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

Enhanced 573 nm yellow emissions of Dy3+ via Tb3+ deactivation in Na2Gd4(MoO4)7 crystal

Open Access Open Access

Abstract

A Dy3+/Tb3+ co-doped Na2Gd4(MoO4)7 yellow laser crystal was successfully grown and analyzed. The use of Tb3+ codoping for enhancing the Dy3+:4F9/26H13/2 yellow emissions was investigated in the Na2Gd4(MoO4)7 crystal for the first time. In comparison to Dy3+ single-doped Na2Gd4(MoO4)7 crystal, the Dy3+/Tb3+ co-doped Na2Gd4(MoO4)7 crystal possessed a higher fluorescence branching ratio (80.2%), transition probability (3704 s−1), and fluorescence emission cross section (1.34×10−20 cm2) corresponding to the laser transition 4F9/26H13/2 of Dy3+. It was found that the introduced Tb3+ enhanced the 573 nm emission by depopulating the population of the laser lower level Dy3+: 6H13/2, and has little influence on the laser upper level Dy3+: 4F9/2 at the same time. These results suggest that Dy3+/Tb3+ co-doped Na2Gd4(MoO4)7 crystal may be an attractive host for developing solid state lasers at around 573 nm under a conventional 450 nm LD.

© 2017 Optical Society of America

1. Introduction

In recent years, visible laser, especially yellow lasers of 560–590 nm, are fascinated by researchers, because it can be widely applied in medical cosmetology, telecommunication, and detection [1–3]. At present, the most reliable technique to generate a continuous wave yellow lasers is using frequency mixing method from two single-frequency Nd:YAG lasers at 1064 nm and 1319 nm by nonlinear frequency-conversion techniques [4, 5]. However, such multiple-cavity systems are still complex to use. Dy3+-doped materials are attractive for directly emitting yellow lasers due to the 4F9/26H13/2 transition (∼570 nm) [6–8]. As is shown in Fig. 1, the energy gap between 4F9/2 and 6F1/2 is large (about 7500 cm−1), which means that multi-phonon relaxation is weak, and hence a high-yield visible emission of Dy3+ is generally expected. Besides, with the development of InGaN laser diodes, the new blue-violet emitting laser diodes are more and more accessible to be applied as optical pumping source. Therefore, potential yellow laser from Dy3+-doped crystal excited by blue laser diode are feasible. For example, yellow laser oscillations in Dy3+:Y3Al5O12 [9], Dy3+:LiLuF4 [10], and Dy3+:LiNbO3 [11] crystals have also been reported in recent years. Moreover, co-doping of deactivation ion, such as Tb3+ ion, has been demonstrated to be an effective method to quench the lower level 6H13/2 of Dy3+, which leads to a fast depopulation of the population in the terminal level, and reduces the pumping threshold [12, 13]. However, as far as we know, there is only one report about Dy3+, Tb3+ co-doped crystal as the gain material for yellow lasers, and limited to fluoride crystal. In 2014, a yellow laser performance with a highest output power of 55 mW was obtained in Dy3+, Tb3+ co-doped LiLuF4 crystal [10].

 figure: Fig. 1

Fig. 1 Simplified energy level diagram of Dy3+ and Tb3+ co-doped system. ET1: energy transfer from Dy3+:4F9/2 to Tb3+:5D4 level; ET2: energy transfer from Dy3+:6H13/2 to Tb3+:7F4 level.

Download Full Size | PDF

Moreover, the 4F9/26H13/2 transition for yellow fluorescence is a non-ground-state transition, the energy gap between 6H13/2 and 6H15/2 is about 3300 cm−1. Therefore, host materials with higher phonon energies may be beneficial to quench the lower level 6H13/2 by nonradiative transition from 6H13/2 to 6H15/2 based on the multiphonon relaxation [14, 15]. As we know, oxide crystals have phonon frequency of 600–1000 cm−1, which is about two times larger than that in fluoride crystals. Among the different oxide crystals, we are now focusing our scientific program on a potential gain medium Na2Gd4(MoO4)7 (NGM) crystal for efficient yellow laser operation.

