Next Article in Journal
Extrusion-Based 3D Printing Applications of PLA Composites: A Review
Next Article in Special Issue
Growth of GaP Layers on Si Substrates in a Standard MOVPE Reactor for Multijunction Solar Cells
Previous Article in Journal
Utilizing of (Zinc Oxide Nano-Spray) for Disinfection against “SARS-CoV-2” and Testing Its Biological Effectiveness on Some Biochemical Parameters during (COVID-19 Pandemic)—”ZnO Nanoparticles Have Antiviral Activity against (SARS-CoV-2)”
Previous Article in Special Issue
Highly Conductive Co-Doped Ga2O3:Si-In Grown by MOCVD
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Study of the Annealing Effect on the γ-Phase Aluminum Oxide Films Prepared by the High-Vacuum MOCVD System

School of Physics and Optoelectronic Engineering, Shandong University of Technology, Zibo 255000, China
*
Authors to whom correspondence should be addressed.
Coatings 2021, 11(4), 389; https://doi.org/10.3390/coatings11040389
Submission received: 27 February 2021 / Revised: 25 March 2021 / Accepted: 25 March 2021 / Published: 29 March 2021

Abstract

:
γ-phase aluminum oxide (γ-Al2O3) films are grown on MgO (100) wafers by metal organic chemical vapor deposition (MOCVD). Post-annealing process is conducted to study the influence of annealing temperature on the properties of the films. Structural analyses indicate that all the deposited and annealed films present a preferred growth orientation of γ-Al2O3 (220) along the MgO (200) direction. And the film annealed at 1100 °C exhibits the best film quality compared with those of the films grown and annealed at other temperatures. Scanning electron microscopy measurements also imply the best surface morphology for the γ-Al2O3 film annealed at 1100 °C, which is in good accordance with the structural analyses. Optical transmittance spectra show good transparency for all the deposited and annealed films in the visible wavelength region with an average transmittance value of 83.5%. The optical bandgaps are estimated to be in the range of 5.56–5.79 eV for the deposited and annealed films. Semiconductor films with high optical transmittance in the visible region as well as wide bandgaps are appropriate for the manufacture of transparent optoelectronic devices and ultraviolet optoelectronic devices.

1. Introduction

Oxide semiconductor materials with wide bandgap have a wide range of applications in short-wavelength light-emitting devices, ultraviolet detectors, solar cells, and thin film transistors [1,2,3,4], which have aroused people’s research interest in recent years. Compared with traditional wide band gap oxide materials, such as ITO [5] and ZnO [6], aluminum oxide (Al2O3) has a wider band gap (~8.7 eV) and relatively stable physical and chemical properties [7]. In addition, Al2O3 also has the advantages of high relative hardness (Mohs scale, ~8.8), high breakdown field strength (6–8 MV/cm), high dielectric constant (~9), easy preparation, and low cost, which can be very suitable for manufacturing ultraviolet optoelectronic devices and transparent optoelectronic devices [8,9].
There are many polymorphs of Al2O3, such as α-phase (hexagonal) [10], γ-phase (cubic) [11], δ-phase (tetragonal or orthogonal) [12], θ-phase (monoclinic) [13] and κ-phase (orthogonal) [14], two of which are the most common crystal forms, γ-Al2O3 and α-Al2O3. The various phases can be converted to each other due to the differences in temperature and preparation methods [15], but almost all other phases are converted into α-Al2O3 at high temperatures above 1300 °C. Al2O3 is widely used in many fields. It can be used as abrasive materials and cutting tools and can also be applied to make high-temperature refractory materials and antireflection coatings on glass surfaces [16,17]. α-phase Al2O3 is usually applied as substrates for growing film materials [18,19]. γ-Al2O3 is also widely used, mainly as coatings, adsorbents, and soft abrasives [20,21]. It plays a more important role in the chemical industry and is often used as a catalyst support material for energy generation and storage [22]. In recent years, the preparations of Al2O3 thin films and their doped materials and the studies of their luminescence properties have attracted widespread interest. Al2O3 can emit fluorescence of different wavelengths through the doping of transition metal elements and rare earth elements [23,24] and has possible application prospects in the fields of optical communications, lasers, integrated optics, and panel displays [25,26].
The common preparation methods of Al2O3 film are including magnetron sputtering [27], plasma deposition [28], sol-gel [29], spray pyrolysis [30], electron beam evaporation [31], pulsed laser deposition [32] and metal organic chemical vapor deposition (MOCVD) [33]. MOCVD is a method of vapor phase epitaxial growth using organic metal thermal decomposition reaction. It has simple reaction equipment, wide growth temperature range, precise control of deposition parameters, and raw material gas which will not affect the film. Advantages such as obvious changes in doping concentration can be achieved along the film growth. Oxide films prepared by MOCVD are widely used in the electronics industry. Therefore, MOCVD method is applied in this work to grow Al2O3 films. The preparation temperature of the crystalline Al2O3 films is generally very high. It is reported in the literature that post-annealing is a very effective way to improve the performance of the film [34]. The structure and optical properties of the films can be greatly improved after annealing [35]. In view of the fact that the maximum deposition temperature of the MOCVD equipment applied in our experiment is below 750 °C, the substrate temperature for preparing the alumina films in the experiment is set to be 700 °C. And then the prepared samples are annealed subsequently. The impact of annealing temperature on the structural and optical properties of the Al2O3 films was studied systematically in this paper.

