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

Near-infrared quantum cutting in Nd3+ and Yb3+ Doped BaGd2ZnO5 phosphors

Open Access Open Access

Abstract

Nd3+ and Yb3+ doped BaGd2ZnO5 phosphors were synthesized via high temperature solid-state method. The composition and structure have been investigated by the X-ray powder diffraction. The near-infrared quantum cutting for BaGd2ZnO5: Nd3+, Yb3+ is proved by the visible and near-infrared excitation, emission spectra and decay curves. Upon 359 nm excitation, visible and near-infrared emission of Nd3+ decrease with Yb3+ concentration increasing, and the intensities of 978 nm near-infrared emission of Yb3+ increase firstly and then decrease because of concentration quenching. The corresponding quantum cutting mechanism has been discussed through the energy level diagram. The maximum energy transfer efficiency calculated approaches 31.9%. Furthermore, the quenching concentration of Yb3+ is 7%. This study has prospect to be applied in the silicon-based solar cells to progress the conversion efficiency.

© 2015 Optical Society of America

1. Introduction

The application of solar cells is of great significance. The conversion from sunlight to electricity using solar cell devices represents a promising approach to green and renewable energy. However, the efficiency limit of the conventional crystalline silicon (c-Si) solar cells was estimated to be 29% in the early 1960s (Shockley Queisser limit) [1]. The main difficulty in improving the conversion efficiency of solar energy is the spectral mismatch between the energy distribution of photons in the incident solar spectrum and the band gap of a semiconductor material [2–4]. Among the different approaches to improve the solar cells efficiency, the quantum cutting (QC) is one of the interesting ways proposed since the thermalization effect of the electrons is a main limitation to the conversion efficiency of solar cells [3,5,6]. Owing to the abundant energy levels and narrow emission spectra lines, rare earth (RE) ions play a great role in the QC down-conversion (DC) process. Yb3+ ion has been extensively applied for its near-infrared (NIR) emission around 980nm (2F5/22F7/2), which is just above the band-gap of the Si-based solar cells. Recently, rare earth (RE) doped DC luminescent materials, such as Er3+-Yb3+ [7–10], Ho3+-Yb3+ [11,12], Tm3+-Yb3+ [13,14], Ce3+-Yb3+ [15], Dy3+-Yb3+ [16–18], Tb3+-Yb3+ [19–23], have been used in solar cells to improve solar energy conversion efficiencies.

In quantum cutting process, the low phonon energy and high chemical stability are the critical factors. Low phonon energy can prevent the non-radiative energy loss of multi-phonon relaxation of doped rare earth ions and enhance the energy transfer efficiency [24,25]. Besides, the chemical stability is an essential condition to the solar cells [25]. The halide and oxide have been proposed in other studies. The halide material has the low phonon energy but chemical stability is poorer. The oxide possesses good chemical stability but their phonon energy is higher. As a new host material, BaGd2ZnO5 has the low phonon energy and high chemical stability. It has been reported in previous studies [26]. The phonon energy of this host material is only 360 cm−1 [27,28].

In this paper, the NIR quantum cutting of Nd3+, Yb3+ doped BaGd2ZnO5 phosphors was studied through the visible and NIR light emission and excitation spectra and energy level diagram. With the measurement of lifetime decay curves, the quantum efficiency and energy transfer efficiency were also calculated.

2. Experimental

2.1Sample preparation

BaGd2ZnO5: Nd3+, Yb3+ were synthesized via high temperature solid-state method. BaCO3,ZnO,Gd2O3,Nd2O3 and Yb2O3 (A.R.) were used as the starting materials. Chemicals were weighed in the stoichiometric ratio and then grinded adequately in the agate mortars. The homogeneous material was calcined at 1300 °C for 3h in the muffle furnace and then cooled down to the room temperature.

