$4)',‘.-‘:.€M‘4 g“. _ ‘vj findnm, , A ‘ v :1” : I" ‘ :fifié’fi .:. 1-;3 has. .1 . ,4 . A '- r.‘ ,, 4-1 ~ V« . a , ,L ,.. p. . r. . :fxg.“ WA? 4 .. "in“ gvxk. . WK, 1 9:735; «3‘ u *1.“ 35sz J?"‘l‘ ' ‘ “K V a 1. vi 7:5: w. d. "K‘t 1‘.‘ 1'3“]? A.“ ‘f Fl 1. v in» w. lllllllllllllllllHllHUllllllllllllWIIIJIIIHIHIIIIIHIII 31293 01051 7971 This is to certify that the thesis entitled CHARACTERIZATION OF THE SOL-GEL AND MOCVD PROCESSES FOR THE DEPOSITION OF ALUMINUM OXIDE THIN FILMS presented by Mark Joseph Waner has been accepted towards fulfillment of the requirements for M- S - degree in Chemig try U y‘é‘professo/ Date @‘/ z/?77y 0-7639 MS U is an Affirmative Action/Equal Opportunity Institution LIBRARY Michigan State University PLACE N RETURN BOX to romovo this chockout from your mood. TO AVOID FINES rotum on or bdoro duo duo. ' DATE DUE DATE DUE DATE DUE 1| I l MSU is An Afflrmotivo Action/EM Opporuanity institution Wanna-m CHARACTERIZATION OF THE SOL-GEL AND MOCVD PROCESSES FOR THE DEPOSITION OF ALUMINUM OXIDE THIN FILMS By Mark Joseph Waner A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of MASTER OF SCIENCE Department of Chemistry 1994 ABSTRACT CHARACTERIZATION OF THE SOL-GEL AND MOCVD PROCESSES FOR THE DEPOSITION OF ALUMINUM OXIDE THIN FILMS By Mark Joseph Waner In this Study the sol-gel solution chemistry of a valerate modified aluminum alkoxide is characterized with FTIR The hydrolysis and condensation of aluminum divalerate species in solution leads to the formation of a gel which is then spin cast and dip coated to form thin films. The films are characterized with FTIR, XPS and SEM, before and after calcination. The layered valerate films form continuous aluminum oxide films when calcined at 400°C. The effect of substrate and carrier gas on the deposition of aluminum oxide thin films from aluminum acetylacetonate was examined. Films deposited in air have less carbon contamination than those deposited under nitrogen. When KBr, quartz and SiC were compared, KBr produced the least contaminated films, while SiC showed the largest amount of aluminum deposited. ACKNOWLEDGEMENTS I would like to thank Dr. Jeffrey Ledford for his guidance, assistance and ideas. Thanks also goes out to my co-workers: Ed, Greg, Jefi‘, Kathy, Mike, Paul, Per and Radu. A final thanks to my family who have supported my studies from the beginning. iii TABLE OF CONTENTS Page List of Tables .................................................................................... vi List of Figures .................................................................................. vii Chapter 1: Introduction to the Preparation and Characterization of Metal Oxide Diffusion Barrier Coatings ............................................................................ l 1.1 Overview of Barrier Coatings for use with Silicon Carbide ......... 1 1.2 Sol-Gel Route to Metal Oxide Thin Films ................................... 2 1.3 Metal Organic Chemical Vapor Deposition (MOCVD) ............... 8 1.4 Materials Characterization .......................................................... 13 1.5 Focus of Research ...................................................................... 26 References ........................................................................................ 28 Chapter 2: Characterization of Aluminum Oxide Coatings Derived from Aluminum Valerate Thin Films ........................................................................... 31 2.1 Introduction ............................................................................... 31 2.2 Experimental .............................................................................. 33 2.3 Results and Discussion ............................................................... 35 2.4 Conclusions ................................................................................. 49 References ........................................................................................ 53 Chapter 3: The Effect of Surface Composition on the MOCVD of Alumina fi'om Aluminum Acetylacetonate ................................................................ 55 3.1 Introduction ............................................................................... 55 3 .2 Experimental .............................................................................. 57 3.3 Results and Discussion ............................................................... 61 3.4 Conclusions ................................................................................. 65 References ........................................................................................ 68 iv TABLE OF CONTENTS Chapter 4: Future Work ................................................................... 4.1 Aluminum Sol-Gel Studies ........................................................ 4.2 Aluminum Oxide MOCVD Studies ............................................ 4.3 Structure of Model Metal/SiC Interfaces ................................... References ....................................................................................... 7O 7O 73 75 76 LIST OF TABLES Table Page Table 2.1: Atomic Ratios for Aluminum Valerate Derived Films ............ 47 Table 2.2: Atomic Ratios for Al(acac)3 MOCVD .................................. 62 vi LIST OF FIGURES Figure Page Figure 1.1: CVD Mechanism ........................................................................ 10 Figure 1.2: Example of Precursor Design in MOCVD .................................. 12 Figure 1.3: XPS Energy Level Diagram ....................................................... 16 Figure 1.4: Inelastic Mean Free Path for Low Energy Electrons ................... 18 Figure 1.5: XPS Chemical Shift Efl‘ect ......................................................... 21 Figure 1.6: Limitation of XPS for Thin Film Analysis ................................... 23 Figure 1.7: SEM Schematic ......................................................................... 24 Figure 1.8: SEM-Surface Morphology ......................................................... 25 Figure 2.1: FTIR Spectra for Reactants ....................................................... 36 Figure 2.2: FTIR Spectra for Aluminum Valerate Sol-Gel Solutions ............ . 38 Figure 2.3: Aluminum Valerate Sol-Gel Solution and Dried Films ................ 42 Figure 2.4: FTIR Spectra : Aluminum Valerate Derived Films ...................... 43 Figure 2.5: FTIR Spectra for Al;uminum Valerate Film Calcination .............. 45 Figure 2.6: Carbon and Oxygen XPS for Films ............................................. 46 Figure 2.7: SEM of as Cast Films ................................................................. 48 Figure 2.8: SEM of as Calcined Films .......................................................... 50 Figure 2.9: SEM of Calcined Dip Coated Film ............................................. 5] vii LIST OF FIGURES Figure Figure 3.1: MOCVD Depositor ................................................................... Figure 3.2: FTIR Spectra for Al(acac)3 Deposition ...................................... . Figure 3.3: SEM Images of SiC and KBr Substrates ..................................... Figure 3.4: SEM Images of Quartz Substrate ................................................ viii Page 60 63 66 67 Chapter 1 Introduction to the Preparation and Characterization of Metal Oxide Diffusion Barrier Coatings 1.1. Overview of Barrier Coatings for use with Silicon Carbide Silicon carbide is an important component of many advanced materials. Its high strength, hardness, and 2700°C melting point makes SiC quite usefiil as a structural material. In addition, SiC has a band gap of 2.2 eV and thus is an ideal candidate for a new generation of high temperature electronics capable of handling large amounts of power at high fiequency [1]. In recent years the use of ceramic fibers as a reinforcing material for interrnetallic matrix composites has received a great deal of attention. The fiber/matrix interface acts as a route for energy dissipation during crack propagation. Ideally, the interaction between the fiber and the matrix should be strong enough for load transfer from the matrix to the fiber to occur, but weak enough that the interface will fail before catastrophic failure of the composite. Silicongcarbide _ fibers have been shown to be effective as reinforcements in a unfinzm..