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

Enhancement of 1.53 μm emission band in NaYF4:Er3+,Yb3+,Ce3+ nanocrystals for polymer-based optical waveguide amplifiers

Open Access Open Access

Abstract

NaYF4:Er3+,Yb3+,Ce3+ nanocrystals (NCs) were synthesized by using a solvothermal approach. Under the excitation of a 980 nm laser, the 1.53 μm emission band of Er3+ ions in the NCs was enhanced by 6 times after codoping Ce3+ ions owing to the efficient energy transfer between Ce3+ and Er3+: 4I11/2 (Er3 +) + 2F5/2 (Ce3+) → 4I13/2 (Er3+) + 2F7/2 (Ce3+). By dispersing the NaYF4:Er3+,Yb3+,Ce3+ NCs into SU-8 2005 polymer matrix, we constructed Er3+-doped polymer-based optical waveguide amplifiers (EDPOWAs) and measured their performances. The measured optical gain of the EDPOWA doped with NaYF4: Er3+,Yb3+,Ce3+ NCs was enhanced by ~2dB in comparison with that doped with NaYF4:Er3+,Yb3+ NCs. These results showed that NaYF4:Er3+,Yb3+,Ce3+ NCs are promising candidates for building high gain EDPOWAs.

©2013 Optical Society of America

1. Introduction

Erbium (Er3+)-doped fiber amplifiers (EDFAs) have been widely employed to overcome losses in long haul silica fiber transmission systems [1]. This is due to the 4I13/24I15/2 transition (~1.53 µm) of Er3+ ions, which matches one of the low loss windows of optical fibers in optical communication networks [24]. However, for access and home network applications, EDFAs are incompatible with miniature and integrated optical devices. Compared to EDFAs, Er3+-doped waveguide amplifiers (EDWAs) could afford high gain in a much smaller device size and be integrated with photonic devices based on silicon substrate [511]. In terms of waveguide materials and fabrication processes, polymers have attracted a great deal of attention because they exhibit many advantages over inorganic glasses or crystals, such as easy processing, permitting fabrication of devices with virtually any shape, and potential low cost [6,7]. Especially, Er3+-doped inorganic nanocrystals (NCs) can be dispersed into polymer matrices and used to construct Er3+-doped polymer-based optical waveguide amplifiers (EDPOWAs) [811].

To obtain high gain EDPOWAs, Er3+-doped inorganic NCs should not only have highly efficient 1.53 µm emission but also good dispersibility in polymer matrices. To date, researchers have developed many approaches to obtain Er3+-doped inorganic NCs with highly efficient 1.53 µm emission and good dispersibility. For example, van Veggel et al. used ammonium di-n-octadecyldithiophosphate as a surface modification agent to obtain 1.53 µm emission of redispersible LaF3:Er3+ and LaF3:Er3+,Yb3+ NCs [8,9]. The dispersibility of NCs depends on the nature of ligands. Zhen et al. reported oleic acid-modified LaF3:Er3+,Yb3+ NCs with excellent dispersibility in common organic solvents and a PMMA matrix, and obtained an optical signal enhancement of 1.77 dB cm−1 from the polymer waveguide [1,10]. Roy et al. synthesized oleic acid-coated NaYF4:Er3+,Yb3+ NCs and constructed EDPOWAs with a optical gain of 4.81 dB cm−1 [11]. It is known that NaYF4 is one of the most suitable nanocrystalline hosts for Er3+ ion doping for photoluminescent applications. In general, when Er3+ ion is excited by 980 nm photon, a considerable amount of electrons are populated at the 4I11/2 level, further promoted to higher excited states, and the transition from the higher excited state to the lower state gives upconversion (UC) emissions. Upconversion (UC) NCs, one kind of Ln3+-doped NCs, can convert low energy radiation to high energy radiation and have been found numerous applications in biological imaging, labeling, therapeutics, and solar cells [1220]. However, the upconversion processes lead a dramatic decrease of the population of 4I13/2 and reduce the emission intensity around 1.53 µm. Previous studies on fluoride glasses clearly indicated that the emission intensity around 1.53 µm could be improved by suppressing UC processes [2125]. This can be realized by doping Ce3+ ions into Er3+-doped inorganic NCs because of the existence of the phonon-assisted energy transfer (ET) between Er3+ and Ce3+, 4I11/24I13/2 (Er3+): 2F5/22F7/2 (Ce3+). To the best of our knowledge, Ce3+ and Er3+ codoped NCs have not yet been investigated, and the effects of Ce3+ ion addition on the emission intensity around 1.53 µm of Er3+ in NCs have not been clarified.

