Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Red photoluminescent Eu3+-doped Y2O3 nanospheres for LED-phosphor applications: Synthesis and characterization

Open Access Open Access

Abstract

Y2O3:Eu3+ nanospheres with sizes of 40-334 nm in diameter were obtained using a low-cost co-precipitation method followed by a thermal annealing process. The sizes of the nanospheres were controlled by tuning the synthesis time and annealing temperature. X-ray diffraction patterns and scanning electron microscopy images were used to determine the structure, shape, and size of the obtained nanoparticles. The optical properties of the nanospheres were investigated with measurements of the photoluminescence excitation and emission spectra. A phase transformation of the nanospheres from an amorphous structure to a cubic crystalline Y2O3 structure was observed when the annealing temperature was higher than 500 °C. Intense red photoluminescence emission and UV excitation of the nanospheres with a crystal structure were identified. In addition, the optimal concentration of dopant (Eu3+) for the red emission was determined to be ~8 mol%. The unique structural and optical properties of the Y2O3:Eu3+ nanospheres could lead to efficient red LED-phosphors for use in generating white light with GaAlN-based UV LEDs.

© 2018 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Europium-doped yttrium oxide (Y2O3:Eu3+) is a well-known red-emitting phosphor widely used in various areas such as fluorescent lamps, plasma display panels, flat-panel and field emission displays, and cathode-ray tubes [1–9]. Y2O3:Eu3+ materials involve abundant 4f electronic configurations and rich energy levels due to the Eu3+ luminescence activator [4,5,10]. Within the host Y2O3 lattice, Eu3+ exhibits efficient luminescence emissions with excellent color quality. In addition, Y2O3:Eu3+ materials have high chemical durability and thermal stability, simple chemical composition, low toxicity, and excellent thermal conductivity [1–9]. Recently, Y2O3:Eu3+ materials have been suggested as a red-emitting photon down converter to generate white light emission with near ultraviolet (UV)- or UV-light-emitting diodes (LEDs) due to the strong light absorption in the UV region of the Eu3+ activator [11–14]. With high emission efficiency group-III nitride-based LEDs (GaN, GaAlN, InGaN, and InGaAlN), which extend the emission spectrum to the green, blue, and ultraviolet regions, development of photon down-conversion materials (LED phosphors), which will combine with blue/UV LEDs to generate high quality white light, has recently been suggested as one of the most important research and development plans for LED-based solid-state lighting technology [15–17]. Development of such phosphor-converted white LEDs would provide a solid-state light source for next-generation lighting industry and display systems, and advance solid-state lighting technology by increasing energy efficiency, which would lead to cheaper lighting and power costs, reduce fossil fuel dependency, and help decrease carbon emissions [15–17]. The synthesis and characterization of high-efficiency Y2O3:Eu3+ LED-phosphors are therefore significant.

While bulk and nano-sized Y2O3:Eu3+ materials with various morphologies have been synthesized [3,6,7,18–45], the most promising Y2O3:Eu3+ LED-phosphor could be Y2O3:Eu3+ nanospheres. With sizes of a few hundred nanometers, it is expected that the application of such nanospheres (a photon down converter) on the top of LEDs may lead to the enhancement of the light extraction efficiency. It is known that, for high-quality luminescence, the preferred morphology of phosphor particles is a perfectly spherical shape as spherical phosphors have high packing densities and low scattering of light [29,40,46–48]. Moreover, recent studies have shown that the deposition of TiO2 arrays with spheres of 500-900 nm in diameter on the top of GaN-based LEDs could enhance the light extraction efficiency by 4-5 times of that of planar LEDs [49,50]. While similar studies with Y2O3:Eu3+ nanospheres and arrays with smaller sizes of nanospheres have not been reported, the mechanism for increasing the light extraction efficiency suggests that Y2O3:Eu3+ nanospheres with a few hundred nanometers in diameter may serve as an efficient LED-phosphor.

Y2O3:Eu3+ nanoparticles with spherical or sphere-like shapes have been synthesized with various approaches [3,6,11,20,22–26,29,32,35,36,40,43]. However, most of the previous syntheses focused on nanoparticles with sizes smaller than 100 nm [3,6,11,22–26,35,36,40,43]. Reports of syntheses of Y2O3:Eu3+ nanospheres with sizes of 100-500 nm are relatively rare [20,29,32]. Li et al. employed a homogeneous precipitation method to synthesize Y2O3:Eu3+ spheres with ~200 nm in diameter [32]. Using a similar approach, Yamagata et al. also obtained nanospheres with sizes of 200-300 nm for potential applications in orthodontic adhesive fluorescent [20]. On the other hand, Yan et al. used a low-temperature reflux method to synthesize Y2O3:Eu3+ nanospheres with sizes of 80-140 nm [29].

Here, we report the synthesis of Y2O3:Eu3+ nanospheres using a modified co-precipitation method. The nanospheres with a wide range of sizes from 40 nm to 334 nm were obtained by tuning the annealing temperature and synthesis time (reaction time). The optimal concentration of Eu3+ was determined. We also report investigations of the structural and luminescent properties of the Y2O3:Eu3+ nanospheres with X-ray diffraction, SEM imaging, and photoluminescence measurements. Strong red emission and UV excitation of the Y2O3:Eu3+ nanospheres synthesized after a phase transformation at high annealing temperatures were identified, indicating a potential of applying Eu3+-doped Y2O3 nanospheres in high-efficiency white LEDs.

2. Experimental details

We employed a modified urea co-precipitation method similar to the one used by Fukushima et al. [51]. According to the desired concentration (5, 8, 10, and 15 mol% of Eu3+), a total of 0.0080 mole of Y2O3 and Eu2O3 were placed with 80 mL of nitric acid in a flask under a hot plate set to 300 °C while stirring magnetically. Once the solution had fully reacted into Y(NO3)3 and Eu(NO3)3, as indicated by a transparent solution, the hot plate was heated to 540 °C until the solution evaporated, forming a light yellow translucent solid. The flask was transferred to an unheated plate with a magnetic stirrer, and 25 g of urea was added. Deionized water (DIW) was added to make a solution of 180 mL. The solution was magnetically stirred until the solid on the bottom had dissolved and the solution was completely transparent. The flask was then placed in an 80 °C water bath for 1, 2, 3, 4, or 5 hours (reaction time) while magnetically stirring. The solution turned translucent milky white as it was heated at 80 °C, eventually reaching a thick white color indicating that the reaction had completed. The resulting solution was centrifuged repeatedly at 15000 rpm for 15 minutes, each time being rinsed (once each with DIW and ethanol) and placed in an ultrasonic cleaner for 10 minutes to remove impurities. This left a white solid that was either left overnight or heated in an oven for around an hour at 80 °C to dry. Finally, the dried powder was annealed at various temperatures ranging from 300 to 900 °C (annealing temperatures) for 1 hour.

Scanning electron microscopy (SEM) images of the obtained samples were taken on the Inspect S50 (FEI Company) SEM. X-ray diffraction (XRD) measurements were performed to study the structure of the samples, using a Rigaku SmartLab diffractometer with Cu Kα1 radiation, λ = 1.54 Å operating at 40 kV and 44 mA. The XRD patterns were obtained at room temperature at an angular range (2θ) of 3°–80° with a step of 0.01°. Photoluminescence spectra were measured using a spectrofluorophotometer (Shimadzu, RF-6000) with a xenon lamp as the excitation source.

3. Results and discussions

Figure 1 shows the SEM images of five samples with the Eu3+ concentration of 5 mol% synthesized at 80 °C for 4 hours and thermally annealed at various temperatures, or not annealed. All five samples are composed of nanospheres, which have a size range of approximately 58-212 nm in diameter. Before annealing, the nanospheres have an average size of ~100 nm and a size range of 85-112 nm [Fig. 1(a)]. After annealing at a relatively low temperature (300 °C), the nanoparticles are still spherical in shape, but the size range becomes broader (58-141 nm) and the average size is slightly smaller [Fig. 1(b)]. However, both larger and smaller nanospheres are observed. Annealing at a temperature of 500 °C resulted in nanospheres with a size range of 87-212 nm and a greater average size of ~130 nm [Fig. 1(c)].

 figure: Fig. 1

Fig. 1 SEM images of nanospheres with the Eu concentration of 5 mol% synthesized at 80 °C for 4 hours, with or without a follow-up annealing process: (a) without annealing, (b)-(f) after annealing for 1 hour at an annealing temperature of 300 °C (b), 500 °C (c), 700 °C (d), and 900 °C (e, f).

Download Full Size | PDF

With a higher annealing temperature (700 °C), the nanospheres have a size range of 87-143 nm, and the average size of ~120 nm is slightly reduced [Fig. 1(d)]. Finally, with a very high annealing temperature (900 °C), the nanospheres have an average size of ~145 nm in diameter, significantly greater than those of the nanospheres obtained with lower annealing temperatures, as shown in Figs. 1(e) and 1(f).

The size variation with the annealing temperature can be understood with information about the structure of the nanospheres. Figure 2 shows the XRD patterns of the five samples. The XRD spectra clearly show that the nanospheres obtained before annealing and with annealing at the temperatures of 300 °C and 500 °C do not exhibit any crystalline features, indicating that the nanospheres are amorphous in structure. The amorphous nanospheres obtained prior to annealing will experience thermal decomposition and recombination during the annealing process. At 300 °C, decomposition results in the formation of small nanospheres [Fig. 1(b)]. Meanwhile, larger nanospheres may be formed by the recombination of small particles. It appears that decomposition is slightly more preferred than recombination at 300 °C, resulting in the formation of nanospheres with a slightly smaller mean size. At 500 °C, recombination dominates, making the nanospheres greater in size [Fig. 1(c)].

 figure: Fig. 2

Fig. 2 XRD spectra of the samples with the Eu3+ concentration of 5 mol% synthesized at 80 °C for 4 hours, both with and without a follow-up annealing process. The annealed samples were obtained after annealing for 1 hour at an annealing temperature of 300 °C, 500 °C, 700 °C, and 900 °C, respectively.

Download Full Size | PDF

On the other hand, the nanospheres undergo a phase transformation from an amorphous structure to a cubic structure when annealing at temperatures above 500 °C (at 700 °C and 900 °C). A similar phase transformation of bulk Y2O3 at an annealing temperature of 600 °C was previously reported [44].

As shown in Fig. 2, the nanospheres that were annealed at 700 °C and 900 °C have a crystal structure. The locations and relative intensities of the XRD intensity peaks shown in Fig. 2 agree with those of the cubic phase of Y2O3 [52,53]. The four major peaks at approximately 29° (2θ), 34°, 48°, and 57° (Fig. 2) correspond to the (222), (400), (440), and (622) diffraction directions of cubic Y2O3, respectively. The other four peaks in the range of ~20-46°, ~20°, 35°, 39°, 42°, and 46°, correspond to the (211), (411), (322), (134), and (125) directions, respectively. In addition, stronger XRD intensities of the Y2O3:Eu3+ nanospheres obtained after annealing at 900 °C suggest that higher annealing temperature may yield higher crystallinity. With an annealing temperature of 700 °C, the phase transformation to the crystalline structure results in a slight decrease in the size of the nanospheres. With a higher annealing temperature of 900 °C, the sintering of small nanoparticles generates larger nanospheres, which exhibit well-ordered cubic structure. The general trend is that higher annealing temperatures yield larger nanospheres, particularly when the nanospheres are with a crystal structure.

Figure 3(a) shows the photoluminescence spectra of the five samples with nanospheres under the excitation of 250-nm UV light. Strong emission intensities are found in the two samples annealed at 700 °C and 900 °C, which are composed of Y2O3:Eu3+ nanospheres with a cubic structure. On the other hand, the three samples with an amorphous structure (unannealed and annealed at 300 °C and 500 °C) all show very weak emission intensities.

 figure: Fig. 3

Fig. 3 (a) Photoluminescence emission spectra of the nanospheres with the Eu3+ concentration of 5 mol% synthesized at 80 °C for 4 hours, with or without a follow-up annealing process (1-hour annealing at four different annealing temperatures). (b) The corresponding excitation spectra of the nanospheres.

Download Full Size | PDF

The strongest emission of each of the two samples with a cubic structure is located at 611.6 nm. This orange-red line is due to the transition of 5𝐷07𝐹2 of the Eu3+ ions that occupy the lattice of cubic Y2O3 [23,45]. It is known that the 5𝐷07𝐹2 transitions of Eu3+ occupying the C2 site in cubic Y2O3 are allowed by the forced electric dipole mechanism [45]. The 611.6-nm line of the Y2O3:Eu3+ nanospheres is slightly longer in wavelength than the corresponding line of bulk Y2O3:Eu3+ nanospheres (611.3 nm) [45]. The other two significant peaks of the 900 °C sample are located at 630 nm (red) and 709.8 nm (deep red). The corresponding peaks of the 700 °C sample are almost the same in location (629.1 nm and 709.7 nm). The former peak is due to a transition from 5𝐷0 to one Stark component of 7𝐹2 while the latter is due to a 5𝐷07𝐹4 transition. Both transitions are electric dipole allowed when the Eu3+ ions occupy the C2 sites in cubic Y2O3 [4,18,45]. The intensity ratio of the 611.6-nm line to the second strongest emission line (the 630-nm line or the 629.1-nm line) is as high as ~7 (the 900 °C sample) or ~5 (700 °C sample).

As shown in Fig. 3(a), there are more weak peaks for each sample. The weak peaks for the 900 °C sample are located at 592.8 nm (orange), 652 nm (red), 662.4 nm (red), and 689.2 nm (red). The corresponding locations of the peaks for the 700 °C sample are 592.8 nm, 653 nm, 661.5 nm, and 689.2 nm, respectively. These peaks also belong to the 5𝐷07𝐹𝐽 (𝐽 = 0, 1, 2, 3, 4) transitions of the Eu3+ ions in the Y2O3:Eu3+ nanospheres [18,45]. The 592.8-nm peak is due to a 5𝐷07𝐹1 transition while the 652-nm and 662.4-nm peaks correspond to the 5𝐷07𝐹3 transitions. The 5𝐷07𝐹1 transitions are weak because they are electric dipole forbidden and magnetic dipole allowed (magnetic dipole transitions usually have much weaker intensities than electric dipole transitions). The 5𝐷07𝐹3 transitions are electric dipole forbidden per the Judd-Ofelt theory and could occur only via J-mixing (the weak interaction between the 4f orbitals and the chemical environment); therefore, the transitions have weak intensities [54,55]. Similar to the deep red line (the 709.7-nm peak), the 689.2-nm peak belongs to the 5𝐷07𝐹4 transitions.

In addition, the emission intensities of the 900 °C sample are observed to be stronger than those of the 700 °C sample. This is due to the fact that the 900 °C sample is composed of larger nanospheres (as shown in the SEM images) and higher crystallinity (as shown in the XRD patterns).

The excitation spectra of the nanospheres were also measured by monitoring the 611-nm emission, and the results are shown in Fig. 3(b). Similar to the emission spectra, the excitation spectra of the nanospheres with a crystalline structure (the 700 °C and 900 °C samples) are significantly stronger in intensity than those of the other three samples that are composed of nanospheres with an amorphous structure. For both the 700 °C and 900 °C samples, the strong excitation bands in the wavelength region below 270 nm are due to the charge transfer transitions [18,39]. The two bands with the peaks at 223 nm and 255 nm for the 900 °C sample are attributed to the O2‒ → Y3+ and O2‒ → Eu3+ transitions, respectively. The corresponding peaks for the 700 °C sample are located at ~220 nm and 256 nm.

The weaker excitation lines in the wavelength range of 350-550 nm are due to the ff transitions of Eu3+ in cubic Y2O3, among them, the 364-nm, 382-nm, 396-nm, 417-nm, 467-nm, and 535-nm lines for the 900 °C sample are due to the 7𝐹05𝐷4, 7𝐹05L7, 7𝐹05L6, 7𝐹05𝐷3, 7𝐹05𝐷2, and 7𝐹05𝐷1 transitions, respectively [4,18]. The corresponding excitation lines for the 700 °C sample are at 363 nm, 382 nm, 395 nm, 414 nm, 467 nm, and 534 nm, respectively. While the corresponding peaks of the two samples are almost at the same locations, the intensities of the transitions for the 900 °C sample are considerably stronger than those for the 700 °C sample. This is similar to the case of emission spectra.

To investigate the role of the synthesis time (the reaction time ‒ the amount of time the urea and the Y/Eu-compounds are in the heated water bath) in forming nanospheres, a set of 5 experiments were carried out with the reaction times of 1, 2, 3, 4, and 5 hours. Figure 4 shows the SEM images of the Y2O3:Eu3+ nanospheres with the Eu3+ concentration of 10 mol% synthesized at 80 °C for 2-5 hours (the SEM images for the 1-hour sample were not taken). Each sample was annealed for 1 hour at 700 °C. The SEM images clearly show that the reaction time plays a significant role in the size of the nanospheres formed. With a reaction time of 2 hours, the size range is ~40-70 nm [Fig. 4(a)]. When the reaction time was increased to 3 hours, the size range is ~40-95 nm [Fig. 4(b)]. Both samples are composed of nanospheres with sizes below 100 nm in diameter. However, when the reaction time reached 4 hours, the nanospheres obtained have a size range from ~100 nm to ~210 nm [Fig. 4(c)]. The largest nanospheres, with a size range of ~164-334 nm, were obtained when the reaction time was 5 hours, as shown in Fig. 4(d). These results suggest that longer reaction time generates large nanospheres.

 figure: Fig. 4

Fig. 4 SEM images of the Y2O3:Eu3+ nanospheres with the Eu3+ concentration of 10 mol% synthesized at 80 °C for (a) 2 hours, (b) 3 hours, (c) 4 hours, and (d) 5 hours. Each sample was annealed for 1 hour at 700 °C.

Download Full Size | PDF

Finally, another set of experiments, with 4 different concentrations of Eu3+, were performed in order to determine the optimal concentration of Eu3+ for photoluminescence emission, and the results for the emission spectra of the samples obtained are shown in Fig. 5. The 8% sample has the strongest emission intensity, followed by the 10% and 5% samples, while the 15% sample shows the weakest intensity. In terms of photoluminescence emission, therefore, the optimal Eu3+ concentration is determined to be ~8%. While it is greater than the 5 mol% optimal concentration reported for the Eu3+-doped Y2O3 nanospheres with small sizes (~5 nm in diameter) [26], the 8 mol% optimal Eu3+ concentration determined in this study is consistent with a recent study [20], in which 8 mol% of Eu3+ was found to be optimal for the nanospheres with larger sizes (200-300 nm).

 figure: Fig. 5

Fig. 5 Photoluminescence emission spectra of the Y2O3:Eu3+ nanospheres with different Eu concentrations (5-15 mol%) synthesized at 80 °C for 4 hours, followed by thermal annealing at 700 °C for 1 hour.

Download Full Size | PDF

The emission spectrum of the 8% sample contains seven obvious peaks at 612 nm (the strongest line), 629.2 nm (the 2nd strongest), 593.3 nm (the 3rd), 709.9 nm (the 4th), 654.1 nm, 660.8 nm, and 691.4 nm. These emissions can be compared with the corresponding lines of the 5% sample discussed earlier: 611.6 nm (strongest), 629.1 nm (the 2nd strongest), 709.7 nm (the 3rd), 592.8 nm (the 4th), 653 nm, 661.5 nm, and 689.2 nm. Overall, compared to the 5% sample, the peaks of the 8% sample shift slightly to the red side.

4. Conclusions

The Eu3+-doped Y2O3 nanospheres with sizes in the range of 40-334 nm in diameter were synthesized by a low-cost urea-assisted co-precipitation method and an annealing process. The sizes of the nanospheres can be controlled by the reaction time and annealing temperature. The general trend is that longer reaction times and higher annealing temperatures result in the formation of larger nanospheres. The nanospheres obtained with an annealing temperature of 500 °C or below exhibit an amorphous structure. However, the nanospheres were found to undergo a phase transformation at an annealing temperature higher than 500 °C, and the obtained nanospheres have a crystalline phase (cubic Y2O3). The nanospheres with a cubic crystal structure exhibit strong red photoluminescent emission, strong UV excitation, and large sizes (100-334 nm) suitable for enhancement of the light extraction efficiency when coated on LEDs, suggesting an application as a potentially efficient red phosphor to generate white light. In addition, the optimal concentration of Eu3+ for the photoluminescent emission and excitation was determined to be ~8 mol%.

Funding

University of Tulsa (Faculty start-up funding), (Faculty Development Summer Fellowship Program).

Acknowledgments

The authors would like to thank Professor Hongyang Zhu for helpful discussions, Professor Alexei Grigoriev for help in the XRD measurement, and Mr. Richard Portman and Mr. Gopi Adhikari for help in the SEM measurements.

References

1. W. H. Yen, S. Shionoya, and H. Yamamoto, Phosphor Handbook (CRC Press, 2006).

2. W. Chen, M. Zhuo, Y. Liu, S. Fu, Y. Liu, Y. Wang, Z. Li, Y. Li, Y. Li, and L. Yu, “Uniform octahedral-shaped Y2O3:Eu3+ submicron single crystals: Solid-state synthesis, formation mechanism and photoluminescence property,” J. Alloys Compd. 656, 764–770 (2016). [CrossRef]  

3. S. Som, S. Das, S. Dutta, H. G. Visser, M. K. Pandey, P. Kumar, R. K. Dubey, and S. K. Sharma, “Synthesis of strong red emitting Y2O3:Eu3+ phosphor by potential chemical routes: comparative investigations on the structural evolutions, photometric properties and Judd-Ofelt analysis,” RSC Advances 5(87), 70887–70898 (2015). [CrossRef]  

4. K. Binnemans, “Interpretation of europium(III) spectra,” Coord. Chem. Rev. 295, 1–45 (2015). [CrossRef]  

5. S. Gai, C. Li, P. Yang, and J. Lin, “Recent progress in rare earth micro/nanocrystals: soft chemical synthesis, luminescent properties, and biomedical applications,” Chem. Rev. 114(4), 2343–2389 (2014). [CrossRef]   [PubMed]  

6. M.-H. Lee, S.-G. Oh, and S.-C. Yi, “Preparation of Eu-doped Y2O3 luminescent nanoparticles in nonionic reverse microemulsions,” J. Colloid Interface Sci. 226(1), 65–70 (2000). [CrossRef]   [PubMed]  

7. D. K. Williams, B. Bihari, and B. M. Tissue, “Preparation and fluorescence spectroscopy of bulk monoclinic Eu3+:Y2O3 Nanocrystals,” J. Phys. Chem. 102, 916–920 (1998). [CrossRef]  

8. Y. Jiang, Z. Wang, F. Zhang, H. Paris, and C. Summers, “Synthesis and characterization of Y2O3:Eu3+ powder phosphor by a hydrolysis technique,” J. Mater. Res. 13(10), 2950–2955 (1988). [CrossRef]  

9. H. T. Hintzen and H. M. van Noort, “Investigation of luminescent Eu-doped sesquioxides Ln2O3 (Ln = In, Sc, Y, La, Gd, Lu) and some mixed oxides by 151Eu Mössbauer spectroscopy,” J. Phys. Chem. Solids 49(8), 873–881 (1998). [CrossRef]  

10. P. S. Peijzel, A. Meijerink, R. T. Wegh, M. F. Reid, and G. W. Burdick, “A complete 4fn energy level diagram for all trivalent lanthanide ions,” J. Solid State Chem. 178(2), 448–453 (2005). [CrossRef]  

11. Q. Dai, M. E. Foley, C. J. Breshike, A. Lita, and G. F. Strouse, “Ligand-passivated Eu:Y2O3 nanocrystals as a phosphor for white light emitting diodes,” J. Am. Chem. Soc. 133(39), 15475–15486 (2011). [CrossRef]   [PubMed]  

12. Y. K. Ooi, C. Liu, and J. Zhang, “Analysis on polarization-dependent light extraction and effect of passivation layer for 230 nm AlGaN nanowire light-emitting diodes,” IEEE Photonics J. 9(4), 4501712 (2017). [CrossRef]  

13. P. Pust, P. J. Schmidt, and W. Schnick, “A revolution in lighting,” Nat. Mater. 14(5), 454–458 (2015). [CrossRef]   [PubMed]  

14. X.-H. Li, T. Detchprohm, T.-T. Kao, M. M. Satter, S.-C. Shen, P. Douglas Yoder, R. D. Dupuis, S. Wang, Y. O. Wei, H. Xie, A. M. Fischer, F. A. Ponce, T. Wernicke, C. Reich, M. Martens, and M. Kneissl, “Low-threshold stimulated emission at 249 nm and 256 nm from AlGaN-based multiple-quantum-well lasers grown on sapphire substrates,” Appl. Phys. Lett. 105(14), 141106 (2014). [CrossRef]  

15. N. Bradsley, M. Hansen, L. Pattison, M. Pattison, K. Stober, V. Taylor, J. Tsao, and M. Yamada, Solid-state lighting R&D plan, edited by J. Brodrick (Solid-State Lighting Program, the U.S. Department of Energy).

16. M. Pattison, N. Bradsley, M. Hansen, L. Pattison, S. Schober, K. Stober, J. Tsao, and M. Yamada, Solid-state lighting2017suggested research topics supplement: Technology and market context, edited by J. Brodrick (Office of Energy Efficiency & Renewable Energy, the U.S. Department of Energy).

17. J. Penning, K. Stober, V. Taylor, and M. Yamada, Energy savings forecast of solid-state lighting in general illumination applications – Prepared for the U.S. Department of Energy Solid-State Lighting Program (Navigant Consulting, Inc. and the U.S. Department of Energy).

18. P. Zhu, W. Wang, H. Zhu, P. Vargas, and A. Bont, “Optical properties of Eu3+-doped Y2O3 nanotubes and nanosheets synthesized by hydrothermal method,” IEEE Photonics J. 10(1), 4500210 (2018). [CrossRef]  

19. J.-H. Wang, Z.-P. Li, B. Liu, and B.-B. Liu, “Local microstructural analysis for Y2O3/Eu3+/Mg2+ nanorods by Raman and photoluminescence spectra under high pressure,” Chin. Phys. B 26(2), 026101 (2017). [CrossRef]  

20. S. Yamagata, Y. Sato, T. Yamamoto, S. Abe, T. Akasaka, Y. Yoshida, and J. Iida, “Evaluation of fluorescent orthodontic adhesives containing Y2O3:Eu3+ particles,” Nano Biomed. 8, 35–40 (2016).

21. G. Chen, W. Qi, Y. Li, C. Yang, and X. Zhao, “Hydrothermal synthesis of Y2O3:Eu3+ nanorods and its growth mechanism and luminescence properties,” J. Mater. Sci. Mater. Electron. 27(6), 5628–5634 (2016). [CrossRef]  

22. M. Kasperczyk, S. Person, D. Ananias, L. D. Carlos, and L. Novotny, “Excitation of magnetic dipole transitions at optical frequencies,” Phys. Rev. Lett. 114(16), 163903 (2015). [CrossRef]   [PubMed]  

23. H. Cui, P. Zhu, H. Zhu, H. Li, and Q. Cui, “Photoluminescence properties and energy transfer in Y2O3:Eu3+ nanophosphors,” Chin. Phys. B 23(5), 057801 (2014). [CrossRef]  

24. A. P. Jadhav, A. U. Pawar, U. Pal, and Y. S. Kang, “Red emitting Y2O3:Eu3+ nanophosphors with >80% down conversion efficiency,” J. Mater. Chem. C Mater. Opt. Electron. Devices 2(3), 496–500 (2014). [CrossRef]  

25. Y. Xiao, D. Wu, Y. Jiang, N. Liu, J. Liu, and K. Jiang, “Nano-sized Y2O3:Eu3+ hollow spheres with enhanced photoluminescence properties,” J. Alloys Compd. 509(19), 5755–5760 (2011). [CrossRef]  

26. B. K. Gupta, D. Haranath, S. Saini, V. N. Singh, and V. Shanker, “Synthesis and characterization of ultra-fine Y2O3:Eu3+ nanophosphors for luminescent security ink applications,” Nanotechnology 21(5), 055607 (2010). [CrossRef]   [PubMed]  

27. H. Zhu, Y. Ma, H. Yang, P. Zhu, J. Du, C. Ji, and D. Hou, “Ultrastable structure and luminescence properties of Y2O3 nanotubes,” Solid State Commun. 150(27-28), 1208–1212 (2010). [CrossRef]  

28. S. Zhong, S. Wang, H. Xu, H. Hou, Z. Wen, P. Li, S. Wang, and R. Xu, “Spindlelike Y2O3:Eu3+ nanorod bundles: hydrothermal synthesis and photoluminescence properties,” J. Mater. Sci. 44(14), 3687–3693 (2009). [CrossRef]  

29. T. Yan, D. Zhang, L. Shi, H. Yang, H. Mai, and J. Fang, “Reflux synthesis, formation mechanism, and photoluminescence performance of monodisperse Y2O3:Eu3+ nanospheres,” Mater. Chem. Phys. 117(1), 234–243 (2009). [CrossRef]  

30. M. K. Devaraju, S. Yin, and T. Sato, “A rapid hydrothermal synthesis of rare earth oxide activated Y (OH)3 and Y2O3 nanotubes,” Nanotechnology 20(30), 305302 (2009). [CrossRef]   [PubMed]  

31. N. Zhang, X. Liu, R. Yi, R. Shi, G. Gao, and G. Qiu, “Selective and controlled synthesis of single-crystalline yttrium hydroxide/oxide nanosheets and nanotubes,” J. Phys. Chem. C 112(46), 17788–17795 (2008). [CrossRef]  

32. J.-G. Li, X. Li, X. Sun, and T. Ishigaki, “Monodispersed colloidal spheres for uniform Y2O3:Eu3+ red-phosphor particles and greatly enhanced luminescence by simultaneous Gd3+ doping,” J. Phys. Chem. C 112(31), 11707–11716 (2008). [CrossRef]  

33. R. Si, Y.-W. Zhang, H.-P. Zhou, L.-D. Sun, and C.-H. Yan, “Controlled-synthesis, self-assembly behavior, and surface-dependent optical properties of high-quality rare-earth oxide nanocrystals,” Chem. Mater. 19(1), 18–27 (2007). [CrossRef]  

34. X. Wu, Y. Tao, F. Gao, L. Dong, and Z. Hu, “Preparation and photoluminescence of yttrium hydroxide and yttrium oxide doped with europium nanowires,” J. Cryst. Growth 277(1-4), 643–649 (2005). [CrossRef]  

35. H. Wang, M. Uehara, H. Nakamura, M. Miyazaki, and H. Maeda, “Synthesis of well-dispersed Y2O3:Eu nanocrystals and self-assembled nanodisks using a simple non-hydrolytic route,” Adv. Mater. 17(20), 2506–2509 (2005). [CrossRef]  

36. J. A. Nelson, E. L. Brant, and M. J. Wagner, “Nanocrystalline Y2O3:Eu phosphors prepared by alkalide reduction,” Chem. Mater. 15(3), 688–693 (2003). [CrossRef]  

37. C. Wu, W. Qin, Q. Qin, D. Zhao, J. Zhang, S. Huang, S. Lü, H. Liu, and Y. Lin, “Photoluminescence from surfactant-assembled Y2O3:Eu nanotubes,” Appl. Phys. Lett. 82(4), 520–522 (2003). [CrossRef]  

38. G. Wakefield, E. Holland, P. J. Dobson, and J. L. Hutchison, “Luminescence properties of nanocrystalline Y2O3:Eu,” Adv. Mater. 13(20), 1557–1560 (2001). [CrossRef]  

39. J. Dhanaraj, R. Jagannathan, T. R. N. Kutty, and C.-H. Lu, “Photoluminescence characteristics of Y2O3:Eu3+ nanophosphors prepared using sol-gel thermolysis,” J. Phys. Chem. 105(45), 11098–11105 (2001). [CrossRef]  

40. M. I. Martinez-Rubio, T. G. Ireland, G. R. Fern, J. Silver, and M. J. Snowden, “A new application for microgels: Novel method for the synthesis of spherical particles of the Y2O3:Eu phosphor using a copolymer microgel of NIPAM and acrylic acid,” Langmuir 17(22), 7145–7149 (2001). [CrossRef]  

41. B. Bihari, H. Eilers, and B. M. Tissue, “Spectra and dynamics of monoclinic Eu2O3 and Eu3+: Y2O3 nanocrystals,” J. Lumin. 75(1), 1–10 (1997). [CrossRef]  

42. T. Ye, Z. Guiwen, Z. Weiping, and Z. Shangda, “Combustion synthesis and photoluminescence of nanocrystalline Y2O3:Eu phosphors,” Mater. Res. Bull. 32(5), 501–506 (1997). [CrossRef]  

43. H. Eilers and B. M. Tissue, “Laser spectroscopy of nanocrystalline Eu2O3 and Eu3+:Y2O3,” Chem. Phys. Lett. 251(1-2), 74–78 (1996). [CrossRef]  

44. T. Sato, S. Imaeda, and K. Sato, “Thermal transformation of yttrium hydroxides to yttrium oxides,” Thermochim. Acta 133, 79–85 (1988). [CrossRef]  

45. N. C. Chang, “Fluorescence and stimulated emission from trivalent europium in yttrium oxide,” J. Appl. Phys. 34(12), 3500–3504 (1963). [CrossRef]  

46. L. S. Wang, Y. H. Zhou, Z. W. Quan, and J. Lin, “Formation mechanisms and morphology dependent luminescence properties of Y2O3:Eu phosphors prepared by spray pyrolysis process,” Mater. Lett. 59(10), 1130–1133 (2005). [CrossRef]  

47. Y. H. Zhou, J. Lin, M. Yu, S. M. Han, S. B. Wang, and H. J. Zhang, “Morphology control and luminescence properties of YAG:Eu phosphors prepared by spray pyrolysis,” Mater. Res. Bull. 38(8), 1289–1299 (2003). [CrossRef]  

48. C. Xu, B. A. Watkins, R. E. Sievers, X. Jing, P. Trowga, C. S. Gibbons, and A. Vechi, “Submicron-sized spherical yttrium oxide based phosphors prepared by supercritical CO2-assisted aerosolization and pyrolysis,” Appl. Phys. Lett. 71(12), 1643–1645 (1997). [CrossRef]  

49. P. Zhu, “Frustrated total internal reflection in organic light-emitting diodes employing sphere cavity embedded in polystyrene,” J. Opt. 18(2), 025403 (2016). [CrossRef]  

50. P. Zhu and N. Tansu, “Effect of packing density and packing geometry on light extraction of III-nitride light-emitting diodes with microsphere arrays,” Photon. Res. 3(4), 184–191 (2015). [CrossRef]  

51. S. Fukushima, T. Furukawa, H. Niioka, M. Ichimiya, T. Sannomiya, J. Miyake, M. Ashida, T. Araki, and M. Hashimoto, “Synthesis of Y2O3 nanophosphors by homogeneous precipitation method using excessive urea for cathodoluminescence and upconversion luminescence bioimaging,” Opt. Mater. Express 6(3), 831–843 (2016). [CrossRef]  

52. JCPDS card No. 41–1105.

53. J. Zhang, H. Cui, P. Zhu, C. Ma, X. Wu, H. Zhu, Y. Ma, and Q. Cui, “Photoluminescence studies of Y2O3:Eu3+ under high pressure,” J. Appl. Phys. 115(2), 023502 (2014). [CrossRef]  

54. B. R. Judd, “Optical absorption intensities of rare-earth ions,” Phys. Rev. 127(3), 750–761 (1962). [CrossRef]  

55. G. S. Ofelt, “Intensities of crystal spectra of rare-earth ions,” J. Chem. Phys. 37(3), 511–520 (1962). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (5)

Fig. 1
Fig. 1 SEM images of nanospheres with the Eu concentration of 5 mol% synthesized at 80 °C for 4 hours, with or without a follow-up annealing process: (a) without annealing, (b)-(f) after annealing for 1 hour at an annealing temperature of 300 °C (b), 500 °C (c), 700 °C (d), and 900 °C (e, f).
Fig. 2
Fig. 2 XRD spectra of the samples with the Eu3+ concentration of 5 mol% synthesized at 80 °C for 4 hours, both with and without a follow-up annealing process. The annealed samples were obtained after annealing for 1 hour at an annealing temperature of 300 °C, 500 °C, 700 °C, and 900 °C, respectively.
Fig. 3
Fig. 3 (a) Photoluminescence emission spectra of the nanospheres with the Eu3+ concentration of 5 mol% synthesized at 80 °C for 4 hours, with or without a follow-up annealing process (1-hour annealing at four different annealing temperatures). (b) The corresponding excitation spectra of the nanospheres.
Fig. 4
Fig. 4 SEM images of the Y2O3:Eu3+ nanospheres with the Eu3+ concentration of 10 mol% synthesized at 80 °C for (a) 2 hours, (b) 3 hours, (c) 4 hours, and (d) 5 hours. Each sample was annealed for 1 hour at 700 °C.
Fig. 5
Fig. 5 Photoluminescence emission spectra of the Y2O3:Eu3+ nanospheres with different Eu concentrations (5-15 mol%) synthesized at 80 °C for 4 hours, followed by thermal annealing at 700 °C for 1 hour.
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.