- ,_._.— .. -... -—;,m, ~4.- , _. "" "h ‘ u .4-......‘.m--J‘ “I" ""-- ~F-Ir'q»...- “1 _,_ .. variety of ceramic (e.g. A1203, ZrOz) and glass mitiiccs [2-5]. Silicon which; received much interest as a reinforcement material for interrnetallic composites. The use of silicon carbide for this application is generally found to be quite limited, however, due to the reactivity of the fiber with the matrix. It has been shown that SiC will react with many metals, such as Ti [6,7], Nb [8,9], Al [10-12], Fe [13-15] and Ni [16-20] at typical composite fabrication temperatures (> 600°C). Depending on the metal involved, metal carbides or metal silicides may form as shown in the following reactions: SiC + 3N1 —~ Ni3Si + C (1) SiC+ Ti—‘TIC +Si (2) In many cases these reactions can consume the fibers, leaving the brittle metal carbides and silicides which degrade the composite performance [21]. The extent of these reactions may be limited if a diffusion barrier is placed between the SiC fiber and the metal matrix. Some of the more effective barrier coatings in use are the refractory metal oxides. A few examples which have shown some chemical compatibility with both the SiC and metal alloy matrix materials are zirconia, yttria stabilized zirconia and alumina [22-25]. 1.2. Sol-Gel Route to Metal Oxide Thin Films One of the most studied low temperature routes to the production of glass and ceramic materials is the solution sol-gel process. Sol-gel chemistry most often involves the formation of metal oxide polymer networks through themcoflntrolled hydrolysis and polycondensation of precursor materials, typically aqueous metal cations or metal alkoxides. Through careful control of network formation one may obtain a wide range of products, from monodispersed colloids to thin films and monoliths. Monoliths produced by sol-gel techniques may be used as optical filters and lenses. One of the major commercial uses of sol-gel derived thin films is in the coating of glass in order to modify its optical properties. Alternating layers of silica and titania have found application as coatings to produce automotive rear view mirrors since the 1950's [26]. This same technology has also been successfully used for the coating of large plate glass windows with a solar reflecting titania layer for use in modern buildings [26]. The techniques of spray or dip coating, or spin casting are often used for the formation of sol-gel derived thin films [26]. Through low temperature sintering processes sol-gel glasses may be converted into amorphous ceramic materials. Crystalline ceramics may be obtained by high temperature processing. The low sintering temperature enables the fabrication of novel glasses and ceramic materials which are not compatible at the high temperatures of conventional ceramic and glass production. In addition to new combinations of glass and ceramic materials, low temperature processing also enables the modification of inorganic networks with organic functional groups. This is currently of interest to the optical chemical sensor field, since inorganic and organic guest molecules can be incorporated into monoliths and thin films of glass and ceramic materials [27-29]. A sol is best described as a suspension or dispersion of colloidal particles ( 10- 1000 A) in solution. The gel state is not well defined, but can be thought of as an amorphous colloidal solid which has some liquid component dispersed throughout. In practice it is not possible to determine precisely the point at which a sol transforms into a gel or to measure an activation energy for the process. Hench and West define a gel as' a material that can support a stress elastically [30]. As the sol becomes a gel, colloidal particles interconnect. Once discrete particles are connected, the newly formed network continues to connect to other particles and networks. Although many early sol-gel processes involved the use of metal nitrates or halides, the hydrolysis of metal alkoxides has received more attention recently. Besides the ease of hydrolysis, metal alkoxides also have physical and chemical properties which make them appropriate for use as sol-gel precursors. First, synthesis of metal alkoxides is, in general, quite facile. Many alkoxides are synthesized simply by reacting the pure metal with an alcohol [31]: I2 Al + 3C2H50H —- Al(OC2H5)3 + %H2 (3) Another widely used synthesis technique is a ligand exchange reaction utilizing the parent alcohol of the desired alkoxide as the solvent [48]: SiCl4 + 4C2H5OH —’ Si(OC2H5)4 + 4HC1 (4) Secondly, most metal alkoxides are easily purified either by distillation or recrystallization [31]. In addition, the metal alkoxides are soluble in a range of organic solvents, most notably in solutions of their parent alcohol. This property is critical to the sol-gel process. Since most sol-gel syntheses involve the hydrolysis of alkoxides, and are often catalyzed by acids or bases, it is important that the alkoxide be soluble in water miscible solvents such as alcohols. The formation of the inorganic network involves two competing types of reaction: hydrolysis and condensation. The reaction scheme is most simply illustrated as: M(OR)n + XOH —. M(OR)n,x(OX)x + ROH The Lewis acidity of the metal center is the major driving force for this reaction scheme. The reactions occur by a 8N2 nucleophilic substitution reaction illustrated below: 5+ H H \5' 5+ \ 6' O M OR —-> O—M—O—R x/ x/ (5) H / —"" X 0 M 0 ——-> XO—M + ROH R X represents hydrogen for a hydrolysis reaction or a metal center in the case of a condensation reaction. In some cases a third type of substitution reaction which may occur is a complexation reaction, where X represents some nucleophilic ligand. The condensation reaction may occur by two different mechanisms, olation and oxolation. The dominant reaction pathway depends both on the degree of hydrolysis and the stability of the various metal species in found in the solution. H + + fl - 6 8 8 M—OH2 + Ho—M —. M—O—M + H20 (6) V Oxolation may occur along with olation when aquo precursors are available, and exclusively when only hydroxyl species are available. Oxolation is a two step process: first a nucleophilic addition of the hydroxyl group followed by a 1,3 proton transfer. a 5‘ H on M—OH + HO—M—* M—O—M (7) 5* 5' 5* 5' H OH H—OH M—o—M -—- M—O—M —- M—o—M + H20 5‘ 5* The reactivity of the alkoxide has a great deal to do both with the nature of the metal center and the alkoxy group used. The reactivity of an alkoxide will increase with the size and inversely with the electronegativity of the metal center [31,32]. Thus alkoxide reactivity should increase as one goes down a group in the periodic table. The alkoxy ligand can also have a dramatic effect on the hydrolysis and condensation reactions [31,32]. The inductive effect of the alkyl group influences the efi‘ective electropositivity of the metal center. In addition the size and the steric bulk of the alkoxy group can be modified in order to control the reactivity of a metal center [31]. While the mechanistic details of silicon sol-gel chemistry are well established [33- 37], the understanding of other main group and transition metal sol-gel chemistry is quite limited. Further, whereas silicon alkoxides hydrolyze slowly and gel formation is quite slow, the rate of