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

Enhanced near-infrared emission from erbium and cerium oxide codoped silica nanocomposite

Open Access Open Access

Abstract

Er3+-doped CeO2-SiO2 nanocomposite was successfully fabricated by a sol-gel route. The homogeneous distribution of CeO2 nanocrystals among amorphous silica matrix was evidenced by the X-ray diffraction, transmission electron microscopy, and Fourier transform infrared spectroscopy. The incorporation of Er3+ ions into CeO2 nanocrystals was revealed by emission spectra and luminescence decay curves. Increasing CeO2 content and raising annealing temperature lead to the enhancement of the 1532 nm Er3+ emission, because CeO2 nanocrystals can efficiently sensitize the emission of Er3+, provide a low phonon energy environment, and defend them from hydroxyl quenching. It is expected that the investigated nanocomposite might be a promising candidate for optical fiber communication.

© 2017 Optical Society of America

1. Introduction

Recently, the development of optoelectronic technology has focused on certain new and more efficient luminescence devices [1–3]. Rare earth (RE) ions doped oxide phosphors have got much attention due to their great potential for application in the luminescence devices such as solid-state lighting, flat plane display, and optical fiber [4,5]. CeO2, which has a cubic fluorite crystal structure, is a favorable kind of host materials for phosphors owing to its unique properties, including high refractive index, optical transparency, and chemical stability [6,7]. More importantly, the incorporation of high concentration of RE ions into CeO2 nanocrystals is possible owing to the analogical radius of the dopant and the host Ce4+ ions [7]. Hence, different RE ions (such as Nd3+, Eu3+, Tm3+, Ho3+ and Yb3+) doped CeO2 nanocrystals have been widely investigated as functional materials [2, 3, 6–8]. However, the high active surface atoms induce aggregation of nanoparticles and consequently degradation of performance [9]. In order to control the size and distribution of the nanoparticles, they have been embedded into inert matrix materials to maintain their excellent performance.

Er3+ is acknowledged as an interesting activator for solid-state laser in the visible to near-infrared range owing to its characteristics of abundant energy level structure. In particular, the 4I13/24I15/2 transition emits at 1.54 µm located in the low-loss optical window for optical fiber communication [10]. Unfortunately, the excitation cross-section of Er3+ is as low as 10−21 cm2 because of the forbidden intra-4f transition [11]. In addition, the photoluminescence (PL) from 4I13/2 can be easily quenched by phonon-assisted molecule collision [12]. Fortunately, the 1.54 µm Er3+ emission can be enhanced by the efficient energy transfer from wide band gap materials, such as In2O3, ZnO [12,13].

In this work, Er3+-doped CeO2-SiO2 nanocomposite (NC) was fabricated by a sol–gel method. The enhanced 1532 nm Er3+ emission induced by the energy transfer from CeO2 nanocrystals was investigated systematically. And the influence of the annealing temperature and CeO2 content on the microstructure and PL properties of the nanocomposite was discussed in detailed.

2. Experimental

Reagent graded chemicals of tetraethylorthosilicate (TEOS), cerium nitrate (Ce(NO3)3), ethyl alcohol (EA, CH3CH2OH), erbium nitrate (Er(NO3)3) and deionized water (H2O) were used as starting materials to synthesis the gel. First, TEOS were dissolved in EA. Ce(NO3)3 and Er(NO3)3 were dissolved in deionized water. Then two parts were mixed slowly, and a trace of nitric acid was added as a catalyst for the hydrolysis. The molar ratio of TEOS:CH3CH2OH:H2O in the final solution was 1:4:10. After stirring for 3 hours, the obtained clear sol was poured into vessels to formed wet gels. The gels were then aged at room temperature for one week and finally dried in steps from 30 °C to 80 °C in one week to produce dry gels (DG) with compositions of xCeO2-(99-x)SiO2:1Er3+ (in mol%, x = 0, 5, 10, 15, 20, denoted at NC0, NC5, NC10, NC15 and NC20, respectively). The NC samples were achieved by annealing these DGs at 600°C, 800 °C, or 1000 °C for 2 h to precipitate the CeO2 nanocrystals.

X-ray diffraction (XRD) analysis was carried out with a powder diffractometer (DMAX2500 RIGAKU) using Cu Kα radiation (λ = 0.154 nm). The microstructure of the sample was characterize by a transmission electron microscope (TEM, JEM-2010). The Fourier transform infrared spectra (FTIR) were recorded in the wavelength range of 400-4000 cm−1 with 4 cm−1 resolution by the KBr method (Spectrum One). FTIR specimen was prepared by diluting the same weight of powder sample with KBr in equal proportions, then pressing into a diameter of 13 mm, thickness of 0.05 mm sheet. The PL spectra and fluorescence decay curves were recorded on an Edinburgh Instruments FLS920 spectrofluorometer.

