Next Article in Journal
Cracking Behavior, Microstructure and Properties of Selective Laser Melted Al-Mn-Mg-Sc-Zr Alloy
Previous Article in Journal
Effect of Heat Treatment on Mechanical Properties and Fracture Behavior of Al-7.0Si-0.3Mg Alloy at Low Temperature
Previous Article in Special Issue
Study on Optical and Electrical Properties of Thermally Evaporated Tin Oxide Thin Films for Perovskite Solar Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Effect of Annealing on the Optoelectronic Properties and Energy State of Amorphous Pyrochlore Y2Ti2O7 Thin Layers by Sol–Gel Synthesis

1
Department of Electrical Engineering, Institute of Electronics, National Yang Ming Chiao Tung University, Hsinchu 30010, Taiwan
2
Department of Mechanical Engineering, Institute of Opto-Mechatronics, National Chung Cheng University, Chiayi 62301, Taiwan
3
Institute of Atomic and Molecular Sciences, Academia Sinica, Taipei 10617, Taiwan
4
Department of Materials Science and Engineering, Feng Chia University, Taichung 40724, Taiwan
5
Advanced Institute of Manufacturing with High-Tech Innovations, National Chung Cheng University, Chiayi 62301, Taiwan
*
Authors to whom correspondence should be addressed.
Submission received: 10 March 2022 / Revised: 12 April 2022 / Accepted: 14 April 2022 / Published: 18 April 2022
(This article belongs to the Special Issue Advances of Metal Halide Perovskite Devices)

Abstract

:
Pyrochlore titanate (Y2Ti2O7) is a promising material for a wide range of applications in optoelectronics and photocatalysis due to its advantageous chemical, mechanical, and optical properties. To enhance its potential for such uses, however, a high-quality and scalable synthesis method is required. We here investigate the crystallization of sol–gel produced Y2Ti2O7 layers. We observe a transition of the amorphous pyrochlore phase at annealing temperatures below 700 °C. The transmittances of the Y2Ti2O7 thin layers annealed at 400 to 700 °C are approximately 92.3%. The refractive indices and packing densities of Y2Ti2O7 thin layers annealed at 400–700 °C/1 h vary from 1.931 to 1.954 and 0.835 to 0.846, respectively. The optical bandgap energies of Y2Ti2O7 thin layers annealed at 400–700 °C/1 h reduce from 4.356 to 4.319 eV because of the Moss–Burstein effect. These good electronic and optical properties make Y2Ti2O7 thin layers a promising host material for many potential applications.

1. Introduction

Pyrochlore titanate (Y2Ti2O7) is an important material because it has high mechanical [1,2], chemical [3], thermal stability [4,5], low phonon energy [4,6], good photocatalytic activity [7], a high refractive index [8], and excellent ion-electron conductivity [9,10], etc. Therefore, it has many applications, such as photocatalyst [11], solid-electrolytes of fuel cells [12,13], gas-sensing materials [14], high-K dielectrics [15,16], H2 storage material [17], nuclear waste storage material [18], transistor device [19], and photovoltaic material [20]. Moreover, Y2Ti2O7 can act as the host matrix of phosphor materials. The rare earth-doped Y2Ti2O7 have attracted many researchers to pay attention to the application of fiber amplifiers [21], integrated electronic devices [22], up-conversion luminescence temperature sensing material [23,24,25], and visible up-conversion luminescent solar converter (LSC) in recent years [26].
To realize these applications, a large-scale synthesis approach is needed. To date, Y2Ti2O7 nanocrystals or thin films have been produced by several approaches, such as coprecipitation [27,28], hydrothermal processes [29], mechanical milling [30,31], the liquid mix technique [32], solid-state reaction [33,34], sol–gel processes [14,35,36], and aerosol-assisted chemical vapor deposition (AACVD) [20]. However, certain processes are not readily scalable due to stringent process conditions and complex preparation.
Sol–gel processes, on the other hand, have demonstrated the potential to produce large scale thin films using simple tools, such as printing and spin-coating. Such films are then annealed at high temperatures to enhance their performance.
We here systematically investigate the effects of annealing temperatures on the phase development, layer morphology, and related optoelectronic properties of Y2Ti2O7 thin layers. We demonstrate that Y2Ti2O7 thin layers have high average transmittance (73.8–75.1%), high refractive index (1.931–1.954 at λ = 550 nm), and high bandgap energy (4.319–4.356 eV), which depends on the variation of annealing temperatures (400–750 °C). Utilizing this understanding, we were able to produce high-quality Y2Ti2O7 thin layers directly on a glass substrate. The good optoelectronic properties and scalable production open up new directions for Y2Ti2O7-based applications.