2.2 Characteristic

X-ray powder diffraction (XRD) is carried out in the 2θ range 10-80° using a Bruker D8 advance X-ray diffracmeter (Bruker Optics, Ettlingen, Germany) with Cu Kα radiation. The emission spectra and excitation spectra were measured by the combined time resolved and steady state fluorescence spectrometers (FLS980, Edinburgh Instruments) equipped with two photomultiplier tubes (PMT) (Hamamatsu R928, U = 900V and a nitrogen cooled (−85°C) R5509-73, U = 1500V), as well as continuous wavelength 450W ozone free xenon (Xe) lamp as excitation sources. Decay curves were measured using FLS920 (Edinburgh Instruments) with a microsecond Xe lamp as the excitation source.

3. Results and discussion

3.1 Crystal structure

Figure 1(a) shows the XRD patterns of Nd3+, Yb3+ doped BaGd2ZnO5 phosphors and the standard card data (JCPDS No.49-0518). The positions and relative intensities of the diffraction peaks accord well with the data reported in the JCPDS standard card (No.49-0518). The radii of Yb3+ and Nd3+ are similar to Gd3+, so the XRD patterns of the samples with different doping concentrations are similar. It indicates that the pure-phase BaGd2ZnO5 phosphors are obtained. Figure 1(b) shows the schematic crystal structure of BaGd2ZnO5 seen from the c axis, which belongs to the orthorhombic. Furthermore, the lattice constants of BaGd2ZnO5 are: a = 7.15731 Å, b = 12.49399 Å, c = 5.77427 Å and V = 516.35 Å3.

 figure: Fig. 1

Fig. 1 (A) XRD patterns of BaGd2ZnO5: 1%Nd3+, x%Yb3+ (x = 0, 1, 3, 7, 10) and standard card data (JCPDS No.49-0518); (B) the crystal structure of BaGd2ZnO5 seen from the c axis.

Download Full Size | PDF

3.2 Fluorescence spectra

In order to investigate the luminescence properties, the visible light photoluminescence of BaGd2ZnO5: 1%Nd3+, x%Yb3+ (0, 1, 3, 5, 7, 10) were measured. Figures 2(a) and 2(b) show the excitation spectra monitored at 423 nm emission light derived from 2P1/24I9/2 transition of Nd3+ and emission spectra excited by 359 nm ultra violet (UV) light derived from 4D3/24I9/2 transition, respectively. The excitation peak is shown in 359 nm, and emission peaks at 423 nm and 457 nm are obtained. The intensity decreases gradually with the concentration of Yb3+ increasing from 0 to 10%, which is shown in the inset. It indicates that the energy transfer (ET) from Nd3+ to Yb3+ occurs, which decreases the population of Nd3+ excited state level.

 figure: Fig. 2

Fig. 2 Visible light photoluminescence of BaGd2ZnO5: 1%Nd3+, x%Yb3+ (x = 0, 1, 3, 5, 7, 10): (A) excitation spectra (λem = 423 nm); (B) emission spectra (λex = 359 nm); inset is the varying of emission peak intensity of 423 nm and 457 nm with Yb3+ concentration.