Na2Gd4(MoO4)7 crystal, whose chemical formula can be rewritten to Na2/7Gd4/71/7MoO4 (∅ stand for vacancies), offering two nonequivalent cationic sites for optically active Re3+ substitution [16]. Consequently, NGM crystal is characterized by a high degree of structural disorder. Thus, the high degree of disorder of NGM crystal inevitably contributes to inhomogeneous broadening of the absorption and fluorescence spectra, bringing a large amount of advantages from the standpoint of applications as active media in LD-pumped solid-state lasers.

To the best of our knowledge, there are no reports on the yellow fluorescence properties in Dy3+, Tb3+ co-doped oxide crystals up to now. In this work, Dy3+, Tb3+ co-doped NGM crystal was successfully grown. Tb3+ was demonstrated to greatly facilitate the Dy3+: 4F9/26H13/2 emission by efficient energy transfer (ET) from Dy3+:6H13/2 to Tb3+: 7F4. The spectroscopic investigation of yellow fluorescence has been also made to future applications in yellow lasers.

2. Experimental section

The Dy3+ single doped and Dy3+/Tb3+ co-doped NGM crystals were grown by the Czochralski method in air. The raw materials were prepared from MoO3(99.99%), Na2CO3(99.99%), Gd2O3(99.99%), Dy2O3(99.99%), and Tb4O7(99.99%). The raw powders were mixed together, grinded and extruded to form chunk, and then placed in the corundum crucible and heated up to 900 C. Then kept for 8 h to make the reaction completely. After the solid-state reaction, these compounds were put into a platinum crucible with diameter of 50 mm and height of 60 mm, and slowly heated up to 1250 °C. The pulling rates and rotating rates were 0.8–2 mm/h and 5–15 r/min. When the pulling procedure ended, the crystal was cooled down to room temperature slowly and annealed at 900 °C in air atmosphere to reduce the color centers inside the crystal. A 1 mm thick crystal cut from the medium and polished was prepared to spectral measurement. The concentrations of Dy3+ and Tb3+ were measured by the inductively coupled plasma-atomic emission spectrometry (ICP–AES) method. The single-doped crystal was measured to be 2.08 at.% (2.65×1020 ions/cm3) of Dy3+. The co-doped crystal was measured to be 2.07 at.% (2.64×1020 ions/cm3) of Dy3+, and 2.0 at.% (2.55×1020 ions/cm3) of Tb3+, respectively.

The room temperature absorption spectra were measured by spectrophotometer (UV-3150, Shimadzu, Japan) in a range 400–1800 nm. Fluorescence spectra were measured by spectrometer (FL920, Edinburgh) under excitation at 454 nm in a range of 500–800 nm. The fluorescence decay curves of the as grown crystals were measured at 573 nm under pulse excitation of 454 nm. All the measurements were taken at room temperature.

3. Experimental results and discussion

The absorption spectrum of Dy3+/Tb3+ co-doped NGM crystal is shown in Fig. 2. It is obvious to see that there are many typical absorption peaks of Dy3+ ion, such as the wavelength near 1703, 1300, 1100, 907, 808, 759 and 454 nm, corresponding to the transitions from the ground state 6H15/2 to the excited state 6H11/2, 6H9/2+6F11/2, 6H7/2+6F9/2, 6H5/2+6F7/2, 6F5/2 and 4I15/2+4F9/2, respectively. In the range of 440–470 nm, the highest absorption of 0.15 cm−1 has been obtained at 454 nm with full width at half maximum (FWHM) of 9.2 nm, which is suitable for being pumped by blue-emitting InGaN laser diodes.

 figure: Fig. 2

Fig. 2 Absorption spectrum of Dy3+/Tb3+ co-doped NGM crystal in the range of 400–1800 nm.