2. Experimental Procedures

2.1. Fabrication of Sol–Gel Solution and Thin Layers

Titanium solution was prepared by mixing acetic acid (HAc, 99.9%, Merck, Darmstadt, Germany), titanium isoproxide (≥99%, Acros, Geel, Belgium), and 2-methoxyethanol (2-MOE, 99.9%, Merck) in a molar ratio of Ti/HAc/2-MOE = 1/15/10. Yttrium solution was produced by mixing methanol (Me, ≥99.9%, Merck), ethylene glycol (EG, ≥99.9%, Alfa, Lancashire, UK), and yttrium acetate (Y(CH3COO)3∙4H2O, 99.9%, Alfa) with molar ratio of Y/Me/EG = 1/50/15 and dissolved into the titanium solution. Homogeneous hydrolysis reaction were completed by stirring the mixture solution for 10 h [37,38].
Subsequently, the Y2Ti2O7 precursor solution was spin-coated on a Corning 7059 glass substrate at 1000 rpm for 10 s and then 3000 rpm for 30 s. Heating to 120 °C was conducted to dry the deposited sol–gel film. The dried sol–gel films were then pyrolyzed at 380 °C for 50 min in ambient atmosphere.
The sol–gel coating process and subsequent pyrolysis process was repeated eight times to obtain films of approximately 430 nm thickness. Afterwards, thin layers were annealed from 400 to 750 °C for 1 h in ambient conditions. The average layer thickness of 400 ± 10 nm and other property calculations (grain size, transmittance, refractive index, and optical bandgap) were obtained by averaging over five samples.

2.2. Characterization of Thin Layers

The α-step profile meter (Alpha-Step IQ, KLA-Tencor, Milpitas, CA, USA) is utilized for the thickness of thin layers’ measurements. A scanning electron microscopy (SEM) (Hitachi, Tokyo, Japan, S4800-I) was performed at an accelerating voltage of 15 kV. A Shimadzu UV-2100 spectrophotometer was used for transmission spectra measurement of the Y2Ti2O7 thin layer in the UV-visible range. X-ray diffraction (XRD) measurements of thin layers were performed by an X-ray diffractometer (Bruker, Billerica, MA, USA, D8 discovery) with CuKα radiation (λ = 0.154 nm).

3. Results and Discussion

3.1. Crystal Structure and Layer Morphology

The XRD patterns in Figure 1 show the effect of annealing at different temperatures (400, 500, 600, 700, and 750 °C) for 1 h on the phase transformation of Y2Ti2O7 thin layers. For the annealing temperatures ≤ 700 °C, Y2Ti2O7 thin layers exhibit a weak broad continuum around 2θ = ~30° ((222) peak) which is the characteristic of an amorphous structure exhibiting short-range order. When the annealing temperature reaches 750 °C, the well-crystallized pyrochlore phase is identified by the (222), (311), (400), and (622) peaks (JCPDS No. 42-0413). No TiO2, Y2O3, and other phases are observed in this system. This transition temperature between amorphous and crystalline Y2Ti2O7 is in the range of previous reports on sol–gel processes, e.g., 725 [39], 750 [40], or 800 °C [35,36]. However, the other sol–gel or AACVD methods can induce the formation of crystallized Y2Ti2O7 phase at 550–570 °C [14,20]. The different starting materials and processes can result in the different crystalized temperatures and crystallinity.
The average grain size (D) of Y2Ti2O7 thin layers is determined by the Scherrer equation [41]:
D = S λ B M B S cos θ
where S is the Scherrer constant (0.9), λ is the wavelength of incident radiation, θ is the Bragg angle corresponding to the XRD peak being considered. BM and BS are the width in radians of one of the sample and standard (Si powder) diffraction peaks at half-maximum, respectively. The instrument broadening (BS) is 0.14 in our system.
After calculation, the average estimated grain sizes of the Y2Ti2O7 thin layers annealed at 600, 700 and 750 °C for 1 h are ~1.4 nm, ~1.5 nm, and ~32 nm, respectively. It is worth noting that the intensity of the broad (222) peak gradually increases when the annealing temperature increases from 400 to 700 °C. Furthermore, the crystals grow up and the XRD peak of the (222) facet becomes sharper at 750 °C. This means that the portion of amorphous Y2Ti2O7 phase reduces as the crystallinity of Y2Ti2O7 phase is enhanced. Due to the glass transition temperature of the Corning 7059 substrate, the annealing temperature range is limited to 750 °C.
Figure 2 represents top-view SEM images of the Y2Ti2O7 thin layers annealed at 400, 500, 600, 700, and 750 °C for 1 h. No crystal facets or signs of grain boundaries were detected when samples were annealed below 750 °C. This absence is expected from our XRD analysis, as the grain size of these films is below the resolution of the SEM. Conversely, crystal structures and grain boundaries can be observed at 750 °C annealing. As the surface of sol–gel spin coating thin films is very smooth, SEM images are not very clear.