In this work, we describe the synthesis of NaYF4:Er3+,Yb3+,Ce3+ NCs coated with oleic acid, which are readily dispersible in organic solvent [26]. With the addition of a suitable amount of Ce3+ ions, the UV and visible UC emissions of Er3+ in the NCs have been reduced greatly, and the NIR emission around 1.53 µm has dramatically enhanced upon the excitation with a 980 nm laser diode. Furthermore, the influence of Ce3+ ions on the emissions of Er3+ in the NaYF4 NCs is discussed in detail here. In addition, we constructed EDPOWAs by using NaYF4:Er3+,Yb3+,Ce3+ NCs as the gain medium. The measured optical gain of the EDPOWA doped with NaYF4:Er3+,Yb3+,Ce3+ NCs was enhanced by ~2dB in comparison with that doped with NaYF4:Er3+,Yb3+ NCs.

2. Results and discussions

The characterizations of double-doped NaYF4:2%Er3+,20%Yb3+ are summarized in Fig. 1 . From the TEM image (Fig. 1A), it is easily seen that the monodisperse NCs has a spherical shape. The samples codoped with Ce3+ ions at different concentrations possess a similar morphology to the double-doped ones. From the corresponding HRTEM image (Fig. 1B), the lattice fringes were indicative of the high crystallinity of these particles, and the measured lattice spacing is 0.275 nm, which corresponds to the (200) planes of α-NaYF4 NCs. The fast Fourier transformation (FFT) pattern (Fig. 1C) reveals that the sample is cubic phase. A histogram of the particle size distribution from TEM images is given in Fig. 1D. The particle sizes range from 14 to 27 nm, the average size is determined to be approximately 20 nm. Moreover, the XRD data is shown in Fig. 1E, which reveals that the sample is pure cubic phase, all the strong peaks are consistent with the calculated pattern for cubic NaYF4 crystal (JCPDS files No. 77-2042), and no other impurity peaks can be detected from the XRD patterns. In addition, FTIR spectrum of the as-prepared NCs was recorded, as shown in Fig. 2 . The band at around 3450 cm−1 can be assigned as O–H stretching. The strong absorption peaks at 2850 and 2917 cm−1 are attributable to the symmetric and asymmetric C–H stretching of the oleic acid coating, respectively. In addition, bands at 1553 and 1462 cm−1 can be assigned to be the asymmetric and symmetric stretching of the carboxylate group (COO) of the oleic acid coating, respectively. The results indicate that the NCs have been coated with oleic acid.

 figure: Fig. 1

Fig. 1 Characterization data for double-doped NaYF4:Er3+,Yb3+ NCs. (A) TEM image; (B) HRTEM image; (C) FFT pattern of a single nanocrystal; (D) histogram of the particle sizes obtained from TEM images of 400 NCs; (E) XRD pattern.

Download Full Size | PDF

 figure: Fig. 2

Fig. 2 The FTIR spectrum of the oleic acid-coated NaYF4:Er3+,Yb3+,Ce3+ NCs.

