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

Energy transfer and visible–infrared quantum cutting photoluminescence modification in Tm-Yb codoped YPO4 inverse opal photonic crystals

Open Access Open Access

Abstract

YPO4:Tm, Yb inverse opal photonic crystals were successfully synthesized by the colloidal crystal templates method, and the visible–infrared quantum cutting (QC) photoluminescence properties of YPO4:Tm, Yb inverse opal photonic crystals were investigated. We obtained tetragonal phase YPO4 in all the samples when the samples sintered at 950°C for 5 h. The visible emission intensity of Tm3+ decreased significantly when the photonic bandgap was located at 650 nm under 480 nm excitation. On the contrary, the QC emission intensity of Yb3+ was enhanced as compared with the no photonic bandgap sample. When the photonic bandgap was located at 480 nm, the Yb3+ and Tm3+ light-emitting intensity weakened at the same time. We demonstrated that the energy transfer between Tm3+ and Yb3+ is enhanced by the suppression of the red emission of Tm3+. Additionally, the mechanisms for the influence of the photonic bandgap on the energy transfer process of the Tm3+, Yb3+ codoped YPO4 inverse opal are discussed.

© 2015 Optical Society of America

1. INTRODUCTION

Recently, much attention has been focused on quantum cutting materials due to its application in Si solar cells. One major energy loss in Si solar cells is thermalization, which is expected to be considerably reduced if the absorbed UV/blue photon (λ<500nm) is cut into two near-infrared (NIR) photons that can be absorbed by Si (λabs<1100nm) [1]. Owing to the abundant energy levels and narrow emission spectra lines, rare earth (RE) ions play a great role in the quantum cutting (QC) downconversion (DC) process. The Yb3+ ion has been extensively used for its near-infrared emission around 980 nm (F5/22F7/22), which is just above the bandgap of Si-based solar cells. Therefore, the QC phosphor based on Yb: rare-earth couples could be realized as a downconversion convertor in front of solar cell panels [2]. The main systems of NIR QC phosphors are Tb3+-Yb3+ [38], Pr3+-Yb3+ [5,913], Tm3+-Yb3+ [5,1318] couples. For these couples, as the excited energy level of each donor ion is located at approximately twice the energy of that of Yb3+, the QC mechanism was supposed to be a second-order cooperative energy transfer (ET) process [5,8,1217]. Therefore, when the couple is absorbed a blue light photon will release two NIR photons. That is to say, the RE3+-Yb3+ couple would theoretically contribute the highest NIR quantum efficiency (QE) of 200%. In the previous works, the calculated NIR QE on the basis of decline lifetime is close to 200% when the Yb3+ ion is in a high doping concentration. However, the highest value before the concentration quenching threshold of the Yb ion is always very low. In the other words, the Yb ion light-emitting intensity achieves maximum with low Yb ion concentration, which is most beneficial for Si-based solar cells. Therefore, there is still a strong visible light emission of spontaneous radiation transition in the donor ions when the Yb ion is in the maximum value doping concentration for ET. As a consequence, the actual energy transfer efficiency and quantum efficiency are not high. In order to better apply this in solar cells, it is necessary to further enhance the NIR light emission of Yb3+. If the spontaneous radiation of sensitizing ions could be suppressed, the energy is likely to further transfer to Yb3+. Thereby, the near-infrared emission of the Yb3+ can be enhanced. The existence of a photonic bandgap of photonic crystal can control spontaneous emission. Hence, to weaken spontaneous radiation of sensitizing ions and enhance the energy transfer efficiency, changing the structure of QC materials is perhaps a promising way.

Photonic crystals have attracted considerable interest since the concept was first independently proposed by Yablonovitch and John [19,20]. Such crystals are three-dimensional periodic dielectric composites in which the distribution of the refractive index varies on the visible wavelength scale. This periodicity in the refractive index may lead to the formation of a photonic bandgap. The existence of a photonic bandgap can manipulate the spontaneous emission from light-emitting materials embedded in photonic crystals [2127].