3.2. Optoelectronic Properties

Figure 3 shows the effect of annealing temperatures on the transmittance spectra of the Y2Ti2O7 thin layers annealed at 400, 500, 600, 700, and 750 °C for 1 h. The small fluctuation at high transmittances is due to the Fabry–Perot interference phenomenon of multiple reflected beams.
For Y2Ti2O7 thin layers annealed at 400, 500, 600, and 700 °C, the regular interference fringes indicate that all of the Y2Ti2O7 thin layers have smooth interfaces and surfaces. The transmittance peaks of difference annealing temperatures slightly shift. Thin layers with different refractive indices result in the different optical path lengths for the constructive/destructive interference. However, Y2Ti2O7 thin layers annealed at 750 °C exhibit more random interference. This effect originates from the surface melting of the glass substrate and other optical properties (transmittance, optical bandgap, and refractive index) of Y2Ti2O7 thin layers annealed at 750 °C cannot be calculated.
The maximum transmittances in the wavelength ranging 200–1100 nm of the Y2Ti2O7 thin layers annealed at 400–700 °C/1 h are approximately 92.3%. These amorphous thin layers have small grain size and low surface roughness resulting in the weak scattering and high transmittance.
The refractive index (n) of the Y2Ti2O7 thin layer can be obtained from the transmittance spectra by fit to the Swanepoel and Cauchy equation [42]. Figure 4 shows the wavelength dependence of refractive indices for Y2Ti2O7 thin layer annealed at 400, 500, 600, and 700 °C for 1 h. As the annealing temperature increases, the refractive index of the thin layer increases. The refractive indices are 1.931, 1.936, 1.941, and 1.954 at λ = 550 nm for Y2Ti2O7 thin layers annealed at 400, 500, 600, and 700 °C for 1 h, respectively. At higher temperatures, the crystallinity and densification of Y2Ti2O7 thin layers are also improved, which leads the higher refractive indices of Y2Ti2O7 thin layers. On the other hand, the refractive index of the amorphous phase is usually lower than that of high crystalline phase (e.g., the refractive index of bulk Y2Ti2O7 is 2.34) [43].
In addition, we calculate the degree of porosity in Y2Ti2O7 thin layers. By following the Bragg–Pippard formula [44], the packing density (P) can be calculated using the following equation:
n f 2 = ( 1 P ) n v 4 + ( 1 + P ) n p 2 n b 2 ( 1 + P ) n v 2 + ( 1 P ) n b 2
where nv, nb and nf are the refractive indices of voids (nv = 1 for empty voids), bulk materials, and porous layers, respectively. The packing densities of the Y2Ti2O7 thin layers annealed at 400, 500, 600, and 700 °C for 1 h are 0.835, 0.837, 0.84, and 0.846, respectively. The packing density increases with annealing temperature, which is attributed to the significant increase in the crystallinity and reduction in the porosity.
Figure 5 plots the relationship of (αhν)2 versus photon energy (E) of the Y2Ti2O7 thin layers annealed at 400–700 °C/1 h, and the extrapolated optical bandgap energies of thin layers were determined.
The optical bandgap energy (Eg) of thin layer can be related to absorption coefficient (α) by Equation (3)
αhν = const. (–Eg)m
where m = 1/2, 3/2, 2 or 3 indicates allowed direct, forbidden direct, allowed indirect, and forbidden indirect electronic transitions, respectively [45]. A better linear dependence can be found in the (αhν)1/m versus E plots at m = 1/2 for all thin layers, which means the Y2Ti2O7 thin layer belongs to direct-transition-type material.
As shown in Figure 5, the optical bandgap energy decrements from 4.356 to 4.319 eV were observed when the Y2Ti2O7 thin layers annealing temperatures increase from 400 to 700 °C. These optical bandgap energies are higher than that of the crystallized Y2Ti2O7 thin films [14]. There are several causes or aspects, such as grain size [46,47], thickness [46,48], stress [49], and defects [50,51] could affect or shift the bandgap energy of thin layer materials. For the amorphous Y2Ti2O7 thin layers annealed from 400 to 700 °C, however, this red shift of bandgap energy could be mainly attributed to defects.
Many defects such as unsaturated bonds, dangling bonds, interstitial atoms, and vacancies can exist in amorphous structure. In general, oxygen vacancies can be observed in many transition metal oxide thin layers such as ZnO [52], BaxSr1-xTiO3 [53,54], CdIn2O4 [55,56], WO3 [57,58], and TiO2 [59]. Moreover, enormous oxygen vacancies can exist in the Y2Ti2O7 lattice because the large unit cell of Y2Ti2O7 allows some of the oxygen ions to move relatively freely, which results in the formation of small polarons [57]. Therefore, oxygen vacancies could be the dominant vacancy defects in amorphous Y2Ti2O7 thin layers.
According to the defect reaction equation, two free charge carriers can be generated from the creation of one oxygen vacancy [60]. Like the Moss–Burstein effect, the lowest state of the conduction band is blocked by these free charge carriers, resulting in an increase in the optical bandgap energy [61,62]. After higher temperature annealing, the crystallization of amorphous Y2Ti2O7 thin layers was improved with the annihilation of oxygen vacancies, which is associated with the reduction in free charge carriers and the decrease in the optical bandgap energy, as shown in Figure 6.