Download Full Size | PDF

With the excitation of a 980 nm laser diode, the UC emission spectra of the triple-doped NaYF4:Er3+,Yb3+,Ce3+ NCs with different Ce3+ concentrations (0 mol%, 2 mol%, 4 mol%, and 10 mol%) were recorded at room temperature, as shown in Fig. 3A . The Yb3+ ions act as sensitizers due to their high absorption coefficient at 980 nm and transfer the absorbed energy to neighboring Er3+ ions. The spectrum of the double-doped NaYF4:Er3+,Yb3+ NCs exhibits several intense UC emission peaks, which are attributed to the 4G11/24I15/2 (~378 nm), 2H9/24I15/2 (~411 nm), 2H11/24I15/2 (~525 nm), 4S3/24I15/2 (~544 nm), and 4F9/24I15/2 (~658 nm) transitions of Er3+ ions, respectively. As shown in Fig. 3A, the UC luminescence of the double-doped NaYF4 sample (without Ce3+ ions) is the strongest one among the four samples. Upon doping small concentration of Ce3+ ions (2%) into the double-doped NaYF4 NCs, the intensities of UV and visible UC emission were all drastically quenched. It is obvious that with the increase of Ce3+ concentration from 2% to 10%, the UC emission intensities of Er3+ are gradually reduced, and when the concentration of Ce3+ reaches 10%, the UC emissions of Er3+ are almost extinguished entirely.

 figure: Fig. 3

Fig. 3 (A) UC emission spectra of the NaYF4:Er3+,Yb3+,Ce3+ NCs with 0%, 2%, 4%, and 10% Ce3+. (B) Plot of relative UC emission intensities vs. codoped Ce3 + ion concentrations. (C) DC emission spectra of the NaYF4:Er3+,Yb3+,Ce3+ NCs with 0%, 2%, 4%, and 10% Ce3 + . (D) Energy level diagrams of Yb3+, Er3+, and Ce3+ ions, and possible processes of populations and emissions.

Download Full Size | PDF

In addition, the dependence of the UC emission intensity on the doping concentration of Ce3+ ions in the NaYF4 sample was investigated, as shown in Fig. 3B. It is clear that the UC emission intensities decrease obviously with adding 2% Ce3+ ions into the double-doped NaYF4 NCs, and the least affected one (~544 nm) has been reduced to ~50% compared to that of the sample without Ce3+ ions, indicating that the energy transfer from Er3+ to Ce3+ ions is highly efficient here since several emission bands of Er3+ ions match well with the f-f absorptions of Ce3+ ions. It obviously exhibits that the quenching of red emission (~658 nm) is stronger than those of the green emissions (~525 nm and ~544 nm) by adding Ce3+ ions and increasing their concentration. When the Ce3+ concentration is 2%, the emission intensities of the green emissions peaked at 525 and 544 nm reduce by 63% and 52%, respectively, while the intensity of the red emission peaked at 658 nm decreases by 78%, compared to those of the sample without Ce3+ ions. The above results show that the UC emissions of Er3+ ions have been suppressed efficiently by codoping Ce3+ ions.

To investigate the effects of codoped Ce3+ ions on the 1.53 µm emission of Er3+ ions, we recorded the infrared emission spectra of the triple-doped NaYF4 NCs with different Ce3+ concentrations (0 mol%, 2 mol%, 4 mol%, and 10 mol%) under 980 nm excitation, as shown in Fig. 3C. The emission of 1535 nm is assigned to the 4I13/24I15/2 transition of Er3+, and the addition of Ce3+ ions influenced the emission intensity of 1535 nm band evidently. By adding 2% Ce3+ ions into the NaYF4:Er3+,Yb3+ NCs, the integrated emission intensity of 1535 nm is about 6 times stronger than that of the sample without Ce3+ ions. However, when a higher concentration (~10%) of Ce3+ ions was used, the intensity of 1535 nm emission decreased relatively, even lower than that of the sample without Ce3+. It could be attribute to the occurrence of the phonon-assisted ET from Er3+ to Ce3+, 4I13/24I15/2: 2F5/22F7/2 when the concentration of Ce3+ ions is too high [27,28]. It is obvious that the near-infrared emission around 1535 nm of Er3+ ions has been drastically enhanced by codoping a certain amount (~2%) of Ce3+ ions into the NaYF4:Er3+,Yb3+ NCs, which means that the performance of EDPOWAs could be improved remarkably by this approach.