2. Experimental Details

The series of Al2O3 films were deposited using a homemade MOCVD system. The oxidant used in the work is O2 [purity: 99.999% (5N)]. The carrier gas is high-purity N2 (5N), which will be purified to 9N by an inert gas purifier before entering the system. The metalorganic (MO) source used in this experiment is trimethylaluminum (TMAl) [Al(CH3)3] with a purity of 6N, which is placed in a stainless steel source bottle in a cold trap. The gas transportation pipeline is made of high-quality stainless steel pipes polished inside and outside, connected by Swagelok VCR joints and special metal gaskets to ensure the airtightness of the gas pipeline. The flow of each pipeline gas is accurately controlled by a mass flow controller and an electrical pressure controller. The reaction chamber adopts a dual-air flow mode. The N2 carries MO steam from the side wall and flows into the top of the substrate, and the other N2 carries O2 into the reaction chamber through a nozzle from directly above the reaction chamber. This dual-air flow mode can not only suppress the thermal convection effect, but also press the reactants to the substrate, improving the deposition rate and film uniformity. The molar flow rate of MO steam is determined jointly by the vapor pressure of the organic source at a specific temperature, the pressure of the MO source bottle, and the flow of the carrier gas carrying the MO vapor. The MgO substrates were cut to the required size (10 mm × 10 mm) by a diamond cutter, then cleaned with an ultrasonic bath of acetone, alcohol, and deionized water, followed by blow-drying with nitrogen flow for use. Upon the growth process (5 h), the deposition temperature and chamber pressure were kept at 700 °C and 35 mbar, respectively. The molar flow rates of the TMAl and O2 were calculated as 2.3 × 10−6 and 2.23 × 10−3 mol/min. After the growth process, the films were taken out of the chamber and post-annealed under air in a tube furnace for 1 h at temperatures of 900, 1000, 1100 °C.
The crystal structure of the films was studied using X-ray diffraction measurement by a Rigaku D/max-rB diffractometer (Rigaku, Tokyo, JPN) with an X-ray wavelength of 1.5406 Å in the 2θ range of 10°–80°. The SEM measurements of the films were taken using FEI Nova Nano SEM-450 (Hillsboro, OR, USA) to study the surface morphology. The thickness of the samples was observed by cross-sectional SEM measurement. Chemical analysis was carried out by X-ray photoelectron spectroscopy (XPS) using a Thermo Scientific ESCALAB 250 XI spectrometer (Waltham, MA, USA). Ultraviolet-visible (UV-VIS) transmission spectroscopy was conducted by using a TU-1901 double-beam UV-VIS spectrophotometer (General Analytical, Beijing, China) to analyze the optical properties of the films.