3. Results and discussion

To identify the crystalline phase, XRD measurement was performed on the NC10 sample annealed at different temperatures. As shown in Fig. 1(a), XRD pattern of the DG one exhibits a typical amorphous feature. After annealed at 600 °C, several diffraction peaks corresponding to cubic CeO2 (JCPDs No. 65-2975) emerge. With increasing annealing temperature from 600 °C to 1000 °C, the peaks become sharp, indicating the growth of cubic CeO2 nanocrystals. Based on the Scherrer equation [14], the mean crystallite sizes of CeO2 in the sample annealed at 600 °C, 800 °C, and 1000 °C are estimated to be about 5.2, 5.6, and 10.8 nm, respectively. To further investigate the influence of the CeO2 content on the microstructure of the obtained nanocomposite, XRD patterns of NC samples with different CeO2 contents annealed at 1000 °C are displayed in Fig. 1(b). Their XRD diffraction peaks are basically coincident, including peak intensities and widths, implying that samples in this CeO2 content range are microstructural stable. The high-resolution TEM (HRTEM) image of the NC10 sample annealed at 1000 °C demonstrates that some particles sized about 4-8 nm are homogeneously dispersed among the amorphous matrix, as shown in Fig. 1(c). The corresponding selected area electron diffraction (SAED) pattern in the inset further confirms the crystalline nature of CeO2 particles.

 figure: Fig. 1

Fig. 1 XRD patterns of the NC10 sample annealed at different temperatures (a) and NC samples with different CeO2 contents annealed at 1000 °C (b). (c) HRTEM image of the NC10 sample annealed at 1000 °C; inset shows corresponding SAED pattern.

Download Full Size | PDF

As is well known, one of the most serious restrictions to RE doped sol-gel silica is the residual OH- which may lead to the quenching of the RE emission [15]. IR spectra of the NC10 sample annealed at different temperatures and DG samples with different CeO2 contents are shown in Fig. 2. The bands at 470 cm−1 and 550 cm−1 can be attributed to the Si-O vibration [15,16] and the peaks at 800 cm−1 and 1080 cm−1 are corresponding to the Si-O-Si bond [15,18]. The intensity of these peaks increases with annealing temperature, indicating the increasing number of [SiO4]4− tetrahedral units and Si-O-Si bonds in the NC sample [17,18]. The small peak at 1380 cm−1 arises from the vibrational mode of NO3- units [19]. As shown in Fig. 2(b), the peak intensifies with increasing the CeO2 content due to the introduction of Ce4+ by the form of Ce(NO3)3. After annealed at 600 °C, the peak disappears due to the decomposition of the group (in Fig. 2(a)). The peak at 1635 cm−1 can be ascribed to the H-O-H bending vibration of the free water or the physically absorbed water [18]. Despite the intensity decreasing with the increase of annealing temperature, the band is still observed, possible owing to the existence of residual free water resided in the micro-pores of the sample. Broad peaks centered around 3465 cm−1 and 954cm−1 originate from the -OH symmetric stretching from the surface hydroxyl group [15]. As a result of surface dehydroxylation during the annealing process, the intensity of these peaks weakens, and finally disappears at 1000 °C, which is beneficial to the Er3+ emission.

 figure: Fig. 2

Fig. 2 IR spectra of the NC10 sample annealed at different temperatures (a) and DG samples with different CeO2 contents (b).

Download Full Size | PDF

In order to explore the energy transfer between CeO2 nanocrystals and Er3+ ions in the obtained nanocomposite, the PL excitation (PLE) spectra of the NC10 sample annealed at different temperatures monitored at 1532 nm, corresponding to the Er3+ 4I13/2-4I15/2 transition, is shown in Fig. 3(a). For comparison, the PLE spectrum of the sample without CeO2 (NC0) annealed at 1000 °C is also displayed in the figure. Three excitation peaks at 490 nm, 522 nm and 654 nm are ascribed to the transitions from 4I15/2 ground state to 4F7/2, 2H11/2, and 4F9/2 excited states of Er3+ ions, respectively. Interestingly, compared with the NC0 sample, for the NC10 one an additional strong excitation band centered at 346 nm was observed, which can be attributed to the charge transfer from O2- valence band to Ce4+ conduction band [14,20], indicating the presence of efficient energy transfer from CeO2 nanocrystals to Er3+ ions. Furthermore, the excitation bands intensify with increasing the annealing temperature because of the precipitation of the CeO2 nanocryscals and the decrease of residual OH- in the silica matrix.