hydrolysis for transition metal alkoxides is many orders of magnitude faster. The two major factors which account for this difference are that most transition metals can have multiple coordinations and, compared to silicon, they are much more electropositive [31]. When multiple coordination numbers are possible, a transition metal complex with a low coordination number may undergo nucleophilic attack through a coordination expansion. The higher electropositivity of the transition metals makes them more susceptible to nucleophilic attack. In general, when dealing with sol-gel processes involving silicon alkoxides the reaction rate must be increased in order to obtain a gel with desired physical properties (i.e. particle size, pore volume, etc). This is usually achieved through the use of acid or base catalysts and small alkoxy ligands such as ethoxide and methoxide. In the case of transition metal sol-gel synthesis, the hydrolysis and condensation reactions often must be slowed down in order to obtain a gel, rather than a precipitate. This control is most often achieved through the use of complexing ligands bound to the metal center. A complexing ligand can lower the reactivity of the metal by decreasing the electropositivity of the metal and/or by steric hindrance. The reagents most often used are organic acids and B- diketones [3 8-47]. Organic acids are more strongly binding ligands than the alkoxides and therefore, make the metal center more stable with respect to hydrolysis. The role of organic acids in the control of titanium and zirconium alkoxide hydrolysis has received some attention in the literature [4143,47-51]. B-diketones are strongly binding chelating ligands which can decrease the rate of nucleophilic attack at the metal center due to the chelation effect and steric hindrance. The hydrolysis of aluminum ions has been investigated for over a century [52]. The classical studies of aluminum sol-gel chemistry involve the hydrolysis of aqueous aluminum ions. In the 1970's Yoldas found that large monoliths of aluminum oxide could be produced through the hydrolysis and polycondensation of aluminum alkoxides [53,54]. Although differing in synthesis, the products obtained by the hydrolysis of aqueous solutions are quite similar to the products obtained by the Yoldas method. Brinker and Scherer note that this similarity may have much to do with the large excess of water in each of the synthesis methods [52]. The study of aluminum sol-gel chemistry in solutions of low water concentration has received very little attention in the literature [52]. The typical sol-gel synthesis involves the hydrolysis of a metal alkoxide in an excess of water (i.e. ~100:l water : alkoxide molar ratio). In some cases small amounts of acid or base are used as a catalyst for the hydrolysis and/or condensation reaction, typically the acid or base to alkoxide ratio is 5 1. The conditions for synthesis are highly dependent on the nature of the alkoxide used. Recently, however, Berglund et al. [48,51] have shown that high quality films may be produced using a large excess of organic acids with titanium and zirconium propoxide systems. In these syntheses low molar ratios of water to alkoxide were used (i.e. ~1.5). One of these studies carefully examined the effect of carboxyl chain length and acid to alkoxide molar ratio on the film quality [48]. For titanium isopropoxide sol-gel it was observed that the best films were obtained when the relatively long chain (valeric and butyric) acids were used, while shorter (propionic) and longer (hexanoic) carboxylic acids gave fairly poor quality films [48]. It was proposed that the shorter chains don't provide sufficient steric or nucleophilic hindrance, while the very long chain acids hinder the polymerization reactions [48]. 1.3. Metal Organic Chemical Vapor Deposition (MOCVD) Chemical vapor deposition (CVD) techniques have been in use for many years as a synthetic route to refractory compounds at reduced temperatures [61]. A typical CVD process involves the reaction of gas phase molecules at a hot substrate surface. A wide variety of reaction types have been utilized, such as pyrolysis, oxidation-reduction, and hydrolysis [55]. The use of reactive pathways to film deposition separates CVD from the physical vapor deposition (PVD) techniques such as evaporation and sputter coating. The classical CVD processes typically involve the reactions of metal halides or hydrides as precursors, and require temperatures in excess of 1000°C. More recently metal-organic chemical vapor deposition (MOCVD) techniques have been widely studied as pathways to coatings at even lower temperatures. An example of the utility of organometallic precursors involves the deposition of silicon nitride. The classical CVD process for deposition of silicon nitride is given by >1000°C _ 8H4 + NH3 _"'" 3131\I4 (8) while Fix et a]. [56] have reported deposition of silicon nitride at temperatures as low as 600°C . 600°C . HXSI(N(CH3)2)4_x—> SI3N4 (9) There are five basic steps in the general CVD mechanism (Figure 1.1), any of which may be the rate determining step for a particular deposition [55]. Once the reactant species are in the gas phase, usually diluted by a carrier gas stream, they must diffuse to , ~___..—u—~--n.;_~u ”Wm ,— the substrate surface. Reactant gases that have diffused through the boundary layer must be adsorbed onto the substrate. The adsorption process is followed by surface chemical interactions which lead directly to the deposition products. Besides various types of reactions, surface interactions may also involve lattice incorporation or surface diffusion [55]. Once the solid reaction products have formed on the substrate, the byproducts of the reaction, which by design should be volatile species, desorb fi'om the surface. The final step in this general mechanism is the diffusion of the byproducts from the surface region into the carrier gas stream. The kinetics of these processes will often determine many of the film characteristics and properties [57,58]. MOCVD processing has become a popular coating technique because of the low temperatures generally involved in the volatilization and decomposition of metal-organic precursors. However, it is the ability to tailor the properties of precursors that has attracted the interest of most chemists. In these methods organo-metallic compounds are designed to provide a core composition around the metal center which is similar in both stoichiometry and structure to the desired coating. Ligands are chosen to enhance the volatility of the precursor and reduce gas phase oligomerization. Figure 1.2 shows an 10 Main flow of reactant gases —-—> -——> —> Gaseous "‘3 ,"" ' by-products V l ; , Boundary G :1. :¢-—@ layer ............ V. .