3. Results and Discussion

The chemical compositions of the Al2O3 films are investigated by XPS measurements. Figure 1a indicates the XPS survey-scan spectrum of the as-deposited Al2O3 film, where the binding energy peaks of C 1s, Al 2s, Al 2p, O 1s, O KLL can be observed. The energy peak of C 1s may be attributed to the adventitious hydrocarbon contamination or the carbon dioxide adsorption on the surface of the film. No signal of energy peaks of elements other than Al, O, C is detected, indicating that the prepared films are aluminum oxide films. The calibration of the peak position is commonly used by the reference of C 1s core level at a binding energy of 284.6 eV [36]. Spectra (b) and (c) are the high-resolution energy peaks of Al 2p and O 1s core levels. As can be seen in Figure 1b, the aluminum core level of Al 2p is located at 74.6 eV, being in accordance with the references related to Al2O3 [29,37]. Furthermore, the Al 2p peak presents a symmetrical spectrum without any obvious Al–Al peak, indicating the absence of dissociative Al ions in our as-deposited film. The O 1s core level shown in Figure 1c can be divided into two peaks by Gaussian fitting. The peak located at a low binding energy of 530.7 eV can be designated as Al–O bonding in Al2O3 lattice. While the other peak located at a high binding energy of 532.3 eV may be attributed to the presence of loosely bound oxygen of adsorbed carbonyl and/or carbon-oxygen compounds on the film surface [38]. Quantitative analyses are studied based on estimated areas of O 1s (fit peak of the low binding energy of 530.7 eV) and Al 2p core levels by the elemental sensitivity factor method. The atomic ratio of O and Al is determined as 1.56, which is consistent with the stoichiometry of pure Al2O3.
Figure 2 shows the XRD patterns of (Figure 2a) the as-grown aluminum oxide film and the post-annealed ones at different annealing temperatures of (Figure 2b) 900 °C, (Figure 2c) 1000 °C, and (Figure 2d) 1100 °C in air, respectively.
It can be seen from the patterns of all the deposited and annealed films, besides the diffraction peaks of the MgO (200) substrates, only signals corresponding to cubic γ-phase Al2O3 (JCPDS #50-0741) are detected with diffractions of (220), (200) and (440) planes. The intensity of the (200) and (440) peaks is rather weak compared with that of the (220) peak, and the (220) peak is the predominant reflection, as is revealed in the spectra. All the films present a preferential growth orientation of γ-Al2O3 (220) along the MgO (200) direction. Note that a slight shift towards lower angle direction of the (220) peak can be observed in Figure 2, implying a systematic lattice expansion, which is also reported previously [39,40]. As we all know, commonly, material has a certain flexibility. The lattice expansion may be attributed to the release of pressure stress in the inner part of the films after heat treatment. As for the annealed films, the intensity of the (220) peaks is more intense, and the full width at half maximum (FWHM) of (220) peaks are measured to be 0.915°, 0.537°, and 0.373°, corresponding to annealing temperatures of 900, 1000, and 1100 °C, respectively, which is rather smaller than that of as deposited film with a value of 1.038°. The FWHM is a reflection of the film crystallinity. Lower FWHM corresponds to higher film crystallinity. The decrease of the FWHM with the increasing annealing temperature reveals the improvement of the film crystallinity, which can be attributed to the recrystallization process in the films at high temperatures [41]. The average particle size (D) of the Al2O3 films is determined by Scherrer equation [42],
D = k λ / b cos θ
where k is a crystalline shape-related constant and usually taken as 0.9, λ is the X-ray wavelength applied in the diffractometer (0.15406 nm), b is the FWHM of γ-Al2O3 (220) plane in radian, and θ is Bragg’s angle of the corresponding (220) peak. The variations of b and D values calculated using Al2O3 (220) plane as a function of annealing temperature are illustrated in Figure 3. As can be seen, the average grain size of the film presents a monotonous increase with the increasing annealing temperature, reaching a maximum value of 22.5 nm for the film annealed at 1100 °C. It is believed that the increase of the average particle size may be attributed to the coalescence of small grains caused by the increasing kinetic energy with the increasing annealing temperature. All the results reveal that the annealing process exhibits a prominent influence on the crystal structure of the film, and the sample annealed at the temperature of 1100 °C exhibits the best crystalline quality.
The surface morphologies of Al2O3 films are studied by SEM with top-view images. Images (a)–(d) of Figure 4 are corresponding to the as-deposited Al2O3 film and the ones post-annealed at 900, 1000, and 1100 °C, respectively. The inset shows the cross-sectional SEM image of the as-deposited Al2O3 film, where the thickness of the deposited film can be obtained to be about 180 nm. As shown in Figure 4a, a leaf-shaped structure is presented for the as-deposited Al2O3 film, exhibiting the coalescence of some small grains. For the annealed films at 900 and 1000 °C, illustrated in Figure 4b,c, the compact textures comprised of both large and small grains are revealed on the MgO substrates. The SEM image of the film annealed at 1100 °C, which is shown in Figure 4d, indicates the homogeneous and oriented grain growth with a clear and regular edge through the film. Closely packed polygonal-shaped grains with larger size have developed on top of the substrate. As clearly observed in the images, the average particle size increases with the increasing annealing temperature, which is consistent with the XRD analyses. The annealing temperature exhibits a remarkable influence on the surface morphology of Al2O3 films.
Figure 5a shows the transmittance spectra of the as-grown and post-annealed Al2O3 films at different temperatures. The mean values of optical transmittance for all the films in the visible wavelength region (380–800 nm) are calculated above 83.5%, which reveals their good optical transparency. The good transmittance exhibited supports the suitability of their applications as flat panel display devices. The oscillations observed in the transmittance spectra are the sign of the interference between two interfaces of air film and film substrate. The absorption edge of the film shifts toward the long wavelength direction with the increasing annealing temperature, indicating the narrowing of the optical bandgap (Eg). The absorption coefficient (α) is related with optical transmittance (T) by the equation [43],
α = ln [ 1 R 2 / T ] / d
where d is the thickness, R stands for the reflectance, which can be neglected (R << 1). The Eg dependence of the α is given by the Tauc equation [44],
α h ν 2 = A h ν E g
where hν indicates the incident photo energy, and A is a constant related to the band edge. Figure 5b shows the plot of (αhν)2 versus hν, where the Eg can be estimated at the crossing point between the line extrapolated from the linear part of the curve and hν-axis of the plot. The optical bandgaps are found to be 5.79, 5.73, 5.65, and 5.56 eV, corresponding to the as-grown and 900, 1000, and 1100 °C annealed samples, respectively. First-principles calculations indicate that the actual optical band gap of γ-phase Al2O3 film is much lower than the theoretical value [45,46]. Some researchers have reported the optical bandgap values of Al2O3 films as 5.36 eV [47], 5.50–5.63 eV [48,49], and 5.80–5.89 eV [30], which is in good accordance with the results obtained in this work. As can be seen, the Eg of the film decreases with the increase of the annealing temperature, which is consistent with the improvement of the structural properties of the film revealed in the XRD results. The explanation may be attributed to the sufficient energy supplied by thermal annealing which can cure the randomness of inner atoms into the films, increasing the degree of atomic ordering [50].