Download Full Size | PDF

According to the Judd–Ofelt theory [17, 18], the intensity parameters Ω2,4,6 of Dy3+ (shown in Table 1) was calculated from the room-temperature absorption spectrum. It can be seen that the Ω2 of Dy3+ in the Dy3+/Tb3+ co-doped NGM crystal is larger than that of Dy3+ single-doped NGM crystal, and much larger than that of other Dy3+ doped crystals. It is well known that the value of Ω2 increases with the decrease in the symmetry [19]. The larger Ω2 of Dy3+ in the Dy3+/Tb3+ co-doped NGM crystal indicates that the codoping of Tb3+ ions would bring about a lower symmetry surrounding Dy3+ ions in NGM crystal.

Tables Icon

Table 1. Judd-Ofelt parameters Ω2,4,6, branching ratio β, emission cross sections σ, life-time of 4F9/2 level for Dy3+ ions (τM and τR are the measured and calculated radiative lifetime), and quantum efficiency η of different Dy3+ doped materials.

Based on the intensity parameters Ω2,4,6 of Dy3+, the spontaneous emission probabilities (A), radiative lifetime (τR) of Dy3+:4F9/2 level, and the fluorescence branching ratio β of Dy3+:4F9/26H13/2 transition can be also calculated, and the results are shown in Table 1. It is clear to see that the fluorescence branching ratio β of Dy3+:4F9/26H13/2 transition for yellow fluorescence emission around 573 nm of Dy3+ single-doped NGM crystal is as high as 75.2%, which is larger than that of the Dy3+ doped fluoride crystals, such as Dy:LiLuF4 (65.4%) [20], and Dy:KYF (59.8%) [21]. Moreover, compared to the Dy3+ single-doped NGM crystal, the Dy3+/Tb3+ co-doped NGM crystal possesses a larger fluorescence branching ratio β of Dy3+:4F9/26H13/2 transition, which is as high as 80.2%. It is well known that the larger fluorescence branching ratio represents the higher possibility of fluorescence emission. The larger β of Dy3+:4F9/26H13/2 transition in the Dy3+/Tb3+ co-doped NGM crystal indicates that the codoping of Tb3+ ions would bring about a more efficient yellow fluorescence emission.

To further study the effect of Tb3+ codoping on the yellow photoluminescence of Dy3+, the corresponding emission cross sections are calculated by the Fuchtbauer–Ladenburg formula [24]:

σem=Aβλ5I(λ)8πcn2λI(λ)dλ
where A is the radiative transition probability, β is the fluorescence branching ratio, c is the speed of light, n is the refractive index of the host, and I(λ)/∫λI(λ)dλ refers to the normalized line shape function of the experimental emission spectrum. And the results are shown in Fig. 3. It is clear to see that an enhanced fluorescence emission band centered at 573 nm corresponding to Dy3+: 4F9/26H13/2 transition is obtained with the codoping of Tb3+ in Dy3+ doped NGM crystal. In particular, the maximum emission cross section of Dy3+/Tb3+ co-doped NGM crystal is 1.34×10−20 cm2 at 573 nm, which is larger than that of Dy3+ single-doped NGM crystal (1.10×10−20 cm2). This enhanced fluorescence emission cross section is supposed to be ascribed to the higher fluorescence branching ratio of the 4F9/26H13/2 transition (from 75.2% to 80.2%) and lower radiative lifetime of the 4F9/2 level (from 255.7 μs to 216.4 μs) with the codoping of Tb3+ ions.

 figure: Fig. 3

Fig. 3 Yellow emission cross-sections of Dy3+ single-doped and Dy3+/Tb3+ co-doped NGM crystals.