The possible UC and DC mechanisms for the triple-doped NaYF4 NCs are discussed on the basis of the energy level diagram presented in Fig. 3D. Under 980 nm excitation, the Yb3+ ions are excited from the 2F7/2 level to the 2F5/2 level and then transfer their energies to the nearby Er3+ ions. The upper levels 4G11/2, 4F7/2, and 4F9/2 of Er3+ ions are mainly populated by the following processes: ETs between Yb3+ and Er3+ (2F5/22F7/2: 4I15/24I11/2, 4I13/24F9/2, and 4I11/24F7/2), the excited state absorption of the pump radiation from the 4I11/2 and 4I13/2 levels, and the cross relaxation (CR) between Er3+ ions, as depicted in Fig. 3D. Additionally, it can be seen from Fig. 3D that the UC emissions are all related to the population of the 4I11/2 level of Er3+ ions. With the addition of Ce3+ ions, the population of the 4I11/2 level decreases due to the phonon-assisted ET from Er3+ to Ce3+, 4I11/24I13/2: 2F5/22F7/2; on the other hand, the ETs from Er3+ to Ce3+ (4S3/24F9/2, 4F9/24I9/2, 4I9/24I11/2, and 4I11/24I13/2: 2F5/22F7/2) make the 4S3/2, 4F9/2, 4I9/2, and 4I11/2 states depopulate to the 4I13/2 state. As a result, the emission intensity around 1.53 µm is enhanced and the UC processes are suppressed by the addition of Ce3+ ions. In addition, since the energy mismatch between 4F9/24I9/2 (Er3+) and 2F5/22F7/2 (Ce3+) is smaller than that between 4S3/24F9/2 (Er3+) and 2F5/22F7/2 (Ce3+), we deduce that the ET (4F9/24I9/2 (Er3+): 2F5/22F7/2 (Ce3+)) is more effective, which causes the red emission from the 4F9/2 level quenched heavier than that of green emissions from the 2H11/2 and 4S3/2 levels, shown in Fig. 3B.

To further investigate the effects of codoped Ce3+ ions on the UC and near-infrared emissions of Er3+, we measured the lifetimes of Er3+ ions in the triple-doped NaYF4 NCs with different Ce3+ concentrations, as shown in Fig. 4 . Each decay curve can be well fitted by using a single exponential function I(t) = I0exp(−t/τ), where I0 is an intensity parameter for t = 0, and τ is the excited state lifetime. Obviously, with increasing the concentration of Ce3+ ions, the lifetimes of 4S3/2, and 4F9/2 states of Er3+ ions in NaYF4 NCs decrease gradually owing to the occurrence of the ETs of, 4S3/24F9/2 (Er3+): 2F5/22F7/2 (Ce3+) and 4F9/24I9/2 (Er3+): 2F5/22F7/2 (Ce3+). Additionally, the lifetime reduction of the 4F9/2 level is more than that of the 4S3/2 level, indicating that the ET 4F9/24I9/2 (Er3+): 2F5/22F7/2 (Ce3+) is more effective than the ET 4S3/24F9/2 (Er3+): 2F5/22F7/2 (Ce3+), which is in good agreement with the measured results of the UC emission intensities.

 figure: Fig. 4

Fig. 4 Emission decay curves of Er3+ in the NaYF4: Er3+,Yb3+,Ce3+ NCs (excited at 980 nm, monitored at 544 nm and 658 nm corresponding to the 4S3/24I15/2 and 4F9/24I15/2 transitions, respectively).

