Next Article in Journal
Sinterability, Mechanical Properties and Wear Behavior of Ti3SiC2 and Cr2AlC MAX Phases
Next Article in Special Issue
Foam-Replicated Diopside/Fluorapatite/Wollastonite-Based Glass–Ceramic Scaffolds
Previous Article in Journal / Special Issue
Three-Dimensional Finite Element Analysis of Different Connector Designs for All-Ceramic Implant-Supported Fixed Dental Prostheses
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Interfacial Reactions between Si and SiO2 with Ceramic Additives

1
Department of Materials Science and Engineering, National Yang Ming Chiao Tung University, Hsinchu 30010, Taiwan
2
Taiwan Semiconductor Research Institute, National Applied Research Laboratories, Hsinchu 300091, Taiwan
*
Author to whom correspondence should be addressed.
Ceramics 2022, 5(1), 44-54; https://doi.org/10.3390/ceramics5010005
Submission received: 3 November 2021 / Revised: 4 January 2022 / Accepted: 25 January 2022 / Published: 28 January 2022
(This article belongs to the Special Issue Advances in Ceramics)

Abstract

:
In this study, 10 wt.% ceramics—Al2O3, La2O3, Y2O3, MgO, and TiO2—were employed as additives for amorphous SiO2 after pressing and annealing at 1300 °C. The amorphous SiO2 changed to cristobalite SiO2. Through X-ray diffraction, scanning electron microscopy, and transmission electron microscopy with energy-dispersive spectrometry, the reaction phases of La2Si2O7, Y2Si2O7, and MgSiO3 (Mg2SiO4) were found in the SiO2 with 10 wt.% La2O3, Y2O3, and MgO additives. Cracks formed in the Si and SiO2–ceramic additive sites because of the difference in the coefficients of thermal expansion among the Si, SiO2, ceramic additives, and reaction phases. After Si came into contact with the SiO2–ceramics, two types of microstructures were found: those with and those without an amorphous SiO2 reaction layer at the interface. Amorphous SiO2 layer formation is due to the replacement of the Si position in SiO2 by Al3+ and Ti4+ impurities, which can break the bonds between Si atoms. The O content in the Si decreased from 6–9 × 1017 atoms/cm3 for SiO2 to less than ~1016 for SiO2–Al2O3 and SiO2–MgO. The average resistivity of the Si was 3 Ω·cm for SiO2 and decreased to 0.12–0.36 Ω·cm for the SiO2 with ceramic additives.

1. Introduction

Modern integrated circuits and electronic devices are mainly manufactured using single-crystal Si wafers produced using the Czochralski (CZ) pulling technique. The CZ method is based on crystal pulling from Si melt, for which a quartz crucible is placed in the hot zone to melt polysilicon. The quartz crucible is softened at high temperatures during the thermal pulling process; thus, the crucible must be secured by a graphite crucible to prevent deformation. Many types of defects are generated in Si ingots after thermal pulling, such as vacancies, interstitial defects, oxidation-induced stacking faults, and pits, because O from the quartz crucible diffuses into the Si ingot [1,2,3,4,5,6,7].
Quartz is a mineral comprising Si and O atoms in a continuous SiO4 tetrahedral framework, with each O atom shared between two tetrahedra, resulting in the overall chemical formula of SiO2. Si inclusions and worm-like cristobalite SiO2 appear locally at the Si–SiO2 interface because of the solution–precipitation mechanism, as was demonstrated through scanning electron microscopy (SEM) [8]. Local cristobalite SiO2 can result in crack formation on the Si surface. When the Ba concentration in silica glass is >30 ppm, cristobalite SiO2 changes to a dense, smooth, continuous layer and results in decreased precipitation of Si inclusions and defects [9,10]. The Ba-doped cristobalite layer is formed through heterogeneous nucleation caused by Ba additives. In addition to Ba additives [9,10], high-purity silica [11], Si3N4 [12,13,14], and SiC [15] coatings on a crucible can considerably affect the lifetime of Si ingots. A high-purity silica and SiC coating on a crucible can act as a diffusion barrier layer and improve ingot quality [11,15]. Most crucibles used for single-crystal Si growth are made from SiO2 (quartz), but alternative ceramic materials for use in Si crystal growth have gained research attention. Lin et al. [16] studied three types of ceramic plates—Al2O3, ZrO2, and quartz (SiO2)—for contact with a Si wafer through annealing at 1450 °C, and they noted that defects appeared at the Si–ceramic interface. Using SEM and transmission electron microscopy (TEM), they found a crack and a dislocation pile-up at the Si–SiO2 interface; two intermetallic compounds, Y2Si2O7 and ZrSi2, at the Si–ZrO2 interface; and no intermetallic compounds and few defects at the Si–Al2O3 interface. Furthermore, the oxygen concentration and electrical resistivity near the interface were high, and they gradually decreased as the distance from the Si–ceramic interface increased [16]. Therefore, the ceramics in contact with Si could result in defect formation, oxygen aggregation, and/or the reaction barrier layers on the surface of Si. It is significant to clarify which types of ceramics affect the defects, oxygen diffusion, and resistivity change at the Si interface after being in contact with Si. The purpose of the present study is to choose different valance electrons of ceramics, such as Mg2+, Al3+, La3+, Y3+, and Ti4+, to place in contact with Si and then evaluate the electrical property of Si due to the charge balance effect and the formation of defects and effective barrier layers on the Si–ceramics. In the present study, 10 wt% ceramics—Al2O3, La2O3, Y2O3, MgO, and TiO2—were separately added to amorphous SiO2 after pressing and annealing at 1300 °C for single-crystal Si growth. The microstructures of the Si/SiO2–additive interfaces were characterized using SEM and TEM in conjunction with energy-dispersive spectroscopy (EDS). In addition, the concentration of O and the resistivity of the Si after contact with the ceramics were measured and are discussed herein.

