www.fgks.org   »   [go: up one dir, main page]

Academia.eduAcademia.edu
Surface & Coatings Technology 201 (2007) 8825 – 8829 www.elsevier.com/locate/surfcoat Homoepitaxial silicon carbide deposition processes via chlorine routes A. Fiorucci, D. Moscatelli, M. Masi ⁎ Dipartimento di Chimica Materiali e Ingegneria Chimica “Giulio Natta”, Politecnico di Milano, via Mancinelli 7, 20131 Milan Italy Available online 3 May 2007 Abstract The use of chlorinated precursors recently emerged as the most promising route in SiC CVD but the chemical mechanisms involved in are still not completely defined. Thus, the homoepitaxial SiC film in horizontal hot wall CVD reactors was here analyzed by considering different processes involving chlorinated species. A general but rather simple deposition mechanism involving the most important chemical species was then developed. Reaction rate constants were estimated through classical kinetic theories while most significant reaction rate constants were refined through quantum mechanical methods. The resulting mechanism was embedded into a simplified reactor model and the simulation results were compared against experimental growth rate data. © 2007 Elsevier B.V. All rights reserved. Keywords: [A] Growth models; [C] Chemical vapour deposition; [D] Silicon carbide; [X] Epitaxial deposition 1. Introduction Silicon carbide is commonly used in power electronics and in optoelectronics, for applications like rectifiers (Schottky diodes) and transistors (MOSFET, MESFET). However, SiC market is still limited by difficulties both in growing highly crystalline bulk material and in depositing good quality epitaxial layers via CVD. Thus, to increase the SiC surface quality up to electronic grade, the deposition of epitaxial layers onto SiC substrates is still necessary. Unfortunately, this process is very difficult to control both in terms of film composition and of “flat area” portion of the reactor. Moreover, the traditional recipe based on silane and light hydrocarbons shows many practical problems related to the formation of particulate that practically limits the growth rate values around the 20 μm/h. To ensure both high quality films and high throughput production processes, the use of chlorinated precursors recently emerged as the most promising route in SiC technology where at least three different pathways can be considered, like adding hydrochloric acid to the standard precursors [1,2], using silicon chlorinated silicon precursors [3] or even carbon chlorinated precursors [4]. However, because the carbon–chlorine bond is the weakest one (i.e., C–Cl ≈ 327 kJ/mol vs. Si–Cl ≈ 381 kJ/mol), these last ⁎ Corresponding author. Tel.: +39 0223993131; fax: +39 0223993180. E-mail address: maurizio.masi@polimi.it (M. Masi). 0257-8972/$ - see front matter © 2007 Elsevier B.V. All rights reserved. doi:10.1016/j.surfcoat.2007.04.110 precursors starts to decompose at low temperatures and therefore are of in interest for low temperature (e.g., around 1300 °C), low growth rates processes [4]. On the other hand, the technology is shifting towards the direct use of SiCl4 or SiHCl3 because of their easier process handling with respect to the HCl addition to the SiH4/hydrocarbons mixture. Accordingly, starting from the previously developed mechanisms involving both the HCl addition [5] and the direct feed of SiHCl3 [6], a refined deposition mechanism was here developed. The detailed mechanism originally developed involving 60 gaseous, 19 adsorbed species connected through 233 reactions (75 surface); in this work the mechanism was reduced because of the necessity to have a simpler figure of the system and thus only the really main species were considered. In [7] part of this work was done, but only the SiH4/ hydrocarbon/HCl/H2 system was examined to define the first lumped mechanisms; here the addition of SiHCl3 and CH3Cl was also addressed. This