- interface (negligible <23 ‘1' f mun...) Diffusion in of reactants through boundary layer Adsorption of reactants on substrate Chemical reaction takes place Desorption of adsorbed species Diffusion out of by-products 9‘99”“? Figure 1.1: CVD Mechanism [55] ll interesting example of this approach, studied by Interrante et al. [59], which involves the use of [(CH3)2AINH2]3 as a precursor to aluminum nitride. The preparation of useful coatings (i.e., desired stoichiometry and morphology) using MOCVD relies on the interaction of many complex factors. The partial pressures of carrier and reactant gases, temperature gradients, fluid flow dynamics of the system and the composition of the precursor and substrate all influence the structure and composition of the final product [61,60]. The mean free path of the vapor phase reactants and the diffusion properties of the substrate surface will play a major role in the coating morphology [58]. Fischer et al. [61] noted that for deposition involving the pyrolysis of an organometallic compound such as alkoxides and B-diketonates, the use of a carrier gas with an oxygen source tends to limit carbon contamination in the deposited films. Although the effect of various precursors, carrier gases and temperatures for many CVD processes have been studied, relatively little attention has been given to the effect of the substrate composition and structure. There has been much investigation of oxide coating deposition by CVD processes [55,62]. Since the discovery of the high Tc Y-Ba-Cu oxide superconductors, however, the MOCVD technique has become the focus of increased effort for the deposition of oxide coatings [63-65]. Larkin et a1. [59] have used the MOCVD process to deposit thin barrier coatings of yttria onto :silicon carbide fiber at 630°C using yttrium B-diketonate precursors Although this was found to be an effective coating for preventing interfacial reactions, the coatings produced were not continuous. Additionally, in many CYD processes problems arise due to the co-deposition of carbon and Slow—igwr‘owth rates of films While high quality and high purity coatings are commercially produced in electronics applications, the equipment and materials used are costly. Through a better understanding of the effect of all CVD parameters, the use of CVD techniques can be improved. 12 [(CH3)2AINH2] 3 ...... Figure 1.2 : Example of Precursor Design in MOCVD 13 1.4. Materials Characterization In order to understand the sol-gel and MOCVD processes we need to obtain detailed molecular level information. For the MOCVD process we need to understand better the chemical processes which occur at the surface of the substrate. In the sol-gel process we would like to understand the reaction process which leads to the thin film formation and to characterize the structure of the oxide films produced. Fourier transform infrared spectroscopy (FTIR) will provide information which describes the molecular level structures involved in sol-gel syntheses and the films produced. X-ray photoelectron spectroscopy (XPS) will be used to provide both structural information as well as stoichiometric information relevant to the films produced. We will also be using scanning electron microscopy (SEM) in order to examine the surface morphology of the coatings and substrates used. X-ray Photoelectron Spectroscopy (XPS) In XPS an X-ray photon transfers its kinetic energy to an atom, causing the ejection of an electron by the photoelectric effect. Conservation of energy requires that Iw+13itot = Elm, +I~‘.f0t (10) where Bio, and EL, are the total energies of the initial and final states, respectively, and Ekin is the kinetic energy of the photoelectron produced. The binding energy of the electron, referenced to the vacuum level is then equal to v _ f i EB-Etot-Etot (11) 14 When (10) and (1 l) are combined they yield the photoelectric equation hv=Ekin+EE (12) When measuring the binding energy of a solid sample the energy scale is referenced to the Fermi level. When a metal sample is examined, an electrical contact is made between the sample and the spectrometer. Since the sample and spectrometer are now in thermodynamic equilibrium, they have the same Fermi level. However, the vacuum level may be different for the sample and the spectrometer. The work function, (I) is used to describe the difference between the Fermi level and the vacuum level. Photoelectrons feel a potential which is equal to the difference between the spectrometer and sample work functions, which leads to the measured kinetic energy being lower than the kinetic energy of the ejected electron. E2338 = Ekin " (q’spec " (psample) (13) The binding energy, referenced to the Fermi level may then be written as EE=hv-199???" .‘0 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. Gitzen, W.H. Alumina as a Ceramic Material, The American Ceramic Society, Columbus, OH, 1970. Yoldas, B.E. J. app]. Chem. Biotechnol, 1973, 23, 803. Yoldas, B.E. Ceram. Bull, 1975, 54(3), 289. Yoldas, HE J. Mat. Sci, 1975, 10, 1856. Pierre, A.C.; Uhlmann, D.R. J. Am. Ceram. Soc., 1987, 70(1), 28. Harris, M.R.; Sing, K.S.W. J. app]. Chem, 1957, 7, 397. Bye, G.C.; Robinson, J.G. Kolloid Z., 1964, 198,53. Brinker, C.J.; Scherer, G.W. Sol-Gel Science: The Physics and Chemistry of So]- Ge] Processing, Academic Press, Inc., London. 1990. Bye, G.C., Robinson, J.G.; Sing, K.S.W. J. app]. Chem, 1967, 17, 138. Aldcroft, D.; Bye, G.C.; Robinson, J.G.; Sing, K.S.W. J. app]. Chem, 1968, 18, 301. Komarneni, S.; Roy, R.; Fyfe, C.A.; Kennedy, G]. J. Am. Ceram. Soc., 1985, 68(9), C243. Olson, W.L., Bauer, L]. in Mat. Res. Soc. Symp. Proc. 73, Brinker, C.J.; Clark, BE; Ulrich, D.R., Eds; Materials Research Society, Pittsburgh, 1986, 187. Nazar, L.F., Klein, L.C. J. Am. Ceram. Soc., 1988, 71(2), C85. Assih, T.; Ayral, A.; Abenoza, M.;Phalippou, J. .1. Mat. Sci, 1988, 23(9), 3326. Gray, V.R.; Alexander, A.E. J. Phys. Colloid Chem, 1949, 53(1), 9. Gray, V.R.; Alexander, A.E. J. Phys. Colloid Chem, 1949, 53(1), 23. Hood, G.C.; Ihde, A]. J. Am. Chem. Soc., 1950, 72, 2094. Harple, W.W.; Wiberley, S.E.; Bauer, W.H. Anal. Chem, 1952, 24(4), 635. Scott, F.A.; Goldenson, J.; Wiberley, S.E.; Bauer, W.H. J. Phys. Chem, 1954, 58, 61. Mehrotra, R.C.; Pande, K.C. J. Inorg. Nucl. Chem, 1956, 2, 60. Pande, K.C.; Mehrotra, R.C. Z. anorg. allg. Chemie., 1956, 286, 291. Leger, A.E.; Haines, R.L.; Hubley, C.E.; Hyde, J.C., Sheffer, H. Can. J. Chem, 1957, 35, 799. Maksimov, V.N.; Semenenko, K.N.; Naumova, T.N.; Novosleova, A.V. Russ. J. Inorg. Chem, 1960, 5(3), 267. Maksimov, V.N.; Grigor'ev, A.I. Russ. J. Inorg. Chem; 1964, 9(4), 559. Ayral, A.; Phalippou, J.; Droguet, J.C.; in Mat. Res. Soc. Symp. Proc. 121, Brinker, C.J.; Clark, D.E.; Ulrich, D.R., Eds; Materials Research Society, Pittsburgh,1988, 239. Ayral, A.; Droguet, J .C. J. Mater. Res, 1989, 4(4), 967. Dislich, H. in Sol-Gel Technology for Thin Films, F ibers, Preforms, Electronics, and Specialty Shapes, Klein, L.C.,Ed., Noyes Publications, Park Ridge, NJ 1988, 50. Gagliardi, C.D.; Dunuwila, D.;Berg1und, K.A. in Mat. Res. Soc. Symp. Proc. 180, Zelinski, B.J.J.; Brinker, C.J.; Clark, D.E.; Ulrich, D.R.,Eds.; Materials Research Society, Pittsburgh, 1990, 801 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 54 Gagliardi, C.D.; Dunuwila, D.; Van Vlierberge-Torgerson, B.A.; Berglund, K.A. in Mat. Res. Soc. Symp. Proc. 271, Hampden-Smith, M.J.; K1emperer,W.G.; Brinker,C.J., Eds; Materials Research Society, Pittsburgh, 1992, 257. Van Vlierberge-Torgerson, B.A.; Dulebohn, J.I.; Berglund, K.A. in Chemical Processes of Advanced Materials, Hench, L.L.; West, J.K., Eds, John Wiley & Sons, Inc., 1992, 415. Briggs, D.; Seah, M.P.; Practical Surface Analysis, Volume 1: Auger and X-ray Photoelectron Spectroscopy, 2nd Ed, John Wiley & Sons Ltd, Chichester, 1990. Wagner, C.D.; Davis, DE; Zeller, M.V.; Taylor, 1A.; Raymond, R.M.; Gale, l. L.H. Surf Interface Anal, 1981, 3, 211. Googly software provided by Dr. Andrew Proctor, University of Pittsburgh. Barraclough, C.G.; Bradley, D.C.; Lewis, J .; Thomas, I.M. J. Chem. Soc., 1961, 2601. Bell, J.V.; Heisler, J .; Tannenbaum, H.; Goldenson, J. Anal. Chem, 1953, 25(11), 1720. Severin, K.G.; Ledford, J .S.; Torgerson, B.A.; Berglund, K.A. Chem. Mater., in press. Colthup, N.B; Daly, L.H.; Wiberley, S.E. Introduction to Infiared and Raman Spectroscopy, 3rd Ed., Academic Press: San Diego, 1990. Silverstein, R.M.; Bassler, C.G.; Morrill, T.C. Spectrometric Identification of Organic Compounds, 3rd Ed., John Wiley & Sons, Inc., New York, 1974. Ross, S.D.; Inorganic Infrared and Ramon Spectra, McGraw Hill, London, 1972. Smith, A.L.; Applied Infrared spectroscopy, Fundamentals, Techniques, and Analytical Problem-solving, John Wiley & Sons, Inc., New York, 1979. Chapter 3 The Effect of Surface Composition on the MOCVD of Alumina From Aluminum Acetylacetonate 3.1. Introduction The production of refractory