4. Conclusions

In general, MOCVD technique has been applied to grow γ-Al2O3 films on MgO (100) substrates, followed by post-annealing at temperatures of 900, 1000, and 1100 °C. XPS study confirmed the stoichiometry of pure Al2O3 without dissociative Al ions. The annealing temperature had a great impact on the structural and optical properties of the films. The best crystallization and surface morphology were observed at the annealing temperature of 1100 °C, with the maximum value of the average particle size of the films. All the films presented a preferred growth orientation of γ-Al2O3 (220)//MgO (200). The average transmittance of the as-grown and post-annealed γ-Al2O3 films in the visible region exceeded 83.5%. The optical bandgaps of the films are found to be ranging from 5.56 to 5.79 eV. Al2O3 film materials with fine performances can be used in many fields such as transparent photoelectrical devices, solar cells, flat panel displays, and UV light detectors.

Author Contributions

Conceptualization, Z.L. and J.X.; methodology, Z.L.; software, Y.L.; formal analysis, J.X.; investigation, Y.X.; data curation, Y.X.; writing—original draft preparation, Z.L.; writing—review and editing, J.X.; project administration, D.Y.; funding acquisition, H.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Shandong Provincial Natural Science Foundation, China (Grant nos. ZR2020QA058 and ZR2020MA088).