Very recently, a few papers reported that spontaneous emission could be modified by photonic bandgap effects. The visible emission of Tm3+ will be suppressed if the photonic bandgap overlaps with the Tm3+ ions visible emission band. Since the energy will release in the another form, the infrared emission with the Yb3+ ions could be increased while the energy transfer between Tm3+ and Yb3+ is enhanced. In this work, the visible–infrared light-emitting photonic bandgap materials (YPO4:Tm, Yb) with an inverse opal structure were prepared by a self-assembly technique in combination with a solgel method. The effect of the photonic bandgap on QC emission of rare earth ions was investigated in inverse opal photonic crystals. In the YPO4:Tm, Yb inverse opal, we not only observed a strong modification of the visible emission spectra of Tm3+ but also obtained an energy transfer enhancement between Tm3+ and Yb3+ by suppression of the red emission of Tm3+.

2. EXPERIMENTAL

Commercial colloidal suspension with a monodispersive polystyrene microsphere (PS) having an average diameter of 360 or 490 nm was used to fabricate opal templates by the vertical deposition process. The single size (360 or 490 nm) PS microsphere suspension or mixed microsphere suspension consisting of 360 and 490 nm PS microspheres were added into the glass containers filled with deionized water. Quartz substrates (4×1cm) were vertically dipped into the glass containers filled with PS microsphere suspension, then placed in a 50°C oven. After the water was evaporated, the unitary order opal and binary disorder templates were formed on the quartz substrates.

The YPO4:Tm, Yb inverse opal was prepared by the template-assisted method. The YPO4 codoped with 1 mol. % Tm and 10 mol. % Yb precursor sol was prepared by using Y2O3, Yb2O3, Tm2O3, and P2O5 as raw materials. Initially, Y2O3, Yb2O3, and Tm2O3 were dissolved in hot nitric acid to form a nitrate solution, which was then evaporated until drying out. The dry nitrates and P2O5 were dissolved in ethanol separately, then mixed together. The prepared YPO4:Tm, Yb precursor solutions were used to infiltrate into the voids of the opal templates. YPO4:Tm, Yb inverse opals were obtained by calcining at 950°C for 5 h in an air furnace. The YPO4:Tm, Yb inverse opals prepared by unitary opal templates constructed with single size microspheres 360 or 490 nm in diameter were denoted as IPC-I and IPC-II, respectively. The YPO4:Tm, Yb inverse opal prepared by the binary disorder template constructed with mixed microsphere 360 and 490 nm in diameter was denoted as the reference sample (RS). The downconversion emission measurements and luminescent decay curves of the inverse opals were carried out on an FLS920 under a 480 nm excitation. Transmittance spectra were measured by a Hitachi U-4100 spectrophotometer. The microstructures of the opal templates and inverse opals were observed by a scanning electron microscope (SEM; FEI QUANTA 200F). X-ray powder diffraction (XRD) data of the inverse opal photonic crystal were obtained on a Rigaku 2200 diffractometer.

3. RESULTS AND DISCUSSION

First, we observed morphologies of photonic crystal materials which we prepared by SEM. Figures 1(a) and 1(b) show the SEM images of a unitary opal template constructed with single microspheres 490 nm in diameter and a binary template constructed with mixed microspheres 360 nm and 490 nm in diameter, respectively. In a unitary opal, the packed microspheres form into a high ordered face-centered cubic structure with a (111) plane parallel to the surface of the glass substrate. The binary template showed that microspheres with two kinds of size were arranged in a completely disordered (random) pattern. Figures 1(c) and 1(d) show the SEM images of IPC-II and the reference sample, respectively. The SEM images of IPC-I are not shown due to its similarity to that of IPC-II. The lighter regions and the darker circles in the SEM images of IPC-II represent the walls of the inverse opals and the air spheres previously occupied by the PS microspheres, respectively. It can be clearly seen that the air spheres possessed a highly ordered hexagonal arrangement in IPC-II, which revealed that the ordered opal template was not destroyed during the sintering process. Inside each hollow region are dark holes corresponding to the air spheres in the sublayer. On the contrary, the air spheres were randomly arranged in the reference sample. The center-to-center distance between the air spheres in IPC-I and IPC-II is about 290 nm and 400 nm, respectively, both of which are about 20% smaller than the sizes of the corresponding PS microspheres used to form the template. Thus, it can be concluded that considerable shrinkage occurs during calcinations.

 figure: Fig. 1.