compounds at low temperatures using chemical vapor deposition (CVD) techniques has been the focus of much attention for the last few decades [1-3]. Today CVD technology is used to produce a wide range of coatings. Currently there is much interest in the deposition of oxides, nitrides, borides and carbides. Although CVD processes generally occur at much lower temperatures than the melting point of the product, the use of metallo-organic compounds as volatile precursors has led to even further reductions in processing temperature for many compounds. One important refractory compound is aluminum oxide. It is commonly used in optical devices, composite materials and as an electrical insulator. The classic CVD process used to obtain aluminum oxide involves the hydrolysis of aluminum halides at 1050°C [1,4]. OSWC I I Organometallic compounds, however, have been shown to deposit alumina at temperatures as low as 300-500°C [5]. The deposits generally obtained at these very low temperatures are of a polycrystalline or amorphous nature. It has been shown that at temperatures above 800°C these films show 7 and y' alumina phases [6]. Much of the utility of organometallic precursors derives fiom the ability to design complexes which mimic both the stoichiometry and the structure of the desired product. This concept of synthetic design may also be applied to the incorporation of substituents which increase the volatility of the complex [7]. 55 56 Since the discovery of high temperature 1-2-3 superconductors in 1987, the interest in MOCVD has grown due to the prospect of producing superconducting thin films via a low temperature route [8,9]. Much of this literature is focused on volatile metal B-diketonates [10-12]. The B-diketonates are chelating ligands which form ring like structures when bound to metal centers. ’1 c=o~x C/ M \c—c/ l. The synthesis of metal B-diketonates often involves the displacement of alkoxy ligands from a metal alkoxide in a B-diketone solvent [13]. The chelating effect of the ligand makes the metal center less sensitive to hydrolysis, and thus easier to handle than an alkoxide. Like the alkoxides, these compounds possess sufficient oxygen and high enough vapor pressures to be suitable for low temperature deposition of oxides. During the 1960's aluminum acetylacetonate was investigated as a CVD precursor. The literature reports that aluminum oxide deposits were formed at temperatures of 350-500°C [14]. Much of the research in deposition processes has focused on the effect of processing parameters, such as temperature, pressure, reactor design and carrier gas flow rate and composition [2]. Much less attention has been given to the effect of the substrate surface structure and composition on the deposition process. In a recent review of CVD techniques, Besmann et a]. [15] mention that the substrate surface has an effect on growth and nucleation of films, and may act as a catalyst for the deposition reaction, but no examples of such effects were discussed. Bradley [16] and Fischer et a]. [17] mention hydroxyl groups as a catalyst for the decomposition of 57 alkoxides. Lindstrom and Johannesson determined that the nucleation of aluminum oxide onto cemented carbide cutting tools could be enhanced by applying a titanium carbide layer prior to the deposition process [18]. Ryabova and Savitskaya have determined that the deposition temperature, growth rate and activation energy for the pyrolysis increase with the conductivity of the substrate [19]. Recently the deposition of high quality YBazCu3O7-x thin films onto various substrates has been investigated. Commonly the deposition of these compounds is done by using single crystal dielectric materials [20]. Chen et al., however, have studied the YBa2Cu3O7,x films grown on silver substrates, using metal B-diketonate precursors [8]. The present study will examine the deposition of aluminum oxide, using aluminum acetylacetonate as the precursor. The effect of the substrate composition on this process will be examined using quartz, silicon carbide and potassium bromide as substrates. The present study will also investigate the effect of the carrier gas composition, specifically nitrogen and air, on the film deposition onto quartz substrates. 3.2. Experimental Substrates. Reaction bonded SiC plates (Goodfellow, Co.) were polished with 45 um and 6 pm Metadi II diamond polish (Buehler). These polished samples were sonicated in hexane and rinsed with methanol. Samples were then placed into a reactor which could be evacuated to <10 Torr. After evacuation the reactor was backfilled with a 100 ml/min hydrogen (90.5%)(AGA Gas Co., Cleveland OH) flow and the temperature was raised to 500°C for 1 hour. Samples were then placed in a 50% HF / 50% water solution for 8 hours. Samples were dried and rinsed in methanol and treated again in hydrogen. Quartz microscope slides which were cut to 1 cm x 1cm were sonicated in methanol prior to deposition. KBr windows (International Crystal, Inc.; 13 mm x 2 mm) were used without any further cleaning. 58 Deposition. Figure 3.1 shows the vertical reactor used for all the depositions. The carrier gas used was varied, and consisted of either: medical grade air, or nitrogen (99.99 %) (AGA Gas Co., Cleveland OH). The gas flow rate was set to 50 cc/min and regulated by mass flow controllers (Porter Instrument Co.). 0.50 g of aluminum acetylacetonate (Aldrich Chemical Co.) was used as the precursor material for production of aluminum oxide coatings. The depositor was evacuated to a pressure of 1 x 10'2 Torr, then backfilled with carrier gas. After a second evacuation and backfilling, the substrate region was heated to 400i1°C. Once the substrate reached deposition temperature, the precursor was heated to 18811°C regulated using an Omega CN 1200 temperature controller. Once the precursor temperature was reached, the deposition was performed for 45 minutes, after which both heaters were shut off. During the deposition the samples were rotated 60° every 2 minutes. Fourier Transform Infrared Spectroscopy (FTIR). A Mattson 3020 infrared spectrometer equipped with a standard DTGS detector was used for each measurement. KBr crystals obtained fiom International Crystal Inc. were used as substrate materials. Each spectrum is an average of 16 scans at 2 cm'1 resolution. The reference spectrum for alumina was obtained from aluminum oxide powder (anhydrous, Fisher Scientific Co.), which was suspended in a hexane solvent and spray coated onto a KBr window. X-ray Photoelectron Spectroscopy (XPS). Samples were analyzed with a Perkin-Elmer surface science instrument equipped with a Mg anode (1253.6 eV) operated at a power of 300 W (15 kV and 20 mA) and a model 10-360 hemispherical energy analyzer with an omnifocus small spot lens. The instrument typically operates at pressures below 1 x 10"8 Torr in the analysis chamber. Spectra were collected with a constant analyzer pass energy of 50 eV using a PC137 board interfaced to a Zeos 386SX computer. Three regions were examined for each sample: C Is, Al 2p, and 0 Is, in addition to K 2p and Br 3d (for depositions onto KBr) and Si 2p (for depositions onto quartz and silicon 59 carbide). Binding energies were referenced to carbon (C Is = 284.6 eV). Empirical sensitivity factors were used to calculate atomic ratios [21]. XPS spectra were fit with 20% Lorentzian-Gaussian mix Voigt functions using a non-linear least squares curve fitting program [22]. Scanning Electron Spectroscopy (SEM). The quartz and KBr samples were mounted on an aluminum stub with a low vapor pressure double sided adhesive tab (ME. Taylor Engineering, Inc.) and coated with gold in an Emscope sputter coater for 3 minutes at 20 mA. Samples were then analyzed at a 39 mm working distance using a JEOL JSM- 35CF Scanning Electron Microscope. The micrographs shown are