new lumped model can accept as entry point the whole system of precursors usually adopted to grow epitaxial SiC. This was possible because, considering that most of the species considered in the detailed mechanism [5,7] are present in concentrations approaching the delectability threshold, the species here included were only those present in significant amount in the process conditions of interest for the epitaxial deposition. In any case, the involved gas phase and surface reactions were all microscopically reversible and their 8826 A. Fiorucci et al. / Surface & Coatings Technology 201 (2007) 8825–8829 Fig. 1. Complete thermodynamic analysis for the systems (A) SiHCl3/C2H4/H2 with H2 = 150 slm, Si/H2 = 0.040%, C/Si = 1.75 and (B7) SiH4/CH3Cl/H2 with H2 = 150 slm, Si/H2 = 0.040%, C/Si = 1.75. reaction rates were estimated through classical kinetic theories (e.g., collisional and transition state theories) [8]. However, the most significant reaction rate constants were refined also through quantum mechanical methods. The resulting mechanism was finally embedded into a simplified reactor model and the simulation results were compared against experimental growth rate data. chlorine side of the mechanism because the literature provides detailed schemes for the typical carbon precursors (e.g., C3H8, C2H4 and CH4) pyrolysis. For all the involved species, thermodynamics data were taken from SANDIA Thermochemistry database [11] while computations were performed through 2. Thermodynamic analysis At typical process temperatures for the epitaxial SiC processes (i.e., around 1600 °C), only a few reaction paths are really important because the system conditions tends to approach the thermodynamics equilibrium ones. The thermodynamic analysis of SiC growth systems involving chlorinated precursors was already performed also in [9,10], but the examined conditions differ significantly from those examined here, by addressing PVT and HTCVD processes. Thus, a preliminary thermodynamics analysis, like the one illustrated in Fig. 1, can be a useful tool to identify the key gaseous species to be included in the kinetics analysis. In particular, this tool is important to analyze the silicon– Fig. 2. Gas phase kinetic scheme. 8827 A. Fiorucci et al. / Surface & Coatings Technology 201 (2007) 8825–8829 Fig. 3. Sketch of the main reaction pathways involved in the deposition mechanism (silicon species adsorbed on surface carbon and vice versa). Cantera software [12]. In particular, two different systems were analyzed here (i.e., SiHCl3/C3H8/H2 and SiH4/CH3Cl/H2) to test the necessity to include different species while changing the chlorine carrying molecule. Generally, by considering the results obtained for the SiHCl3/ C3H8/H2 system shown in Fig. 1A, the most important species gaseous species on the solid SiC are H2, HCl, H, C2H2, CH4, CH3, C2H4, Si, SiH2, SiCl and SiCl2. It is also very important to notice that most abundant silicon–carbon species (i.e., Si2C and SiC2) show molar fractions below 10− 7; however, because these are the key species for SiC high temperature bulk growth they were still included in the mechanism. Because these results are coherent with other works [13], all of these species were thus included into the lumped scheme developed below. Moreover, the relative species abundance was checked also for the SiH4/ CH3Cl/H2 system, obtaining results comparable with those of the previous system for analogous C/Si/Cl/H molar ratios, as shown in Fig. 1B. also the presence of adsorbed chlorine and chlorinated species, reflecting more the chemistry of the silicon deposition from chlorosilanes [14]. A particular feature of this surface mechanism is to consider all the possible surface reactions among the adsorbed species and not only a “sticking coefficient” incorporation in the film of the gas phase ones. There, all the gaseous species adsorb dissociatively, producing the following adsorbed species H, C, CH, SiH2, SiH, Si, SiCl, Cl where carbon containing species stick on a silicon atom and vice versa for the silicon containing ones, while chlorine can be bonded to both silicon and carbon. Then, these adsorbed species react to desorb gaseous stable species or to form the three solid species SiC, Si2C and SiC2, where, the last two are fictitious ones representing the silicon rich and the carbon rich lattices. 