Download Full Size | PDF

The time-resolved decays of the Dy3+:4F9/2 level for the Dy3+ single-doped and Dy3+/Tb3+ co-doped NGM crystals were measured, and shown in Fig. 4. The measured lifetime is 114 μs for the Dy3+/Tb3+ co-doped NGM crystal, which is a little shorter than that of Dy3+ single-doped NGM crystal (125 μs). As shown in Fig. 1, firstly ions of Dy3+:6H15/2 state are excited to Dy3+:4I15/2 state by a 450 nm LD, then decay nonradiatively to Dy3+:4F9/2 level by the multiphonon relaxation. One part of the Dy3+ ions on the 4F9/2 level will mainly decay radiatively to 6H15/2 state and 6H13/2 state with ~480 nm and the ∼573 nm emissions, respectively. Other Dy3+ ions on 4F9/2 level will undergo the energy transfer 1 (ET1) to the Tb3+ ion on the 5D4 level, which would go against the yellow fluorescence emission of Dy3+. The ET1 efficiency ηET from Dy3+ to Tb3+ can be described as ηET =1-τDy−−Tb/τDy, where τDy−−Tb and τDy are lifetime of Dy3+:4F9/2 level of Dy3+ doped NGM crystal with and without Tb3+, and the calculated result is as low as 8.8%. This confirms that the codoping of Tb3+ ions has little influence on the Dy3+:4F9/2 level. Moreover, the quantum efficiency, evaluated by η=τM/τR, where τM and τR are the measured and calculated (by Judd–Ofelt theory) radiative lifetime of 4F9/2 level for Dy3+ ions in Dy3+ doped crystals. Therefore, the value of η in the Dy3+/Tb3+ co-doped NGM crystal was calculated to be 52.7%, which is larger than that of Dy3+ single-doped NGM crystal (48.9%), and larger than that of the other Dy3+ doped crystals (shown in Table 1), indicating that Dy3+/Tb3+ co-doped NGM crystal is a promising candidate for yellow laser operation.

 figure: Fig. 4

Fig. 4 Fluorescence decay curves of Dy3+ single-doped and Dy3+/Tb3+ co-doped NGM crystals for the 4F9/2 mainfold.

Download Full Size | PDF

As shown in Fig. 1, after the yellow fluorescence emission based on the decay radiatively from Dy3+:4F9/2 level to Dy3+:6H13/2 level, the ions in the Dy3+:6H13/2 level will undergo the ET2 process to Tb3+:7F4 level, leading to an accelerated depletion of the population in the laser terminal level 6H13/2. In this way, the lifetime of the lower laser level 6H13/2 can be effectively decreased, which makes population inversion for the Dy3+:4F9/26H13/2 transition much easier and enhances the yellow fluorescence emission. To further confirm the energy interaction mechanism, the time-resolved decays of the 6H13/2 level of the Dy3+ single-doped and Dy3+/Tb3+ co-doped NGM crystals were measured, and the results are shown in Fig. 5. It is obvious to see that the measured lifetime value of the 6H13/2 level in the crystal co-doped with Tb3+ is 0.96 ms, which is 33.3% shorter when compared with that of the single-doped crystal (1.44 ms). This shortening of the measured lifetime indicates that there is a strong interaction between the Dy3+ and Tb3+, and confirms that Tb3+ ions are able to effectively depopulate the Dy3+:6H13/2 for ∼573 nm emission in NGM crystal by energy transition from Dy3+:6H13/2 to Tb3+:7F4, which may induce the population inversion and facilitate laser operation.

 figure: Fig. 5

Fig. 5 Fluorescence decay curves of Dy3+ single-doped and Dy3+/Tb3+ co-doped NGM crystals for the 6H13/2 mainfold.

Download Full Size | PDF

4. Conclusion

In conclusion, Dy3+ single- and Dy3+/Tb3+ co-doped Na2Gd4(MoO4)7 crystals were successfully grown. Compared with the Dy3+ single-doped Na2Gd4(MoO4)7 crystal, the Dy3+/Tb3+ co-doped Na2Gd4(MoO4)7 crystal has higher fluorescence branching ratio (80.2%), transition probability (3704 s−1), and fluorescence emission cross section (1.34×10−20 cm2) corresponding to the laser transition 4F9/26H13/2 of Dy3+ around the yellow bands. It was also demonstrated that the codoping of Tb3+ depopulates the population of the laser lower level Dy3+: 6H13/2, and has little influence on the laser upper level Dy3+: 4F9/2 at the same time, which is benefited to the possible population inversion for the Dy3+:4F9/26H13/2 transition, and enhances the fluorescence emission. These results suggests that the Dy3+/Tb3+ co-doped Na2Gd4(MoO4)7 crystal is a promising material for yellow laser applications under being pumped by a conventional 450 nm LD.