Download Full Size | PDF

For further studying the properties of ET from Nd3+ to Yb3+, the excitation and emission spectra of x%Yb3+ doped BaGd2ZnO5: 1%Nd3+ (x = 0, 1, 3, 5, 7, 10) were measured and analyzed. Figure 3(a) shows the excitation spectra monitored at 978 nm derived from the 2F5/22F7/2 transition of Yb3+. The excitation peaks at 336 nm, 359 nm, 437 nm and 480 nm are obtained, and the intensities of 978 nm emission increase firstly and then decrease. The difference is that the maximums of 336 nm, 359 nm, 437 nm and 480 nm peaks correspond to 5%Yb3+, 7%Yb3+, 1%Yb3+ and 1%Yb3+, respectively, which are shown in the inset of Fig. 3(a). It indicates that the quenching concentrations for these four peaks are different. Furthermore, the shape of 359 nm excitation peak monitored at 978 nm is similar to that monitored at 423 nm in Fig. 2(a). The quenching concentration for 359 nm excitation peak is up to7%, so it is chosen as the excitation light source to measure the emission spectra. Figure 3(b) shows the emission spectra of BaGd2ZnO5: 1%Nd3+, x%Yb3+ex = 359 nm), and the NIR emission peaks are located at 978 nm and 1071 nm, which are derived from 2F5/22F7/2 transition of Yb3+ and 4F3/24I11/2 transition of Nd3+, respectively. In Nd3+ singly doped BaGd2ZnO5 phosphors, the 978 nm (Yb3+ 2F5/22F7/2) emission light is not observed. However, intensities of these peaks increase when adding Yb3+ into BaGd2ZnO5: 1%Nd3+ phosphors. The intensity of 978 nm emission increases firstly and then decreases because of concentration quenching, but that of 1071 nm emission decreases gradually with the Yb3+ concentration increasing. It further indicates that the ET from Nd3+ to Yb3+ occurs from all above. Furthermore, the quenching concentration of Yb3+ under the excitation of 359 nm is 7%, so it is a useful highly doped material for silicon-based solar cells. Interestingly, the intensities variety of 336 nm, 437 nm and 480 nm excitation peaks proves that the ET also occurs under the excitation of these lights. However, the 978 nm emission intensities excited by these peaks are weak compared with 359 nm. So, we mainly studied the 359 nm excitation peak in this paper.

 figure: Fig. 3

Fig. 3 Near-infrared photoluminescence of BaGd2ZnO5: 1%Nd3+, x%Yb3+ (x = 0, 1, 3, 5, 7, 10): (A) excitation spectra (λem = 978 nm), inset shows the excitation peaks intensities varying with Yb3+ concentrations; (B) emission spectra (λex = 359 nm); inset is the relation between emission peak intensity of 423 nm and 457 nm and Yb3+ concentration.

Download Full Size | PDF

For studying the ET mechanism, the energy level diagram of Nd3+ and Yb3+ is shown in Fig. 4.Upon 359 nm excitation, the 4D3/24I9/2 transition occurs, and then the electrons jump to the 2P1/2 by multi-phonon relaxation (MPR). After that, the 2P1/24I9/2 transition occurs, and 423 nm blue light is obtained. Part of electrons in 2P1/2 jump to 4G11/2 by MPR, and the 457 nm blue light is observed. From Fig. 2 and Fig. 3, the ET from Nd3+ to Yb3+ has been proved. The energies of 423 nm and 457 nm are approximately twice the energy of the Yb3+ 2F5/22F7/2 transition, so the 978 nm NIR emission comes from 423 nm and 457 nm photons of Nd3+ as the way of quantum cutting, which can be described as follows:

P21/2(Nd3+)+2F27/2(Yb3+)I49/2(Nd3+)+2F25/2(Yb3+)
2F25/2(Yb3+)2F27/2(Yb3+)
Figure 5 shows the luminescence decay curves of BaGd2ZnO5: 1%Nd3+, x%Yb3+ (x = 0, 1, 3, 5, 7, 10), which corresponds to the 423 nm emission of 2P1/24I9/2 transition, and the excitation wavelength is 359 nm. With Yb3+ concentration increasing, the lifetime gradually decreases. The inset clearly shows the lifetime varying process upon Yb3+ contents. Combining the photoluminescence properties proposed above, the decrease of lifetime of 423 nm emission comes from the other path that ET from Nd3+ to Yb3+. With the Yb3+ concentration increasing, more energy is transferred to Yb3+, so the lifetime of 2P1/2 energy level decreases. So, the ET from Nd3+ to Yb3+ is proved again. Furthermore, the lifetime decay curves of 457 nm derived from 4G11/24I9/2 transition are given in Fig. 6, and the similar lifetime varying tendency is observed. The reason is that the electrons of 4G11/2 energy level come from the MPR of 2P1/2 level, and the energy of this level is also transferred to Yb3+.

 figure: Fig. 4

Fig. 4 The energy level diagram of Nd3+ and Yb3+ and the ET mechanism between them.