2. Experimental Procedures

Al2O3 (aluminum oxide power, Sigma-Aldrich, St. Louis, MO, USA, ≤10 μm, 99.5%), La2O3 (lanthanum oxide, Cerac, Inc., Milwaukee, WI, USA 325 mesh, 99.9%), Y2O3 (yttrium oxide, Sigma-Aldrich, St. Louis, MO, USA, 99.999%), MgO (magnesium oxide nano powder, Inframat Advanced Materials, Manchester, CT, USA, 30-nm, 99.9%), and TiO2 (titanium oxide, NOAH Technologies, San Antonio, TX, USA, 325 mesh, 99.9%) powders (10 wt.%) were separately added to amorphous SiO2 (quartz (silicon dioxide), UniRegion Bio-Tech, Hsinchu, Taiwan, 40 nm, 99.9%) powder and pressed at 5 tons for 1 min to obtain one-inch discs (thickness ≈ 2 mm). These green bodies were placed in an atmospheric furnace after sintering at 1300 °C for 2 h (Figure 1). Figure 1 shows a photograph of the green bodies and sintered SiO2 and SiO2 with ceramic additives, showing the volume shrinkage and color change in all the samples after sintering. Linear shrinkage was calculated using the formula Ls = [(d0 − d)/d0) × 100%], where Ls is the linear shrinkage, d0 is the diameter of the green body, and d is the diameter of the sintered sample. The linear shrinkage of all sintered samples was similar: 27.36% for SiO2, 24.97% for SiO2 with 10 wt.% Al2O3, 26.66% for SiO2 with 10 wt.% La2O3, 27.77% for SiO2 with 10 wt.% Y2O3, 24.52% for SiO2 with 10 wt.% MgO, and 27.40% for SiO2 with 10 wt.% TiO2. The densities of all sintered samples were calculated using the Archimedes method, and the relative densities were ~90%. Next, the sintered samples were interfaced with Si (p-type-doped boron, CZ growth method, orientation of [100], resistivity ≈ 15–25 Ω·cm, thickness = 655–695 μm, O content = 12.00–15.00 ppm) to evaluate the reaction between the Si and the SiO2 with ceramic additives. Then, the sintered SiO2–ceramics were ground, polished, and interfaced with Si in a rectangular graphite crucible with inner dimensions of 23 × 23 × 20 mm3 and outer dimensions of 33 × 33 × 25 mm3, as shown in Figure 2a,b. The crucibles containing the Si/SiO2–ceramics were placed in a vacuum furnace, vacuumed to approximately 10−4 Torr, and annealed at 1450 °C (slightly higher than the melting temperature of Si: 1414 °C) under an Ar atmosphere. After holding this temperature for 30 min, the furnace was cooled to room temperature; then, the graphite crucibles were cut and removed, as shown in Figure 2c. For the O concentration and Si resistivity measurements after the reactions between the Si and SiO2–ceramics, several pieces of Si were cut longitudinally with respect to the interface, as shown in Figure 2c. For the phase identification and microstructural characterization of the SiO2–ceramics, SEM (Model JSM 6500F, JEOL Ltd., Tokyo, Japan), TEM (Model JEM 2010F, JEOL Ltd., Tokyo, Japan), and X-ray diffraction (XRD; X’Pert Pro MRD, Malvern Panalytical Ltd., Malvern, UK) were performed before contact with Si. The crystal structure of the reaction products at the interface was characterized by selected area diffraction patterns (SADPs) of TEM and EDS. The Inorganic Crystal Structure Database (ICSD), Joint Committee on Powder diffraction Standards (JCPDS) database and crystallographic software (Diamond version 3.0 and CaRIne Crystallography 3.1) were used for the identification of the crystal structures of the phases. The cross-sectional TEM specimens of Si–ceramic joints were prepared by conventional mechanical polishing and focused ion beam (FIB, FEI NovaLab 600). The quantitative composition analyses were performed based on the principle of the Cliff–Lorimer [17] standard-less method. Additionally, in order to accurately calculate the lattice parameters of reaction products, the image magnification of TEM was calibrated using the MAG*I*CAL reference standard sample (Norrox Scientific Ltd., Beaver Pond, ON, Canada). The pieces of Si from Si–ceramic joints were cut along a direction longitudinal to the interface for the residual oxygen and resistivity measurements. The concentration evaluation of the residual oxygen in Si by Fourier-transform infrared spectroscopy (FTIR, Bruker VERTEX 70) was performed based on the ASTM F 1188 method [18], and the resistivity of Si was measured by four-point probe resistance meter (RT-80, Napson Corporation, Tokyo, Japan).