Fig. 1. (a) and (b) SEM images of unitary opal templates constructed with polystyrene microspheres 490 nm in diameter and binary templates constructed with polystyrene microspheres 490 nm and 360 nm in diameter, respectively. (c) and (d) SEM images of IPC-II and the reference sample, respectively.

Download Full Size | PDF

The structures of the inverse opal samples and the reference sample were also characterized. Figure 2 shows the XRD pattern of IPC-I, IPC-II, and the reference sample on a quartz substrate. The broadband range from 17° to 25° in the XRD pattern was considered as the diffraction of the quartz substrate. The tetragonal phase YPO4:Tm, Yb can be obtained in the three sample types when sintered at 950°C for 5 h. It is indicated that the inverse opal structure did not changed the crystals type. No impurity peaks were observed, implying that inverse opal is YPO4 in a pure tetragonal phase. The ordered YPO4:Tm, Yb inverse opals display iridescent regions, which can be easily observed with the naked eye. The observed bright iridescent color corresponds to the Bragg reflection from the ordered porous structure.

 figure: Fig. 2.

Fig. 2. XRD patterns of IPC-I, IPC-II, and the RS.

Download Full Size | PDF

To investigate the three-dimensional ordered structural effect in photonic crystals, the photonic stop band of inverse opal arising from Bragg diffraction of the inverse opal order structure should be measured. Figure 3 shows the transmittance spectra collected from the reference sample and IPC-I and IPC-II at the normal direction of the (111) plane. The photonic bandgaps for IPC-I and IPC-II are at 480 nm and 650 nm, respectively, while there was no photonic bandgap in the reference sample. Figure 4 shows the visible emission and the NIR emission spectra of IPC-I, IPC-II, and the RS, respectively. All samples show typical Tm3+ emission bands, corresponding to the G41F43 transitions. In the NIR region, the emission band near 1000 nm is attributed to the Yb3+:F5/22F7/22 transition. Compared with the visible emission intensity of the Tm3+ emission band in the reference sample, the visible emission intensity of the Tm3+ was decreased in IPC-II due to the inhibition of the Tm3+ spontaneous emission in the inverse opal. On the contrary, the QC emission intensity of the Yb3+ in IPC-II were enhanced in comparison with this in the reference sample. These results were attributed to the enhancement of energy transfer from Tm3+ to Yb3+. When the bandgap is located at 480 nm, Yb3+ and Tm3+ light-emitting intensity weakened at the same time. The absorption of Tm3+ will diminish when the photonic bandgap overlaps the Tm3+ excitation light wavelength range. This phenomenon strongly suggests the existence of energy transfer from Tm3+ to Yb3+.

 figure: Fig. 3.

Fig. 3. Transmittance spectra of IPC-I, IPC-II, and the RS.

Download Full Size | PDF

 figure: Fig. 4.

Fig. 4. Visible emission spectra and quantum cutting emission spectra of IPC-I, IPC-II, and the RS under 480 nm excitation.

Download Full Size | PDF

The luminescence decay curves of Tm3+:G41F43 emission at 650 nm were also observed to explain the changing of the emission spectra, as shown in Fig. 5. We prepared singly doped Y0.99Tm0.01PO4 with no photonic bandgap to demonstrate the energy transfer efficiency. It can be clearly seen that the luminescence lifetimes at 650 nm for singly doped Y0.99Tm0.01PO4, the RS, and IPC-II are 0.275, 0.203, and 0.138 ms under the excitation of 480 nm, respectively. The luminescence lifetime at 650 nm for IPC-I is identical with the RS. The Yb3+ doping concentration in the RS, IPC-I, and IPC-II was 10 mol. %. The decline in the lifetime as a function of Yb3+ concentration can be explained by the introduction of extra decay pathways due to the Yb3+ doping: ET from Tm3+:G41 to Yb3+ enhances the Tm3+:G41 decay rate. Compared with the decay curves of IPC-II and the RS, the decline with IPC-II revealed that the energy transfer efficiency was further enhanced. In the photonic crystals, spontaneous emission will be inhibited because the optical electromagnetic modes do not exist within the photonic bandgap frequency range. While the 650 nm emission was suppressed, the photons released another way. We supposed that the energy transfer rate became much faster in IPC-II which caused the luminescence lifetime at 650 nm to decline. By dividing the effective decay time of Tm3+ in the RS and IPC-II by that in the Yb3+ free counterpart, the energy transfer efficiency is calculated to be 26.2% and 49.8% for the RS and IPC-II, respectively, by the equation ηT=1τRSorIPC-II/τ0 [28], which indicates a rather efficient ET in photonic bandgap samples. Here, τ0 is the decay lifetime of Tm3+:G41 in the singly doped samples. The quantum efficiency is calculated to be 126.2% and 149.8% for the RS and IPC-II, respectively. The relation between the ET efficiency and the total QE is defined as [3] ηQE=ηTm(1ηRSorIPC-II)+2ηRSorIPC-II, where QE for the Tm3+ ions, ηTm, is set to 1.