Download Full Size | PDF

 figure: Fig. 5

Fig. 5 Luminescence decay curves of BaGd2ZnO5: 1%Nd3+, x%Yb3+ (x = 0, 1, 3, 5, 7, 10) (λex = 359 nm, λem = 423 nm), the inset is the lifetime varying process upon Yb3+ concentration.

Download Full Size | PDF

 figure: Fig. 6

Fig. 6 Luminescence decay curves of BaGd2ZnO5: 1%Nd3+, x%Yb3+ (x = 0, 1, 3, 5, 7, 10) (λex = 359 nm, λem = 457nm).

Download Full Size | PDF

Table 1 shows the lifetime, energy transfer efficiency (ETE) and quantum efficiency (QE) of BaGd2ZnO5: 1%Nd3+, x%Yb3+ (x = 0, 1, 3, 5, 7, 10) (λex = 359 nm, λem = 423 nm). The ETE and QE are calculated by lifetime as follows [29–31]:

τ=+I(t)I0dt
ηET=1τx%Yb3+τ0Yb3+
ηQE=ηNdβ(1ηET)+2ηYbηET
where x % represents the Yb3+ concentration and I denotes the decay intensity. τ is the lifetime and ηET is the energy transfer efficiency. ηNd and ηYb stand for the luminescent quantum efficiency of the 4F3/2 level of Nd3+ and the 2F5/2 level of Yb3+, respectively. β for the population ratio of the energy levels 4F3/2 to 2P1/2. Assuming that all the excited Yb3+ ions and the residual excited Nd3+ ions decay radiatively, i.e., ηNd = ηYb = β = 1 [15, 30, 32], which is an ideal scenario, the upper limited values of the total down-conversion QE are calculated to be 100% to 131.9% for samples with Yb3+ contents from 0% to 10%.

Tables Icon

Table 1. Lifetime (τ), energy transfer efficiency (ηET) and quantum efficiency (ηQE) of BaGd2ZnO5: 1%Nd3+, x%Yb3+ (x = 0, 1, 3, 5, 7, 10) (λex = 359 nm, λem = 423 nm).

4. Conclusion

In this paper, Nd3+ and Yb3+ doped BaGd2ZnO5 phosphors were synthesized and investigated as NIR quantum cutting luminescence materials. The composition and structure have been studied by the X-ray powder diffraction. Upon the excitation of 359 nm light, the 423 nm, 457 nm and 1071 nm light of Nd3+ are obtained, and the intensities of these emission light all decrease with Yb3+ concentration increasing. With Yb3+ adding into the BaGd2ZnO5: Nd3+, the 978 nm NIR light of Yb3+ is also obtained under the 359 nm excitation. The intensities of 978 nm light increase firstly and then decrease because of concentration quenching, and the high quenching concentration is 7%. It indicates that the ET from Nd3+ to Yb3+ occurs. The maximum of ET is 31.9%. This material has a potential application in solar cells to progress the conversation efficiency.

Acknowledgments

This work was supported by National Science Foundation of China (No. 11474083), Hebei Province Department of Education Fund (ZD2014069).

References and links

1. W. Shockley and H. J. Queisser, “Detailed Balance Limit of Efficiency of p-n Junction Solar Cells,” J. Appl. Phys. 32(3), 510–519 (1961). [CrossRef]  

2. B. M. van der Ende, L. Aarts, and A. Meijerink, “Lanthanide ions as spectral converters for solar cells,” Phys. Chem. Chem. Phys. 11(47), 11081–11095 (2009). [CrossRef]   [PubMed]  

3. C. Strümpel, M. McCann, G. Beaucarne, V. Arkhipov, A. Slaoui, V. S. Vrcek, C. del Cañizo, and I. Tobias, “Modifying the solar spectrum to enhance silicon solar cell efficiency—An overview of available materials,” Sol. Energy Mater. Sol. Cells 91(4), 238–249 (2007). [CrossRef]  