Download Full Size | PDF

By dispersing NaYF4:Er3+,Yb3+,Ce3+ NCs into SU-8 2005 polymer matrix, we constructed EDPOWA and measured their performances. The SU-8 2005 polymer was diluted in toluene. NaYF4 NCs (1 wt %) were added into the above solution, and it was dissolved at room temperature by ultrasonic treatment for 20 min. Waveguides of dimensions, 4 μm height by 8 μm width were fabricated by standard photolithography and wet etching technology of a thin silicon dioxide layer based on a silicon substrate. The SU-8 2005 polymer dispersed with NCs was spin-coated on the silicon dioxide layer and pre-annealed at 90 °C for 20 min. Then, the waveguide channels were cured by the photo mask using UV light at a 365 nm for 8 s and then baked at 95 °C for 10 min. Finally, a thin PMMA-GMA was used as the top cladding. The refractive indices of the materials were measured using ellipsometry method (J. A. Woollam., Co. M2000). The refractive indices of the core and cladding were 1.578 and 1.495 at 1535 nm wavelength, respectively. Figure 5A is a SEM image of the NaYF4 NCs dispersed SU-8 2005 polymer rectangular waveguide amplifier (without top cladding), indicating that the size of the waveguide amplifier is 4 μm high and 8 μm wide.

 figure: Fig. 5

Fig. 5 (A) Experimental setup for measuring the optical gain of the waveguide amplifier; (B) SEM image of the NaYF4:Er3+,Yb3+,Ce3+ NCs-doped polymer waveguide amplifier; (C) Relative gain as a function of pump power (980 nm) with 0.1 mW input signal powers (1535 nm) in an 4 μm high, 8 μm wide and 1.3 cm long NaYF4:Er3+,Yb3+,Ce3+ and NaYF4:Er3+,Yb3+ NCs-dispersed polymer waveguide.

Download Full Size | PDF

Figure 5B shows the schematics of the experimental setup for the optical gain measurement. Gain measurement of the NaYF4:Er3+,Yb3+,Ce3+ and NaYF4:Er3+,Yb3+ NCs dispersed polymer waveguide amplifier was carried out as the signal light source in the wavelength range of 1510-1590 nm (Santec TSL-210) as the signal source and a 976 nm laser diode as the pump source. Both the pump and signal sources were coupled into a single output optical fiber by a 980/1535 nm wavelength division multiplexer and together butt coupled to a single output fiber. The relative gain was determined from the ratio of the out-put signal observed on the optical spectrum analyzer (OSA, Ando AQ-6315A) when both the pump and signal beams are coupled to the polymer waveguide to the signal power without the pump.

Figure 5C shows the measured relative gain as a function of pump power in a 1.3 cm long waveguide. The relative gain gradually increases with the increase of pump power. When the pump power is 210 mW and the input signal power is 0.1 mW, the maximum gain of about 4.0 dB (3.08 dB cm−1) at 1535 nm was obtained. In addition, the optical gain of the EDPOWA doped with NaYF4:Er3+,Yb3+,Ce3+ NCs was enhanced by ~2dB in comparison with that doped with NaYF4:Er3+,Yb3+ NCs. This result shows that the improved 1.53 µm emission is very important for the EDPOWA.

3. Conclusion

In summary, we synthesized NaYF4:Er3+,Yb3+,Ce3+ NCs via a solvothermal method, and the NCs could be easily dispersed in organic solvents and polymer matrices. The UC and DC emission spectra of the NCs with different concentration of Ce3+ ions were investigated in detail. The results showed that the addition of a certain amount of Ce3+ ions enhanced the 1.53 µm emission of Er3+ by ~6 times via the phonon-assisted ET (4I11/24I13/2 (Er3+): 2F5/22F7/2 (Ce3+)) while the UC emissions of Er3+ were quenched effectively. Using NaYF4:Er3+,Yb3+,Ce3+NCs doped SU-8 2005 polymer as gain medium, we constructed an EDPOWA and measured its performance. A relative optical gain of the EDPOWA doped with NaYF4:Er3+,Yb3+,Ce3+ NCs was enhanced by ~2dB in comparison with that doped with NaYF4:Er3+,Yb3+ NCs in 1.3 cm long waveguide. This work would provide a new strategy to improve the performance of EDPOWAs effectively.

Acknowledgments

This work was supported by the NSFC (grants 51072065, 61178073, 61177027, 61077041, 60908031, 60908001, and 61077033), the Program for NCET in University (No: NCET-08-0243), the Opened Fund of the State Key Laboratory on Integrated Optoelectronics, and Tsinghua National Laboratory for Information Science and Technology (TNList) Cross-discipline Foundation.

References and links

1. S. Bo, J. Wang, H. Zhao, H. Ren, Q. Wang, G. Xu, X. Zhang, X. Liu, and Z. Zhen, “LaF3: Er,Yb doped sol–gel polymeric optical waveguide amplifiers,” Appl. Phys. B 91(1), 79–83 (2008). [CrossRef]  

2. W. H. Wong, E. Y. B. Pun, and K. S. Chan, “Er3+–Yb3+ codoped polymeric optical waveguide amplifiers,” Appl. Phys. Lett. 84(2), 176–178 (2004). [CrossRef]  

3. J. A. Dorman, J. H. Choi, G. Kuzmanich, and J. P. Chang, “Elucidating the effects of a rare-earth oxide shell on the luminescence dynamics of Er3+:Y2O3 nanoparticles,” J. Phys. Chem. C 116(18), 10333–10340 (2012). [CrossRef]  

4. Y. B. Mao, J. Y. Huang, R. Ostroumov, K. L. Wang, and J. P. Chang, “Synthesis and Luminescence properties of erbium-doped Y2O3 nanotubes,” J. Phys. Chem. C 112(7), 2278–2285 (2008). [CrossRef]  

5. S. H. Bo, J. Hu, Z. Chen, Q. Wang, G. M. Xu, X. H. Liu, and Z. Zhen, “Core-shell LaF3:Er,Yb nanocrystal doped sol–gel materials as waveguide amplifiers,” Appl. Phys. B 97(3), 665–669 (2009). [CrossRef]  

6. C. Chen, D. Zhang, T. Li, D. Zhang, L. Song, and Z. Zhen, “Erbium-ytterbium codoped waveguide amplifier fabricated with solution-processable complex,” Appl. Phys. Lett. 94(4), 041119 (2009). [CrossRef]  

7. D. Zhang, C. Chen, C. Chen, C. Ma, D. Zhang, S. Bo, and Z. Zhen, “Optical gain at 1535 nm in LaF3:Er,Yb nanoparticle-doped organic-inorganic hybrid material waveguide,” Appl. Phys. Lett. 91(16), 161109 (2007). [CrossRef]  

8. J. W. Stouwdam, G. A. Hebbink, J. Huskens, and F. van Veggel, “Lanthanide-doped nanoparticles with excellent luminescent properties in organic media,” Chem. Mater. 15(24), 4604–4616 (2003). [CrossRef]  

9. J. W. Stouwdam and F. C. van Veggel, “Improvement in the luminescence properties and processability of LaF3/Ln and LaPO4/Ln nanoparticles by surface modification,” Langmuir 20(26), 11763–11771 (2004). [CrossRef]   [PubMed]  

10. J. Wang, J. Hu, D. Tang, X. Liu, and Z. Zhen, “Oleic acid (OA)-modified LaF3:Er,Yb nanocrystals and their polymer hybrid materials for potential optical-amplification applications,” J. Mater. Chem. 17(16), 1597–1601 (2007). [CrossRef]  

11. K. L. Lei, C. F. Chow, K. C. Tsang, E. N. Y. Lei, V. A. L. Roy, M. H. W. Lam, C. S. Lee, E. Y. B. Pun, and J. S. Li, “Long aliphatic chain coated rare-earth nanocrystal as polymer-based optical waveguide amplifiers,” J. Mater. Chem. 20(35), 7526–7529 (2010). [CrossRef]  

12. F. Wang, Y. Han, C. S. Lim, Y. H. Lu, J. Wang, J. Xu, H. Y. Chen, C. Zhang, M. H. Hong, and X. G. Liu, “Simultaneous phase and size control of upconversion nanocrystals through lanthanide doping,” Nature 463(7284), 1061–1065 (2010). [CrossRef]   [PubMed]  

13. J. C. Boyer, F. Vetrone, L. A. Cuccia, and J. A. Capobianco, “Synthesis of colloidal upconverting NaYF4 nanocrystals doped with Er3+, Yb3+ and Tm3+, Yb3+ via thermal decomposition of lanthanide trifluoroacetate precursors,” J. Am. Chem. Soc. 128(23), 7444–7445 (2006). [CrossRef]   [PubMed]  

14. Y. S. Liu, D. T. Tu, H. M. Zhu, R. F. Li, W. Q. Luo, and X. Y. Chen, “A strategy to achieve efficient dual-mode luminescence of Eu3+ in lanthanides doped multifunctional NaGdF4 nanocrystals,” Adv. Mater. 22(30), 3266–3271 (2010). [CrossRef]  

15. L. Li, C. K. Tsung, Z. Yang, G. D. Stucky, L. D. Sun, J. F. Wang, and C. H. Yan, “Rare-earth-doped nanocrystalline titania microspheres emitting luminescence via energy transfer,” Adv. Mater. 20(5), 903–908 (2008). [CrossRef]  

16. W. Qin, D. Zhang, D. Zhao, L. Wang, and K. Zheng, “Near-infrared photocatalysis based on YF3: Yb3+,Tm3+/TiO2 core/shell nanoparticles,” Chem. Commun. (Camb.) 46(13), 2304–2306 (2010). [CrossRef]   [PubMed]  

17. N. Liu, W. Qin, G. Qin, T. Jiang, and D. Zhao, “Highly plasmon-enhanced upconversion emissions from Au@β-NaYF4:Yb,Tm hybrid nanostructures,” Chem. Commun. (Camb.) 47(27), 7671–7673 (2011). [CrossRef]   [PubMed]  

18. H. Q. Wang, M. Batentschuk, A. Osvet, L. Pinna, and C. J. Brabec, “Rare-earth ion doped up-conversion materials for photovoltaic applications,” Adv. Mater. 23(22-23), 2675–2680 (2011). [CrossRef]   [PubMed]  

19. T. Trupke, M. A. Green, and P. Wurfel, “Improving solar cell efficiencies by up-conversion of sub-band-gap light,” J. Appl. Phys. 92(7), 4117–4122 (2002). [CrossRef]  

20. B. Henke, B. Ahrens, J. A. Johnson, P. T. Miclea, and S. Schweizer, “Upconverted fluorescence in Er-doped ZBLAN glasses for high efficiency solar cells,” Proc. SPIE 7411, 74110E (2009). [CrossRef]  

21. Z. Meng, T. Yoshimura, K. Fukue, M. Higashihata, Y. Nakata, and T. Okada, “Large improvement in quantum fluorescence yield of Er3+-doped fluorozirconate and fluoroindate glasses by Ce3+ codoping,” J. Appl. Phys. 88(5), 2187–2190 (2000). [CrossRef]  

22. K. Nagamatsu, S. Nagaoka, M. Higashihata, N. J. Vasa, Z. Meng, S. Buddhudu, T. Okada, Y. Kubota, N. Nishimura, and T. Teshima, “Influence of Yb3+ and Ce3+ codoping on fluorescence characteristics of Er3+-doped fluoride glass under 980 nm excitation,” Opt. Mater. 27(2), 337–342 (2004). [CrossRef]  