3. Results and Discussion

Figure 3 shows the XRD spectra of the pure SiO2 and the SiO2 with 10 wt.% Al2O3, La2O3, Y2O3, MgO, and TiO2. Before sintering, the crystal structure of the SiO2 tended to be in the amorphous phase (broad peak at ~20°) and then changed to a combination of the amorphous phase and cristobalite tetragonal phase (reference code: 01-076-0940, crystal system: tetragonal, space group: P41212(92), a = b = 4.9964 Å, c = 7.0169 Å, α = β = γ = 90°) after sintering at 1300 °C, as indicated by the sharp peak at ~20° in Figure 3a. The Al2O3 and TiO2 additives did not react with the SiO2 to form other phases (Figure 3b,f). The other ceramic additives, La2O3, Y2O3, and MgO, reacted with the SiO2 to form other reaction phases in the samples. For example, La2Si2O7, Y2Si2O7, and MgSiO3 (Mg2SiO4) phases were found with the addition of 10 wt.% La2O3 (Figure 3c), Y2O3 (Figure 3d), and MgO (Figure 3e), respectively. The corresponding chemical formulas are 2SiO2 + La2O3 = La2Si2O7, 2SiO2 + Y2O3 = Y2Si2O7, and SiO2 + MgO = MgSiO3 or SiO2 + 2MgO = Mg2SiO4.
Figure 4 shows the SEM and backscattered electron images (BEIs) for the SiO2 and the SiO2 with 10 wt.% Al2O3, La2O3, Y2O3, MgO, and TiO2 additives, showing that no other reaction phase was formed after the addition of Al2O3 (Figure 4b) or TiO2 (Figure 4f). However, the reaction phases La2Si2O7, Y2Si2O7, and MgSiO3 were found with the La2O3 (Figure 4c), Y2O3 (Figure 4d), and MgO (Figure 4e) additives. The reaction phases were further confirmed using TEM. Figure 5 shows the TEM bright-field images (BFIs) and the selected area diffraction patterns (SADPs) of the Si with 10 wt.% Al2O3, La2O3, Y2O3, MgO, and TiO2 additives after sintering at 1300 °C in Ar. Through SADP indexing and EDS and TEM, the composition of Al2O3, La2Si2O7, Y2Si2O7, MgSiO3, and TiO2 (Figure 6f) and the reaction phases (La2Si2O7, Y2Si2O7, and MgSiO3) in the SiO2 with ceramic additives were identified.
Through XRD (Figure 3), SEM (Figure 4), and TEM (Figure 5), the microstructures and reaction phases were thoroughly examined in the sintered SiO2–ceramic additives. After the reaction between the Si and the sintered SiO2–ceramic additives, two types of microstructures were identified: those with and those without a reaction layer at the interface. Figure 6 shows the BEIs/SEM images for the Si/SiO2 and Si/SiO2–La2O3, Si/SiO2–Y2O3, and Si/SiO2–MgO interfaces, showing the absence of a reaction layer. Cracks formed in the Si and SiO2–ceramics because of a coefficient of thermal expansion (CTE) mismatch. XRD (Figure 3) revealed amorphous SiO2 (broad peak) and cristobalite SiO2 in the SiO2 and SiO2–ceramics. The difference in the CTEs of amorphous SiO2 (0.5 × 10−6/K) [19] and cristobalite SiO2 (14.5 × 10−6/K) [20] is large and, thus, resulted in cracks in the SiO2 (Figure 6a). In Figure 6b–d, cracks are observable in the SiO2–ceramics (with 10 wt.% La2O3, Y2O3, and MgO) because of the CTE mismatch between the SiO2 and the reaction phases La2Si2O7 (6 × 10−6/K) [21], Y2Si2O7 (3.9 × 10−6/K) [22], and MgSiO3 (47.7 × 10−6/K) [23]. The CTEs of cristobalite SiO2, La2Si2O7, Y2Si2O7, and MgSiO3 are larger than those of Si (2.6 × 10−6/K) [24]. Therefore, the Si endured compressive stress at the SiO2 site, resulting in cracks near the interface. Furthermore, the CTE of MgSiO3 is high (~47.7 × 10−6/K) and, therefore, resulted in severe cracks in the Si near the interface (Figure 6d).
O in Si severely hinders Si ingot growth [3,4]. To evaluate the O content in the Si near and far from the Si–SiO2 interface, EDS line scanning of TEM was performed. Figure 7 shows the BFIs/TEM images and EDS line scans of the Si/–SiO2, Si/SiO2–La2O3, Si/SiO2–Y2O3, and Si/SiO2–MgO interfaces, showing the elemental distribution at the SiO2 and Si sites. The O content was approximately 10 at.% in the Si near the interface for the SiO2 with no ceramic additives (Figure 7a). When ceramics were added, the O content in the Si near the interface decreased to approximately 3 at.% for the La2O3 additive (Figure 7b) and approximately 8 at.% for the Y2O3 and MgO additives (Figure 7c,d).
A reaction layer formed at the interface of the Si with SiO2–Al2O3 and SiO2–TiO2, as shown in the BEIs/SEM images in Figure 8. TEM was used to analyze the reaction layer at the interface (Figure 9). Figure 9a,b show the BFIs/TEM images, EDS line scans, and SADPs for the Si/SiO2–Al2O3 and Si/SiO2–TiO2 interfaces, respectively. No diffraction spots were found in the SADP of the reaction layer or the cristobalite SiO2, as shown by the spot patterns (white arrows) in the SADP shown in Figure 9a,b. The reaction layer mainly comprised Si and O, as well as small quantities of Al (~8 at.%) and Ti (~2 at.%), as shown in the EDS images in Figure 9a,b, respectively. From the SADP and EDS results, the reaction layer could be identified as amorphous SiO2 with small quantities of dissolved Al or Ti. Furthermore, the amorphous SiO2 reaction layer effectively restricted O diffusion into the Si, as indicated by the O content of approximately 2 at.% and <0.8 at.% in the EDS line scans in Figure 9a,b, respectively. The mechanism underlying the formation of the amorphous SiO2 reaction layer is shown in Figure 10. Figure 10a shows a schematic of crystalline SiO2, revealing the tetrahedral arrangement of one Si atom bonded to four O atoms; in the image, most O atoms are bridged to two Si atoms, and two tetrahedra are joined at a corner to form the regular crystal structure of quartz. When the SiO2–Al2O3 and SiO2–TiO2 came into contact with the Si substrate, the Si, O, Al, and Ti atoms diffused toward the Si site. The O from the SiO2 site easily diffused into the Si substrate; this finding is supported by the diffusivity of the Al (1.073 × 10−11 cm2/s) [25], Ti (1.162 × 10−8 cm2/s) [26], Si (8.610 × 10−14 cm2/s) [27], and O (1.429 × 10−10 cm2/s) [28] in the Si at 1200 °C. The results indicate that Al atoms could remain at the SiO2 site because of their lower diffusivity than that of the O near the interface. Another possible explanation is the formation of native oxynitride on the Si surface, which blocks Al and Ti diffusion into the Si [29]. Thus, numerous Al and Ti atoms aggregated at the SiO2 near the interface and resulted in the formation of amorphous SiO2 because of the replacement of the Si position in the SiO2 with Al3+ and Ti4+ impurities, which can break the bonds between Si and O, as shown in Figure 10b [30].
For the O content and resistivity measurements, several (~800 µm thick) pieces of Si were cut longitudinally with respect to the interface after contact between the Si and SiO2 with additives, as shown in Figure 2d. Figure 11a shows the distribution of the O concentration in the Si pieces and the distance from the Si/SiO2–ceramic interface, as measured through FTIR spectroscopy. The O concentration of the Si was 10.5–13.2 × 1017 atoms/cm3 (12–15 ppm) before contact with the SiO2. The O concentration of the Si after contact with SiO2 decreased to ~6–9 × 1017 atoms/cm3, probably because of SiO evaporation during the heat treatment. However, the O concentrations of all the SiO2–ceramics were lower than those of the pure SiO2. Among them, the O concentrations of SiO2–Al2O3 and SiO2–MgO were less than 1016, implying that ceramic additives in SiO2 affect the diffusion of O into the Si substrate. Figure 11b shows the electrical resistivity of the Si pieces at various distances from the interface of the Si/SiO2–ceramics, as measured using a four-point probe resistance meter. Similarly, the electrical resistivity of the Si near the interface was high and gradually decreased away from the interface for all samples. For example, the resistivity of the Si near the interface was approximately 5 Ω·cm and decreased to as low as 3 Ω·cm far from the Si/SiO2 interface. The resistivity of the Si was 21.85 Ω·cm before contact with the SiO2. As anticipated, the resistivity of the Si both near and far from the Si–ceramic interface was lower than that of pure Si. Hull [31] indicated that unintentional O diffusion into Si usually manifests as n-type doping (with SiO4 as the donor) and results in a decrease in the resistivity of Si. When Si comes into contact with SiO2, the O from the SiO2 diffuses into the Si, causing a decrease in resistivity; this is supported by our resistivity measurements shown in Figure 11b. The resistivity of the Si with all additives (0.12, 0.24, 0.21, 0.36, and 0.30 Ω·cm for Y2O3, La2O3, TiO2, Al2O3, and MgO, respectively) was less than that of the Si (~3 Ω·cm) in contact with the pure SiO2. When Al3+, La3+, and Y3+ ions diffused and replaced one or two (for Mg2+) Si atoms in the substrate, the remaining valence electrons were insufficient to satisfy the four covalent neighboring bonds of the Si, resulting in the formation of holes. Thus, the Al, La, and Y(IIIA) atoms became acceptors. As for the diffusion of the Ti and the Si originating from the SiO2 into the Si substrate, the electrical property of the Si substrate could not be changed because of the charge balance effect (Si4+ = Ti4+). However, the dissolution of Al, Mg, Y, La, Ti, Y, and O impurities in the Si substrate resulted in interstitial or vacancy defects, thus affecting the resistivity.

4. Conclusions

In summary, 10 wt% ceramics—Al2O3, La2O3, Y2O3, MgO, and TiO2—were separately added to amorphous SiO2 after pressing and annealing at 1300 °C, resulting in the amorphous SiO2 changing to cristobalite SiO2. The reaction phases La2Si2O7, Y2Si2O7, and MgSiO3 (Mg2SiO4) in the SiO2 with 10 wt.% La2O3, Y2O3, and MgO additives were identified through XRD, SEM, and TEM. After Si came into contact with SiO2–ceramics, observed by SEM and TEM, two types of microstructures were found: those with and those without an amorphous SiO2 reaction layer at the interface. No reaction layer was found with the La2O3, Y2O3, and MgO additives; by contrast, a reaction layer was found with the Al2O3 and TiO2 additives. The formation of the amorphous SiO2 reaction layer was due to the substitution of the Si position of the SiO2 with Al3+ and Ti4+ impurities, which can break the bonds between Si atoms. The O content in the Si far from the interface decreased to 6–9 × 1017 atoms/cm3 for pure SiO2 and less than 1016 atoms/cm3 for SiO2–Al2O3 and SiO2–MgO, implying that ceramic additives in SiO2 affect the diffusion of O into the Si substrate. The resistivity of the Si in the SiO2 decreased from 3 to 0.12–0.36 Ω·cm with the ceramic additives using a four-point probe resistance meter due to charge balance effect.

Author Contributions

Conceptualization, Y.-H.C. and K.-L.L.; Investigation, Y.-H.C. and K.-L.L.; Methodology, Y.-H.C. and K.-L.L.; Project administration, Y.-H.C.; Supervision, K.-L.L. and C.-C.L.; Writing—original draft, Y.-H.C. and K.-L.L.; Writing—review & editing, K.-L.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Ministry of Science and Technology, Taiwan, grant number MOST 110-2221-E-492-002 and MOST 110-2119-M-002-018–MBK.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