 figure: Fig. 3

Fig. 3 (a) PLE spectra of NC10 samples annealed at different temperatures. (b) Normalized visible and near-infrared PL spectra of the NC10 sample annealed at 1000 °C, excited at 346 nm and 520 nm respectively. The inset shows the normalized emitted photons ratio at the 4I13/24I15/2 transition for two excitation wavelength. (c) PL spectra of the NC10 sample annealed at 1000 °C excited at different wavelength. For comparison, excitation and emission spectra of the NC0 sample annealed at 1000 °C are also displayed in (a) and (b).

Download Full Size | PDF

Normally, the energy transfer efficiency between semiconductor and RE ions is weak owing to the limit solubility of RE in the host and consequently weak interaction between them [12]. In the obtained nanocomposite, the incorporation of most Er3+ ions into CeO2 nanocrystals is expected due to the similar radius between the dopant and the host Ce4+ ions, which may induce the efficient energy transfer from nanocrystals to dopants. To reveal the distribution of Er3+ in the nanocomposite, visible and near-infrared (NIR) PL spectra of the NC10 sample annealed at 1000 °C upon indirect excitation at 346 nm and direct excitation at 520 nm were measured. As shown in Fig. 3(b), under direct excitation at 520 nm (in this case all Er3+ ions might be excited), the PL spectrum shows well-resolved peaks, indicating that most of Er3+ ions are incorporated into CeO2 nanocrystals by substituting Ce4+ ions due to the similar radius of the Er3+ ions (0.089 nm) and the host Ce4+ ions (0.087 nm). Under indirect excitation at 346 nm, the line-shape of the spectrum is different from that excited at 520 nm because only the Er3+ ions located in/near nano-crystals could benefit from the energy transfer [20]. In comparison, the PL spectrum of the NC0 sample annealed at 1000 °C shows a typical Er3+ emission in silica. Under the identical measured conditions and corrected for the excitation intensity at two excitation wavelength, the efficiency of the two excitation schemes is displayed in the inset of Fig. 3 (b) [21]. Obviously, the efficiency of the indirect excitation is enhanced with respect to the direct one. Moreover, PL spectra of the NC10 sample excited in the range from 270 nm to 510 nm are shown in Fig. 3(c). Excited within the region of the CeO2 band-to-band transition, spectra show structured peaks, further implying the efficient energy transfer from CeO2 to Er3+ ions located in/near nanocrytals.

To evaluate the influence of the annealing temperature on the PL properties, PL spectra and decay curves of the NC10 sample annealed at different temperatures are measured and exhibited in Fig. 4, under the excitation of 346 nm. The emission intensifies with increasing annealing temperature. The mean decay lifetimes (τ) of the emission at 1532 nm can be evaluated by the equation:

τ=I(t)/I0dt
where I(t) is the PL intensity at time t and I0 the maximal PL intensity at time 0. The mean PL lifetime exhibits a monotonic increase from 0.30 ms to 0.62 ms, when the annealing temperature arises from 160 °C (the DG sample) to 1000 °C. The enhancement of the emission intensity and the lengthening of the PL lifetime are induced by the elimination of residual OH- in silica matrix, taking into account the results of FTIR spectra. Especially, the intensity of the sample annealed at 1000 °C far exceeds that of the one annealed at 800 °C possible due to the complete disappearance of the surface hydroxyl group. Moreover, more and more Er3+ ions are incorporated into CeO2 nanocrystals with the increase of annealing temperature, also resulting in the enhanced Er3+ emission [20].

 figure: Fig. 4

Fig. 4 (a) PL spectra and (b) decay curves of the NC10 sample annealed at different temperatures excited at 346 nm.

Download Full Size | PDF

Figure 5 presents PL spectra and luminescence decay curves of NC samples with different CeO2 contents annealed at 1000 °C. The characteristic NIR emission of Er3+ is hardly detected in the NC0 sample (without CeO2 content). The emission intensity increases monotonously with the CeO2 content. Impressively, the intensity of the 1532 nm emission in the sample with 20 mol % CeO2 is about 12 times as high as that in the sample with 5 mol% CeO2. The mean PL lifetime evaluated from the Eq. (1) increases monotonically from 0.20 ms to 0.99 ms, when the CeO2 content increases from 0 to 20 mol%. As the CeO2 content adds, more Er3+ ions can be partitioned into the nanocrystals which defend Er3+ emission from the residual OH- quenching, provide a low phonon energy environment to reduce the nonradiative recombination, and sensitize the Er3+ emission [20].

 figure: Fig. 5

Fig. 5 (a) PL spectra and (b) decay curves of the samples with different CeO2 contents excited at 346 nm. Inset in (a) shows the dependence of 1532 nm emission intensity on the CeO2 content. The decay curve for the NC0 sample is excited at 520 nm.