4. Conclusions

Y2Ti2O7 thin layers with a thickness of ~400 nm thin layers were fabricated by the sol–gel technique. At annealing temperatures below 700 °C, all thin layers maintain an amorphous structure. The maximum transmittance of amorphous thin layers is approximately 92.3%. The transmittance is high because of small grain size, low surface roughness, and weak scattering. The refractive indices and optical bandgap energies of Y2Ti2O7 thin layers are strongly related to the annealing temperatures. The enlarged refractive indices with the increase in annealing temperatures are attributed to thin layers with the improved densification and crystallinity. The refractive indices (n at λ = 550 nm) of Y2Ti2O7 thin layers can be altered from 1.931 to 1.954 as the annealing temperature raises from 400 to 700 °C/1 h. The amorphous Y2Ti2O7 thin layers annealed at higher temperatures possess smaller optical bandgap energy (4.319 eV), which could be attributed to the Burstein–Moss shift.

Author Contributions

Conceptualization, C.-C.T.; Formal analysis, H.-A.T., Y.-Y.C., Z.-M.L. and C.-C.T.; Investigation, H.-A.T., Y.-Y.C., Z.-M.L. and C.-C.T.; Methodology, H.-A.T., Y.-Y.C. and Z.-M.L.; Supervision, C.-C.T.; Validation, C.-C.T.; Writing—original draft, H.-A.T., Y.-Y.C., Z.-M.L. and C.-C.T.; Writing—review & editing, Y.-P.H., S.-K.C. and C.-C.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research was founded by the Ministry of Science and Technology, Taiwan, the Republic of China under Contract No. MOST 108-2221-E-194-001.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Suganya, M.; Ganesan, K.; Vijayakumar, P.; Gill, A.S.; Ramaseshan, R.; Ganesamoorthy, S. Structural, optical and mechanical properties of Y2Ti2O7 single crystal. Scr. Mater. 2020, 187, 227–231. [Google Scholar] [CrossRef]
  2. Zhao, H.; Liu, T.; Bai, Z.; Wang, L.; Gao, W.; Zhang, L. Corrosion behavior of 14Cr ODS steel in supercritical water: The influence of substituting Y2O3 with Y2Ti2O7 nanoparticles. Corros. Sci. 2020, 163, 108272. [Google Scholar] [CrossRef]
  3. Badjeck, V.; Walls, M.; Chaffron, L.; Malaplate, J.; March, K. New insights into the chemical structure of Y2Ti2O7− δ nanoparticles in oxide dispersion-strengthened steels designed for sodium fast reactors by electron energy-loss spectroscopy. J. Nucl. Mater. 2015, 456, 292–301. [Google Scholar] [CrossRef]
  4. Lai, F.; Cheng, H.; Jiang, Z.; Liu, C.; Liang, T.; You, W. Manipulation of upconversion property and enhancement of sensitivity by tailoring the local structure. J. Am. Ceram. Soc. 2021, 104, 369–382. [Google Scholar] [CrossRef]
  5. Dai, P.; Zhang, X.; Zhou, M.; Li, X.; Yang, J.; Sun, P.; Xu, C.; Liu, Y. Thermally stable pyrochlore Y2Ti2O7: Eu3+ orange–red emitting phosphors. J. Am. Ceram. Soc. 2012, 95, 658–662. [Google Scholar] [CrossRef]
  6. Yu, X.; Zhao, T.; Wang, T.; Bao, W.; Zhang, H.; Su, C. Up-conversion luminescence properties of Ho3+-Yb3+ Co-doped transparent glass ceramics containing Y2Ti2O7. J. Non-Cryst. Solids 2021, 574, 121163. [Google Scholar] [CrossRef]
  7. Merka, O.; Bahnemann, D.W.; Wark, M. Improved photocatalytic hydrogen production by structure optimized nonstoichiometric Y2Ti2O7. ChemCatChem 2012, 4, 1819–1827. [Google Scholar] [CrossRef]
  8. Wang, Z.; Wang, X.; Zhou, G.; Xie, J.; Wang, S. Highly transparent yttrium titanate (Y2Ti2O7) ceramics from co-precipitated powders. J. Eur. Ceram. Soc. 2019, 39, 3229–3234. [Google Scholar] [CrossRef]
  9. Yamaguchi, S.; Kobayashi, K.; Abe, K.; Yamazaki, S.; Iguchi, Y. Electrical conductivity and thermoelectric power measurements of Y2Ti2O7. Solid State Ion. 1998, 113, 393–402. [Google Scholar] [CrossRef]
  10. Gill, J.K.; Pandey, O.; Singh, K. Ionic conductivity, structural and thermal properties of pure and Sr2+ doped Y2Ti2O7 pyrochlores for SOFC. Solid State Sci. 2011, 13, 1960–1966. [Google Scholar] [CrossRef]
  11. Abe, R.; Higashi, M.; Zou, Z.; Sayama, K.; Abe, Y. Photocatalytic water splitting into H2 and O2 over R2Ti2O7 (R = Y, rare earth) with pyrochlore structure. Chem. Lett. 2004, 33, 954–955. [Google Scholar] [CrossRef]
  12. Gill, J.K.; Pandey, O.; Singh, K. Ionic conductivity, structural and thermal properties of Ca2+ doped Y2Ti2O7 pyrochlores for SOFC. Int. J. Hydrogen Energy 2012, 37, 3857–3864. [Google Scholar] [CrossRef]
  13. Higashi, M.; Abe, R.; Sayama, K.; Sugihara, H.; Abe, Y. Improvement of photocatalytic activity of titanate pyrochlore Y2Ti2O7 by addition of excess Y. Chem. Lett. 2005, 34, 1122–1123. [Google Scholar] [CrossRef]
  14. Mahapatra, A.; Subudhi, S.; Swain, S.; Sahu, R.; Negi, R.; Samanta, B.; Kumar, P. Electrical and optical properties of yttrium titanate thin films synthesized by Sol-Gel technique. Integr. Ferroelectr. 2019, 203, 43–51. [Google Scholar] [CrossRef]
  15. Wen, Q.; Zhou, W.; Gao, H.; Zhou, Y.; Luo, F.; Zhu, D.; Huang, Z.; Qing, Y. Enhanced dielectric and microwave absorption properties of Y2Ti2O7 ceramics by Sr doping. Appl. Phys. A 2019, 125, 413. [Google Scholar] [CrossRef]
  16. Öztürk, E.; Sarılmaz, E. The investigation of the photoluminescent and piezoelectric effect of Eu3+ doped Y2Ti2O7 and Sm2Ti2O7 host crystals. Mater. Chem. Phys. 2020, 239, 122085. [Google Scholar] [CrossRef]
  17. Li, W.; Chuah, C.Y.; Yang, Y.; Bae, T.-H. Nanocomposites formed by in situ growth of NiDOBDC nanoparticles on graphene oxide sheets for enhanced CO2 and H2 storage. Microporous Mesoporous Mater. 2018, 265, 35–42. [Google Scholar] [CrossRef]
  18. Pace, S.; Cannillo, V.; Wu, J.; Boccaccini, D.; Seglem, S.; Boccaccini, A. Processing glass–pyrochlore composites for nuclear waste encapsulation. J. Nucl. Mater. 2005, 341, 12–18. [Google Scholar] [CrossRef]
  19. Goodenough, J.; Castellano, R. Defect pyrochlores as catalyst supports. J. Solid State Chem. 1982, 44, 108–112. [Google Scholar] [CrossRef]
  20. Munawar, K.; Mansoor, M.A.; Olmstead, M.M.; Yusof, F.B.; Misran, M.B.; Basirun, W.J.; Mazhar, M. Pyrochlore-structured Y2Ti2O7–2TiO2 composite thin films for photovoltaic applications. J. Aust. Ceram. Soc. 2019, 55, 921–932. [Google Scholar] [CrossRef]
  21. Mondal, K.; Hartman, K.; Dasgupta, D.; Trifon, G.; Dasari, M. Synthesis and characterization of Y2Ti2O7 and ErxY2−xTi2O7 nanofibers. J. Sol-Gel Sci. Technol. 2015, 73, 265–269. [Google Scholar] [CrossRef]
  22. Jenouvrier, P.; Boccardi, G.; Fick, J.; Jurdyc, A.-M.; Langlet, M. Up-conversion emission in rare earth-doped Y2Ti2O7 sol–gel thin films. J. Lumin. 2005, 113, 291–300. [Google Scholar] [CrossRef]
  23. Jenouvrier, P.R.; Langlet, M.; Fick, J. High photoluminescence in a new nanocrystalline active phase: YETO. In Proceedings of the Sol-Gel Optics VI. International Symposium on Optical Science and Technology, Seattle, WA, USA, 23 October 2002; pp. 52–59. [Google Scholar]
  24. Tu, X.; Xu, J.; Li, M.; Xie, T.; Lei, R.; Wang, H.; Xu, S. Color-tunable upconversion luminescence and temperature sensing behavior of Tm3+/Yb3+ codoped Y2Ti2O7 phosphors. Mater. Res. Bull. 2019, 112, 77–83. [Google Scholar] [CrossRef]
  25. Huang, X.; Huang, K.; Chen, L.; Chen, N.; Lei, R.; Zhao, S.; Xu, S. Effect of Li+/Mg2+ co-doping and optical temperature sensing behavior in Y2Ti2O7: Er3+/Yb3+ upconverting phosphors. Opt. Mater. 2020, 107, 110114. [Google Scholar] [CrossRef]
  26. Reisfeld, R.; Saraidarov, T.; Panzer, G.; Levchenko, V.; Gaft, M. New optical material europium EDTA complex in polyvinyl pyrrolidone films with fluorescence enhanced by silver plasmons. Opt. Mater. 2011, 34, 351–354. [Google Scholar] [CrossRef]
  27. Henkes, A.E.; Bauer, J.C.; Sra, A.K.; Johnson, R.D.; Cable, R.E.; Schaak, R.E. Low-temperature nanoparticle-directed solid-state synthesis of ternary and quaternary transition metal oxides. Chem. Mater. 2006, 18, 567–571. [Google Scholar] [CrossRef]
  28. Lee, W.J.; Bae, D.S. Synthesis and Characterization of Y2Ti2O7 Photocatalytic Powders by Thermal Assist Process. Defect Diffus. Forum 2017, 380, 86–91. [Google Scholar] [CrossRef]
  29. Gadipelly, T.; Dasgupta, A.; Ghosh, C.; Krupa, V.; Sornadurai, D.; Sahu, B.K.; Dhara, S. Synthesis and structural characterisation of Y2Ti2O7 using microwave hydrothermal route. J. Alloys Compd. 2020, 814, 152273. [Google Scholar] [CrossRef]
  30. Fuentes, A.F.; Boulahya, K.; Maczka, M.; Hanuza, J.; Amador, U. Synthesis of disordered pyrochlores, A2Ti2O7 (A = Y, Gd and Dy), by mechanical milling of constituent oxides. Solid State Sci. 2005, 7, 343–353. [Google Scholar] [CrossRef]
  31. Simondon, E.; Giroux, P.-F.; Chaffron, L.; Fitch, A.; Castany, P.; Gloriant, T. Mechanical synthesis of nanostructured Y2Ti2O7 pyrochlore oxides. Solid State Sci. 2018, 85, 54–59. [Google Scholar] [CrossRef]
  32. Dimesso, L. Pechini processes: An alternate approach of the sol–gel method, preparation, properties, and applications. Handb. Sol-Gel Sci. Technol. 2016, 2, 1–22. [Google Scholar]
  33. Wang, S.; Wang, L.; Ewing, R.; Kutty, K.G. Ion irradiation of rare-earth-and yttrium-titanate-pyrochlores. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2000, 169, 135–140. [Google Scholar] [CrossRef]
  34. Vishwakarma, P.; Shahi, P.; Rai, S.; Bahadur, A. Low temperature optical sensor based on non-thermally coupled level of Ho3+ and defect level of Zn2+ in Yb3+: Y2Ti2O7 phosphor. J. Phys. Chem. Solids 2020, 142, 109445. [Google Scholar] [CrossRef]
  35. Chen, Z.; Gong, W.; Chen, T.; Li, S.; Wang, D.; Wang, Q. Preparation and upconversion luminescence of Er3+/Yb3+ codoped Y2Ti2O7 nanocrystals. Mater. Lett. 2012, 68, 137–139. [Google Scholar] [CrossRef]
  36. Pavitra, E.; Raju, G.S.R.; Yu, J.S. Solvothermal synthesis and luminescent properties of Y2Ti2O7: Eu3+ spheres. Phys. Status Solidi Rapid Res. Lett. 2013, 7, 224–227. [Google Scholar] [CrossRef]
  37. Zhang, X.; Yang, H.; Tang, A. Optical, electrochemical and hydrophilic properties of Y2O3 doped TiO2 nanocomposite films. J. Phys. Chem. B 2008, 112, 16271–16279. [Google Scholar] [CrossRef]
  38. Guo, J.; Li, J.; Kou, H. Chemical preparation of advanced ceramic materials. In Modern Inorganic Synthetic Chemistry; Elsevier: Amsterdam, The Netherlands, 2011; pp. 429–454. [Google Scholar]
  39. Chen, Z.S.; Gong, W.P.; Chen, T.F.; Xiong, G.X.; Huang, G.L. Preparation of Y2Ti2O7 nanocrystal by sol-gel method and its characterization. Adv. Mater. Res. 2010, 97, 2175–2179. [Google Scholar] [CrossRef]
  40. Jenouvrier, P.; Langlet, M.; Rimet, R.; Fick, J. Influence of crystallisation on the photoluminescence properties of Y2−xErxTi2O7 sol-gel thin films. Appl. Phys. A 2003, 77, 687–692. [Google Scholar] [CrossRef]
  41. Watanabe, T. Nano-Plating: Microstructure Control Theory of Plated Film and Data Base of Plated Film Microstructure; Elsevier: Amsterdam, The Netherlands, 2004. [Google Scholar]
  42. Swanepoel, R. Determination of the thickness and optical constants of amorphous silicon. J. Phys. E Sci. Instrum. 1983, 16, 1214. [Google Scholar] [CrossRef]
  43. Ting, C.-C.; Chiu, Y.-S.; Chang, C.-W.; Chuang, L.-C. Visible and infrared luminescence properties of Er3+-doped Y2Ti2O7 nanocrystals. J. Solid State Chem. 2011, 184, 563–571. [Google Scholar] [CrossRef]
  44. Prathap, P.; Revathi, N.; Subbaiah, Y.V.; Reddy, K.R. Thickness effect on the microstructure, morphology and optoelectronic properties of ZnS films. J. Phys. Condens. Matter 2007, 20, 035205. [Google Scholar] [CrossRef]
  45. Sinha, G.; Adhikary, K.; Chaudhuri, A. Sol–gel derived phase pure α-Ga2O3 nanocrystalline thin film and its optical properties. J. Cryst. Growth 2005, 276, 204–207. [Google Scholar] [CrossRef]
  46. Lee, S.-M.; Joo, Y.-H.; Kim, C.-I. Influences of film thickness and annealing temperature on properties of sol–gel derived ZnO–SnO2 nanocomposite thin film. Appl. Surf. Sci. 2014, 320, 494–501. [Google Scholar] [CrossRef]
  47. Guang-Lei, T.; Hong-Bo, H.; Jian-Da, S. Effect of microstructure of TiO2 thin films on optical band gap energy. Chin. Phys. Lett. 2005, 22, 1787. [Google Scholar] [CrossRef]
  48. Bao, D.; Yao, X.; Wakiya, N.; Shinozaki, K.; Mizutani, N. Band-gap energies of sol-gel-derived SrTiO3 thin films. Appl. Phys. Lett. 2001, 79, 3767–3769. [Google Scholar] [CrossRef]
  49. Peng, L.; Fang, L.; Yang, X.; Li, Y.; Huang, Q.; Wu, F.; Kong, C. Effect of annealing temperature on the structure and optical properties of In-doped ZnO thin films. J. Alloys Compd. 2009, 484, 575–579. [Google Scholar] [CrossRef]
  50. Xin, G.; Guo, W.; Ma, T. Effect of annealing temperature on the photocatalytic activity of WO3 for O2 evolution. Appl. Surf. Sci. 2009, 256, 165–169. [Google Scholar] [CrossRef]
  51. Kumar, M.; Hazra, S.; Som, T. Role of metallic-like conductivity in unusual temperature-dependent transport in n-ZnO: Al/p-Si heterojunction diode. J. Phys. D Appl. Phys. 2015, 48, 455301. [Google Scholar] [CrossRef]
  52. Shinde, V.; Lokhande, C.; Mane, R.; Han, S.-H. Hydrophobic and textured ZnO films deposited by chemical bath deposition: Annealing effect. Appl. Surf. Sci. 2005, 245, 407–413. [Google Scholar] [CrossRef]
  53. Roy, S.C.; Sharma, G.; Bhatnagar, M. Large blue shift in the optical band-gap of sol–gel derived Ba0.5Sr0.5TiO3 thin films. Solid State Commun. 2007, 141, 243–247. [Google Scholar] [CrossRef]
  54. Saravanan, K.V.; Sudheendran, K.; Krishna, M.G.; Raju, K.J. Effect of the amorphous-to-crystalline transition in Ba0.5Sr0.5TiO3 thin films on optical and microwave dielectric properties. J. Phys. D Appl. Phys. 2009, 42, 045401. [Google Scholar] [CrossRef]
  55. Wang, Z.; Zou, T.; Xing, X.; Zhao, R.; Wang, Z.; Yang, Y.; Wang, Y. CdIn2O4 nanoporous thin film gas-sensor for formaldehyde detection. Phys. E Low-Dimens. Syst. Nanostruct. 2018, 103, 18–24. [Google Scholar] [CrossRef]
  56. Deokate, R.; Bhosale, C.; Rajpure, K. Synthesis and characterization of CdIn2O4 thin films by spray pyrolysis technique. J. Alloys Compd. 2009, 473, L20–L24. [Google Scholar] [CrossRef]
  57. Chatten, R.; Chadwick, A.V.; Rougier, A.; Lindan, P.J. The oxygen vacancy in crystal phases of WO3. J. Phys. Chem. B 2005, 109, 3146–3156. [Google Scholar] [CrossRef] [PubMed]
  58. Gillet, M.; Lemire, C.; Gillet, E.; Aguir, K. The role of surface oxygen vacancies upon WO3 conductivity. Surf. Sci. 2003, 532, 519–525. [Google Scholar] [CrossRef]
  59. Iijima, K.; Goto, M.; Enomoto, S.; Kunugita, H.; Ema, K.; Tsukamoto, M.; Ichikawa, N.; Sakama, H. Influence of oxygen vacancies on optical properties of anatase TiO2 thin films. J. Lumin. 2008, 128, 911–913. [Google Scholar] [CrossRef]
  60. Sudo, H.; Nakamura, K.; Maeda, S.; Okamura, H.; Izunome, K.; Sueoka, K. Point Defect Reaction in Silicon Wafers by Rapid Thermal Processing at More Than 1300 °C Using an Oxidation Ambient. ECS J. Solid State Sci. Technol. 2019, 8, P35. [Google Scholar] [CrossRef]
  61. Peelaers, H.; Van de Walle, C.G. Sub-band-gap absorption in Ga2O3. Appl. Phys. Lett. 2017, 111, 182104. [Google Scholar] [CrossRef]
  62. Li, K.; Gao, Q.; Zhao, L.; Liu, Q. Electrical and Optical Properties of Nb-doped SrSnO3 Epitaxial Films Deposited by Pulsed Laser Deposition. Nanoscale Res. Lett. 2020, 15, 164. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of the Y2Ti2O7 thin layers annealed at the temperatures of 400 to 750 °C for 1 h. The 2θ positions for bulk Y2Ti2O7 are shown on the plot for the reference.
Figure 1. XRD patterns of the Y2Ti2O7 thin layers annealed at the temperatures of 400 to 750 °C for 1 h. The 2θ positions for bulk Y2Ti2O7 are shown on the plot for the reference.
Crystals 12 00564 g001
Figure 2. Top-view SEM imagines of the Y2Ti2O7 thin layers annealed at 400, 500, 600, 700, and 750 °C for 1 h (magnification: 200 K—400–750 °C, 300 K—750 °C).
Figure 2. Top-view SEM imagines of the Y2Ti2O7 thin layers annealed at 400, 500, 600, 700, and 750 °C for 1 h (magnification: 200 K—400–750 °C, 300 K—750 °C).
Crystals 12 00564 g002
Figure 3. Transmittance spectra of the Y2Ti2O7 thin layers annealed at the temperatures of 400 to 750 °C for 1 h.
Figure 3. Transmittance spectra of the Y2Ti2O7 thin layers annealed at the temperatures of 400 to 750 °C for 1 h.
Crystals 12 00564 g003
Figure 4. Wavelength dependence of the refractive index for the Y2Ti2O7 thin layers annealed at 400, 500, 600, and 700 °C for 1 h.
Figure 4. Wavelength dependence of the refractive index for the Y2Ti2O7 thin layers annealed at 400, 500, 600, and 700 °C for 1 h.
Crystals 12 00564 g004
Figure 5. (a) (αhν)2 as a function of photon energy (E) of the Y2Ti2O7 thin layers annealed at 400–700 °C/1 h. (b) Enlarged diagram of (a). (c) Optical bandgap as a function of annealing temperatures. The error bars represent the statistical deviation among five samples.
Figure 5. (a) (αhν)2 as a function of photon energy (E) of the Y2Ti2O7 thin layers annealed at 400–700 °C/1 h. (b) Enlarged diagram of (a). (c) Optical bandgap as a function of annealing temperatures. The error bars represent the statistical deviation among five samples.
Crystals 12 00564 g005
Figure 6. Schematic representation of the Burstein–Moss (B-M) shift for the Y2Ti2O7 thin layers annealed at 400 and 700 °C for 1 h.
Figure 6. Schematic representation of the Burstein–Moss (B-M) shift for the Y2Ti2O7 thin layers annealed at 400 and 700 °C for 1 h.
Crystals 12 00564 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ting, H.-A.; Chen, Y.-Y.; Li, Z.-M.; Hsieh, Y.-P.; Chiu, S.-K.; Ting, C.-C. The Effect of Annealing on the Optoelectronic Properties and Energy State of Amorphous Pyrochlore Y2Ti2O7 Thin Layers by Sol–Gel Synthesis. Crystals 2022, 12, 564. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst12040564

AMA Style

Ting H-A, Chen Y-Y, Li Z-M, Hsieh Y-P, Chiu S-K, Ting C-C. The Effect of Annealing on the Optoelectronic Properties and Energy State of Amorphous Pyrochlore Y2Ti2O7 Thin Layers by Sol–Gel Synthesis. Crystals. 2022; 12(4):564. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst12040564

Chicago/Turabian Style

Ting, Hsiang-An, Yong-Yu Chen, Zong-Ming Li, Ya-Ping Hsieh, Sheng-Kuei Chiu, and Chu-Chi Ting. 2022. "The Effect of Annealing on the Optoelectronic Properties and Energy State of Amorphous Pyrochlore Y2Ti2O7 Thin Layers by Sol–Gel Synthesis" Crystals 12, no. 4: 564. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst12040564

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