineering, ETH Zurich, Wolfgang-Pauli str. 10, 8093 Zurich, Switzerland. E-mail: wendelin.stark@chem.ethz.ch; athanssiou@chem.ethz.ch

discovered several decades ago and motivated the development of new nanoparticles. Carbon black, silica, and titania, are readily available from commercial gas-phase-based processes in the order of several million metric tons per year (Kammler et al. 2001; Osterwalder et al. 2006; Stark and Pratsinis 2002). The wide success of these materials as well as a better understanding of the process dynamics favored the analysis and synthesis of a much broader class of materials than the traditional metal oxides, such as metal salt (Grass and Stark 2005), suldes (Kruis et al. 1998b) and pure metal nanoparticles (Grass and Stark 2006a) and lms (Rosner 2005). Nanoparticles of such a wide range of materials can be prepared by a variety of methods. Nevertheless, gas-phase-based aerosol processes have dominated over wet-based methods (Park et al. 2004) and top-down approaches, as they provide a scalable manufacturing of nanomaterials exhibiting high purity and relatively narrow particle size distributions. The scope of this review is, therefore, to give an overview on the use of aerosol based processes for the production of nanoparticles. We would like to highlight the evolutionary development of gas phase aerosol processes from mostly physical investigations into chemical processing, which has allowed the synthesis of new materials beyond oxides. Without the physical understanding of particle dynamics and combustion aerosol formation, later inclusion of chemical control in aerosol processing would have not been feasible (Friedlander and Wang 1966; Kruis et al. 1993; Landgrebe and Pratsinis 1990; Zachariah and Carrier 1999). The detailed understanding of particle nucleation, coagulation and aggregation provided the fundament to engineer the production of nanoparticles under controlled combustion conditions. These mainly physical studies have enabled the production of titania and silica, which today serve as reference studies while the physical principles can be implemented to any material class (oxides, salts, metals). In the past few years, chemical aerosol engineering established itself on the interface between classical aerosol science and the discipline of chemical engineering. This implies that aerosol production units can be considered as chemical reactors, where ultra-fast chemical reactions take place during particle formation or, in other words, next to physical (coagulation, sintering, agglomeration) chemical processes (reaction, 161

162

E. K. ATHANASSIOU ET AL.

Downloaded by [203.250.80.20] at 18:04 22 December 2011

reduction, solid-state chemistry) contribute to the particle formation process. First indications show that simple thermodynamic considerations and calculations follow very accurately the experimental ndings. This indicates that the gas composition and temperature of the aerosol comprising gas stream after the combustion can be used for the further development of aerosol processes to prepare new classes of materials such as metals, salts, sulde and other compositions of nanoparticles. Nevertheless, it remains a challenge to incorporate these chemical thermodynamic considerations in the modeling of gas phase processes and to analyze the competition of different chemical reactions (reduction vs. oxidation), the inuence of the concentration of combustion gases (H2 O, H2 , CO2 , CO) and the characteristic time of each chemical reaction step in comparison to the particle dynamics. Another important factor is the validity of thermodynamics concerning the turbulent or diffuse conditions of the produced aerosol. This will provide a deeper understanding of nanomaterial synthesis from gas phase processes and further contribute to controlled scale-up of reactors for the manufacturing of more complicated materials than oxides. The review gives an overview on the different classes of nanomaterials derived from aerosols and highlights how chemical aerosol engineering triggered this rapid expansion of aerosol derived nanoparticle products. It further analyzes some of the key future challenges for aerosol engineering, which arise at the interface of this established technology with other engineering disciplines.

disperse, usually spherical nanomaterials derived by gas phase processes. Compared to the liquid phase, the gas phase manufacturing of nanoparticles excels in terms of high rate and low volume requirements. Reaction and residence times are generally in the region of milliseconds and rarely exceed one second for the whole process from nanoparticle formation to separation. Most aerosol gas phase processes are characterized by the use of high temperatures, often above 1000 C. The common denominator of the individual processes is the formation of a molecular nuclei, either from condensation or from chemical reaction at high temperatures. The nuclei then quickly collide by Brownian motion and thermophoresis and grow by coalescence in high temperature regions (physical description). In order to limit the growth and aggregation of the nanoparticles, the residence time needs to be kept as short as possible or the process needs to exhibit extremely fast cooling rates (>1000 K s1 ). Aerosol gas phase processes vary in their energy or heat supply and nucleation source. The main, industrially relevant, gasphase processes exhibiting high temperatures in combination with high cooling rates are plasma and laser-based processing with electricity as an energy source and combustion processes which rely on fossil energy. Laser-Based Aerosol Production These aerosol reactors involve a laser source to vaporize chemical precursors followed by condensation and/or chemical reaction to form nanoparticles. Kato (1976) showed the preparation of several oxides such as SiO2 , Fe3 O4 , CaTiO3 , and Mg2 SiO4 by laser ablation and the use of a continuouswave CO2 laser. A more suitable large-scale synthesis method is laser-ablation of microspheres. This process proceeds by the transport of suitable microspheres (e.g., metals, ceramics, oxides, semiconductors) to the reaction zone where the spheres are irradiated with a laser beam. If the energy of the laser is sufciently high, the particles vaporize (ablate) without going through the liquid state and nanoparticles are formed from the cooling vapors (Hergenroder 2006). Illumination of gaseous precursors by a laser results in a rapid temperature increase triggering the precursor decomposition. Examples include silicon nanoparticles from SiH4 gas (Cannon et al. 1982) and iron or iron oxide nanoparticles from gaseous Fe(CO)5 (Majima et al. 1994; Majima et al. 1993; Martelli et al. 2000). Advanced process reactor systems have been described by Gardner et al. (2003) or Bi and Kambe (2001) and allow the use of liquid, vapor and aerosol precursor systems enabling the fabrication of nanomaterials of various compositions and relatively low cost. This approach, therefore, enables the manufacture of nanoparticles at high production scales (Friedlander and Wang 1966; Kambe 2001). In spite of the steady progress in laser technology, lasers are still expensive and process efciency remains low, constraining the commercial application of laser-based aerosol manufacturing methods in an industrial scale.

GAS PHASE PROCESSES While top-down approaches, such as ball milling (Schiotz et al. 1998; Suryanarayana 2001) are generally straightforward and simple, they are of very low energy efciency (Osterwalder et al. 2006) and mostly result in products with inhomogeneous particle sizes. Bottom-up synthesis methods result in well-controlled nanoparticles, which are produced by reaction of a metal precursor, followed by a controlled aggregation. The bottom-up methods are divided in liquid or gas phase methods. Compared with the gas phase, reactions in liquid media are easily conducted in a laboratory setup, usually involve moderate temperatures but often require complex chemicals at moderate prices. Due to the smaller diffusion coefcient in liquids the chemical reactions are much slower, resulting in low production rates (Osterwalder et al. 2006). Separation of the produced nanomaterials is also a major issue in comparison to the more easily powder ltration of an aerosol gas phase process. Nevertheless the slower chemical reaction rates allow better control of the experimental conditions and parameters determining the physical formation processes such as nucleation, growth, and aggregation of nanoparticles in contrast to the often turbulent conditions of gas phase processes. This has as an effect that wet phase methods result in the preparation of functional nanomaterials with uniform (monodisperse) sizes and size distributions and also enable the easier access to nanomaterials with various anisotropic shapes or structures comparing to the poly-