Download Full Size | PDF

4. Conclusions

In summary, Er3+-doped CeO2-SiO2 composite was successfully prepared by the sol-gel process. The influence of the annealing temperature and CeO2 content on the photoluminescence properties was systematically investigated. As the annealing temperature and the CeO2 content increase, the amount of Er3+ ions locating in the CeO2 lattices increases, resulting in the remarkable intensification of the Er3+ emission, which can be attributed to the effective energy transfer process due to the formation of CeO2 nanocrystals as well as the elimination of residual OH- in the silica matrix. It demonstrates that the CeO2 may be one of the most effective sensitizer to improve the photoluminescence behaviors.

Funding

National Natural Science Foundation of China (51202244, 51503036); Natural Science Foundation of Fujian Province (2015J01632, 2016J05109); Scientific Project of Fujian Education Department (JA15534, JA15536, JZ160484).

References and links

1. M. Garai and B. Karmakar, “Rare earth ion controlled crystallization of mica glass-ceramics,” J. Alloys Compd. 678(9), 360–369 (2016). [CrossRef]  

2. C. Lv, C. Zhu, C. Wang, Y. Gao, X. Ma, and D. Yang, “Electroluminescence from metal-oxide-semiconductor devices with erbium-doped CeO2 films on silicon,” Appl. Phys. Lett. 106(14), 141102 (2015). [CrossRef]  

3. J. Malleshappa, H. Nagabhushana, S. C. Sharma, D. V. Sunitha, N. Dhananjaya, C. Shivakumara, and B. M. Nagabhushana, “Self propagating combustion synthesis and luminescent properties of nanocrystalline CeO2:Tb3+ (1-10 mol%) phosphors,” J. Alloys Compd. 590(590), 131–139 (2014). [CrossRef]  

4. X. Zhao, S. He, and M. Tan, “Design of infrared-emitting rare earth doped nanoparticles and nanostructured composites,” J. Mater. Chem. C Mater. Opt. Electron. Devices 36(36), 8349–8372 (2016). [CrossRef]  

5. T. Ninjbadgar, G. Garnweitner, A. Börger, L. M. Goldenberg, O. V. Sakhno, and J. Stumpe, “Synthesis of luminescent ZrO2:Eu3+ nanoparticles and their holographic sub-micrometer patterning in polymer composites,” Adv. Funct. Mater. 19(11), 1819–1825 (2009). [CrossRef]  

6. S. Fujiharaa and M. Oikawa, “Structure and luminescent properties of CeO2: rare earth (RE=Eu3+ and Sm3+) thin films,” J. Appl. Phys. 95(12), 8002–8006 (2004). [CrossRef]  

7. D. Avram, I. Porosnicu, B. Cojocaru, M. Florea, and C. Tiseanu, “Time-gated down-/up-conversion emission of Ho-CeO2 and Ho, Yb-CeO2 nanoparticles,” J. Lumin. 179(11), 265–271 (2016). [CrossRef]  

8. J. Malleshappa, H. Nagabhushana, D. Kavyashree, S. C. Prashantha, S. C. Sharma, H. B. Premkumar, and C. Shivakumara, “Shape tailored green synthesis of CeO2:Ho3+ nanopowders, its structural, photoluminescence and gamma radiation sensing properties,” Spectrochim. Acta A Mol. Biomol. Spectrosc. 145, 63–75 (2015). [CrossRef]   [PubMed]  

9. H. Zhang, J. Han, and B. Yang, “Structural fabrication and functional modulation of nanoparticle-polymer composites,” Adv. Funct. Mater. 20(10), 1533 (2010). [CrossRef]  

10. X. Zhang, S. Lin, T. Lin, P. Zhang, J. Xu, L. Xu, and K. Chen, “Improved sensitization efficiency in Er3+ ions and SnO2 nanocrystals co-doped silica thin films,” Phys. Chem. Chem. Phys. 17(18), 11974–11980 (2015). [CrossRef]   [PubMed]  

11. S. Manna, R. Aluguri, R. Bar, S. Das, N. Prtljaga, L. Pavesi, and S. K. Ray, “Enhancement of photoluminescence intensity of erbium doped silica containing Ge nanocrystals: distance dependent interactions,” Nanotechnology 26(4), 045202 (2015). [CrossRef]   [PubMed]  