3. Kinetic scheme Reaction As underlined before, the aim of this work is to propose a simple kinetic scheme for SiC growth via chlorinated precursors that allows the prediction of growth rate and main gas phase species. The proposed mechanism includes 19 gaseous and 9 surface species linked by means of 25 gas phase and of 61 surface reactions, as sketched in Figs. 2 and 3, whose reactions and reaction rate constants are summarized in Tables 1 and 2 for the gaseous and the surface reactions, respectively. Carbon precursor reactions start from propane or ethylene because they are the most commonly used precursors. However, the pathways to produce C2H2 or CH4, that are the most abundant carbon containing species, are similar for all the light hydrocarbons. Main silicon precursors examined were SiHCl3 and SiH4, producing SiHX and SiClX (X = 0, 1, 2) as their main fragments. These species, in particular dichlorosilylene (SiCl2) and gaseous Si, are the main responsible for silicon adsorption on the surface and consequently for SiC growth when they interact with the C-containing molecules. The same species are also responsible for the formation of gaseous Si2C and SiC2. On the surface side, the original mechanism, already developed for the silane–hydrocarbons system [5,6], was upgraded to include Table 1 Gas phase kinetic for SiC crystal growth; K = A·Tb·e− E/RT, with parameters coherent with rates in [mol/cm3/s], gas concentrations in [mol/cm3] and activation energy in [cal/mol] 2H + M_H2 + M C3H8_C2H5 + CH3 C2H5_C2H4 + H C2H4 + H_C2H5 C2H4_C2H2 + H2 CH4_CH3 + H CH3 + H2_CH4 + H 2CH3_C2H5 + H CH4 + CH3_C2H5 + H2 SiHCl3_SICl2 + HCl SiH4_SiH2 + H2 Si + H2_SiH2 SiH2 + H_SiH + H2 Si + H_SiH SiH_Si + H 2Si + CH4_Si2C + 2H2 2SiH + CH4_Si2C + 3H2 Si + C2H2_SiC2 + H2 SiH2 + C2H4_SiC2 + 3H Si + HCl_SiCl + H SiCl + HCl_SiCl2 + H SiH3Cl_HCl + SiH2 SiH2 + HCl_SiCl + 3H CH3Cl + H2_CH3 + HCl R R F F R R R R R F R R R F F F F F F F F R F F Log10A β Ea [cal/mol] Reference 18.73 22.42 10.8 8.92 26.5 16.92 2.361 21.06 13.01 14.64 28.2 12.08 14.05 13.3 6.18 19.75 21.3 15.1 13.35 15.4 12.75 14.99 13.27 15.1 −1.3 −1.8 0.4 1.49 −3.5 0 3.12 0 0 0 −4.79 0.5 0 0 0 0.23 0 0 0 −0.2 0 0 0 0 0 88,693.72 39,200 991.513 88,759.29 103,810 8711 26,377.4 22,850 73,683.9 60,450.0 0 15,000 20,000 32,135 5000 0 30,000 0 38,510 17,190 76,000 16,090 75,000 [5] [5] Modified Modified [5] [5] [5] Modified [5] Modified [5] Modified [5] [5] [5] [5] [5] [5] [5] [5] [5] [5] Modified Modified Gas reactions are equilibrium (R) or forward (F). Reaction constants obtained by previous models or modified. 8828 A. Fiorucci et al. / Surface & Coatings Technology 201 (2007) 8825–8829 Table 2 Surface kinetic scheme for SiC crystal growth ( is Si surface site, $2 is C surface site, X⁎ are adsorbed species); K = A·T b·e− E/RT, with parameters coherent with rates in [mol/cm2/s], surface concentrations in [mol/cm2] and activation energy in [cal/mol] Reaction Log10A β Ea [cal/mol] Reference CH4 + $1_C⁎ + 2H2 CH3 + $1_CH⁎ + H2 C2H5 + 2$1_C⁎ + CH⁎ + 2H2 C2H4 + 2$1_2C⁎ + 2H2 C2H2 + 2$1_C⁎ + H2 SiH2 + $2_SiH2⁎ SiH4 + $2_SiH2⁎ + H2 SiH + $2_SiH⁎ Si + $2_Si⁎ SiHCl3 + 2$1 + 2$2_ SiCl⁎ + H⁎ + 2Cl⁎S SiCl2 + 2$2_SiCl⁎ + Cl⁎C SiCl2 + $2 + $1_SiCl⁎ + Cl⁎S HCl + $1 + $2_CH⁎ + Cl⁎S H2 + 2$2_2H⁎ 2CH⁎_C⁎ + H2 H + CH⁎_C⁎ + H2 H + C⁎_CH⁎ CH⁎_0.5H2 + C⁎ CH⁎ + CH⁎_C2H2 + 2$1 CH⁎ + H2_$1 + CH3 CH⁎ + H⁎_H2 + $2 + C⁎ CH⁎ + 0.5H2_C⁎ + H2 2SiH⁎_2Si⁎ + H2 SiH2⁎_Si⁎ + H2 Si⁎_$2 + Si SiCl⁎ + 0.5H2_Si⁎ + HCl SiCl⁎ + H_Si⁎ + HCl SiCl⁎ + HCl_SiCl2 + H + $2 SiCl⁎ + Cl⁎C_SiCl2 + 2$2 SiCl⁎ + Cl⁎S_SiCl2 + $1 + $2 2SiCl⁎_SiCl2 + Si⁎ + 2$2 SiCl⁎ + CH⁎_SiC + $1 + $2 H⁎ + SiCl⁎_HCl + Si⁎ + $2 2Cl⁎C + H2_2HCl + 2$2 Cl⁎C + H_HCl + 2$2 2Cl⁎C + SiCl2_SiCl4 + 2$2 2Cl⁎S + H2_2HCl + 2$2 Cl⁎S + H_HCl + $1 Cl⁎C + Cl⁎S + H2_2HCl + $2 + $1 Cl⁎C + Cl⁎S + SiCl2_SiCl4 + $2$1 CH⁎ + Cl⁎S_HCl + C⁎ + $1 Si⁎ + Cl⁎S_SiCl⁎ + $1 H⁎ + Cl⁎S_HCl + $2 + $1 CH⁎ + Cl⁎C_HCl + C⁎ + $2 Si⁎ + Cl⁎C_SiCl⁎ + $2 H⁎ + Cl⁎C_HCl + 2$2 H⁎ + H⁎_H2 + 2$2 HCl + SiC_SiCl⁎ + CH⁎ Si⁎ + C⁎_SiC + $1 + $2 2Si⁎ + C⁎_Si2C(s) + 2$2 + $1 Si⁎ + 2C⁎_SiC2(s) + $2 + 2$1 SiCl + $2_SiCl⁎ SiC2 + $1 + $2_2C⁎ + Si⁎ 2C⁎ + Si⁎_SiC2 2Si⁎ + C⁎_Si2C Si⁎ + CH⁎_SiC + $1 + $2 + H SiH⁎ + CH⁎_SiC + $1 + $2 + 2H SiH⁎ + C⁎_SiC + $1 + $2 + H Si2C + 2$2 + $1_2Si⁎ + C⁎ SiH3Cl + 2$2_SiCl⁎ + H⁎ + H2 9.38 11.93 20.76 17.97 19.08 11.78 10.5 11.79 11.8 16.42 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0 0 0 0 0 0 18,678 0 0 0 Modified [5] [5] Modified [5] [5] [5] [5] [5] Modified 19.49 19.49 12.55 11.36 23 12.55 12.55 23 23 11.36 23 12.36 25 25 13 15.37 15.52 10.55 19 19 19 17 19 12.37 12.5 11.51 12.37 12.5 12.37 10.51 19 17 19 19 17 19 24 10.55 17 23 23 11.62 20.1 13 13 19 19 19 19.57 16.42 0.5 0.5 0.5 0.5 0 0.5 0.5 0 0 0.5 0 0.5 0 0 0 0.5 0.5 0.5 0 0 0 0 0 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0 0 0 0 0 0 0 0.5 0 0 0 0.5 0.5 0 0 0 0 0 0.5 0.5 0 0 0 0 61,000 0 0 57,100 87,954 87,954 61,000 0 61,000 61,000 40,500 60,000 60,000 0 20,095 89,806 89,806 0 70,006 78,580 0 25,000 89,806 0 84,190 25,000 89,806 0 89,806 84,190 0 84,190 61,000 0 0 0 0 0 0 82,262 55,834 0 0 0 0 0 Modified Modified [5] Modified [5] [5] [5] [5] [5] [5] [5] [5] Modified Modified [5] [5] Modified [5] [5] [5] [5] Modified Modified [5] [5] [5] [5] [5] [5] Modified [5] [5] Modified Modified Modified Modified Modified [5] Modified Modified Modified Modified Modified [5] [5] [5] [5] [5] [5] [5] Table 2 (continued) Reaction Log10A β Ea [cal/mol] SiH⁎ + $2_Si⁎ + H⁎ 17 0 0 Reference [5] All surface are forward reactions. Reaction constants obtained by previous models or modified. The final solid composition was then obtained by considering their relative abundance. 4. Reactor model The developed kinetic scheme was embedded into a simplified CVD reactor model [15,16]. This model solves the mass balances for any involved species and the energy balance, in a heterogeneous ideal plug flow reactor scheme, to evaluate the thermal and the composition profile along the reactor axis. Its use is justified by considering the resulting very high Peclet value (about 1000) in the experimental reactor where the validation data are taken. Here, the mass transport coefficient was estimated by the combination of that calculated for flow in rectangular ducts and that induced by the susceptor rotation. Because the slow rotation speed usually adopted, the former phenomenon dominates and thus the susceptor rotation has its main effect in the azimuthal averaging of the growth rate. 5. Results and discussion The experimental data taken for the model validation were obtained in LPE ACISOne reactor. This reactor holds 6 × 2″ or 3 × 3″ wafers, with an inlet carrier flow rate of 150 slm [3]. The experimental runs were performed at different temperatures (between 1550 and 1650 °C) and with different values of C/Si, Cl/Si and Si/H2 inlet ratios as summarized in Tables 3 and 4 where all the reported runs exhibit mirror like surfaces and lower defect density, compared with traditional processes [3,17]. Runs examined both SiH4/C2H4/HCl/H2 and SiHCl3/ C2H4/H2 systems. As illustrated in Fig. 4, the calculated growth rate values matches the experimental ones within the 5% error, thus well inside the experimental error (e.g., about 10%). It is important to remark that with respect to the starting lumped mechanism [5,7] no corrections were introduced in the reaction rate values besides the addition of the SiHCl3 decomposition to SiCl2, that is a well known reaction with published rate constant [18], and the CH3Cl decomposition, here estimated. The rate Table 3 Main experimental conditions tested in the ACISOne reactor 1 2 3 4 5 0.00012 0.00022 0.00040 0.00045 0.00063 Si/H2 C/Si 2.7 2.5 1.75 1.5 1.5 150 150 150 150 H2 flow (slm) 150 Pressure 100 100 100 100 100 (mbar) Temperature 1550 1600 1600 1650 1650 (°C) Process operated with SiHCl3 + C2H4 + H2. 