CHEMICAL AEROSOL ENGINEERING

163

Downloaded by [203.250.80.20] at 18:04 22 December 2011

Plasma-Based Aerosol Production Similar to nanoparticle formation by laser, plasma synthesis can use various types of precursor materials. Molecular nuclei are formed by electrically heating the precursor up to 104 C, decomposing the reactants into partially ionic species. When exiting the plasma zone the temperature of the gas drops and nanoparticles are formed. Again, steep homogeneous cooling gradients are necessary for the fabrication of materials with a narrow particle size distribution. Thermal plasmas applicable for nanoparticle synthesis include direct current plasma jets, direct current arc plasmas and radio frequency induction plasmas. Large production rates have been achieved for nickel nanoparticles (70 g h1 ) using an anodic arc plasma with a power consumption of above 5 kW and, a very recent application, with a DC thermal plasma for the synthesis of ZnO from commercial zinc powders at production rates exceeding 1 kg h1 (Ko et al. 2006). The high energy consumption of up to 70 kW and the required inert gas for the plasma formation may limit the commercial applicability of this process. A potentially more energy efcient process is the pulsed wire discharge synthesis. This rather exible and simple system has been used for the synthesis of metal oxides (Kinemuchi et al. 2002; Kinemuchi et al. 2001; Sangurai et al. 2001; Suzuki et al. 2001), metals (Jiang and Yatsui 1998), and metal nitrides (Lee et al. 2006) and offers aerosol gas phase nanoparticle synthesis at an energy consumption of down to 20 MJ kg1 (Jiang and Yatsui 1998). Flame Synthesis Flame synthesis has its roots in the formation of carbon black (soot). For nearly a century this method has been used for the large-scale preparation of carbon black of over 8 million metric tons per year, mostly covering applications in the production of tires and rubber (Stark and Pratsinis 2002). The ame technology was later adapted for the production of silica and titania by the oxidation of chlorides (SiCl4 /TiCl4 ) in high temperature ames (Osterwalder et al. 2006). TiCl4 + O2 TiO2 + 2Cl2 SiCl4 + O2 SiO2 + 2Cl2 Today this is the major synthesis route for the manufacturing of pigmentary titania (5 Mt/year) and fumed silica nanoparticles (4 Mt/year) (SRI 2001), which are used as powder owing aids and in cosmetics and in the fabrication of optical bers for telecommunication. However, most nanoparticles cannot be derived from the metal chloride oxidation process. This aerosolbased process only works for a limited set of elements and requires volatile metal chlorides, which can be fed to a ame reactor and reacted exothermally with oxygen. At applicable temperatures this is only possible for silicon-, aluminum-, titanium-, vanadium- (Stark et al. 2001), and zirconium-chloride, all currently used for the large scale synthesis of nanoparticles in ame processes. In order to manufacture other interesting nanomate-

rials (there are more than 100 chemical elements) another alternative metal feed system for ame reactors was required. As solids are usually not easily combustible, liquid systems are preferred. Water based precursor systems were introduced by Zachariah and Huzarewicz (1991a) as well as Matsoukas and Friedlander (1991), where an aqueous precursor was atomized, dried and then fed to a burning ame, resulting in the decomposition of the precursor and the formation of nanoparticles. Using this technology, complex mixed oxide nanoparticles, such as superconducting yttrium-barium-copper oxides, could be formed. The use of aqueous starting materials resulted in a broader window of materials, but the often larger and less homogeneous particles suggest the preferred use of organic and combustible precursors. Optimized metal-organic precursors consist of an organic liquid of high metal loading (530 wt%; achieve high mass production), which is combustible at a high gross caloric value. This allows the complete combustion of the organic constituents to water and carbon dioxide and the nucleation of the metal constituents in the hot-zone of the ames. Various precursor systems have been applied, including metal alkoxides (Mueller et al. 2004a), acetates (Tani et al. 2003), triethanolamines (Kim et al. 2004; Marchal et al. 2004), and metal carboxylates (Stark et al. 2003; Stark et al. 2004; Stark and Pratsinis 2004). The metal carboxylate system appears promising, as nearly all elements are available in low-cost napthenate-metal salts. Further, these precursors are generally stable in air and against humidity, allowing the fabrication of mixed-oxide nanoparticles with excellent chemical homogeneity and relatively narrow particle size distributions.

PHYSICAL UNDERSTANDING In spite of its apparent simplicity the long history of aerosol process development has been characterized by evolutionary trial and error research. This experimental early knowledge has later been supported by an in-depth understanding of the process dynamics. A typical sequence of the basic steps illustrating particle formation in a gas stream is given in Figure 1. A metal loaded precursor is injected as a gas or liquid spray into the ow. The sprayed precursor is evaporated due to rapid heating through surrounding gases depending on the process (source of the thermal energy: laser, plasma, or ame). The additionally generated heat of combustion (ames) favors the evaporation of the solvent and precursor, ideally resulting in the formation of a homogeneous gas phase monomer. Consumption of the gaseous precursors proceeds either by gas phase and or by surface reactions (Spicer et al. 2002). This is typically followed by the formation of some rst nuclei clusters. The thermal stability of these nuclei strongly inuences the evolving particle formation process. According to classical nucleation theory based on bulk material properties, the critical cluster size is often smaller than the dimension of the idealized gas phase monomers. The cluster further grows either by collision with additional monomers and particles with sticking coefcients often assumed to be one or

164

E. K. ATHANASSIOU ET AL.

Downloaded by [203.250.80.20] at 18:04 22 December 2011

FIG. 1. In high temperature gas phase processes nanoparticles grow from monomers or nuclei by fast agglomeration forming fractal aggregates (transition region). Sintering and coalescence then leads to the formation of spherical particles that continue to collide by Brownian motion. If the temperature is insufcient for full coalescence to spheres, hard agglomerates are formed. In the cold zone of the process and during ltration soft agglomerates are formed, which are only interconnected by Van-der Waals forces. Chemical aerosol engineering allowed considering ame spray synthesis as chemical reactors. The product composition can be controlled by the combustion and reaction conditions, resulting in a wide variety of nanoparticles (present example: iron).

by the addition of gas phase monomers on the clusters surface. The coalescence of cluster ensembles is normally very fast, resulting in compact, dense and nearly spherical particles. The characteristic time of the gas-phase reactions and the coalescence is usually much shorter than the characteristic time for subsequent evolution of the particle shape (i.e., sintering). When typical particle diameters become large enough (several nanometers), the further development of a particle is often determined by surface growth and by the interdependence of coagulation and coalescence (Tsantilis and Pratsinis 2000). Fractal structures start to form by Brownian motion and coagulation and can merge (collapse) again into spheres by coalescence (Matsoukas and Friedlander 1991). This nally results in the formation of fractal aggregates. These groups of particles are called soft if the primary particles are interconnected (Figure 1) by van-der-Walls forces and hard, if sinter necks exist (Grass et al. 2006; Tsantilis and Pratsinis 2004). Physical material properties, residence time of the particles, core and temperature of the gas stream determines the nal morphology and crystallinity of the agglomerates. This understanding of the particle/aggregate formation in an aerosol process was initiated with the observation of selfpreserving particle size distributions (Friedlander and Wang 1966). It was later discovered (Schaefer and Hurd 1990; Xiong and Pratsinis 1993) that ame-made particles had a fractal

structure with a fractal dimension of 1.71.9, representing a very open aggregate structure. To get real experimental insight into the dynamics of the ame process Arabi-Katbi et al. (2001) used thermophoretic sampling to remove relevant samples from different positions in the ame. Analysis by electron microscopy conrmed that the earlier derived physical models of the ame process indeed correctly predicted the nanoparticle growth process. While more complex and exact population balance models are now available in literature (Koch and Friedlander 1990; Tsantilis et al. 2002; Xiong et al. 1993; Xiong and Pratsinis 1993), a simple model based on particle nucleation, agglomeration by Brownian motion and coalescence has been proposed by Kruis et al. (1993). Despite the often rough assumptions made, the model offers a good quantitative understanding of many ame processes and is routinely used in combination with computational uid dynamics (CFD) packages for the improved design of ame reactors (Johannessen et al. 2000, 2001; Muhlenweg et al. 2002), yielding nanoparticles with a desired size and morphology (Grass et al. 2006; Heine and Pratsinis 2006; Tsantilis and Pratsinis 2004). This detailed understanding has allowed the scale-up, design and engineering of industrial aerosol reactors for nanoparticle synthesis. Laboratory set-ups of ame-spray reactors usually operate now in the region of 10100 g/hr and small pilot plants have been shown to produce silica, zirconia, and yttrium stabilized zirconia at