the secondary electron images obtained with a beam energy of 10-15 kV. 6O Rotary Feed Exhaust/Pump I: Tube Furnace Substrate Region Precursor 1:. Gas Inlet Vessel Heating Mantle Figure 3.1 :MOCVD Depositor I— 61 3.3. Results and Discussion FTIR Spectra. Figure 3.2 shows the FTIR spectra for the aluminum acetylacetonate precursor, fihn deposited at 400°C on KBr and aluminum oxide powder. The IR band assignments for aluminum acetylacetonate were published by Nakamoto et a]. [2’]. It should be noted that the spectrum collected is missing the C=O stretch at 1545 cm‘l, C-O stretch and C-H bend at 1466 cm'l, and the out of plane bending mode at 425 cm'1 reported by Nakamoto eta]. [23]. In addition, some features are observed which are not present in the reference spectrum for aluminum acetylacetonate, namely those at 1446, 1363, 1018, 951, 812, 800 and 786 cm'l. All of these, except for the 800 and 786 cm'1 bands, are weak and appear as shoulders on the aluminum B-diketonate bands. The spectrum measured for the film deposited at 400°C (Figure 32b) shows bands characteristic for the Al-O stretches, as observed for aluminum oxide powder, and a feature at 1384 cm'1 which can be attributed to the incomplete pyrolysis of the aluminum acetylacetonate. XPS Spectra. The binding energy measured 0 Is for depositions on SiC and KBr substrates (531.3 eV) is lower than for the depositions onto quartz (5317-5323 eV). This is due to the contribution of 0 1s photoelectrons from the quartz substrate. The C Is binding energy, due to oxidized carbon, measured for the deposition on quartz under nitrogen (287. 1-287 .4 eV) is lower than that measured for the other depositions (287.6- 288.0 eV). The Al 2p binding energies measured for films deposited onto quartz (74.5- 75.2 eV) are higher than those measured for the depositions onto KBr and SiC (73.8-74.2 eV). Table 3.1 gives the XPS derived atomic ratios for the coatings deposited on quartz, silicon carbide and potassium bromide substrates, in different gas atmospheres. The oxygen to aluminum ratio was corrected to account for the 0 Is photoelectrons coming from the substrate. To do this the oxygen to silicon (potassium) ratio, measured 62 before deposition was divided by the aluminum to silicon ratio (potassium). The surface enrichment of oxygen may be due in part to the incomplete pyrolysis of aluminum acetylacetonate, or the adsorption of atmospheric water or C02. Table 3.1 : Atomic Ratios for Al(acac)3 MOCVD Gas / Substrate OzAI C:A1 C(ox):A1 Al:Si N2 /SiC 1.9-5.0 $4.9 50.6 18.0-37.2 N2 / SiOz $6.2 4.0-7.2 0.5-1.1 0.4-4.2 Air / 810; 2.4-3.5 2.6—3.8 0.3-0.9 3.2-10.6 C(ox):K AI:K Air IKBr 1.7-2.4 0.7-1.1 0.2-0.3 5.6-6.5 63 l I l (C) 8 S “E 8 '8 Cb) (a) I l l L 4000 3000 2000 1000 frequency (cm' I) (a) Aluminum acetylacetonate (b) 400°C deposition (c) Aluminum oxide powder Figure 3.2: FTIR spectra for Al(acac)3 deposition The carbon to aluminum ratio for the deposition on quartz is lower when air is used as the carrier gas, compared to when nitrogen is used. This efi‘ect of the carrier gas has been seen before for depositions using B-diketonate precursors [17]. The lowest carbon to metal ratio, 0.7-1.1, was observed for the film produced on KBr in an air atmosphere. The oxidized carbon to aluminum ratio measured for the deposition on KBr under air is found to be lower than that of the deposition onto quartz under nitrogen. These ratios indicate that 95% of the oxidized carbon of aluminum acetylacetonate is removed in the deposition on KBr in air, while for quartz under nitrogen only 80-90% removed. This finding may be due more to the carrier gas than the substrate. For the depositions on quartz under air and SiC in nitrogen the C(ox)/A1 ratio indicates 85-95+% removal of oxidized carbon. This carbon may have been removed in the form of acetone, as suggested by Politycki and Hieber [24] O a, II Al(()2CsH7):i W Alxoy + H3C—C-CH3 + Con and/or due to the breaking of the carbon-oxygen bond in the precursor. The aluminum to silicon and aluminum to potassium ratios give an indication as to the relative amount of film growth; although it does not indicate whether the film is continuous and thin, or thick and agglomerated. The Al/SiCK) atomic ratios for the depositions on quartz and KBr were found to lie in the 04-106 range. The aluminum to silicon ratio measured for the deposition on silicon carbide in nitrogen was 18-3 7, possibly indicating a more efficient film grth process. At present the origin of this effect is not known. It is interesting to note, however, that the silicon carbide is a fairly covalent compound, whereas quartz and KBr are more ionic in nature. Another factor which may play a role in this effect is the sample morphology. *r 65 SEM Micrographs. The SEM micrographs of the three substrate materials used for depositions are shown in Figures 3.3 and 3.4. The silicon carbide was found to have a fairly rough surface, with pore sizes of 1-10 pm. When imaged at twice the magnification, the KBr is found to have small particulates, on the order of 0.3 um, and grooves less than 0.1um wide. The quartz substrate is smooth and non-porous, with stepped terrace features. 3.4. Conclusions The FTIR and XPS results indicate that the films deposited from aluminum acetylacetonate at 400°C are structurally comparable to oxygen enriched aluminum oxide. This study has shown that the deposition of aluminum oxide from aluminum acetylacetonate produces cleaner films when air is used as a carrier gas. This study found that the cleanest films were obtained by deposition onto the ionic KBr crystal lattice under an air atmosphere. The most efficient deposition, in terms of relative amount of aluminum deposited, however, was found on the more covalent silicon carbide surface under nitrogen. 66 , . ‘)\ . stucxzaae near 10.8U c5093 ;8HU H4888 8888 Figure 3.3: SEM Images of SiC and KBr Substrates 67 318888 8889 Figure 3.4: SEM Images of Quartz Substrate References 1. Powell, C.; Oxley, J .; Blocher, J. M., Jr. Vapor Deposition, John Wiley & Sons, New York, 1966. 2. Pierson, H.O. Handbook of Chemical Vapor Deposition: Principles, Technology andApplications, Noyes Publications, Park Ridge, NJ, 1992. 3. Vossen, J .L.; Kern, W. Thin Film Processes, Academic Press, New York, 1978 4. Messier, D.R.; Wong, P. J. Electrochem. Soc., 1971, 118, 772. 5. Matsushita, M. ;, Yoga, Y. Electrochem. Soc. Extend Abstr. N0. 90, Spring Meeting, 1968, 230. 6. Aboaf, J. A.; J. Electrochem. Soc., 1967, 114, 948. 7. Interrante, L.V.; Sigel, G.A.; Garbauskas, M.; Hehna, C.; Slack, G.A. Inorg. Chem, 1989, 28(2), 252. 8. Chen, L.; Piazza, T. W.; Schmidt, BE; Kelsey, J. E.; Kaloteros, A. E.; Hazelton, D. W., Walker, M. S.; Luo, L.; Dye, R. C.; Maggiore, C. J.; Wilkins, D. J.; Knorr, D. B. .1. App]. Phys, 1993, 73 (11), 7563. 9. Che, C.; Hwang, D.; No, K.; Chun, J. S.; Kim, S. .1. Mat. Sci, 1993, 28, 2915. 10. Gardiner, R.; Brown, D. W.; Kirlin, P. S.; Rheingold, A. L. Chem. Mater., 1991, 3,1053. 11. Snezhko, N.; Moroz, 8.; Petchurova, N. Mat. Sci. Eng, 1993, 818, 230. 12. Buriak, J. M.; Cheatham, L. K.; Gordon, R. G.; Graham, J. J .; Barron, A. R. Eur. J. Solid State Inorg. Chem, 1992, 29, 43. 13. Mehrotra, R. C.; Bohra, R.; Gaur, D. P. Metal ,6-Diketonates andAllied Derivatives, Academic Press, London, 1978. 14. Politycki, A.; Hieber, K. in Science and Technology of Surface Coating: a NA T 0 Advanced Study Institute April 1972, Ed. Chapman, B. N.; Anderson, J .C., Academic Press, London, 1974. 15. Besmann, T.M.; Stinton, D.P.; Lowden, R.A. MRS Bull., 1988, Nov., 45 16. Bradley, D.C. Chem. Rev., 1989, 89, 1317. 17. Fischer, HE; Larkin, D.J.; Interrante, L.V. MRS Bull., Apr. 1991, 59. 