Conflicts of Interest

There are no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Sun, Y.R.; Giebink, N.C.; Kanno, H.; Ma, B.W.; Thompson, M.E.; Forrest, S.R. Management of singlet and triplet excitons for efficient white organic light-emitting devices. Nat. Cell Biol. 2006, 440, 908–912. [Google Scholar] [CrossRef] [Green Version]
  2. Peng, W.; He, Y.; Wen, C.; Ma, K. Surface acoustic wave ultraviolet detector based on zinc oxide nanowire sensing layer. Sens. Actuators A Phys. 2012, 184, 34–40. [Google Scholar] [CrossRef]
  3. Jeong, S.; Jeong, Y.; Moon, J. Solution-processed zinc tin oxide semiconductor for thin-film transistors. J. Phys. Chem. C 2008, 112, 11082–11085. [Google Scholar] [CrossRef]
  4. Hara, K.; Horiguchi, T.; Kinoshita, T.; Sayama, K.; Sugihara, H.; Arakawa, H. Highly efficient photon-to-electron conversion with mercurochrome-sensitized nanoporous oxide semiconductor solar cells. Sol. Energy Mater. Sol. Cells 2000, 64, 115–134. [Google Scholar] [CrossRef]
  5. Kim, K.; Park, J.H.; Kim, H.; Kim, J.K.; Schubert, E.F.; Cho, J. Energy bandgap variation in oblique angle-deposited indium tin oxide. Appl. Phys. Lett. 2016, 108, 041910. [Google Scholar] [CrossRef]
  6. Ansari, S.A.; Khan, M.M.; Kalathil, S.; Nisar, A.; Lee, J.; Cho, M.H. Oxygen vacancy induced band gap narrowing of ZnO nanostructures by an electrochemically active biofilm. Nanoscale 2013, 5, 9238–9246. [Google Scholar] [CrossRef] [PubMed]
  7. Othman, H.; Valiev, D.; Polisadova, E. Structural and mechanical properties of zinc aluminoborate glasses with different content of aluminum oxide. J. Mater. Sci. Mater. Electron. 2017, 28, 4647–4653. [Google Scholar] [CrossRef]
  8. Lee, B.R.; Lee, T.H.; Son, K.-R.; Kim, T.G. ITO/Ag/AlN/Al2O3 multilayer electrodes with conductive channels: Application in ultraviolet light-emitting diodes. J. Alloys Compd. 2018, 741, 21–27. [Google Scholar] [CrossRef]
  9. Zhang, G.Z.; Wu, H.; Chen, C.; Wang, T.; Wang, P.Y.; Mai, L.Q.; Yue, J.; Liu, C. Transparent capacitors based on nanolaminate Al2O3/TiO2/Al2O3 with H2O and O3 as oxidizers. Appl. Phys. Lett. 2014, 104, 163503. [Google Scholar] [CrossRef]
  10. Eng, P.J.; Trainor, T.P.; Brown, G.E., Jr.; Waychunas, G.A.; Newville, M.; Sutton, S.R.; Rivers, M.L. Structure of the hydrated α-Al2O3 (0001) surface. Science 2000, 288, 1029–1033. [Google Scholar] [CrossRef]
  11. Chou, T.C.; Nieh, T.G. Nucleation and concurrent anomalous grain growth of α-Al2O3 during γ-α phase transformation. J. Am. Ceram. Soc. 1991, 74, 2270–2279. [Google Scholar] [CrossRef]
  12. Paunović, N.; Dohčević-Mitrović, Z.; Scurtu, R.; Aškrabić, S.; Prekajski, M.; Matović, B.; Popović, Z.V. Suppression of inherent ferromagnetism in Pr-doped CeO2 nanocrystals. Nanoscale 2012, 4, 5469–5476. [Google Scholar] [CrossRef]
  13. Jbara, A.S.; Othaman, Z.; Saeed, M. Structural, morphological and optical investigations of θ-Al2O3 ultrafine powder. J. Alloys Compd. 2017, 718, 1–6. [Google Scholar] [CrossRef]
  14. Belonoshko, A.B.; Ahuja, R.; Johansson, B. Mechanism for the κ−Al2O3 to the α−Al2O3 transition and the stability of κ-Al2O3 under volume expansion. Phys. Rev. B 2000, 61, 3131. [Google Scholar] [CrossRef]
  15. Santos, P.S.; Santos, H.S.; Toledo, S.P. Standard transition aluminas. Electron microscopy studies. Mater. Res. 2000, 3, 104–114. [Google Scholar] [CrossRef]
  16. Musante, L.; Munoz, V.; Labadie, M.H.; Tomba-Martinez, A.G. High temperature mechanical behavior of Al2O3–MgO–C refractories for steelmaking use. Ceram. Int. 2011, 37, 1473–1483. [Google Scholar] [CrossRef]