8829 A. Fiorucci et al. / Surface & Coatings Technology 201 (2007) 8825–8829 Table 4 Main experimental conditions tested in the ACISOne reactor Si/H2 Cl/Si C/Si H2 flow (slm) Pressure (mbar) Temperature (°C) 1 2 3 4 5 6 7 0.0010 2 1.5 150 100 1650 0.0020 1 1 150 100 1650 0.0030 1 1 150 100 1650 0.0035 1 1 150 100 1650 0.0040 1 1 150 100 1650 0.0045 1 0.8 150 100 1650 0.0060 1 0.8 150 100 1650 Process operated with SiH4 + HCl + C2H2 + H2. constant basis set was obtained by means of consolidated reaction rate estimation methods while for the most critical ones (e.g., for reactions forming gaseous SiCl2) a quantum chemical based refinement was performed [19]. It is important to notice that the HCl addition produces processes maintaining high quality surfaces even at higher growth rates than the pure chlorosilanes feed. This point can be related to the reduction in the process degrees of freedom about the Cl/Si ratio. Thus, it is advisable that the industrial process will shift towards the SiHCl3/HCl recipe. The obtained results start to produce a moderate optimism about the understanding of the main features for the deposition mechanism involving chlorinated precursors. Both the addition of HCl and the direct introduction of SiHCl3 were already tested. Because the growth recipes reported in the literature are not significantly different from those here examined it is reasonable to forecast that the reported mechanism will match also experimental data of different sources. Unfortunately, up to now, these data were not available in the open literature and we hope a further test of this mechanism by also other authors. In particular the extension to the HTCVD conditions would be very interesting. Acknowledgments The financial support of LPE Epitaxial Technology is kindly acknowledged. References [1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] [14] [15] [16] [17] [18] [19] Fig. 4. Experimental growth rate values (♦) and calculated ones (———) for the points of (A) Table 3. SiHCl3 + C2H4 + H2 system. (B) Table 4. SiH4 + HCl + C2H2 + H2 system; experimental growth rate data were collected along the whole susceptor and then averaged; however, uniformity for showed points was always below 10%; the same averaging procedure was adopted also for simulated points. O. Kordina, et al., Phys. Status Solidi, B Basic Res. 202 (1) (1997) 321. D. Crippa, et al., Silicon Carbide and Related Materials 2004, 2005, p. 67. S. Leone, et al., Mater. Sci. Forum 527529 (2006) 179. Y. Koshka, et al., J. Cryst. Growth 294 (2) (2006) 260. A. Veneroni, M. Masi, Chem. Vap. Deposition 12 (8–9) (2006) 562. A. Veneroni, F. Omarini, M. Masi, Cryst. Res. Technol. 40 (10–11) (2005) 967. M. Masi, A. Veneroni, ECS Transac. 2 (7) (2007) 11. S.W. Benson, Thermochemical Kinetics, Wiley Interscience, New York, 1976. D. Chaussende, et al., Chem. Vap. Deposition 12 (8–9) (2006) 541. S. Nigam, et al., J. Cryst. Growth 284 (1–2) (2005) 112. SANDIA Thermochemistry database, www.sandia.gov. www.cantera.org. D.D. Avrov, et al., J. Cryst. Growth 199 (1999) 1011. G. Valente, et al., J. Cryst. Growth 230 (1–2) (2001) 247. S. Carra, M. Masi, Prog. Cryst. Growth Charact. Mater. 37 (1) (1998) 1. M. Masi, S. Kommu, Silicon Epitaxy, 2001, p. 185. F. La Via, et al., Chem. Vap. Deposition 12 (8–9) (2006) 509. S.P. Walch, C.E. Dateo, J. Phys. Chem., A 105 (10) (2001) 2015. K. Raghavachari, J.B. Anderson, J. Phys. Chem. 100 (31) (1996) 12960.