 figure: Fig. 5.

Fig. 5. Photoluminescence decays from G41 of Tm3+ in singly doped Y0.99Tm0.01PO4, the RS, and IPC-II. The excitation and emission wavelengths are 480 nm and 650 nm, respectively.

Download Full Size | PDF

In order to understand the luminescence modification mechanisms of Tm-Yb codoped YPO4 inverse opal photonic crystals, the schematic energy level diagram of Tm3+ and Yb3+ ions is shown in Fig. 6. The G41 excited state Tm3+ returns to the intermediate state G41 through two mechanisms: spontaneous fluorescence emission of Tm3+ and energy transfer. The energy transfer and the spontaneous emission of the Tm3+ compete with each other. In IPC-II, the photonic bandgap at 650 nm inhibits the 650 nm fluorescence emission of the donor Tm3+. Therefore, the energy transfer of the other mechanism is enhanced through the route (Tm3+,G14F43)(Yb3+,F27/2F5/22). The G41 excited state of Tm3+ returns to the ground state by (Tm3+,G14F43)(Yb3+,F27/2F5/22), making energy transfer much more readily. Therefore, the energy transfer between Tm3+ and Yb3+ is greatly enhanced by the suppression of the red spontaneous emission of Tm3+. Thus, the QC emission from Yb3+ is considerably improved in IPC-II.

 figure: Fig. 6.

Fig. 6. Simplified energy level scheme of the quantum cutting emission mechanism in photonic crystals. The photonic bandgap in the photonic crystals is at 650 nm.

Download Full Size | PDF

4. CONCLUSION

In this paper, the Tm3+ sensitized quantum cutting emission inverse opal of YPO4:Tm, Yb was successfully fabricated. The effect of the photonic bandgap on the spontaneous emission of Tm3+ and the energy transfer efficiency of Tm3+ and Yb3+ was investigated. With the inhibition of Tm3+ emission by the photonic bandgap in the inverse photonic crystals, the Yb3+ NIR emission was enhanced at the same time and the fluorescence decay of Tm3+ was reduced. By monitoring the steady-state photoluminescence (PL) and the PL decay of the intermediate G41 level of Tm3+ as a function of the photonic bandgap, we demonstrated that the energy transfer efficiency and the quantum efficiency were increased to 49.8% and 149.8% by suppression of the red emission of Tm3+, respectively. The results demonstrate that the energy transfer efficiency and quantum efficiency is greatly improved. We believe that the present work will be valuable for not only the foundational study of quantum cutting emission modification but also for new optical devices in Si solar cells.

Funding

National Natural Science Foundation of China (NSFC) (51272097, 61265004, 61265004); Nature and Science Fund (KKJA201432042).

REFERENCES

1. T. Trupke, M. A. Green, and P. Wurfel, “Improving solar cell efficiencies by down-conversion of high-energy photons,” J. Appl. Phys. 92, 1668–1674 (2002). [CrossRef]  

2. B. S. Richards, “Luminescent layers for enhanced silicon solar cell performance: down-conversion,” Sol. Energy Mater. Sol. Cells 90, 1189–1207 (2006).