18. Lindstrom, J.N.; Johannesson, R.T. in Proceedings of the Conference on Chamical Vapor Deposition, 5th International Conference 1974, Blocher, J.M., jr.; Hintermann, HE; Hall, L.H., Eds; The Electrochemical Society, 1974, 453 19. Ryabova, L. A.; Sasvitskaya, J. Vac. Sci. T echno]., 6, 934, 1969 20. Chen, L; Piazza, T.W.; Schmidt, BE; Kelsey, J .E.; Kaloyeros, A.E.; Hazelton, D.W.; Walker, M.S.; Luo, L.; Dye, R.C.; Maggiore, C.J.; Wilkins, D.J.; Knorr, D.B. J. Appl. Phys, 1993, 73(11), 7563. 21. Briggs, D.; Seah, M.P.; Practical Surface Analysis, Volume 1: Auger and X-ray Photoelectron Spectroscopy, 2nd Ed, John Wiley & Sons Ltd, Chichester, 1990. Wagner, C.D.; Davis, L.E.; Zeller, M.V.; Taylor, J .A.; Raymond, R.M.; Gale, L.H. Surf Interface Anal, 1981, 3, 211. 22. Googly software provided by Dr. Andrew Proctor, University of Pittsburgh. 23. Nakamoto, K.; McCarthy, P.J.; Ruby, A.; Martell, A.E. J. Amer. Chem. Soc., 68 1961, 83, 1066. 69 24. Politycki, A.; Hieber, K., in Science on Technology of Surface Coatings: A NA T 0 Advanced Stua)I Institute April 1972, Chapman, B.N.; Anderson, J.C., Eds; Academic Press, London & New York, 1974. Chapter 4 Future Work 4.1. Aluminum Sol-Gel Studies Film Synthesis. The air dried films produced in this study appeared to be layered, since a cracked and peeling layer was observed (Figure 2.8), but the substrate was not observed using XPS. In order to produce higher quality oxide films a more complete investigation of the synthesis and processing parameters is necessary. One of the most important factors is the rate of hydrolysis. The chemistry of alkoxides is well known, especially for the common alkoxides of aluminum such as the ethoxide, isopropoxide and sec-butoxide. In general the reactivity of the metal center in an alkoxide is found to decrease with an increase of alkyl chain length and branching. This is due to both the electron donating capabilities of the alkoxy group and also to steric hindrance. This would indicate that the hydrolysis process should proceed more slowly with the sec-butoxide. This will allow us to determine the effect of the residual alkoxide group on the aluminum sol-gel chemistry. Another way to alter the hydrolysis chemistry of the sol-gel process is through the use of complexing agents. In the present work it was shown that valerate ligands were complexed to the aluminum centers. The use of acetic acid, such as in the work of Ayral et a]. [1,2], could give information about the effect of the carboxylate chain length on the hydrolysis rate. Gagliardi et a]. [3,4] have observed a reduction in the hydrolysis rate with the use of longer chain carboxylate species on titanium centers. It would also be interesting to see how the presence of a shorter carboxylate chain would effect the carbon content of the calcined films. B-diketonates are also used as a complexing agent in order to slow down the hydrolysis of transition metals [5-11]. Acetylacetone 7O 71 could be used in place of, or in conjunction with, the carboxylic acids in order to better control the reaction kinetics of hydrolysis and condensation. The hydrolysis rate is expected to be directly proportional to the concentration of water present. In the present experiment water was present as a by product of the esterification reaction between valeric acid and iSOpropanol as well as from the addition of water in the final step of the synthesis. The synthesis could be performed without the addition of water, simply relying on the generation of water fi'om the esterification reaction. This technique has recently been used by Laaziz et a]. [12] for the hydrolysis of zirconium n-propoxide. The present study used a 9:1 molar ratio of carboxylic acid to alkoxide; however, Gagliardi et a]. [3] have shown that this ratio has dramatic efl‘ects on the properties of titanium films. In future studies, a variety of acid ratios could be investigated as a means of better controlling the film properties. Characterization. Although the FTIR and XPS experiments of the current study have proven to be useful in the determination of structural information, they cannot describe the complete picture. For instance the FT IR data was not able to show the presence of any monodentate ligands due to overlapping bands. The FTIR and XPS techniques are also not able to determine the aluminum coordination number or the macromolecular structure of the aluminum valerate network. Multinuclear NMR techniques have proven to be quite useful for the investigation of inorganic polymer gels. The formation of silica gels by the sol-gel route has been widely studied with 29Si NMR [13-16], and the structure of these systems is well known. 27A1 NMR has been used to investigate the structure of alumina gels obtained by the hydrolysis of alkoxides by excess water [17-20]. These studies have shown the effectiveness of this method for probing the coordination of aluminum centers in sol-gel derived gels. Additionally, the use of 1H and 13C NMR might be useful in determining the more precisely the binding of the valerate ligands. Although the FTIR experiments 72 indicate the formation of bridging and chelating valerate ligands, no quantitative information as to the relative abundance of each type of ligand exists. Further, the determination of monodentate species cannot be made with the FTIR experiments due to overlap of the their C-O stretches with other intense bands. Through NMR techniques it may be possible to determine if this type of binding is present. Another technique which has been used to follow the sol-gel process is Raman spectroscopy. Gagliardi et a]. [4] have examined the kinetics of the hydrolysis and condensation processes for titanium sol-gel systems using this technique. Raman spectroscopy could be used to study the time evolution of the aluminum sol-gel process. The present work has focused on gaining molecular level structural information; however, the long range order of the polymer chains in the gels and the calcined films has not been investigated. In the case of crystalline samples, where there is very long range order, diffraction techniques such as X-ray diffraction (XRD) may be used. Although they are amorphous in nature, the gels and films obtained prior to calcination have some long range structure due to the presence metal-oxygen polymer chains. The use of small angle scattering of X-rays (SAXS) has been shown to be quite useful for probing the structure of gels [14-16]. Several theoretical models have been developed for the interpretation of SAXS data for polymer like chains [21,22]. Information related to the polymer branching and fi'actal dimension could be obtained from this study. Effect of Substrate Modification on the Sol-Gel Deposition. One technique used in the synthesis of heterogeneous catalysts involves the use of grafting reactions [23]. This technique typically involves the reaction of a metal alkoxide with a hydroxyl group on a solid surface. This is similar to the alkoxide sol-gel chemistry involved in the current study. It is expected that metal alkoxide derived sol-gel films might adhere better to a hydroxylated glass surface than to a silicon carbide surface. The surface chemistry of silicon carbide has been examined by a number of workers [24-28]. These studies have found that oxidation of the silicon carbide surface 73 produces a surface layer of silica, whose thickness may be tailored by temperature and oxygen partial pressure. Further, the presence of silanol groups on silicon carbide has been observed with FTIR [29]. Sol-gel derived films could be deposited on clean, oxidized and hydroxylated silicon carbide substrates and their adhesion properties examined. SEM and transmission electron microscopy (TEM) investigations of the interface structure could be performed. 4.2 Aluminum Oxide MOCVD Studies Characterization of Surface Coverage. Although some basic structural information has been obtained using XPS and FTIR, the surface coverage of the aluminum oxide films is not known. The deposited films may be very thin, but continuous or they may be thicker, but agglomerated. The lateral resolution of our XPS instrument ( ~3mm) and the escape depth of the photoelectrons (~ 10-100A) limits the use of XPS in this case. If the films are very thin, (< 100A) analysis with SEM is limited. However, the scanning Auger microprobe (SAM) has the same surface sensitivity as XPS combined with lateral resolution approaching 100 nm. Thus, the use of SAM could provide some insight into the surface coverage and nucleation of the films. Another technique which could indicate the surface coverage of the films is ion scattering spectroscopy, which has monolayer surface sensitivity. Evaluation of the MOCVD Process. An extension of the current study is necessary in order to better examine the issue of substrate morphology on the deposition process. Prior to the SEM study of the substrate structure, it was hypothesized that an explanation for the low aluminum content observed for the silica substrates might be explained by a porous surface, which could fill with the aluminum oxide deposit but be too deep within the surface to be observed by the XPS. In fact it is the silicon carbide substrate which is the most porous surface, yet has the most aluminum relative to silicon 74 after deposition. The deposition could be studied as a fimction of surface morphology simply by polishing with 1 um and .25 um diamond polish. The present study found differences between the films deposited on silicon carbide and quartz. As mentioned previously, silicon carbide surfaces can be modified with layers of silica or silanol functions. An extension of the present study would be a study on the effect of surface Si-O and Si-OH groups on the deposition process. The ability to alter the deposition process by simple gas phase treatment could lead to the production of better diffusion barriers. Presently we do not know the composition of the gas phase during the deposition process. For instance, the aluminum acetylacetonate may be partially pyrolyzed before it gets to the substrate surface. A carefiil study of the effects of deposition temperature and carrier gas composition may provide some insight into this. An interesting comparison to the present study would be to use aluminum alkoxides as precursors. The deposition process could be studied as a function of the alkoxide chain length. This could provide a better understanding of the pyrolysis chemistry. Precursor-Substrate Model Studies. Model studies of single crystal substrates might provide some insight into the chemistry involved in the CVD process. Simple gas phase treatments could be used in order to modify single crystal SiC surfaces. After the evaporation of small numbers (sub-monolayer) of precursor molecules (eg. Al(acac)3), electron energy loss spectroscopy (EELS) could be used to probe surface vibrations. This type of information could provide details about the binding mode of precursor to the substrate. This may also lead to further information on the role of the substrate in the deposition process. Temperature programmed desorption (TPD) experiments, in conjunction with EELS could then be used to examine the changes occurring as the precursor decomposes on the surface. 75 4.3 Structure of Model Metal/SiC Interfaces In order to evaluate the effectiveness of the oxide coating as a diffusion barrier, model interfaces could be prepared using Ni, Al, Nb, Fe and Ti. Interfaces could be prepared by hot pressing metal foils onto oxide coated silicon carbide substrates. After annealing at composite processing temperatures (~1000°C) these samples could be sectioned and examined with SEM, SAM and electron microprobe analysis (EPMA). The SEM could be used to evaluate the extent of reaction at both the metal/oxide and oxide/SiC interfaces. The SAM and EPMA could be used to determine structural and compositional information about the reaction zones. The progress of the reactions at the interfacial region could be studied with an in situ experiment. Thin (<20A) layers of metal would be evaporated onto the coated silicon carbide samples in a UHV surface science chamber. XPS would be used to follow the reaction chemistry occurring at different temperatures. 76 References 1. u.) 89‘ 09° 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. Ayral, A.; Phalippou, J.; Droguet, J.C.; in Mat. Res. Soc. Symp. Proc. 121, Brinker, C.J.; Clark, DE; Ulrich, D.R., Eds; Materials Research Society, Pittsburgh” 1988, 239. Ayral, A.; Droguet, J.C. J. Mater. Res, 1989, 4(4), 967 . Gagliardi, C.D.; Dunuwila, D.; Berglund, K.A. in Mat. Res. Soc. Symp. Proc. 180; Zelinski, B.J.J.; Brinker, C.J.; Clark, DE; Ulrich, D.R.,Eds.; Materials Research Society, Pittsburgh, 1990, 801. Gagliardi, C.D.; Dunuwila, D.; Van Vlierberge-Torgerson, B.A.; Berglund, K.A. in Mat. Res. Soc. Symp. Proc. 271, Hampden-Smith, M.J.; K1emperer,W.G.; Brinker,C.J., Eds; Materials Research Society, Pittsburgh, 1992, 257. Mehrotra, R.C.; Gaur, D.P.; Bohra, R. Metal fi-diketonates andAllied Derivatives, Academic Press, London, 1978. Mehrotra, R.C.; Bohra, R. Metal Carboxylates, Academic Press, London, 1983. Sanchez, 0; Livage, J.; Henry, M.; Babonneau, F. J. Non-Cryst. Solids, 1988, 100, 65. Doeuff, S.; Henry, M.; Sanchez, C.; Livage, J. J. Non-Cryst. Solids, 1987, 89, 206. Atik, M.; Aegerter, M.A. J. Non-Cryst. Solids, 1992, 147&148, 813. Chaumont, D.; Craievich, A.; Zarzycki, J. J. Non-Cryst. Solids, 1992, 147&148, 127. Debsikar, J.C. J. Non-Cryst. Solids, 1986, 87, 343. Laaziz, I.; Larbot, A.; Guizard, C.; Julbe, A.; Cot, L. in Mat. Res. Soc. Symp. Proc. 271, Hampden-Smith, M.J.; K1emperer,W.G.; Brinker,C.J., Eds; Materials Research Society, Pittsburgh, 1992. Brinker, 0].; Clark, DE; Ulrich, D.R., Eds; Mat. Res. Soc. Symp. Proc. 32, Materials Research Society, Pittsburgh, 1984. Brinker, C.J.; Clark, DE; Ulrich, D.R., Eds; Mat. Res. Soc. Symp. Proc. 73, Materials Research Society, Pittsburgh, 1986. Brinker, C.J.; Clark, DE; Ulrich, D.R., Eds; Mat. Res. Soc. Symp. Proc. 121, Materials Research Society, Pittsburgh, 1988. Zelinski, B.J.J.; Brinker, C.J.; Clark, DE; Ulrich, D.R.,Eds.; Mat. Res. Soc. Symp. Proc. 180, Materials Research Society, Pittsburgh, 1990 Komarneni, S.; Roy, R.; Fyfe, C.A.; Kennedy, G.J. J. Am. Ceram. Soc., 1985, 68(9), C243. Olson, W.L.; Bauer, L]. in Mat. Res. Soc. Symp. Pros 73, Brinker, C.J.; Clark, DE; Ulrich, D.R., Eds; Materials Research Society, Pittsburgh, 1986. Nazar, L.F.; Klein, L.C. J. Am. Ceram. Soc., 1988, 71(2), C85. Assih, T.; Ayral, A.; Abenoza, M.; Phalippou, J. J. Mat. Sci, 1988, 23(9), 3326. Feigin, L.A.; Svergun, D.I. Structure Analysis by Small-Angle X-ray and Neutron Scattering, Plenum Press, New York: 1987. Schaefer, D.W.; Keefer, K.D. Phys. Rev. Lett., 1984, 53(14), 1383. Baiker, A.; Wokaun, A. Naturwiss, 76, 168, 1989. Costello, J.A.; Tressler, RE J. Am. Ceram. Soc., 1986, 69(9), 674. 25. ~ 26." / 27. 28. 29. 77 Muehlhoff, L.; Bozack, M.J.; Choyke, W.J.; Yates, J.T., jr. J. Appl. Phys, 1986, 60(7), 2558. Bellina, J.J., jr.; Ferrante, J.; Zeller, M.V. J. Vac. Sci. Technol., 1986, A4(3), 1692. Powers, J .M.; Somorjai, G.A. Surf Sci, 1991, 244, 39. Ledford, J .S.; Thelen, M.S.; Waner, M.J. Unpublished results. Ramis, G.; Quintard, P.; Cauchetier, M.; Busca, G.; Lorenzelli, V. J. Am. Ceram. Soc., 1989, 72(9), 1692. re. - «r... *' '1‘st . ’“ 135777325 ' with: 1‘52“ #- ""‘r' P gent-1‘ .2‘ .. _,.; w . . cc .. c. .3 t":.~...‘,r‘- j - J'It‘rfi ‘ » ‘35; L 4, ,Ir a. grant? .136 . I‘m, ' r’ I» "I w- «v . 4 .«p 75: Jn...'"1;' 1,. .’ '47?"th . .. V, ,1 : ‘ J' -‘~'I‘y~‘l); m fuuwv—il m ~- . a s I s .I rm: UM ~7- ,x ?' - as i (sh: I. vb? ; a. ,8}: ex.“ 4’ 4'2:er g “$333.”: I" ‘ ( I guilt ”I ”if: a «L @3135.“ ‘0 .. ’84 ~ we}. _ . alder, .7 n: n :- «rs. -.--i : cavI-TWV‘W' H‘ .1 . sauna-sec: 4 .‘m g ‘3‘ -. «vs-«waver»