4. W. A. Tisdale, K. J. Williams, B. A. Timp, D. J. Norris, E. S. Aydil, and X. Y. Zhu, “Hot-electron transfer from semiconductor nanocrystals,” Science 328(5985), 1543–1547 (2010). [CrossRef]   [PubMed]  

5. Y. T. An, C. Labbé, J. Cardin, M. Morales, and F. Gourbilleau, “Highly Efficient Infrared Quantum Cutting in Tb3+−Yb3+ Codoped Silicon Oxynitride for Solar Cell Applications,” Adv. Optical Mater. 1(11), 855–862 (2013). [CrossRef]  

6. M. D. Prèa, I. Morrow, D. J. Martin, M. Mos, A. D. Negro, S. Padovani, and A. Martucci, “Preparation and characterization of down shifting ZnS:Mn/PMMA nanocomposites for improving photovoltaic silicon solar cell efficiency,” Mater. Chem. Phys. 139(2-3), 531–536 (2013). [CrossRef]  

7. V. K. Tikhomirov, V. D. Rodríguez, J. Méndez-Ramos, J. del-Castillo, D. Kirilenko, G. Van Tendeloo, and V. V. Moshchalkov, “Optimizing Er/Yb ratio and content in Er–Yb co-doped glass-ceramics for enhancement of the up- and down-conversion luminescence,” Sol. Energy Mater. Sol. Cells 100, 209–215 (2012). [CrossRef]  

8. S. Xiao, X. Yang, and J. W. Ding, “Red and near infrared down-conversion in Er3+/Yb3+ co-doped YF3 performed by quantum cutting,” Appl. Phys. B 99(4), 769–773 (2010). [CrossRef]  

9. B. Fan, C. Chlique, O. M. Conanec, X. H. Zhang, and X. P. Fan, “Near-Infrared Quantum Cutting Material Er3+/Yb3+ Doped La2O2S with an External Quantum Yield Higher than 100%,” J. Phys. Chem. C 116(21), 11652–11657 (2012). [CrossRef]  

10. J. J. Eilers, D. Biner, J. T. van Wijngaarden, K. Krämer, H. U. Güdel, and A. Meijerink, “Efficient visible to infrared quantum cutting through downconversion with the Er3+-Yb3+ couple in Cs3Y2Br9,” Appl. Phys. Lett. 96(15), 151106 (2010). [CrossRef]  

11. H. Lin, D. Chen, Y. L. Yu, A. P. Yang, and Y. S. Wang, “Near-infrared quantum cutting in Ho3+/Yb3+ codoped nanostructured glass ceramic,” Opt. Lett. 36(6), 876–878 (2011). [CrossRef]   [PubMed]  

12. L. N. Guo, Y. Wang, J. Zhang, Y. Z. Wang, and P. Y. Dong, “Near-infrared quantum cutting in Ho3+, Yb3+-codoped BaGdF5 nanoparticles via first- and second-order energy transfers,” Nanoscale Res. Lett. 7(1), 636 (2012). [CrossRef]   [PubMed]  

13. 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(3), 654–657 (2010). [CrossRef]  

14. 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(19), 6974–7180 (2012). [CrossRef]  

15. S. F. Zou, Z. L. Zhang, F. Zhang, and Y. L. Mao, “High efficient quantum cutting in Ce3+/Yb3+co-doped oxyfluoride glasses,” J. Alloy. Comp. 572, 110–112 (2013). [CrossRef]  

16. W. L. Zhou, J. Yang, J. Wang, Y. Li, X. J. Kuang, J. K. Tang, and H. B. Liang, “Study on the effects of 5d energy locations of Ce³⁺ ions on NIR quantum cutting process in Y₂SiO₅: Ce³⁺, Yb³⁺,” Opt. Express 20(S4Suppl 4), A510–A518 (2012). [CrossRef]   [PubMed]  

17. Q. Yan, J. Ren, Y. Tong, and G. Chen, “Near-infrared quantum cutting of Eu2+/Yb3+ codoped chalcohalide glasses,” J. Am. Ceram. Soc. 96(5), 1349–1351 (2013). [CrossRef]  

18. Y. Dwivedi and S. B. Rai, “Spectroscopic study of Dy3+ and Dy3+/Yb3+ ions co-doped in barium fluoroborate glass,” Opt. Mater. 31(10), 1472–1477 (2009). [CrossRef]  

19. Q. Q. Duan, F. Qin, D. Wang, W. Xu, J. M. Cheng, Z. Zhang, and W. Cao, “Quantum cutting mechanism in Tb3+-Yb3+ co-doped oxyfluoride glass,” J. Appl. Phys. 110(11), 113503 (2011). [CrossRef]  

20. I. R. Martín, A. C. Yanes, J. Méndez-Ramos, M. E. Torres, and V. D. Rodríguez, “Cooperative energy transfer in Yb3+–Tb3+ codoped silica sol-gel glasses,” J. Appl. Phys. 89(5), 2520 (2001). [CrossRef]  

21. S. Ye, B. Zhu, J. X. Chen, J. Luo, and J. R. Qiu, “Infrared quantum cutting in Tb3+,Yb3+ codoped transparent glass ceramics containing CaF2 nanocrystals,” Appl. Phys. Lett. 92(14), 141112 (2008). [CrossRef]  

22. Y. H. Wang, L. C. Xie, and H. J. Zhang, “Cooperative near-infrared quantum cutting in Tb3+, Yb3+ codoped polyborates La0.99−xYbxBaB9O16:Tb0.01,” J. Appl. Phys. 105, 023528 (2009). [CrossRef]  

23. J. J. Zhou, Y. Teng, S. Ye, Y. X. Zhuang, and J. R. Qiu, “Enhanced downconversion luminescence by co-doping Ce3+ in Tb3+–Yb3+ doped borate glasses,” Chem. Phys. Lett. 486(4-6), 116–118 (2010). [CrossRef]  

24. K. Deng, X. Wei, X. Wang, Y. Chen, and M. Yin, “Near-infrared quantum cutting via resonant energy transfer from Pr3+ to Yb3+ in LaF3,” Appl. Phys. B 102(3), 555–558 (2011). [CrossRef]  

25. Z. G. Xia, Y. Luo, M. Guan, and L. B. Liao, “Near-infrared luminescence and energy transfer studies of LaOBr:Nd3+/ Yb3+,” Opt. Express 20(S5Suppl 5), A722–A728 (2012). [CrossRef]   [PubMed]  

26. Y. M. Yang, F. Y. Jiao, H. X. Su, Z. Q. Li, J. P. Jiao, Y. F. Liu, and Z. Q. Li, “Synthesis and up-conversion luminescence efficiency of Yb3+, Er3+ co-doped BaGd2ZnO5 Phosphors,” Chin. J. Lumin. 33, 1319–1323 (2012). [CrossRef]  

27. B. Tian, B. Chen, Y. Tian, J. Sun, X. P. Li, J. S. Zhang, H. Y. Zhong, L. H. Cheng, Z. L. Wu, and R. N. Hua, “Visible quantum cutting in BaGd2ZnO5: Eu3+ phosphor,” Ceram. Int. 38(5), 3537–3540 (2012). [CrossRef]  

28. B. N. Tian, B. J. Chen, Y. Tian, J. S. Sun, X. P. Li, J. S. Zhang, H. Y. Zhong, L. H. Cheng, and R. N. Hua, “Concentration and temperature quenching mechanisms of Dy3+ luminescence in BaGd2ZnO5 phosphors,” J. Phys. Chem. Solids 73(11), 1314–1319 (2012). [CrossRef]  

29. J. Sun, Y. Sun, J. Zeng, and H. Du, “Near-infrared quantum cutting in Eu2+, Yb3+ co-doped Sr3Gd(PO4)3 phosphor,” Opt. Mater. 35(6), 1276–1278 (2013). [CrossRef]  

30. J. D. Chen, H. Zhang, F. Li, and H. Guo, “High efficient near-infrared quantum cutting in Ce3+,Yb3+ co-doped LuBO3 phosphors,” Mater. Chem. Phys. 128(1-2), 191–194 (2011). [CrossRef]  

31. S. Z. Cai, L. L. Liu, M. M. Li, F. Yu, C. Mi, X. D. Li, J. Zhang, Y. Z. Liu, and Y. M. Yang, “Synthesis and down-conversion luminescence study of Er3+, Yb3+ co-doped NaYF4 phosphors,” Chin. J. Lumin. 35(9), 1058–1064 (2014). [CrossRef]  

32. Y. M. Yang, B. J. Chen, C. Wang, H. Y. Zhong, L. H. Cheng, J. S. Sun, Y. Peng, and X. Q. Zhang, “Investigation on structure and optical properties of Er3+, Eu3+ single-doped Na2O–ZnO–B2O3–TeO2 glasses,” Opt. Mater. 31(2), 445–450 (2008). [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) XRD patterns of BaGd2ZnO5: 1%Nd3+, x%Yb3+ (x = 0, 1, 3, 7, 10) and standard card data (JCPDS No.49-0518); (B) the crystal structure of BaGd2ZnO5 seen from the c axis.
Fig. 2
Fig. 2 Visible light photoluminescence of BaGd2ZnO5: 1%Nd3+, x%Yb3+ (x = 0, 1, 3, 5, 7, 10): (A) excitation spectra (λem = 423 nm); (B) emission spectra (λex = 359 nm); inset is the varying of emission peak intensity of 423 nm and 457 nm with Yb3+ concentration.
Fig. 3
Fig. 3 Near-infrared photoluminescence of BaGd2ZnO5: 1%Nd3+, x%Yb3+ (x = 0, 1, 3, 5, 7, 10): (A) excitation spectra (λem = 978 nm), inset shows the excitation peaks intensities varying with Yb3+ concentrations; (B) emission spectra (λex = 359 nm); inset is the relation between emission peak intensity of 423 nm and 457 nm and Yb3+ concentration.
Fig. 4
Fig. 4 The energy level diagram of Nd3+ and Yb3+ and the ET mechanism between them.
Fig. 5
Fig. 5 Luminescence decay curves of BaGd2ZnO5: 1%Nd3+, x%Yb3+ (x = 0, 1, 3, 5, 7, 10) (λex = 359 nm, λem = 423 nm), the inset is the lifetime varying process upon Yb3+ concentration.
Fig. 6
Fig. 6 Luminescence decay curves of BaGd2ZnO5: 1%Nd3+, x%Yb3+ (x = 0, 1, 3, 5, 7, 10) (λex = 359 nm, λem = 457nm).

Tables (1)

Tables Icon

Table 1 Lifetime (τ), energy transfer efficiency (ηET) and quantum efficiency (ηQE) of BaGd2ZnO5: 1%Nd3+, x%Yb3+ (x = 0, 1, 3, 5, 7, 10) (λex = 359 nm, λem = 423 nm).

Equations (5)

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

P 2 1 / 2 ( N d 3 + ) + 2 F 2 7 / 2 ( Y b 3 + ) I 4 9 / 2 ( N d 3 + ) + 2 F 2 5 / 2 ( Y b 3 + )
2 F 2 5 / 2 ( Y b 3 + ) 2 F 2 7 / 2 ( Y b 3 + )
τ = + I ( t ) I 0 d t
η E T = 1 τ x % Y b 3 + τ 0 Y b 3 +
η Q E = η N d β ( 1 η E T ) + 2 η Y b η E T
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.