3. P. Vergeer, T. Vlugt, M. Kox, M. den Hertog, J. van der Eerden, and A. Meijerink, “Quantum cutting by cooperative energy transfer in YbxY1-xPO4: Tb3+,” Phys. Rev. B 71, 014119 (2005). [CrossRef]  

4. Q. Y. Zhang, C. H. Yang, Z. H. Jiang, and X. H. Ji, “Concentration-dependent near-infrared quantum cutting in GdBO3:Tb3+,Yb3+ nanophosphors,” Appl. Phys. Lett. 90, 061914 (2007). [CrossRef]  

5. Q. Y. Zhang, G. F. Yang, and Z. H. Jiang, “Cooperative downconversion in GdAl3(BO3)4:RE3+, Yb3+ (RE = Pr, Tb, and Tm),” Appl. Phys. Lett. 91, 051903 (2007). [CrossRef]  

6. S. Ye, B. Zhu, J. Chen, J. Luo, and J. R. Qiu, “Infrared quantum cutting in Tb3+, Yb3+ codoped transparent glass ceramics containing CaF2 nanocrystals,” Appl. Phys. Lett. 92, 141112 (2008). [CrossRef]  

7. D. Chen, Y. Yu, Y. Wang, P. Huang, and F. Weng, “Cooperative energy transfer up-conversion and quantum cutting down-conversion in Yb3+: TbF3 nanocrystals embedded glass ceramics,” J. Phys. Chem. C 113, 6406–6410 (2009). [CrossRef]  

8. Z. F. Wang, Y. H. Wang, Y. Z. Li, and H. J. Zhang, “Near-infrared quantum cutting in Tb3+, Yb3+ co-doped calcium tungstate via second-order downconversion,” J. Mater. Res. 26, 693–696 (2011). [CrossRef]  

9. B. M. van der Ende, L. Aarts, and A. Meijerink, “Near-infrared quantum cutting for photovoltaics,” Adv. Mater. 21, 3073–3077 (2009). [CrossRef]  

10. J. T. van Wijngaarden, S. Scheidelaar, T. J. H. Vlugt, M. F. Reid, and A. Meijerink, “Energy transfer mechanism for downconversion in the (Pr3+, Yb3+) couple,” Phys. Rev. B 81, 155112 (2010). [CrossRef]  

11. Y. Katayama and S. Tanabe, “Near infrared downconversion in Pr3+–Yb3+ codoped oxyfluoride glass ceramics,” Opt. Mater. 33, 176–179 (2010).

12. K. Y. Li, L. Y. Liu, R. Z. Wang, S. G. Xiao, H. Zhou, and H. Yan, “Broadband sensitization of downconversion phosphor YPO4 by optimizing TiO2 substitution in host lattice co-doped with Pr3+-Yb3+ ion-couple,” J. Appl. Phys. 115, 123103 (2014). [CrossRef]  

13. X. Liu, Y. Qiao, G. Dong, S. Ye, B. Zhu, G. Lakshminarayana, D. Chen, and J. Qiu, “Cooperative downconversion in Yb3+-RE3+ (RE = Tm or Pr) codoped lanthanum borogermanate glasses,” Opt. Lett. 33, 2858–2860 (2008). [CrossRef]  

14. L. Xie, Y. Wang, and H. Zhang, “Near-infrared quantum cutting in YPO4: Yb3+, Tm3+ via cooperative energy transfer,” Appl. Phys. Lett. 94, 061905 (2009). [CrossRef]  

15. S. Ye, B. Zhu, J. Luo, J. Chen, G. Lakshminarayana, and J. Qiu, “Enhanced cooperative quantum cutting in Tm3+-Yb3+ codoped glass ceramics containing LaF3 nanocrystals,” Opt. Express 16, 8989–8994 (2008). [CrossRef]  

16. Q. Zhang, B. Zhu, Y. X. Zhuang, G. R. Chen, X. F. Liu, G. Zhang, J. R. Qiu, and D. P. Chen, “Quantum cutting in Tm3+/Yb3+ -codoped lanthanum aluminum germanate glasses,” J. Am. Ceram. Soc. 93, 654–657 (2010). [CrossRef]  

17. G. C. Jiang, X. T. Wei, Y. H. Cheng, C. K. Duan, and M. Yin, “Broadband downconversion in YVO4:Tm3+, Yb3+ phosphors,” J. Rare Earths 31, 27–31 (2013) [CrossRef]  