Funding

The National Key Research and Development Program of China (2017YFB1104500); National Natural Science Foundation of China (NSFC)(51702124, 614750671, 61605062); Guangdong Project of Science and Technology Grants (201508010021, 2016B090917002, 2016B090926004); Guangzhou Union Project of Science and Technology Grants (201604040006, 201604040007).

References and links

1. H. M. Pask and J. A. Piper, “Efficient all-solid-state yellow laser source producing 1.2-W average power,” Opt. Lett. 24, 1490–1492 (1999). [CrossRef]  

2. Y. F. Chen and S. W. Tsai, “Diode-pumped Q-switched Nd:YVO4 yellow laser with intracavity sum-frequency mixing,” Opt. Lett. 27, 397–399 (2003). [CrossRef]  

3. A. Kananovich, A. Demidovich, M. Danailov, A. Grabtchikov, and V. Orlovich, “All-solid-state quasi-CW yellow laser with intracavity self-Raman conversion and sum frequency generation,” Laser Phys. Lett. 7, 573 (2010). [CrossRef]  

4. E. Granados, H.M. Pask, and D.J. Spence, “Synchronously pumped continuous-wave mode-locked yellow Raman laser at 559 nm,” Opt. Express. 17, 569–574 (2009). [CrossRef]   [PubMed]  

5. F. G. Yang, Z. Y. You, Z. J. Zhu, Y. Wang, J. F. Li, and C. Y. Tu, “End-pumped continuous-wave intracavity yellow Raman laser at 590 nm with SrWO4 Raman crystal,” Laser Phys. Lett. 7, 14 (2010). [CrossRef]  

6. A. Kaminskii, U. Hommerich, D. Temple, J. T. Seo, K. Ueda, S. Bagayev, and A. Pavlyulk, “Visible laser action of Dy3+ ions in monoclinic KY(WO4)2 and KGd(WO4)2 crystals under Xe-flashlamp pumping,” Jpn. J. Appl. Phys. 2(39), L554 (2000).

7. J. H. Huang, Y. F. Lin, X. H. Gong, Y. J. Chen, Z. D. Luo, and Y. D. Huang, “Spectroscopic properties of Dy3+-doped NaGd(MoO4)2 crystal,” J. Alloy. Compd. 664, 266–271 (2016). [CrossRef]  

8. W. Ryba-Romanowski, G. Dominiak-Dzik, P. Solarz, and R. Lisiecki, “Transition intensities and excited state relaxation dynamics of Dy3+ in crystals and glasses: A comparative study,” Opt. Mater. 31, 1547–1554 (2009). [CrossRef]  

9. S. R. Bowman, S. O. Connor, and N. J. Condon, “Diode pumped yellow dysprosium lasers,” Opt. Express 20, 12906–12911 (2012). [CrossRef]   [PubMed]  

10. G. Bolognesi, D. Parisi, D. Calonico, G.A. Costanzo, F. Levi, P. W. Metz, C. Krankel, G. Huber, and M. Tonelli, “Yellow laser performance of Dy3+ in co-doped Dy, Tb:LiLuF4,” Opt. Lett. 39, 6628–6631 (2014). [CrossRef]   [PubMed]  

11. P. Haro-Gonzalez, L. L. Martin, I. R. Martin, G. G. Dominiak-Dzik, and W. Ryba-Romanowski, “Pump and probe measurements of optical amplification at 584 nm in dysprosium doped lithium niobate crystal,” Opt. Mater. 33, 196–199 (2010). [CrossRef]  

12. D. Parisi, A. Toncelli, M. Tonelli, E. Cavalli, E. Bovero, and A. Belletti, “Optical spectroscopy of BaY2F8:Dy3+,” J. Phys. Condens. Matter. 17, 2783 (2005). [CrossRef]  

13. L. Beauzamy, B. Moine, and P. Gredin, “Energy transfers between dysprosium and terbium in YF3,” J. Lumin. 127, 568 (2007). [CrossRef]  

14. P. W. Metz, D. T. Marzahl, A. Majid, C. Krankel, and G. Huber, “Efficient continuous wave laser operation of Tb3+-doped fluoride crystals in the green and yellow spectral regions,” Laser Photonics Rev. 10(2), 335–344 (2016). [CrossRef]  

15. C. Krankel, D. Marzahl, F. Moglia, G. Huber, and P. W. Metz, “Out of the blue: semiconductor laser pumped visible rare-earth doped lasers,” Laser Photonics Rev. 10(4), 548–568 (2016). [CrossRef]  

16. V. Morozov, A. Arakcheeva, B. Redkin, V. Sinitsyn, S. Khasanov, E. Kudrenko, M. Raskina, O. Lebedev, and G. V. Tendeloo, “Na2/7Gd4/7MoO4: a Modulated Scheelite-Type Structure and Conductivity Properties,” Inorg Chem. 51, 5313–5324 (2012). [CrossRef]   [PubMed]  

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

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

19. C. K. Jorgensen and R. Reisfeld, “Judd-Ofelt parameters and chemical bonding,” J. Less-Common Met. 93(1), 107–112 (1983). [CrossRef]  

20. M. G. Brik, T. Ishii, A. M. Tkachuk, S. E. Ivanova, and I. K. Razumova, “Calculations of the transitions intensities in the optical spectra of Dy3+:LiYF4,” J. Alloy. Compd. 374, 63–68 (2004). [CrossRef]  

21. S. Bigotta, M. Tonelli, E. Cavalli, and A. Belletti, “Optical spectra of Dy3+ in KY3F10 and LiLuF4 crystalline fibers,” J. Lumin. 130, 13–17 (2010). [CrossRef]  

22. A. Lupei, V. Lupei, C. Gheorghe, A. Ikesue, and M. Enculescu, “Spectroscopic characteristics of Dy3+ doped Y3Al5O12 transparent ceramics,” J. Appl. Phys. 110, 083120 (2011). [CrossRef]  

23. X. Xu, Z. Hu, R. Li, D. Li, J. Di, L. Su, Q. Yang, Q. Sai, H. Tang, Q. Wang, A. Strze, and J. Xu, “Optical spectroscopy of Dy3+-doped CaGdAlO4 single crystal for potential use in solid-state yellow lasers,” Opt. Mater. 66, 469–473(2017). [CrossRef]  

24. B. F. Aull and H. P. Jenssen, “Vibronic interactions in Nd:YAG resulting in nonreciprocity of absorption and stimulated emission cross sections,” IEEE J. Quantum Elect. 18(5), 925–930 (1982). [CrossRef]  

Cited By

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

Alert me when this article is cited.


Figures (5)

Fig. 1
Fig. 1 Simplified energy level diagram of Dy3+ and Tb3+ co-doped system. ET1: energy transfer from Dy3+:4F9/2 to Tb3+:5D4 level; ET2: energy transfer from Dy3+:6H13/2 to Tb3+:7F4 level.
Fig. 2
Fig. 2 Absorption spectrum of Dy3+/Tb3+ co-doped NGM crystal in the range of 400–1800 nm.
Fig. 3
Fig. 3 Yellow emission cross-sections of Dy3+ single-doped and Dy3+/Tb3+ co-doped NGM crystals.
Fig. 4
Fig. 4 Fluorescence decay curves of Dy3+ single-doped and Dy3+/Tb3+ co-doped NGM crystals for the 4F9/2 mainfold.
Fig. 5
Fig. 5 Fluorescence decay curves of Dy3+ single-doped and Dy3+/Tb3+ co-doped NGM crystals for the 6H13/2 mainfold.

Tables (1)

Tables Icon

Table 1 Judd-Ofelt parameters Ω2,4,6, branching ratio β, emission cross sections σ, life-time of 4F9/2 level for Dy3+ ions (τM and τR are the measured and calculated radiative lifetime), and quantum efficiency η of different Dy3+ doped materials.

Equations (1)

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

σ e m = A β λ 5 I ( λ ) 8 π c n 2 λ I ( λ ) d λ
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.