  17. Kauppinen, C.; Isakov, K.; Sopanen, M. Grass-like alumina with low refractive index for scalable, broadband, omnidirectional antireflection coatings on glass using atomic layer deposition. ACS Appl. Mater. Interfaces 2017, 9, 15038–15043. [Google Scholar] [CrossRef] [PubMed]
  18. Wojewoda-Budka, J.; Sobczak, N.; Morgiel, J.; Nowak, R. Microstructure characteristics of the reaction product region formed due to the high temperature contact of molten aluminium and ZnO single crystal. Solid State Phenom. 2011, 172–174, 1267–1272. [Google Scholar] [CrossRef]
  19. Shen, P.; Fujii, H.; Matsumoto, T.; Nogi, K. Surface orientation and wetting phenomena in Si/α-alumina system at 1723 K. J. Am. Ceram. Soc. 2005, 88, 912–917. [Google Scholar] [CrossRef]
  20. Jiang, Z.; Yang, X.; Liang, Y.; Hao, G.; Zhang, H.; Lin, J. Favorable deposition of γ-Al2O3 coatings by cathode plasma electrolysis for high-temperature application of Ti-45Al-8.5Nb alloys. Surf. Coat. Technol. 2018, 333, 187–194. [Google Scholar] [CrossRef]
  21. Tabesh, S.; Davar, F.; Loghman-Estarki, M.R. Preparation of γ-Al2O3 nanoparticles using modified sol-gel method and its use for the adsorption of lead and cadmium ions. J. Alloys Compd. 2018, 730, 441–449. [Google Scholar] [CrossRef]
  22. Lelo, R.V.; Maron, G.K.; Thesing, A.; Alano, J.H.; Rodrigues, L.D.S.; Noremberg, B.D.S.; Escote, M.T.; Valentini, A.; Probst, L.F.D.; Carreño, N.L.V. Vanadium effect over γ-Al2O3-supported Ni catalysts for valorization of glycerol. Fuel Process. Technol. 2021, 216, 106773. [Google Scholar] [CrossRef]
  23. He, Y.; He, J.; Zhang, H.; Liu, Y.; Lei, B. Luminescent properties and energy transfer of luminescent carbon dots assembled mesoporous Al2O3: Eu3+ co-doped materials for temperature sensing. J. Colloid Interface Sci. 2017, 496, 8–15. [Google Scholar] [CrossRef]
  24. Wang, X.; Hu, L.; Meng, X.; Li, H.; Wang, S. Effect of Al2O3 and La2O3 on structure and spectroscopic properties of Nd-doped sol–gel silica glasses. J. Lumin. 2018, 204, 554–559. [Google Scholar] [CrossRef]
  25. Lakshminarayana, G.; Ruan, J.; Qiu, J. NIR luminescence from Er–Yb, Bi–Yb and Bi–Nd codoped germanate glasses for optical amplification. J. Alloys Compd. 2009, 476, 878–883. [Google Scholar] [CrossRef]
  26. He, L.; Pan, L.; Li, W.; Dong, Q.; Sun, W. Spectral response characteristics of Eu3+ doped YAG-Al2O3 composite nanofibers reinforced aluminum matrix composites. Opt. Mater. 2020, 104, 109845. [Google Scholar] [CrossRef]
  27. Ding, J.C.; Zhang, T.F.; Mane, R.S.; Kim, K.-H.; Kang, M.C.; Zou, C.W.; Wang, Q.M. Low-temperature deposition of nanocrystalline Al2O3 films by ion source-assisted magnetron sputtering. Vacuum 2018, 149, 284–290. [Google Scholar] [CrossRef]
  28. Haiying, W.E.; Hongge, G.U.; Lijun, S.A.; Xingcun, L.I.; Qiang, C.H. Study on deposition of Al2O3 films by plasma-assisted atomic layer with different plasma sources. Plasma Sci. Technol. 2018, 20, 065508. [Google Scholar]
  29. Hu, B.; Yao, M.; Xiao, R.; Chen, J.; Yao, X. Optical properties of amorphous Al2O3 thin films prepared by a sol–gel process. Ceram. Int. 2014, 40, 14133–14139. [Google Scholar] [CrossRef]
  30. Dhonge, B.P.; Mathews, T.; Sundari, S.T.; Thinaharan, C.; Kamruddin, M.; Dash, S.; Tyagi, A. Spray pyrolytic deposition of transparent aluminum oxide (Al2O3) films. Appl. Surf. Sci. 2011, 258, 1091–1096. [Google Scholar] [CrossRef]