The authors would like to thank the Ministry of Science and Technology, Taiwan, for financially supporting this research under contract NO. MOST 110-2221-E-492-002 and MOST 110-2119-M-002-018–MBK.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Válek, L.; Śik, J. Defect engineering during Czochralski crystal growth and silicon wafer manufacturing. In Modern Aspects of Bulk Crystal and Thin Film Preparation; Kolesnikov, N., Ed.; InTech: London, UK, 2012; ISBN 978-953-307-610-2. [Google Scholar] [CrossRef] [Green Version]
  2. Hwang, D.H.; Lee, B.Y.; Yoo, H.D.; Kwon, O.J. Anomalous oxygen precipitation near the vacancy and interstitial boundary in CZ-Si wafers. J. Cryst. Growth 2000, 213, 57–62. [Google Scholar] [CrossRef]
  3. Itsumi, M. Octahedral void defects in Czochralski silicon. J. Cryst. Growth 2002, 237, 1773–1778. [Google Scholar] [CrossRef]
  4. Furukawa, J.; Tanaka, H.; Nakada, Y.; Ono, N.; Shiraki, H. Investigation on grown-in defects in CZ-Si crystal under slow pulling rate. J. Cryst. Growth 2000, 210, 26–30. [Google Scholar] [CrossRef]
  5. Minami, T.; Maeda, S.; Higasa, M.; Kashima, K. In-situ observation of bubble formation at silicon melt–silica glass interface. J. Cryst. Growth 2011, 318, 196–199. [Google Scholar] [CrossRef]
  6. Xu, J.; Wang, W.; Yang, D.; Moeller, H.J. Transmission electron microscopy investigation of the micro-defects in Czochralski silicon. J. Alloy. Compd. 2009, 478, 758–762. [Google Scholar] [CrossRef]
  7. Lanterne, A.; Gaspar, G.; Hu, Y.; Øvrelid, E.; Di Sabatino, M. Characterization of the loss of the dislocation-free growth during Czochralski silicon pulling. J. Cryst. Growth 2017, 458, 120–127. [Google Scholar] [CrossRef]
  8. Schnurre, S.M.; Schmid-Fetzer, R. Reactions at the liquid silicon/silica glass interface. J. Cryst. Growth 2013, 250, 370–381. [Google Scholar] [CrossRef]
  9. Huang, X.; Koh, S.; Wu, K.; Chen, M.; Hoshikawa, T.; Hoshikawa, K.; Uda, S. Reaction at the interface between Si melt and a Ba-doped silica crucible. J. Cryst. Growth 2003, 277, 154–161. [Google Scholar] [CrossRef]
  10. Huang, X.; Hoshikawa, T.; Uda, S. Analysis of the reaction at the interface between Si melt and Ba-doped silica glass. J. Cryst. Growth 2007, 306, 422–427. [Google Scholar] [CrossRef]
  11. Kvande, R.; Arnberg, L.; Martin, C. Influence of crucible and coating quality on the properties of multicrystalline silicon for solar cells. J. Cryst. Growth 2009, 311, 765–768. [Google Scholar] [CrossRef]
  12. Nakajima, K.; Murai, R.; Morishita, K.; Kutsukake, K. Growth of Si single bulk crystals inside Si melts by the noncontact crucible method using silica crucibles without coating Si3N4 particles. In Proceedings of the 2013 IEEE 39th Photovoltaic Specialists Conference (PVSC), Tampa, FL, USA, 16–21 June 2013; pp. 0174–0176. [Google Scholar]
  13. Huguet, C.; Dechamp, C.; Camel, D.; Drevet, B.; Eustathopoulos, N. Eustathopoulos. Study of interactions between silicon and coated graphite for application to photovoltaic silicon processing. J. Mater. Sci. 2019, 54, 11546–11555. [Google Scholar] [CrossRef]
  14. Hendawi, R.; Ciftja, A.; Arnberg, L.; Di Sabatino, M. Kinetics of silicon nitride coatings degradation and its influence on liquid infiltration in PV silicon crystallization processes. Sol. Energy Mater. Sol. Cells 2021, 15, 111190. [Google Scholar] [CrossRef]
  15. Camel, D.; Cierniak, E.; Drevet, B.; Cabal, R.; Ponthenier, D.; Eustathopoulos, N. Eustathopoulos. Directional solidification of photovoltaic silicon in re-useable graphite crucibles. Sol. Energy Mater Sol. Cells 2020, 215, 110637. [Google Scholar] [CrossRef]