CHEMICAL AEROSOL ENGINEERING

165

production rates of up to 1 kg/hr (Jossen et al. 2005; Mueller et al. 2004b; Mueller et al. 2003). CHEMISTRY IN AEROSOL PROCESSING OF MATERIALS Single Oxides The understanding of the particle formation dynamics and combustion science did not only facilitate a better reactor design but it also resulted in the synthesis of more complex materials. Besides silica, many other single oxides have been synthesized including TiO2 , Al2 O3 GeO2 , V2 O5 , SnO2 , ZrO2 , CeO2 , ZnO, and WO2 . The most important parameters determining particle morphology are the concentration of the precursor, residence time and temperature next to ame gas composition. The controlled adjustment of high temperature residence time and the high cooling rates have given access to the formation of metal oxides with various crystallinities and particle morphologies. The controlled formation of maghemite (Fe2 O3 ) (Grimm et al. 1997), rutile or anatase TiO2 (Pratsinis and Vemury 1994), monoclinic or cubic Y2 O3 (Camenzind et al. 2005), and tetragonal ZrO2 (Karthikeyan et al. 1997) are examples of some high temperature phases (crystal polymorphs) obtained by aerosol ame technology. These are often materials which are difcult to prepare by conventional wet phase (low temperature preparation method in water) and subsequent slow bulk calcination (large sintered area, loss of nanoparticle size). In a further extension of ame spray synthesis Laine et al. (2006) has also shown the transformation of suspended -Al2 O3 particles into -Al2 O3 by a short additional heat treatment in a ame aerosol process. The predominantly physical aspects of morphology reects its dependence on the above mentioned physical process parameters. The rst chemical modication investigated was the use of dopants. The addition of specic elements (mostly metals) may change the surface energy resulting of nanoparticles and consequently inuence the product morphology. The typical polyhedral CeO2 nanoparticles from ame processing become spherical upon doping with TiO2 and Al2 O3 (Feng et al. 2006). In an even more astonishing example, the addition of indium during ZnO formation promoted one-directional crystal growth and allowed aerosol synthesis of elongated particles (Height et al. 2006). Single metal oxides have found broad applications in the eld of catalysis and photocatalysis since the 1970s. Several excellent reviews (Astruc et al. 2005; Daniel and Astruc 2004; Strobel et al. 2006a) describe these fascinating applications of derived aerosol single oxides both as catalytic material or a support. One of the rst commercial materials from a ame process has been a specic TiO2 grade (Degussa P25) for catalyst substrates and rapidly became the benchmark material in photocatalysis. More recently single oxides have found numerous applications as sensing materials (Madler et al. 2006; Sahm et al. 2004). Among them, TiO2 and SnO2 are widely applied as semiconducting gas sensors. Tricoli et al. (2008) even demonstrated a one-step preparation of dense SiO2 doped SnO2 lm on

wired silicon substrates with the ability to quickly sense ethanol vapor.

Mixed Oxides and Oxide Glasses A decisive advantage of many ame spray methods is the use of low-cost metal precursors with high processing and handling stability. This array of starting materials can often be combined at will and results in an unlimited number of precursor combinations and stoichiometric ratios, enabling the preparation of complicated mixed oxides. As in the case of single metal oxides, the literature has become very broad and only selected examples will be highlighted. General control of the composition (stoichiometric ratio) can be illustrated in the case of solid solutions of CeO2 -ZrO2 ceramics for automotive catalysts (Stark et al. 2003). Still one of the most complicated classes of mixed metal oxides are perovskites for catalysis or for cathode (Gd-doped ceria (Jud et al. 2006), Sm0.5 Sr0.5 CoO3x Sm0.1 Ce0.9 O3x (Liu et al. 2004)) and anode materials in solid oxide fuel cells. The high thermal stability of ame derived perovskites such as NiMnO3 (Kriegel et al. 1994) and BaTiO3 (Brewster and Kodas 1996) enabled maintaining a very high catalytic activity and deactivation-resistance at harsher conditions than conventional materials (cold preparation and sintering). Additionally, mixed oxides have been investigated for electroceramics. Superconducting YBa2 Cu3 O7x lms (Zachariah and Huzarewicz 1991b) and lithium-based spinels (Ernst et al. 2007) have been prepared by ame spray synthesis. The latter exhibited high initial discharge capacity even after several charge/discharge cycles. The ability of metal carboxylates to homogeneously mix and combust even allows materials with four to ve different elements and enables the preparation of amorphous, bioactive glasses (Brunner et al. 2006). Different compositions of SiO2 -CaO-P2 O5 -Na2 O-F nanoparticles were prepared in ames and yielded a highly bioactive material. The combination of bioactivity, biocompatibility, and reactivity motivated for the application of bioglass nanoparticles for remineralization of dentin (the basal tissue for teeth) (Vollenweider et al. 2007) or as a dental ller for root canal treatments (Waltimo et al. 2009).

Downloaded by [203.250.80.20] at 18:04 22 December 2011

From Oxides to Salt Nanoparticles Classical aerosol based processes were used for the manufacturing of metal oxides. More recently, Loher et al. (2005) demonstrated how a ame can be considered a chemical reactor (Figure 1). Chemical reactions, usually preformed in solid-state synthesis, can be readily adapted for the aerosol preparation of salt nanoparticles opening the number of accessible compounds by several orders of magnitude. With the aim of producing calcium phosphate nanoparticles for biomedical applications (bone surgery) a ame spray set-up was fed with a mixture of calcium carboxylate and a combustible phosphate source,

166

E. K. ATHANASSIOU ET AL.

Downloaded by [203.250.80.20] at 18:04 22 December 2011

FIG. 2. Salt nanoparticles can be highly biocompatible and reactive. (a) Transmission electron microscopy image of as-prepared amorphous tricalcium phosphate (Ca3 (PO4 )2 ). The biodegradable nanoparticles can be incorporated in a medically used polymer (PLGA) and processed into millimeter-sized long bers by electrospinning. (b) Such nanocomposites are highly bioactive, promoting hydroxyapatite formation on the ber surface after immersion in simulating bodyuid. This biomineralization enables use of such composites in bone surgery. (c) The morphology of the brous PLGA/TCP scaffold is similar to cotton wool, which greatly facilitates clinical handling.

tributylphosphate: Ca(C8 H15 O2 )2 + 22O2 (CaO)np + 16CO2 + 15H2 O 2(C4 H9 )3 PO4 + 36O2 (P2 O5 )g + 24CO2 + 27H2 O 3 (CaO)np + (P2 O5 )g (Ca3 (PO4 )2 )np The calcium oxide formed in the hot zone of the ame reacted with the simultaneously forming phosphate precursor, affording a dry synthesis of calcium phosphate nanoparticles with particle sizes below 50 nm. By varying the calcium to phosphate ratio in the precursor, a broad range of calcium phosphate phases, including the major biomaterials tricalciumphosphate (Figure 2a) and hydroxapatite, were accessible (Loher et al. 2006). By implementing amorphous tricalcium phosphate (TCP) particles in poly(lactide-co-glucolide), a exible and cotton-wool like biomaterial was produced by electrospinning (Figure 2b) (Schneider et al. 2008). This novel composite conserved the very high bioactivity and absence of cytotoxic effects of TCP nanoparticles and is being studied as bone substitute material for complex bone defects (Figure 2c). This early study indicated that otherwise chemical reactions (typical solid-state synthesis take hours to react) can take place in a very hot zone within milliseconds, mainly ignoring the ongoing physical processes (growth, coalescence, and agglomeration), and paved the way to other salt nanomaterials. Grass and Stark (2005) introduced combustible uoride and chloride sources (chloro- and uorobenzene) to nanoparticle forming ames and obtained the corresponding barium-, calcium-, strontium- uorides as well as sodium chloride, the common kitchen salt. The particle size distribution of these nanoparticles was preserved and could be described as log normal with a geometric standard deviation of = 1.33. This value is consistent with free molecular regime, diffusion-limited coagulation of non-fractal particles as shown from modelling results for

the traditional metal oxide synthesis (Dekkers and Friedlander 2002; Vemury and Pratsinis 1995). This observation conrmed that the physical process remained unaffected even if the chemical changes to the system were large. Besides access to phosphates and halides the concept of exchanging the anion (oxide vs. salt) produced carbonates, such as nano-limestone (Huber et al. 2005), sulfates, nano-gypsum (Osterwalder et al. 2007), and barium carbonates (Strobel et al. 2006b).

Reducing Flames: From Oxides to Metal Nanoparticles The possibility to produce complex salt nanoparticles has showed us that ame reactors can well be considered as chemical reactors operating at higher temperatures. This observation obviously suggested to not stop at anion exchange but fundamentally altering the chemistry of these reactors (Figure 1). Depending on the reaction and combustion conditions in the ame a complex gas mixture can be altered from CO2 /H2 O (oxidizing, traditional ame) to CO/H2 /H2 O (reducing conditions). Noble metal nanoparticles such as Pt, Au, Ag, or their alloys can be obtained readily in oxygen-rich ames but the production of non-noble metals (i.e., most technically important metals) requires reducing conditions. Forsman et al. (2008) showed that cobalt and nickel nanoparticles can be accessed by hydrogen reduction from cobalt or nickel chloride precursor vapor in nitrogen carrier gas. The combustion of cobalt or bismuth organic precursors (cobalt(II)-, bismuth(III)-2-ethylhexanoate) in controlled atmosphere (O2 < 100 ppm) with a high fuel to oxygen ratio (Figure 3) allows the rapid preparation of pure Co and Bi metal nanoparticles by simple extension of the traditional ame process (Grass and Stark 2006a; b). The reducing conditions provoke the full consumption of the available oxygen by H2 and CO and induce the reduction of oxides to metals (Figure 3). After removal from the protective atmosphere, the particles covered

CHEMICAL AEROSOL ENGINEERING

167

Downloaded by [203.250.80.20] at 18:04 22 December 2011

FIG. 3. Classical (left) and reducing ame synthesis lab unit (right) for the production of oxide or metal nanoparticles up to 20 g per hour, respectively. The ame spray pyrolysis is surrounded by a porous tube and placed into a glove box with an inert atmosphere. Controlling the gas ow rates allows highly reducing conditions (O2 < 100 ppm). While traditional reactors mostly yield oxides, this modied process allows the one-step synthesis of metal or alloy or carbon coated metal nanoparticles, or metal/ceramic composites.

core-shell materials showed an excellent air and acid stability and non-traditional electronic properties. The thickness of the carbon layers (23 layers of carbon, about 1 nm or particles with 610 layers, about 3 nm) could be adjusted by the simultaneous feed of acetylene during combustion. When pressed into pills by uniaxial compression this core-shell material exhibited a pronounced temperature and pressure dependent conductivity. The temperature dependence, given by the negative temperature coefcient was similar or even enhanced when compared to commercial piezoelectric materials (Athanassiou et al. 2006). The tunneling based conduction mechanism (Figure 4) showed that core/shell conductive/insulator materials offered a new and simple production route of highly sensitive pressure and temperature sensors (Athanassiou et al. 2008; Athanassiou et al. 2007b). Cu(C8 H15 O2 )2 + 22O2 (CuO)np + 16CO2 + 15H2 O (CuO)np + H2 (Cu)np + H2 O (Cu)np (C/Cu)np Luechinger et al. (2007) implemented carbon/copper nanoparticles in water based dispersions. These aerosol derived metal inks could be used either as humidity sensors (Luechinger et al. 2007) or printed onto exible polymer substrate by conventional ink-jet printing (Luechinger et al. 2008a). The in-situ carbon encapsulation of the metal nanoparticles within an aerosol process gives access to numerous functional materials. The possibility to link organic chemistry to nanoparticles makes aerosol-derived particles amenable to the atom-precision control of synthetic chemistry. The carbon serves as a platform, where covalent surface functionalization with different organic molecular groups via radical chemistry is possible (Grass et al. 2007b; Herrmann et al. 2009). Carbon coated cobalt nanoparticles have been correspondingly functionalized resulting in the preparation of magnetic catalysts (Schatz et al. 2008), magnetic chelating agents for water purication from heavy metal contamination (Koehler et al. 2009) or magnetic actuators (Fuhrer et al. 2009) with muscle-like exibility and magnetic melt-processable polymers (Luechinger et al. 2008b). The successful synthesis of non-noble metal nanoparticles in a ame aerosol based process motivates the use of thermodynamics as a guiding principle to predict possible materials before attempting their preparation by an aerosol route. Assuming equilibrium conditions is reasonable in most cases as chemical gas-phase reactions are exceptionally fast, the gas composition can then be calculated from known thermodynamic properties (Figure 5). The use of straightforward tools, such as the Ellingham diagram, conrms these recent experimental ndings and helps rationalizing how ame synthesis could be extended to reducing or reactive conditions, ultimately enabling the formation of metal alloys such as nickel-molybdenum superalloy nanoparticles with enhanced mechanical properties (Athanassiou et al. 2007a).
CO/H2

themselves with a very thin protective oxide layer enabling further processing of the particles. The combustion of carboxylate precursors under inert atmospheres has shown that the particle size distribution was preserved as in the case of metal oxide and salt nanoparticles and could be described as log-normal resulting in spherical particles with standard geometric deviations of = 1.5 (Grass and Stark 2006a) indicating that the inert (low oxygen concentrations) conditions do not affect the physical processes of nucleation, growth and aggregation. Latter has been further conrmed by the experimental observation of Grass and Stark (2006b) where the combustion of a bismuth precursor under various precursor ow rates and constant fuel to oxygen ratio resulted in metal nanoparticles with unchanged sizes similar to previous studies on the formation of metal oxides. Only dilution of the precursor resulted in a signicant decrease of the powder mean particle size due to smaller particle concentration in the ame. As the problem of oxidation is inevitable, this limits the use of highly pyrophoric materials consisting of elements with a high negative reduction potential such as silicon, aluminum. Therefore, different strategies have been developed to prepare air-stable core/shell metal nanoparticles and composites. The most common techniques are to passivate the surface or to encapsulate it with carbon. Aizawa and Buriak (2007) have incorporated copper nanoparticles in a diblock polymer in a two step process to avoid oxidation, whereas Forsman et al. (2008) suggested that a thin HCl layer on the metal nanoparticles protects them from oxidation. Athanassiou et al. (2006) have shown that the combustion of copper(II)-2-ethylhexanoate under highly reducing conditions resulted in the production of metallic copper nanoparticles covered with thin (1 nm) carbon layers. These

168

E. K. ATHANASSIOU ET AL.

Downloaded by [203.250.80.20] at 18:04 22 December 2011

FIG. 4. An example of a non-traditional aerosol derived nanomaterial. Two core/shell metal/insulator (carbon/coated) nanoparticles in close vicinity can be viewed as a classical tunneling barrier. The energy gap arises from the different Fermi levels in the metal cores and the shell material (graphene bi- or trilayers, right). Electron transport through such an assembly follows multiple parallel paths along (virtual) particle chains. The macroscopic conductivity is a result of a large number of parallel and serial tunneling gaps along the conduction path. Such non-traditional materials mimic the properties of traditional piezoelectric materials (spinels).

Simultaneous Synthesis of Oxides and Metal Nanoparticles Resolves the Mixing Gap Problem in Nanocomposite Preparation The thermodynamic concept outlined above (Athanassiou et al. 2007a) has paved the way to further experimental investigations combining bismuth and cerium containing precursors. This resulted in the simultaneous formation of metal and ceramic nanoparticles that could be compacted to non-traditional composites. The possibility of preparing a homogeneous mixture of metallic bismuth and cerium oxide nanoparticles in a single process step circumvents the otherwise difcult or impossible dry mixing of non-wetting nanoparticles. Pressed pills

of the as-prepared powder resulted in bulk samples with ceramic loadings as high as 25 vol %. For comparison, classical oxide reinforced steel contains 0.51 wt% oxides. These high-oxide loaded nanocomposites combine metallic properties (electrical conductivity, high gloss) and ceramic properties (Vickers hardness comparable to steel, coarse grained steel 100120 HV) while pure bismuth (no oxide) can be indented with a nger nail (Grass et al. 2007a). The successful preparation of salt and metal alloy nanoparticles indicate that aerosol processes can indeed be used as a versatile chemical reactor (Figure 5). Naturally, the basic principles of particle dynamics dominate the morphology and

FIG. 5. Thermodynamic calculations in the form of an Ellingham diagram allow discussions on accessible alloys and metals. The lines which lie below the ame process operating line indicate accessible metals at the corresponding temperature (copper, cobalt, nickel, and iron above 1500 C). Others, such as cerium, manganese are not accessible in the current CO/H2 system.

CHEMICAL AEROSOL ENGINEERING

169

crystallinity of the nanoparticles. Since the characteristic time scale of chemical reactions is often faster, the traditional assumption of constant sintering kinetic parameters may loose its validity in complex chemical transformations. The Arrhenius relationship implies that the chemical reactions proceed very fast at these typically high temperatures. The composition of the combustion gases in the different temperature regimes inuences the nal outcome of the combustion. In addition, the usually (but not always) applied turbulent conditions favor a good mixing of the gas-phase reactants. For all these reasons the problem of considering chemical aerosol engineering in the modeling of gas-phase particle processes is very complicated. A deeper understanding of what exactly happens during particle formation will ultimately provide access to more complex multicomponent functional materials and will facilitate the up scaling of gas-phase aerosol reactors for the fabrication of more complicated nanomaterials beyond oxides such as titania or silica. In order to incorporate chemical processes in modeling and up scaling it is important to implement or develop adequate analytical techniques. Such instrumentation needs to be more accurate than existing methods determining the physical properties of the formed nanoparticles, such as their diameter. Ideally they have to exhibit a high resolution in order to analyze chemical reactions in the micro-millisecond range. Unfortunately, this is not straightforward. Current methods such as Fourier Transform Infrared Spectroscopy have already been applied to measure in situ the temperature during aerosol production. If such methods could be used at better resolution or combined with other analytical methods based on X-ray spectroscopy or diffraction, we could get more phase/chemical information. In situ applications of such methods have already been successfully applied in heterogeneous catalysis where they assist the investigation of complex chemical mechanisms. A space resolved measurement of the gas-phase composition via mass spectroscopy in a gas-phase reactor would enable understanding the gas phase kinetics, which determine the nal chemical composition of nanoparticles. In combination with the temperature prole of the reactor zone this will give a great insight in the complex chemical and physical aspects of aerosol processing, allowing the aerosol community to model and predict the adequate combustion conditions (ow rate of dispersion gas, ow rate of precursor, temperature prole of the reactor zone), for the production of novel nanomaterials with predened physical and chemical properties (particle size distribution, mean particle size, crystallinity, surface area, composition, morphology).

SUMMARY AND OUTLOOK Aerosol engineering has contributed signicantly to the tremendous success of nanomaterials during the past decades. Today, our discipline offers reliable and scalable gas phase methods and some of them have found their way to industrial implementation. The deep and valuable understanding of physical particle dynamics and combustion engineering has enabled the

production of novel and functional materials. The eld continues to expand particularly in the direction of functionalization and chemistry, where we nd plenty of room for more complex materials. This may ultimately require reconsidering fundamental process assumptions, such as constant sintering time or separation of physical and chemical processes. Chemical reaction kinetics will greatly inuence the nal design and operation of next generation aerosol reactors. To achieve that adequate, reliable measurement techniques need to be found or even developed. The wealth of now available nanoparticle-forming aerosol processes have provoked novel, technical nanosolutions. At present many of them demand large amounts of energy, either in the form of electrical energy, fossil sources or by the involvement of large amounts of solvents and chemicals, which have to be puried for reuse or discarded. Therefore, new and improved processes for the synthesis of nanoparticles need to be engineered. Besides obvious properties such as quality of the products, additional focus should be placed on overall energy, resource and recycling if possible (Osterwalder et al. 2006) as well as powder processing. More energy efcient processes directly reduce the costs of nanoparticles and enable more widespread applications. Again, a deeper understanding of the (physical and chemical) fundamentals during nanoparticle formation will provoke optimization of current production methods. Chemical and aerosol engineering still initially provide dry powders. In most technical applications these raw products cannot be directly applied in a nal product. Further processing and derivatization is required. To achieve an optimal engineering solution fundamental interactions of nanoparticles with their surroundings or at interfaces must be investigated at more detail. Most applications require particles dispersed in solvents, polymers or other materials, which typically demand functional modication of the particles surface. The design of such secondstep processes relies on the understanding of the interactions and forces at the nanoscale and how they are inuenced by crystallinity, morphology, size, chemical surface properties, and solubility. An optimum engineering approach considers nanomaterials as functional building blocks. The demand for post processing and derivatization results in rapidly growing applications and initiates the need for understanding the potential health risks related to these novel nanosolutions. Therefore, the physico-chemical interactions of the nanoparticles with living organisms and the environment need to be extensively studied. This will provide the key to the design of inherently safer and sustainable handling and implementation of nanoparticles (Limbach et al. 2009). In summary, the striving for an improved understanding and the combination of physical and chemical engineering in aerosol processes and unit operations will provide access to:

Downloaded by [203.250.80.20] at 18:04 22 December 2011

more complex multicomponent functional materials, more efcient scale-up and development of energyefcient nanoparticle reactor designs,

170

E. K. ATHANASSIOU ET AL. Feng, X. D., Sayle, D. C., Wang, Z. L., Paras, M. S., Santora, B., Sutorik, A. C., Sayle, T. X. T., Yang, Y., Ding, Y., Wang, X. D., and Her, Y. S. (2006). Converting Ceria Polyhedral Nanoparticles Into Single-Crystal Nanospheres. Science 312:15041508. Forsman, J., Tapper, U., Auvinen, A., and Jokiniemi, J. (2008). Production of Cobalt and Nickel Particles by Hydrogen Reduction. J. Nanopart. Res. 10:745759. Friedlander, S. K., and Wang, C. S. (1966). Self-Preserving Particle Size Distribution for Coagulation by Brownian Motion. J. Colloid Interface Sci. 22:126 132. Fuhrer, R., Athanassiou, E. K., Luechinger, N. A., and Stark, W. J. (2009). Crosslinking Metal Nanoparticles Into the Polymer Backbone of Hydrogels Enables Preparation of Soft, Magnetic Field-Driven Actuators with MuscleLike Flexibility. Small 5:383388. Gardner, J., Kumar, S., Cornell, R. M., Mosso, R. J., and Bi, X. (2003). Reactant Delivery Apparatuses, Us2003127316. Gleiter, H. (1989). Nanocrystalline Materials. Prog. Mater. Sci. 33:223315. Grass, R. N., Albrecht, T. F., Krumeich, F., and Stark, W. J. (2007a). LargeScale Preparation of Ceria/Bismuth Metal-Matrix Nano-Composites with a Hardness Comparable to Steel. J. Mater. Chem. 17:14851490. Grass, R. N., Athanassiou, E. K., and Stark, W. J. (2007b). Covalently Functionalized Cobalt Nanoparticles as a Platform for Magnetic Separations in Organic Synthesis. Angew. Chem. Int. Edit. 46:49094912. Grass, R. N., and Stark, W. J. (2005). Flame Synthesis of Calcium-, Strontium-, Barium Fluoride Nanoparticles and Sodium Chloride. Chem. Commun. 1767 1769. Grass, R. N., and Stark, W. J. (2006a). Gas Phase Synthesis of FCC-Cobalt Nanoparticles. J. Mater. Chem. 16:18251830. Grass, R. N., and Stark, W. J. (2006b). Flame Spray Synthesis Under a NonOxidizing Atmosphere: Preparation of Metallic Bismuth Nanoparticles and Nanocrystalline Bulk Bismuth Metal. J. Nanopart. Res. 8:729736. Grass, R. N., Tsantilis, S., and Pratsinis, S. E. (2006). Design of HighTemperature, Gas-Phase Synthesis of Hard or Soft Tio2 Agglomerates. Aiche J. 52:13181325. Grimm, S., Schultz, M., Barth, S., and Muller, R. (1997). Flame PyrolysisA Preparation Route for Ultrane Pure Gamma-Fe2O3 Powders and the Control of Their Particle Size and Properties. J. Mater. Sci. 32:10831092. Height, M. J., Madler, L., Pratsinis, S. E., and Krumeich, F. (2006). Nanorods of Zno Made by Flame Spray Pyrolysis. Chem. Mater. 18:572578. Heine, M. C., and Pratsinis, S. E. (2006). High Concentration Agglomerate Dynamics at High Temperatures. Langmuir 22:1023810245. Hergenroder, R. (2006). Laser-Generated Aerosols in Laser Ablation for Inductively Coupled Plasma Spectrometry. Spectroc. Acta Pt. BAtom. Spectr. 61:284300. Herrmann, I. K., Grass, R. N., Mazunin, D., and Stark, W. J. (2009). Synthesis and Covalent Surface Functionalization of Nonoxidic Iron Core-Shell Nanomagnets. Chem. Mater. 21:32753281. Huber, M., Stark, W. J., Loher, S., Maciejewski, M., Krumeich, F., and Baiker, A. (2005). Flame Synthesis of Calcium Carbonate Nanoparticles. Chem. Commun. 648650. Jiang, W. H., and Yatsui, K. (1998). Pulsed Wire Discharge for Nanosize Powder Synthesis. IEEE Trans. Plasma Sci. 26:14981501. Johannessen, T., Pratsinis, S. E., and Livbjerg, H. (2000). Computational FluidParticle Dynamics for the Flame Synthesis of Alumina Particles. Chem. Eng. Sci. 55:177191. Johannessen, T., Pratsinis, S. E., and Livbjerg, H. (2001). Computational Analysis of Coagulation and Coalescence in the Flame Synthesis of Titania Particles. Powder Technol. 118:242250. Jossen, R., Mueller, R., Pratsinis, S. E., Watson, M., and Akhtar, M. K. (2005). Morphology and Composition of Spray-Flame-Made Yttria-Stabilized Zirconia Nanoparticles. Nanotechnology 16:S609S617. Jud, E., Gauckler, L., Halim, S., and Stark, W. (2006). Sintering Behavior of In Situ Cobalt Oxide-Doped Cerium-Gadolinium Oxide Prepared by Flame Spray Pyrolysis. J. Am. Ceram. Soc. 89:29702973.

development of analytical methods to characterize insitu chemical kinetics during aerosol formation, a better understanding of the inuence of physicochemical parameters on the interface and its interaction with the surrounding material or organism. The challenges and the opportunities for producing these fascinating materials are still huge, in spite of the respectable development during the past thirty years. Most of the future work, however, will happen in areas where the concepts of aerosol material science meet with other engineering disciplines and even with medicine. REFERENCES
Aizawa, M., and Buriak, J. M. (2007). Block Copolymer Templated Chemistry for the Formation of Metallic Nanoparticle Arrays on Semiconductor Surfaces. Chem. Mater. 19:50905101. Arabi-Katbi, O. I., Pratsinis, S. E., Morrison, P. W., and Megaridis, C. M. (2001). Monitoring the Flame Synthesis of Tio2 Particles by In-Situ FTIR Spectroscopy and Thermophoretic Sampling. Combust. Flame 124:560572. Astruc, D., Lu, F., and Aranzaes, J. R. (2005). Nanoparticles as Recyclable Catalysts: The Frontier Between Homogeneous and Heterogeneous Catalysis. Angew. Chem. Int. Edit. 44:78527872. Athanassiou, E. K., Grass, R. N., Osterwalder, N., and Stark, W. J. (2007a). Preparation of Homogeneous, Bulk Nanocrystalline Ni/Mo Alloys With Tripled Vickers Hardness Using Flame-Made Metal Nanoparticles. Chem. Mater. 19:48474854. Athanassiou, E. K., Grass, R. N., and Stark, W. J. (2006). Large-Scale Production of Carbon-Coated Copper Nanoparticles for Sensor Applications. Nanotechnology 17:16681673. Athanassiou, E. K., Krumeich, F., Grass, R. N., and Stark, W. J. (2008). Advanced Piezoresistance of Extended Metal-Insulator Core-Shell Nanoparticle Assemblies. Phys. Rev. Lett. 101:166804. Athanassiou, E. K., Mensing, C., and Stark, W. J. (2007b). Insulator Coated Metal Nanoparticles with a Core/Shell Geometry Exhibit a Temperature Sensitivity Similar to Advanced Spinels. Sens. Actuator A Phys. 138:120129. Bi, X., and Kambe, N. (2001). Efcient Production of Particles by Chemical Reaction, Us6248216. Brewster, J. H., and Kodas, T. T. (1996). Generation of Unagglomerated, Dense, Batio3 Particles by Flame-Spray Pyrolysis, in Aiche Topical Conference On Advanced Ceramics Processing As Part of the 5Th World Congress of Chemical Engineering, American Institute of Chemical Engineers, San Diego, CA, 26652669. Brunner, T. J., Grass, R. N., and Stark, W. J. (2006). Glass and Bioglass Nanopowders by Flame Synthesis. Chem. Commun. 13841386. Camenzind, A., Strobel, R., and Pratsinis, S. E. (2005). Cubic or Monoclinic Y2O3: Eu3+ Nanoparticles by One Step Flame Spray Pyrolysis. Chem. Phys. Lett. 415:193197. Cannon, W. R., Danforth, S. C., Flint, J. H., Haggerty, J. S., and Marra, R. A. (1982). Sinterable Ceramic Powders from Laser-Driven Reactions. 1. Process Description and Modeling. J. Am. Ceram. Soc. 65:324330. Daniel, M. C., and Astruc, D. (2004). Gold Nanoparticles: Assembly, Supramolecular Chemistry, Quantum-Size-Related Properties, and Applications Toward Biology, Catalysis, and Nanotechnology. Chem. Rev. 104:293 346. Dekkers, P. J., and Friedlander, S. K. (2002). The Self-Preserving Size Distribution Theory I. Effects of the Knudsen Number on Aerosol Agglomerate Growth. J. Colloid Interface Sci. 248:295305. Ernst, F. O., Kammler, H. K., Roessler, A., Pratsinis, S. E., Stark, W. J., Ufheil, J., and Novak, P. (2007). Electrochemically Active Flame-Made Nanosized Spinels: Limn2O4, Li4Ti5O12 and Life5O8. Mater. Chem. Phys. 101:372 378.

Downloaded by [203.250.80.20] at 18:04 22 December 2011

CHEMICAL AEROSOL ENGINEERING Kambe, N. (2001). Highly-Uniform Nano-Structured Building Blocks of Metal(O, C, N, S) and Their Complex Compounds. Scripta Materialia 44:1671 1675. Kammler, H. K., Madler, L., and Pratsinis, S. E. (2001). Flame Synthesis of Nanoparticles. Chem. Eng. Technol. 24:583596. Karthikeyan, J., Berndt, C. C., Tikkanen, J., Wang, J. Y., King, A. H., and Herman, H. (1997). Nanomaterial Powders and Deposits Prepared by Flame Spray Processing of Liquid Precursors. Nanostruct. Mater. 8:6174. Kato, M. (1976). Preparation of Ultrane Particles of Refractory Oxides by Gas Evaporation Method. Jpn. J. Appl. Phys. 15:757760. Kim, S., Gislason, J. J., Morton, R. W., Pan, X. Q., Sun, H. P., and Laine, R. M. (2004). Liquid-Feed Flame Spray Pyrolysis of Nanopowders in the Alumina-Titania System. Chem. Mater. 16:23362343. Kinemuchi, Y., Ikeuchi, T., Suzuki, T., Suematsu, H., Jiang, W. H., and Yatsui, K. (2002). Synthesis of Nanosize Pzt Powders by Pulsed Wire Discharge. IEEE Trans. Plasma Sci. 30:18581862. Kinemuchi, Y., Ishizaka, K., Suematsu, H., Jiang, W. H., and Yatsui, K. (2001). Magnetic Properties of Nanosize Nife2O4 Particles Synthesized by Pulsed Wire Discharge, in 14Th Symposium on Plasma Science for Materials (Spsm14), Elsevier Science Sa, Tokyo, Japan, 109113. Ko, T. S., Yang, S., Hsu, H. C., Chu, C. P., Lin, H. F., Liao, S. C., Lu, T. C., Kuo, H. C., Hsieh, W. F., and Wang, S. C. (2006). Zno Nanopowders Fabricated by DC Thermal Plasma Synthesis. Mater. Sci. Eng. B 134:5458. Koch, W., and Friedlander, S. K. (1990). The Effect of Particle Coalesence on the Surface-Area of a Coagulating Aerosol. J. Colloid Interface Sci. 140:419 427. Koehler, F. M., Rossier, M., Waelle, M., Athanassiou, E. K., Limbach, L. K., Grass, R. N., Gunther, D., and Stark, W. J. (2009). Magnetic EDTA: Coupling Heavy Metal Chelators to Metal Nanomagnets for Rapid Removal of Cadmium, Lead and Copper from Contaminated Water. Chem. Commun. 48624864. Kriegel, R., Topfer, J., Preuss, N., Grimm, S., and Boer, J. (1994). Flame PyrolysisA Preparation Route for Ulrane Powders of Metastable BetaSrmno3 and Nimn2O4. J. Mater. Sci. Lett. 13:11111113. Kruis, F. E., Fissan, H., and Peled, A. (1998a). Synthesis of Nanoparticles in the Gas Phase for Electronic, Optical and Magnetic ApplicationsA Review. J. Aerosol. Sci. 29:511535. Kruis, F. E., Kusters, K. A., Pratsinis, S. E., and Scarlett, B. (1993a). A Simple Model for the Evolution of the Characteristics of Aggregate Particles Undergoing Cogulation and Sintering. Aerosol Sci. Technol. 19:514526. Kruis, F. E., Nielsch, K., Fissan, H., Rellinghaus, B., and Wassermann, E. F. (1998b). Preparation of Size-Classied Pbs Nanoparticles in the Gas Phase. Appl. Phys. Lett. 73:547549. Laine, R. M., Marchal, J. C., Sun, H. P., and Pan, X. Q. (2006). Nano-AlphaAl2O3 by Liquid-Feed Flame Spray Pyrolysis. Nat. Mater. 5:710712. Landgrebe, J. D., and Pratsinis, S. E. (1990). A Discrete Section Model for Particulate Production by Gas-Phase Chemical Reaction and Aerosol Coagulation in the Free Molecule Regime. J. Colloid Interface Sci. 139:6386. Lee, P., Suematsu, H., Jiang, W. H., and Yatsui, K. (2006). Synthesis of Al2O3-Zro2 Nanocomposite Powders by Pulsed Wire Discharge. IEEE Trans. Plasma Sci. 34:11901194. Limbach, L. K., Grass, R. N., and Stark, W. J. (2009). Physico-Chemical Differences Between Particle- and Molecule-Derived Toxicity: Can We Make Inherently Safe Nanoparticles? Chimia 63:3843. Liu, Y., Zha, S. W., and Liu, M. L. (2004). Novel Nanostructured Electrodes for Solid Oxide Fuel Cells Fabricated by Combustion Chemical Vapor Deposition (Cvd). Adv. Mater. 16:256260. Loher, S., Reboul, V., Brunner, T. J., Simonet, M., Dora, C., Neuenschwander, P., and Stark, W. J. (2006). Improved Degradation and Bioactivity of Amorphous Aerosol Derived Tricalcium Phosphate Nanoparticles in Poly(Lactide-CoGlycolide). Nanotechnology 17:20542061. Loher, S., Stark, W. J., Maciejewski, M., Baiker, A., Pratsinis, S. E., Reichardt, D., Maspero, F., Krumeich, F., and Gunther, D. (2005). Fluoro-Apatite and Calcium Phosphate Nanoparticles by Flame Synthesis. Chem. Mater. 17:36 42.

171

Luechinger, N. A., Athanassiou, E. K., and Stark, W. J. (2008a). GrapheneStabilized Copper Nanoparticles as an Air-Stable Substitute for Silver and Gold in Low-Cost Ink-Jet Printable Electronics. Nanotechnology 19: 445201. Luechinger, N. A., Booth, N., Heness, G., Bandyopadhyay, S., Grass, R. N., and Stark, W. J. (2008b). Surfactant-Free, Melt-Processable Metal-Polymer Hybrid Materials: Use of Graphene as a Dispersing Agent. Adv. Mater. 20:3044 3049. Luechinger, N. A., Loher, S., Athanassiou, E. K., Grass, R. N., and Stark, W. J. (2007). Highly Sensitive Optical Detection of Humidity on Polymer/Metal Nanoparticle Hybrid Films. Langmuir 23:34733477. Madler, L., Roessler, A., Pratsinis, S. E., Sahm, T., Gurlo, A., Barsan, N., and Weimar, U. (2006). Direct Formation of Highly Porous Gas-Sensing Films by in Situ Thermophoretic Deposition of Flame-Made Pt/Sno2 Nanoparticles. Sens. Actuator B-Chem. 114:283295. Majima, T., Miyahara, T., Haneda, K., Ishii, T., and Takami, M. (1994). Preparation of Iron Ultrane Particles by the Dielectric-Breaksdown of Fe(Co)5 Using a Transversely Excited Atmospheric Co2-Laser and Their Characteristics. Jpn. J. Appl. Phys. Part 1 33:47594763. Majima, T., Miyahara, T., Haneda, K., and Takami, M. (1993). Synthesis of IronCobalt Ultrane Particles by Decomposition of Fe(Co)5 -Co(Co)3 No Using a Tea Co2 Lasser, in Xvith International Conference on Photochemistry, Elsevier Science Sa Lausanne, Vancouver, Canada, 423427. Marchal, J., John, T., Baranwal, R., Hinklin, T., and Laine, R. M. (2004). Yttrium Aluminum Garnet Nanopowders Produced by Liquid-Feed Flame Spray Pyrolysis (Lf-Fsp) of Metalloorganic Precursors. Chem. Mater. 16:822 831. Martelli, S., Mancini, A., Giorgi, R., Alexandrescu, R., Cojocaru, S., Crunteanu, A., Voicu, I., Balu, M., and Morjan, I. (2000). Production of Iron-Oxide Nanoparticles by Laser-Induced Pyrolysis of Gaseous Precursors. Appl. Surf. Sci. 154:353359. Matsoukas, T., and Friedlander, S. K. (1991). Dynamics of Aerosol Agglomerate Formation. J. Colloid Interface Sci. 146:495506. Mueller, R., Jossen, R., Kammler, H. K., and Pratsinis, S. E. (2004a). Growth of Zirconia Particles Made by Flame Spray Pyrolysis. Aiche J. 50: 30853094. Mueller, R., Jossen, R., Pratsinis, S. E., Watson, M., and Akhtar, M. K. (2004b). Zirconia Nanoparticles Made in Spray Flames at High Production Rates. J. Am. Ceram. Soc. 87:197202. Mueller, R., Madler, L., and Pratsinis, S. E. (2003). Nanoparticle Synthesis at High Production Rates by Flame Spray Pyrolysis. Chem. Eng. Sci. 58:1969 1976. Muhlenweg, H., Gutsch, A., Schild, A., and Pratsinis, S. E. (2002). Process Simulation of Gas-to-Particle-Synthesis Via Population Balances: Investigation of Three Models. Chem. Eng. Sci. 57:23052322. Osterwalder, N., Capello, C., Hungerbuhler, K., and Stark, W. J. (2006). Energy Consumption During Nanoparticle Production: How Economic is Dry Synthesis? J. Nanopart. Res. 8:19. Osterwalder, N., Loher, S., Grass, R. N., Brunner, T. J., Limbach, L. K., Halim, S. C., and Stark, W. J. (2007). Preparation of Nano-Gypsum from Anhydrite Nanoparticles: Strongly Increased Vickers Hardness and Formation of Calcium Sulfate Nano-Needles. J. Nanopart. Res. 9:275281. Park, J., An, K. J., Hwang, Y. S., Park, J. G., Noh, H. J., Kim, J. Y., Park, J. H., Hwang, N. M., and Hyeon, T. (2004). Ultra-Large-Scale Syntheses of Monodisperse Nanocrystals. Nat. Mater. 3:891895. Pratsinis, S. E. (1998). Flame Aerosol Synthesis of Ceramic Powders. Prog. Energy Combust. Sci. 24:197219. Pratsinis, S. E., and Vemury, S. (1994). Particle Formation in Gases: A Review, in 1st International Particle Technology Forum, Elsevier Science Sa Lausanne, Denver, Co, 267273. Rosner, D. E. (2005). Flame Synthesis of Valuable Nanoparticles: Recent Progress/Current Needs in Areas of Rate Laws, Population Dynamics, and Characterization. Ind. Eng. Chem. Res. 44:60456055. Roth, P. (2007). Particle Synthesis in Flames. Proc. Combust. Inst. 31:1773 1788.

Downloaded by [203.250.80.20] at 18:04 22 December 2011

172

E. K. ATHANASSIOU ET AL. Suzuki, T., Keawchai, K., Jiang, W. H., and Yatsui, K. (2001). Nanosize Al2O3 Powder Production by Pulsed Wire Discharge. Jpn. J. Appl. Phys. Part 1. 40:10731075. Swihart, M. T. (2003). Vapor-Phase Synthesis of Nanoparticles. Curr. Opin. Colloid Interface Sci. 8:127133. Tani, T., Watanabe, N., and Takatori, K. (2003). Emulsion Combustion and Flame Spray Synthesis of Zinc Oxide/Silica Particles. J. Nanopart. Res. 5:3946. Tricoli, A., Graf, M., and Pratsinis, S. E. (2008). Optimal Doping for Enhanced Sno2 Sensitivity and Thermal Stability. Adv. Funct. Mater. 18: 19691976. Tsantilis, S., Kammler, H. K., and Pratsinis, S. E. (2002). Population Balance Modeling of Flame Synthesis of Titania Nanoparticles. Chem. Eng. Sci. 57:21392156. Tsantilis, S., and Pratsinis, S. E. (2000). Evolution of Primary and Aggregate Particle-Size Distributions by Coagulation and Sintering. Aiche J. 46:407 415. Tsantilis, S., and Pratsinis, S. E. (2004). Soft- and Hard-Agglomerate Aerosols Made at High Temperatures. Langmuir 20:59335939. Vemury, S., and Pratsinis, S. E. (1995). Self-Preserving Size Distributions of Agglomerates. J. Aerosol. Sci. 26:701701. Vollenweider, M., Brunner, T. J., Knecht, S., Grass, R. N., Zehnder, M., Imfeld, T., and Stark, W. J. (2007). Remineralization of Human Dentin Using Ultrane Bioactive Glass Particles. Acta Biomater. 3:936943. Waltimo, T., Mohn, D., Paque, F., Brunner, T. J., Stark, W. J., Imfeld, T., Schatzle, M., and Zehnder, M. (2009). Fine-Tuning of Bioactive Glass for Root Canal Disinfection. J. Dent. Res. 88:235238. Xiong, Y., Akhtar, M. K., and Pratsinis, S. E. (1993). Formation of Agglomerate Oarticles by Coagulation and Sintering. 2. The Evolution of the Morphology of Aerosol-Made Titania, Silica and Silica-Doped Titania Powders. J. Aerosol. Sci. 24:301313. Xiong, Y., and Pratsinis, S. E. (1993). Formation of Agglomerate Particles by Coagulation and Sintering. 1. A 2-dimensional Solution of the Population balance equation. J. Aerosol. Sci. 24:283300. Zachariah, M. R., and Carrier, M. J. (1999). Molecular Dynamics Computation of Gas-Phase Nanoparticle Sintering: A Comparison with Phenomenological Models. J. Aerosol. Sci. 30:11391151. Zachariah, M. R., and Huzarewicz, S. (1991a). Flame Synthesis of High-Tc Superconductors. Combust. Flame 87:100103. Zachariah, M. R., and Huzarewicz, S. (1991b). Aerosol Processing of YBaCuO Superconductors in a Flame Reactor. J. Mater. Res. 6:264269.

Downloaded by [203.250.80.20] at 18:04 22 December 2011

Sahm, T., Madler, L., Gurlo, A., Barsan, N., Pratsinis, S. E., and Weimar, U. (2004). Flame Spray Synthesis of Tin Dioxide Nanoparticles for Gas Sensing. Sens. Actuator B-Chem. 98:148153. Sangurai, C., Kinemuchi, Y., Suzuki, T., Jiang, W. H., and Yatsui, K. (2001). Synthesis of Nanosize Powders of Aluminum Nitride by Pulsed Wire Discharge. Jpn. J. Appl. Phys. Part 1 40:10701072. Schaefer, D. W., and Hurd, A. J. (1990). Growth and Structure of Combustion AerosolsFumed Silica. Aerosol Sci. Technol. 12:876890. Schatz, A., Grass, R. N., Stark, W. J., and Reiser, O. (2008). Tempo Supported on Magnetic C/Co-Nanoparticles: A Highly Active and Recyclable Organocatalyst. Chem. Eur. J. 14:82628266. Schiotz, J., Di Tolla, F. D., and Jacobsen, K. W. (1998). Softening of Nanocrystalline Metals at Very Small Grain Sizes. Nature 391:561563. Schneider, O. D., Loher, S., Brunner, T. J., Uebersax, L., Simonet, M., Grass, R. N., Merkle, H. P., and Stark, W. J. (2008). Cotton Wool-Like Nanocomposite Biomaterials Prepared by Electrospinning: In Vitro Bioactivity and Osteogenic Differentiation of Human Mesenchymal Stem Cells. J. Biomed. Mater. Res. Part B 84B:350362. Spicer, P. T., Chaoul, O., Tsantilis, S., and Pratsinis, S. E. (2002). Titania Formation by Ticl4 Gas Phase Oxidation, Surface Growth and Coagulation. J. Aerosol. Sci. 33:1734. SRI. (2001). Chemical Economics Handbook, Merlo Park, Sri International. Stark, W. J., Madler, L., Maciejewski, M., Pratsinis, S. E., and Baiker, A. (2003). Flame Synthesis of Nanocrystalline Ceria-Zirconia: Effect of Carrier Liquid. Chem. Commun.: 588589. Stark, W. J., M adler, L., and Pratsinis, S. E. (2004). Metal Oxides Prepared by Flame Spray Pyrolysis. WO2004005184. Stark, W. J., and Pratsinis, S. E. (2004). Metal Delivery System for Nanoparticle Manufacture. WO2004103900. Stark, W. J., and Pratsinis, S. E. (2002). Aerosol Flame Reactors for Manufacture of Nanoparticles. Powder Technol. 126:103108. Stark, W. J., Wegner, K., Pratsinis, S. E., and Baiker, A. (2001). Flame Aerosol Synthesis of Vanadia-Titania Nanoparticles: Structural and Catalytic Properties in the Selective Catalytic Reduction of No by Nh3. J. Catal. 197:182191. Strobel, R., Baiker, A., and Pratsinis, S. E. (2006a). Aerosol Flame Synthesis of Catalysts. Adv. Powder Technol. 17:457480. Strobel, R., Madler, L., Piacentini, M., Maciejewski, M., Baiker, A., and Pratsinis, S. E. (2006b). Two-Nozzle Flame Synthesis of Pt/Ba/Al2O3 for Nox Storage. Chem. Mater. 18:25322537. Suryanarayana, C. (2001). Mechanical Alloying and Milling. Prog. Mater. Sci. 46:1184.

Das könnte Ihnen auch gefallen