  31. Shuzhen, S.; Lei, C.; Haihong, H.; Kui, Y.; Zhengxiu, F.; Jianda, S. Annealing effects on electron-beam evaporated Al2O3 films. Appl. Surf. Sci. 2005, 242, 437–442. [Google Scholar] [CrossRef]
  32. Ion, M.; Berbecaru, C.; Iftimie, S.; Filipescu, M.; Dinescu, M.; Antohe, S. PLD deposited Al2O3 thin films for transparent electronics. Dig. J. Nanomater. Biostruct. 2012, 7, 1609–1614. [Google Scholar]
  33. Ito, A.; Tu, R.; Goto, T. Amorphous-like nanocrystalline γ-Al2O3 films prepared by MOCVD. Surf. Coat. Technol. 2010, 204, 2170–2174. [Google Scholar] [CrossRef]
  34. Ahn, C.H.; Kim, S.H.; Kim, Y.K.; Lee, H.S.; Cho, H.K. Effect of post-annealing temperatures on thin-film transistors with ZnO/Al2O3 su-perlattice channels. Thin Solid Films 2015, 584, 336–340. [Google Scholar] [CrossRef]
  35. Tahir, D.; Kwon, H.L.; Shin, H.C.; Oh, S.K.; Kang, H.J.; Heo, S.; Chung, J.G.; Lee, J.C.; Tougaard, S. Electronic and optical properties of Al2O3/SiO2 thin films grown on Si substrate. J. Phys. D Appl. Phys. 2010, 43, 255301. [Google Scholar] [CrossRef] [Green Version]
  36. Greczynski, G.; Hultman, L. X-ray photoelectron spectroscopy: Towards reliable binding energy referencing. Prog. Mater. Sci. 2020, 107, 100591. [Google Scholar] [CrossRef]
  37. Usman, M.; Arshad, M.; Suvanam, S.S.; Hallén, A. Influence of annealing environment on the ALD-Al2O3/4H-SiC interface studied through XPS. J. Phys. D Appl. Phys. 2018, 51, 105111. [Google Scholar] [CrossRef]
  38. Guan, S.; Yamawaki, L.; Zhang, P.; Zhao, X. Charge-transfer effect of GZO film on photochemical water splitting of transparent ZnO@ GZO films by RF magnetron sputtering. Top. Catal. 2018, 61, 1585–1590. [Google Scholar] [CrossRef]
  39. Qiao, Z.; Latz, R.; Mergel, D. Thickness dependence of In2O3:Sn film growth. Thin Solid Films 2004, 466, 250–258. [Google Scholar] [CrossRef]
  40. Bouhdjer, A.; Attaf, A.; Saidi, H.; Benkhetta, Y.; Aida, M.; Bouhaf, I.; Rhil, A. Influence of annealing temperature on In2O3 properties grown by an ultrasonic spray CVD process. Optik 2016, 127, 6329–6333. [Google Scholar] [CrossRef]
  41. Raj, A.M.E.; Lalithambika, K.; Vidhya, V.; Rajagopal, G.; Thayumanavan, A.; Jayachandran, M.; Sanjeeviraja, C. Growth mechanism and optoelectronic properties of nanocrystalline In2O3 films prepared by chemical spray pyrolysis of metal-organic precursor. Phys. B Condens. Matter 2008, 403, 544–554. [Google Scholar] [CrossRef]
  42. Prabeesh, P.; Selvam, I.P.; Potty, S. Effect of annealing temperature on a single step processed Cu2ZnSnS4 thin film via solution method. Thin Solid Films 2016, 606, 94–98. [Google Scholar] [CrossRef]
  43. Kavitha, V.; Suresh, S.; Chalana, S.; Pillai, V.M. Luminescent Ta doped WO3 thin films as a probable candidate for excitonic solar cell applications. Appl. Surf. Sci. 2019, 466, 289–300. [Google Scholar] [CrossRef]
  44. Deng, S.; Chen, R.; Li, G.; Xia, Z.; Zhang, M.; Zhou, W.; Wong, M.; Kwok, H.-S. High-performance staggered top-gate thin-film transistors with hybrid-phase microstructural ITO-stabilized ZnO channels. Appl. Phys. Lett. 2016, 109, 182105. [Google Scholar] [CrossRef] [Green Version]
  45. Menéndez-Proupin, E.; Gutiérrez, G. Electronic properties of bulk γ−Al2O3. Phys. Rev. B 2005, 72, 035116. [Google Scholar] [CrossRef]
  46. Costina, I.; Franchy, R. Band gap of amorphous and well-ordered Al2O3 on Ni3Al(100). Appl. Phys. Lett. 2001, 78, 4139–4141. [Google Scholar] [CrossRef] [Green Version]