  16. Lin, C.Y.; Lin, K.L.; Lin, C.C. Reactions between Si melt and various ceramics. Process. Appl. Ceram. 2019, 13, 115–123. [Google Scholar] [CrossRef] [Green Version]
  17. Cliff, G.; Lorimer, G.W. The quantitative analysis of thin specimens. J. Microsc. 1975, 130, 203–207. [Google Scholar] [CrossRef]
  18. ASTM Standard F 1188; Standard Test Method for Interstitial Atomic Oxygen Content of Silicon by Infrared Absorption. ASTM International: West Conshohocken, PA, USA, 2003.
  19. Souder, W.; Hidnert, P. Measurements on the Thermal Expansion of Fused Silica; Scientific Papers of the Bureau of Standards No. 524; National Bureau of Standards: Washington, DC, USA, 1926; pp. 1–23.
  20. Sweeney, W.T. Cristobalite for dental investment. J. Am. Dent. Assn. 1933, 20, 108–119. [Google Scholar]
  21. Fukuda, K.; Asaka, T.; Uchida, T.K. Thermal expansion of lanthanum silicate oxyapatite (La9.33+2x(SiO4)6O2+3x), lanthanum oxyorthosilicate (La2SiO5) and lanthanum sorosilicate (La2Si2O7). J. Solid State Chem. 2021, 194, 157–161. [Google Scholar] [CrossRef]
  22. Sun, Z.; Zhou, Y.; Wang, J.; Li, M. Thermal properties and thermal shock resistance of γ-Y2Si2O7. J. Am. Ceram. Soc. 2008, 91, 2623–2629. [Google Scholar] [CrossRef]
  23. Hugh-Jones, D. Thermal expansion of MgSiO3 and FeSiO3 ortho- and clinopyroxenes. Am. Mineral. 1977, 82, 689–696. [Google Scholar] [CrossRef]
  24. Zhou, S.Q.; Vantomme, A.; Zhang, B.S.; Yang, H.; Wu, M.F. Comparison of the properties of GaN grown on complex Si-based structures. Appl. Phys. Lett. 2005, 86, 081912. [Google Scholar] [CrossRef] [Green Version]
  25. Navon, D.; Chernyshov, V. Retrograde solubility of aluminum in silicon. J. Appl. Phys. 1957, 28, 823–824. [Google Scholar] [CrossRef]
  26. Kuge, S.; Nakashima, H. Solubility and diffusion coefficient of electrically active titanium in silicon. Jpn. J. Appl. Phys. 1991, 30, 2659–2663. [Google Scholar] [CrossRef]
  27. Peart, R.F. Self diffusion in intrinsic silicon. Phys. Status Solidi. 1966, 15, K119–K122. [Google Scholar] [CrossRef]
  28. Logan, R.A.; Peters, A.J. Diffusion of oxygen in silicon. J. Appl. Phys. 1959, 30, 1627–1630. [Google Scholar] [CrossRef]
  29. Guha, S.; Gusev, E.P.; Okorn-Schmidt, H.; Copel, M.; Ragnarsson, L.Å.; Bojarczuk, N.A.; Ronsheim, P.S. High temperature stability of Al2O3 dielectrics on Si: Interfacial metal diffusion and mobility degradation. Appl. Phys. Lett. 2002, 81, 2956–2958. [Google Scholar] [CrossRef]
  30. Rottier, B.; Rezeau, H.; Casanova, V.; Kouzmanov, K.; Moritz, R.; Schlöglova, K.; Wälle, M.; Fontboté, L. Trace element diffusion and incorporation in quartz during heating experiments. Contrib. Mineral. Petrol. 2017, 172, 23. [Google Scholar] [CrossRef]
  31. Hull, R. Properties of Crystalline Silicon; The Institution of Electrical Engineers: London, UK, 1999. [Google Scholar]
Figure 1. Photograph of green bodies and sintered (1300 °C for 2 h) samples of SiO2 and SiO2 with ceramic additives.
Figure 1. Photograph of green bodies and sintered (1300 °C for 2 h) samples of SiO2 and SiO2 with ceramic additives.
Ceramics 05 00005 g001
Figure 2. Schematic of (a) graphite crucible and (b) Si and SiO2–ceramic additive interface. (c) Cross-sectional photograph of Si contact with SiO2–TiO2. (d) Diagram of several pieces of Si cut longitudinally with respect to the interface after contact for evaluation of O concentration and electrical resistivity.