18. W. Zheng, H. M. Zhu, R. F. Li, D. T. Tu, Y. S. Liu, W. Q. Luo, and X. Y. Chen, “Visible-to-infrared quantum cutting by phonon-assisted energy transfer in YPO4:Tm3+, Yb3+ phosphors,” Phys. Chem. Chem. Phys. 14, 6974–6980 (2012). [CrossRef]  

19. E. Yablonovitch, “Spontaneous emission in solid-state physics and electronics,” Phys. Rev. Lett. 58, 2059–2062 (1987). [CrossRef]  

20. S. John, “Strong localization of photons in certain disordered dielectric superlattices,” Phys. Rev. Lett. 58, 2486–2489 (1987). [CrossRef]  

21. J. Zhou, Y. Zhou, S. Buddhudu, S. L. Ng, Y. L. Lam, and C. H. Kam, “Photoluminescence of ZnS:Mn embedded in three-dimensional photonic crystals of submicron polymer spheres,” Appl. Phys. Lett. 76, 3513–3515 (2000). [CrossRef]  

22. Z. W. Yang, J. Zhou, X. G. Huang, G. Yang, Q. Xie, L. Sun, B. Li, and L. Li, “Photonic band gap and photoluminescence properties of LaPO4:Tb inverse opal,” Chem. Phys. Lett. 455, 55–58 (2008). [CrossRef]  

23. R. F. Nabiev, P. Yeh, and J. J. Sanchez-Mondragon, “Dynamics of the spontaneous emission of an atom into the photon-density-of-states gap: solvable quantum-electrodynamical model,” Phys. Rev. A 47, 3380–3384 (1993). [CrossRef]  

24. S. John and T. Quang, “Spontaneous emission near the edge of a photonic band gap,” Phys. Rev. A 50, 1764–1769 (1994). [CrossRef]  

25. X. H. Wang, R. Z. Wang, B. Y. Gu, and G. Z. Yang, “Decay distribution of spontaneous emission from an assembly of atoms in photonic crystals with pseudogaps,” Phys. Rev. Lett. 88, 093902 (2002). [CrossRef]  

26. Z. X. Li, L. L. Li, H. P. Zhou, Q. Yuan, C. Chen, L. D. Sun, and C. H. Yan, “Colour modification action of an upconversion photonic crystal,” Chem. Commun. 43, 6616–6618 (2009).

27. Z. W. Yang, K. Zhu, Z. G. Song, D. C. Zhou, Z. Y. Yin, and J. B. Qiu, “Energy transfer and photoluminescence modification in Yb–Er–Tm triply doped Y2Ti2O7 upconversion inverse opal,” Appl. Opt. 50, 287–290 (2011). [CrossRef]  

28. P. I. Paulose, G. Jose, V. Thomas, N. V. Unnikrishnan, and M. K. R. Warrier, “Sensitized fluorescence of Ce3+/Mn2+ system in phosphate glass,” J. Phys. Chem. Solids 64, 841–846 (2003). [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 (6)

Fig. 1.
Fig. 1. (a) and (b) SEM images of unitary opal templates constructed with polystyrene microspheres 490 nm in diameter and binary templates constructed with polystyrene microspheres 490 nm and 360 nm in diameter, respectively. (c) and (d) SEM images of IPC-II and the reference sample, respectively.
Fig. 2.
Fig. 2. XRD patterns of IPC-I, IPC-II, and the RS.
Fig. 3.
Fig. 3. Transmittance spectra of IPC-I, IPC-II, and the RS.
Fig. 4.
Fig. 4. Visible emission spectra and quantum cutting emission spectra of IPC-I, IPC-II, and the RS under 480 nm excitation.
Fig. 5.
Fig. 5. Photoluminescence decays from G 4 1 of Tm 3 + in singly doped Y 0.99 Tm 0.01 PO 4 , the RS, and IPC-II. The excitation and emission wavelengths are 480 nm and 650 nm, respectively.
Fig. 6.
Fig. 6. Simplified energy level scheme of the quantum cutting emission mechanism in photonic crystals. The photonic bandgap in the photonic crystals is at 650 nm.
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.