12. F. Xiao, R. Chen, Y. Shen, Z. Dong, H. Wang, Q. Zhang, and H. Sun, “Efficient energy transfer and enhanced infrared emission in Er-doped ZnO-SiO2 composites,” J. Phys. Chem. C 116(24), 13458–13462 (2012). [CrossRef]  

13. T. Lin, X. Zhang, J. Xu, X. Liu, M. T. Swihart, L. Xu, and K. Chen, “Strong energy-transfer-induced enhancement of Er3+ luminescence in In2O3 nanocrystal codoped silica films,” Appl. Phys. Lett. 103(18), 181906 (2013). [CrossRef]  

14. L. Li, H. Yang, B. K. Moon, Z. Fu, C. Guo, J. H. Jeong, S. S. Yi, K. Jang, and H. S. Lee, “Photoluminescence properties of CeO2: Eu3+ nanoparticles synthesized by a Sol-Gel method,” J. Phys. Chem. C 113(2), 610–617 (2009). [CrossRef]  

15. T. Lin, X. Ding, J. Xu, N. Wan, L. Xu, and K. Chen, “Influences of doping and annealing conditions on the photoluminescence from In2O3 nanocrystals and Eu3+ ions co-doped sol-gel SiO2 films,” J. Appl. Phys. 109(8), 083512 (2011). [CrossRef]  

16. P. J. Hyun and D. J. Min, “H, S, Song, “FT-IR spectroscopic study on structure of CaO-SiO2 and CaO-SiO2CaF2 slags,” Transactions of the Iron & Steel Institute of Japan 42(4), 344–351 (2002). [CrossRef]  

17. I. Coroiu, E. Culea, and A. Darabont, “Magnetic and structural behaviour of the sol–gel-derived iron aluminosilicate glass-ceramics,” J. Magn. Magn. Mater. 290-291, 997–1000 (2005). [CrossRef]  

18. Y. Yu, D. Chen, Y. Wang, W. Luo, Y. Zheng, Y. Cheng, and L. Zhou, “Structural evolution and its influence on luminescence of SiO2-SrF2-ErF3 glass ceramics prepared by sol–gel method,” Mater. Chem. Phys. 100(2-3), 241–245 (2006). [CrossRef]  

19. J. T. Kloprogge, D. Wharton, L. Hickey, and R. L. Frost, “Infrared and raman study of interlayer anions CO3-2 NO3- SO42- CIO4-, ” Am. Mineral. 87(5–6), 623–629 (2002). [CrossRef]  

20. Y. Yu, D. Chen, P. Huang, H. Lin, A. Yang, and Y. Wang, “Sensitization and protection of Eu3+ luminescence by CeO2 in nano-composite,” J. Alloys Compd. 513(4), 626–629 (2012). [CrossRef]  

21. S. Brovelli, A. Chiodini, A. Lauria, F, Meinardi, and A, Paleari, “Energy transfer to erbium ions from wide-band-gap SnO2 nanocrystals in silica,” Physical Review B 73 (73), 07340 (2006).

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 XRD patterns of the NC10 sample annealed at different temperatures (a) and NC samples with different CeO2 contents annealed at 1000 °C (b). (c) HRTEM image of the NC10 sample annealed at 1000 °C; inset shows corresponding SAED pattern.
Fig. 2
Fig. 2 IR spectra of the NC10 sample annealed at different temperatures (a) and DG samples with different CeO2 contents (b).
Fig. 3
Fig. 3 (a) PLE spectra of NC10 samples annealed at different temperatures. (b) Normalized visible and near-infrared PL spectra of the NC10 sample annealed at 1000 °C, excited at 346 nm and 520 nm respectively. The inset shows the normalized emitted photons ratio at the 4I13/24I15/2 transition for two excitation wavelength. (c) PL spectra of the NC10 sample annealed at 1000 °C excited at different wavelength. For comparison, excitation and emission spectra of the NC0 sample annealed at 1000 °C are also displayed in (a) and (b).
Fig. 4
Fig. 4 (a) PL spectra and (b) decay curves of the NC10 sample annealed at different temperatures excited at 346 nm.
Fig. 5
Fig. 5 (a) PL spectra and (b) decay curves of the samples with different CeO2 contents excited at 346 nm. Inset in (a) shows the dependence of 1532 nm emission intensity on the CeO2 content. The decay curve for the NC0 sample is excited at 520 nm.

Equations (1)

Equations on this page are rendered with MathJax. Learn more.

τ= I(t) / I 0 dt
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.