Figure 2. Schematic of (a) graphite crucible and (b) Si and SiO2–ceramic additive interface. (c) Cross-sectional photograph of Si contact with SiO2–TiO2. (d) Diagram of several pieces of Si cut longitudinally with respect to the interface after contact for evaluation of O concentration and electrical resistivity.
Ceramics 05 00005 g002
Figure 3. XRD spectra of (a) SiO2 and (bf) SiO2 with 10 wt.% Al2O3, La2O3, Y2O3, MgO, and TiO2, respectively.
Figure 3. XRD spectra of (a) SiO2 and (bf) SiO2 with 10 wt.% Al2O3, La2O3, Y2O3, MgO, and TiO2, respectively.
Ceramics 05 00005 g003
Figure 4. BEIs/SEM images of (a) SiO2 and (bf) SiO2 with 10 wt.% Al2O3, La2O3, Y2O3, MgO, and TiO2, respectively.
Figure 4. BEIs/SEM images of (a) SiO2 and (bf) SiO2 with 10 wt.% Al2O3, La2O3, Y2O3, MgO, and TiO2, respectively.
Ceramics 05 00005 g004
Figure 5. (ae) BFIs/TEM images and inserted SADP of Si–10 wt.% Al2O3, La2O3, Y2O3, MgO, and TiO2 interfaces after sintering at 1300 °C in Ar, respectively. (f) EDS summary of reaction phases (La2Si2O7, Y2Si2O7, and MgSiO3) and ceramic additives (Al2O3 and TiO2) in SiO2.
Figure 5. (ae) BFIs/TEM images and inserted SADP of Si–10 wt.% Al2O3, La2O3, Y2O3, MgO, and TiO2 interfaces after sintering at 1300 °C in Ar, respectively. (f) EDS summary of reaction phases (La2Si2O7, Y2Si2O7, and MgSiO3) and ceramic additives (Al2O3 and TiO2) in SiO2.
Ceramics 05 00005 g005
Figure 6. BEIs/SEM images of (a) Si/SiO2, (b) Si/SiO2–10 wt.% La2O3, (c) Si/SiO2–10 wt.% Y2O3, and (d) Si/SiO2–10 wt.% MgO interfaces.
Figure 6. BEIs/SEM images of (a) Si/SiO2, (b) Si/SiO2–10 wt.% La2O3, (c) Si/SiO2–10 wt.% Y2O3, and (d) Si/SiO2–10 wt.% MgO interfaces.
Ceramics 05 00005 g006
Figure 7. BFIs/TEM images and EDS line scans of (a) Si/Si–SiO2, (b) Si/SiO2–10 wt.% La2O3, (c) Si/SiO2–10 wt.% Y2O3, and (d) Si/SiO2–10 wt.% MgO interfaces.
Figure 7. BFIs/TEM images and EDS line scans of (a) Si/Si–SiO2, (b) Si/SiO2–10 wt.% La2O3, (c) Si/SiO2–10 wt.% Y2O3, and (d) Si/SiO2–10 wt.% MgO interfaces.
Ceramics 05 00005 g007
Figure 8. BEIs/SEM images of (a) Si/SiO2–10 wt.% Al2O3 and (b) Si/SiO2–10 wt.% TiO2 interfaces.
Figure 8. BEIs/SEM images of (a) Si/SiO2–10 wt.% Al2O3 and (b) Si/SiO2–10 wt.% TiO2 interfaces.
Ceramics 05 00005 g008
Figure 9. BFIs/TEM images, EDS line scans, and SADPs of (a) Si/SiO2–10 wt.% Al2O3 and (b) Si/SiO2–10 wt.% TiO2 interfaces.
Figure 9. BFIs/TEM images, EDS line scans, and SADPs of (a) Si/SiO2–10 wt.% Al2O3 and (b) Si/SiO2–10 wt.% TiO2 interfaces.
Ceramics 05 00005 g009
Figure 10. Schematic of (a) crystalline SiO2 and (b) amorphous SiO2 induced by adding impurities (Al3+ and Ti4+).
Figure 10. Schematic of (a) crystalline SiO2 and (b) amorphous SiO2 induced by adding impurities (Al3+ and Ti4+).
Ceramics 05 00005 g010
Figure 11. (a) Vertical O concentration and (b) electrical resistivity distributions of Si pieces away from the interface of the Si/SiO2–ceramic additives.
Figure 11. (a) Vertical O concentration and (b) electrical resistivity distributions of Si pieces away from the interface of the Si/SiO2–ceramic additives.
Ceramics 05 00005 g011
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chen, Y.-H.; Lin, K.-L.; Lin, C.-C. Interfacial Reactions between Si and SiO2 with Ceramic Additives. Ceramics 2022, 5, 44-54. https://doi.org/10.3390/ceramics5010005

AMA Style

Chen Y-H, Lin K-L, Lin C-C. Interfacial Reactions between Si and SiO2 with Ceramic Additives. Ceramics. 2022; 5(1):44-54. https://doi.org/10.3390/ceramics5010005

Chicago/Turabian Style

Chen, Yu-Hsiang, Kun-Lin Lin, and Chien-Cheng Lin. 2022. "Interfacial Reactions between Si and SiO2 with Ceramic Additives" Ceramics 5, no. 1: 44-54. https://doi.org/10.3390/ceramics5010005

Article Metrics

Back to TopTop