23. D. Q. Chen, Y. S. Wang, Y. L. Yu, E. Ma, and Z. J. Hu, “Spectroscopic properties of Er3+ ions in transparent oxyfluoride glass ceramics containing CaF2 nanocrystals,” J. Phys. Condens. Matter 17(41), 6545–6557 (2005). [CrossRef]  

24. Y. L. Yu, Z. F. Shan, D. Q. Chen, P. Huang, H. Lin, and Y. S. Wang, “Improving Er3+ 1.53 μm luminescence by CeF3 nanocrystallization in aluminosilicate glass,” J. Appl. Phys. 108(12), 123523 (2010). [CrossRef]  

25. D. Q. Chen, Y. S. Wang, Y. L. Yu, E. Ma, F. Bao, Z. J. Hu, and Y. Cheng, “Luminescence at 1.53 μm for a new Er3+-doped transparent oxyfluoride glass ceramic,” Mater. Res. Bull. 41(6), 1112–1117 (2006). [CrossRef]  

26. M. Wang, C. C. Mi, Y. X. Zhang, J. L. Liu, F. Li, C. B. Mao, and S. K. Xu, “NIR-responsive silica-coated NaYbF4:Er/Tm/Ho upconversion fluorescent nanoparticles with tunable emission colors and their applications in immunolabeling and fluorescent imaging of cancer cells,” J. Phys. Chem. C 113(44), 19021–19027 (2009). [CrossRef]  

27. S. X. Shen, B. Richards, and A. Jha, “Enhancement in pump inversion efficiency at 980 nm in Er3+/Eu3+ and Er3+/Ce3+ doped tellurite glass fibers,” Opt. Express 14(12), 5050–5054 (2006). [CrossRef]  

28. E. Sani, A. Toncelli, and M. Tonelli, “Effect of Cerium codoping on Er:BaY2F8 crystals,” Opt. Express 13(22), 8980–8992 (2005). [CrossRef]   [PubMed]  

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 Characterization data for double-doped NaYF4:Er3+,Yb3+ NCs. (A) TEM image; (B) HRTEM image; (C) FFT pattern of a single nanocrystal; (D) histogram of the particle sizes obtained from TEM images of 400 NCs; (E) XRD pattern.
Fig. 2
Fig. 2 The FTIR spectrum of the oleic acid-coated NaYF4:Er3+,Yb3+,Ce3+ NCs.
Fig. 3
Fig. 3 (A) UC emission spectra of the NaYF4:Er3+,Yb3+,Ce3+ NCs with 0%, 2%, 4%, and 10% Ce3+. (B) Plot of relative UC emission intensities vs. codoped Ce3 + ion concentrations. (C) DC emission spectra of the NaYF4:Er3+,Yb3+,Ce3+ NCs with 0%, 2%, 4%, and 10% Ce3 + . (D) Energy level diagrams of Yb3+, Er3+, and Ce3+ ions, and possible processes of populations and emissions.
Fig. 4
Fig. 4 Emission decay curves of Er3+ in the NaYF4: Er3+,Yb3+,Ce3+ NCs (excited at 980 nm, monitored at 544 nm and 658 nm corresponding to the 4S3/24I15/2 and 4F9/24I15/2 transitions, respectively).
Fig. 5
Fig. 5 (A) Experimental setup for measuring the optical gain of the waveguide amplifier; (B) SEM image of the NaYF4:Er3+,Yb3+,Ce3+ NCs-doped polymer waveguide amplifier; (C) Relative gain as a function of pump power (980 nm) with 0.1 mW input signal powers (1535 nm) in an 4 μm high, 8 μm wide and 1.3 cm long NaYF4:Er3+,Yb3+,Ce3+ and NaYF4:Er3+,Yb3+ NCs-dispersed polymer waveguide.
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.