  47. Amirsalari, A.; Shayesteh, S.F. Effects of pH and calcination temperature on structural and optical properties of alumina nanoparticles. Superlattices Microstruct. 2015, 82, 507–524. [Google Scholar] [CrossRef]
  48. Adamopoulos, G.; Thomas, S.; Bradley, D.D.C.; McLachlan, M.A.; Anthopoulos, T.D. Low-voltage ZnO thin-film transistors based on Y2O3 and Al2O3 high-k dielectrics deposited by spray pyrolysis in air. Appl. Phys. Lett. 2011, 98, 123503. [Google Scholar] [CrossRef]
  49. Li, Z.; Ma, J.; Feng, X.; Du, X.; Wang, W.; Wang, M. Effect of thermal annealing on the optical and structural properties of γ-Al2O3 films prepared on MgO substrates by MOCVD. Ceram. Int. 2016, 42, 551–558. [Google Scholar] [CrossRef]
  50. Dai, S.; Wang, T.; Li, R.; Wang, Q.; Ma, Y.; Tian, L.; Su, J.; Wang, Y.; Zhou, D.; Zhang, X.; et al. Preparation and electrical properties of N-doped ZnSnO thin film transistors. J. Alloys Compd. 2018, 745, 256–261. [Google Scholar] [CrossRef]
Figure 1. X-ray photoelectron spectroscopy (XPS) spectra of the as-deposited Al2O3 film: (a) survey spectrum, high-resolution spectra of (b) Al 2p and (c) O 1s.
Figure 1. X-ray photoelectron spectroscopy (XPS) spectra of the as-deposited Al2O3 film: (a) survey spectrum, high-resolution spectra of (b) Al 2p and (c) O 1s.
Coatings 11 00389 g001
Figure 2. XRD spectra of Al2O3 films of (a) the as-grown and post-annealed at temperatures of (b) 900 °C, (c) 1000 °C and (d) 1100 °C.
Figure 2. XRD spectra of Al2O3 films of (a) the as-grown and post-annealed at temperatures of (b) 900 °C, (c) 1000 °C and (d) 1100 °C.
Coatings 11 00389 g002
Figure 3. The full width at half maximum (b) and average grain size (D) values as a function of annealing temperature.
Figure 3. The full width at half maximum (b) and average grain size (D) values as a function of annealing temperature.
Coatings 11 00389 g003
Figure 4. Top-view SEM images of (a) the as-deposited Al2O3 film, and post-annealed ones at temperatures of (b) 900 °C, (c) 1000 °C and (d) 1100 °C, with the cross-sectional SEM image of the as-deposited Al2O3 film in the inset.
Figure 4. Top-view SEM images of (a) the as-deposited Al2O3 film, and post-annealed ones at temperatures of (b) 900 °C, (c) 1000 °C and (d) 1100 °C, with the cross-sectional SEM image of the as-deposited Al2O3 film in the inset.
Coatings 11 00389 g004
Figure 5. (a) Optical transmittance spectra of the as-grown Al2O3 film and the post-annealed ones at temperatures of 900, 1000 and 1100 °C, (b) (αhν)2 vs. hν of the corresponding samples.
Figure 5. (a) Optical transmittance spectra of the as-grown Al2O3 film and the post-annealed ones at temperatures of 900, 1000 and 1100 °C, (b) (αhν)2 vs. hν of the corresponding samples.
Coatings 11 00389 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, Z.; Xin, Y.; Liu, Y.; Liu, H.; Yu, D.; Xiu, J. Study of the Annealing Effect on the γ-Phase Aluminum Oxide Films Prepared by the High-Vacuum MOCVD System. Coatings 2021, 11, 389. https://doi.org/10.3390/coatings11040389

AMA Style

Li Z, Xin Y, Liu Y, Liu H, Yu D, Xiu J. Study of the Annealing Effect on the γ-Phase Aluminum Oxide Films Prepared by the High-Vacuum MOCVD System. Coatings. 2021; 11(4):389. https://doi.org/10.3390/coatings11040389

Chicago/Turabian Style

Li, Zhao, Yangmei Xin, Yunyan Liu, Huiqiang Liu, Dan Yu, and Junshan Xiu. 2021. "Study of the Annealing Effect on the γ-Phase Aluminum Oxide Films Prepared by the High-Vacuum MOCVD System" Coatings 11, no. 4: 389. https://doi.org/10.3390/coatings11040389

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop