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

Activation and deactivation effects to Ho3+ at ~2.8 μm MIR emission by Yb3+ and Pr3+ ions in YAG crystal

Open Access Open Access

Abstract

The use of Yb3+ and Pr3+ co-doping with Ho3+ to enhance the Ho3+:5I65I7 ~2.8μm emissions are investigated in the YAG crystal for the first time. Absorption spectra of the Yb:Ho:YAG and Yb:Ho:Pr:YAG crystals find that doping of Yb3+ increase absorption intensity within 850-1050nm and make the crystals suitable to high-power InGaAs LDs. Fluorescence spectra show that ~2.8μm lasers may be realized on both crystals at different wavelengths, with the maximum emission cross section of 1.69 × 10−20 cm2 at 2780 nm for Yb:Ho:Pr:YAG and 1.23 × 10−20 cm2 at 2848 nm for Yb:Ho:YAG. Lifetime results and energy transfer analyses indicate that activator Yb3+ ions can prolong the higher laser level of Ho3+: 5I6 and deactivator Pr3+ ions are able to depopulate the lower laser level of Ho3+: 5I7, both beneficial to population inversion and contributed to facilitate ~2.8μm laser operations. All these results indicate that Yb:Ho:Pr:YAG crystal is a promising laser medium for LD-pumped ~2.8μm laser applications.

© 2016 Optical Society of America

1. Introduction

In recent years, ~2.8μm mid-IR (MIR) lasers have attracted much attention for their applications in medical, biological as well as sensing technologies due to the strong vibrational absorption band of water in this spectral region [1–3]. Furthermore, ~2.8μm lasers can also act as efficient pumping sources for longer wavelength MIR lasers and optical parametric oscillators (OPO) [4].

Ho3+ is a well-known candidate for ~2.8μm lasers with the transition of 5I65I7, which has already been discussed and demonstrated in several hosts [5–7]. However, efficient ~2.8μm laser operations of Ho3+-singly-doped hosts have been limited by two bottle-necks. One is that the intrinsic absorption of Ho3+ can’t match an efficient, currently well-developed laser diodes (LDs) or flash lamp, and the other is that the fluorescence lifetime of the lower 5I7 level is considerably longer than that of the upper 5I6 level. To overcome these deficiencies, on one hand, Yb3+ has been considered favorable to improve the absorption efficiencies of Ho3+ with commercially available 940 or 970 nm pumping sources, which has been demonstrated in Yb:Ho:YSGG, Yb:Ho:GGG, etc [8–10]. On the other hand, Pr3+ has been explored to be feasible to quench the lower level of Ho3+: 5I7, while hardly depopulate the upper level of Ho3+: 5I6, as good results of ~2.8μm emissions and lasers achieved in studies of Ho,Pr-doped floride fiber [11] and Ho:Pr:LiLuF4 crystal [12]. Therefore, it is believed that co-doping of both Yb3+ and Pr3+ with Ho3+ could open up the possibility of efficient, high-power ∼2.8μm lasing from Ho3+ by suitable high-power diode pumping, which is recently studied in Yb:Ho:Pr:YAP [13].

Host material is another factor that should be considered to get powerful ~2.8μm emissions from Ho3+. Among the various crystals, YAG is one of the most popular laser crystals due to its excellent optical, thermal, mechanical as well as physicochemical properties. Optical characterization and laser operations focused on ~2μm emissions have been studied in Ho:YAG and Yb:Ho:YAG [14–16], but to our knowledge, there is still no report on the growth of Yb:Ho:Pr:YAG crystal or the ~2.8μm emission in this crystal up to now. In this letter, the Yb3+, Ho3+-doped YAG and Yb3+, Ho3+, Pr3+-doped YAG laser crystals were successfully grown. Yb3+ and Pr3+ were demonstrated to greatly enhance the Ho3+: 5I65I7 ~2.8μm emission by efficient energy transfer from Yb3+: 2F5/2 to Ho3+: 5I6 and Ho3+: 5I7 to Pr3+: 3F2, respectively. The absorption spectra, emission spectra, fluorescence decay curve and transfer mechanisms among the ions were investigated for future applications in MIR lasers.

2. Experiments

The 15 at.% Yb3+/ 2.5 at.% Ho3+-co-doped YAG and 15 at.% Yb3+/ 2.5 at.% Ho3+/ 1 at.% Pr3+-co-doped YAG were grown along the <111> crystal orientation by the Czochralski method with an intermediate frequency induction heating system. Oxide powders of Yb2O3 (5N), Ho2O3 (5N), Pr6O11 (5N), Y2O3 (5N) and Al2O3 (4N) were used as starting materials, which were mixed adequately for 12h and pressed into disks, followed by heating in air for 10h at 1200 °C. Then, the polycrystalline materials were loaded into an Iridium crucible of Φ60 mm for crystal growth in an atmosphere of Ar, which could effectively protect the Ir crucible from oxidization, with a pulling rate of 1.1 mm/h and a rotation speed of 10 rpm. After growth, the crystals were annealed at 1200 °C in air for 20h to oxidize some Yb2+ ions, which were formed during growth and made the crystal green, into Yb3+. Samples with dimension of 10 × 10 × 3 mm3 were cut from the annealed crystals and then polished on both faces for spectroscopic measurements. The concentrations of Yb, Ho and Pr ions were detected by inductively coupled plasma-atomic emission spectrometry (ICP-AES) analysis. The absorption spectra in the range of 300nm-2250nm were recorded by Perkin Elmer Lambda 900. The fluorescence spectra in range of 2700 nm to 3000 nm were measured under 915 nm LD pumping. The fluorescence decay curves were recorded under excitation at 915nm by OPO pulse lasers. All the measurements were taken at room temperature and at the same condition for two crystals.

3. Results and discussions

The doping concentrations in the top of both crystals were measured to be Yb of 15.5at.% and Ho of 2.06at.% for the Yb:Ho:YAG crystal, and Yb of 15.6at.%, Ho of 2.04at.% and Pr of 0.18at. % for the Yb:Ho:Pr:YAG crystal, respectively. Therefor the segregation coefficient of Yb, Ho and Pr in the YAG crystal are approximate to 1.04, 0.82 and 0.18.

The room-temperature absorption spectra of Yb:Ho:YAG and Yb:Ho:Pr:YAG crystals were shown as absorption coefficient in Fig. 1. The absorption peaks of these two crystals are very similar, except the bands around 600nm and in the range of 1340-1570nm, which correspond to the transitions of Pr3+: 3H41D2 and Pr3+: 3H43F3 + 3F4, respectively. The broad and strong absorption bands within 850-1050nm centered at 940nm correspond to the transition of Yb3+: 2F7∕22F5∕2, which make both crystals suitable to be pumped by commercially available 940 high-power InGaAs LDs. The absorption cross-sections of Yb3+ ions can be calculated from

σabs(λ)=OD(λ)L×N0×loge,
where OD(λ) is the measured absorption optical density as a function of wavelength, Lis the thickness of sample and N0 is the number of Yb3+ ions per cm3. The results are shown in the right inset of Fig. 1, and the maximum absorption cross-sections are 4.93 × 10−21cm2 and 5.68 × 10−21cm2 at 936 nm for the Yb:Ho:YAG and Yb:Ho:Pr:YAG, respectively. It is no accident that the peak absorption cross-section of Yb:Ho:Pr:YAG is a little greater than that of Yb:Ho:YAG, because the testing sample of Yb:Ho:Pr:YAG is cut from an upper position of the as-grown crystal than Yb:Ho:YAG, as shown in the left inset of Fig. 1. Since the segregation coefficient of Yb is a little greater than 1, the Yb concertration of the test sample of Yb:Ho:Pr:YAG can be a little larger than that of Yb:Ho:YAG.

 figure: Fig. 1

Fig. 1 Absorption spectra of the Yb:Ho:YAG and Yb:Ho:Pr:YAG crystals.

Download Full Size | PDF

The absorption spectra were used to calculate the Judd–Ofelt [17,18] intensity parameters Ω2;4;6 of Ho3+. The obtained Ω2;4;6 values are shown in Table 1 with comparison of Ho:YAG. As Ω2 reflects the asymmetry and high-covalence of ligands, it is reasonable that introduction of Yb3+ and Pr3+ ions leads to a higher Ω2 value, which indicates a lower symmetry surrounding Ho3+ ions [12,19]. With the Ω2;4;6, several optical parameters, such as the fluorescence branching ratio β and the radiative lifetime τr are calculated and listed in Table 1 for further calculation. It can be seen that the fluorescence branching ratio β of transition 5I65I7 for ~2.8μm emission in Yb:Ho:YAG and Yb:Ho:Pr:YAG are approximately 17%, which is higher than that in Ho:YAG. Furthermore, the radiative lifetime τr of 5I6 manifold in Yb:Ho:YAG and Yb:Ho:Pr:YAG are approximately 4.3ms, which is longer than that in Ho:YAG. The higher fluorescence branching ratio and longer radiative lifetime both indicate that Yb:Ho:YAG and Yb:Ho:Pr:YAG may more easily induce the ~2.8μm emissions. It should be noticed that, the J-O theory does not take into account non-radiative decay, so the τr[5I7] of Yb:Ho:YAG and Yb:Ho:Pr:YAG are quite similar in calculation.

Tables Icon

Table 1. Judd–Ofelt parameters, fluorescence branching ratio and lifetimes of Ho-doped materials.

The fluorescence spectra of the Yb:Ho:YAG and Yb:Ho:Pr:YAG crystals are shown in Fig. 2. For both crystals, there are three main emission bands around 2780, 2848, and 2944 nm assigned to the Ho3+: 5I65I7. It is evident that the emission intensity around 2780nm of Yb:Ho:Pr:YAG is quite strong, which is nearly 2.8 times that of Yb:Ho:YAG. Although the emission intensity around 2848 of Yb:Ho: Pr:YAG is a little lower than that of Yb:Ho:YAG, the emission intensities around 2944nm of both crystals are comparable. In addition, the left inset of Fig. 2 shows that the ~2μm emission (Ho3+: 5I75I8) of Yb:Ho:Pr:YAG is much weaker than that of Yb:Ho:YAG and the 2780nm emission of Yb:Ho:Pr:YAG, which indicates that co-doping Pr weaken the ~2μm emission and benefit the ~2.8μm emission. Thus, output of ~2.78μm lasers may be realized on Yb:Ho:Pr:YAG.

 figure: Fig. 2

Fig. 2 Fluorescence spectra of the Yb:Ho:YAG and Yb:Ho:Pr:YAG crystals.

Download Full Size | PDF

The emission cross sections can be subsequently calculated by the Fuchtbauer–Ladenburg equation [21]

σem(λ)=βλ5I(λ)8πcn2τrλI(λ)dλ,
where I(λ) is the measured fluorescence intensity at wavelength λ, β is the fluorescence branching ratio as β[5I65I7] in Table 1, is the speed of light, n is the crystal refractive index and τr is the radiative lifetime as τr[5I6] in Table 1. The results are shown as the right inset of Fig. 2, and the maximum emission cross section of Yb:Ho:Pr:YAG crystal is estimated to be 1.69 × 10−20 cm2 at 2780 nm, and that of the Yb:Ho:YAG is 1.23 × 10−20 cm2 at 2848 nm.

The fluorescence decay curves of Ho3+: 5I6 and 5I7 multiplets for the Yb:Ho:YAG and Yb:Ho:Pr:YAG crystals are shown in Fig. 3. The four experiment curves all exhibit a single exponential decaying behavior, and the fitted values are shown in Table 1 as τmeas. On one hand, lifetime of 5I6 manifold in Yb:Ho:Pr:YAG is 0.106ms, which is 5% shorter compared with that of 0.112ms in Yb:Ho:YAG, indicating that co-doping of Pr3+ has little influence on the higher laser level of Ho3+: 5I6. On the other hand, lifetime of 5I7 manifold in Yb:Ho:Pr:YAG is 1.83ms, which is 83% shorter when compared with that of 10.63ms in Yb:Ho:YAG. This remarkable decrease confirms that Pr3+ ions are able to depopulate the lower laser level of Ho3+: 5I7 for ~2.8μm emission in YAG crystal, which may induce the population inversion and facilitate laser operations. In addition, comparing the τmeas of Ho:YAG of [20] and the Yb:Ho:YAG of this study, it can tell that Yb3+ may efficiently populate the Ho3+: 5I6 while rarely influence the Ho3+: 5I7. Lifetime of 5I6 in Yb:Ho:YAG is 149% longer than that in Ho:YAG while lifetime of 5I7 in Yb:Ho:YAG is only 52% longer than that in Ho:YAG. It indicates that due to ET process of Yb3+: 2F5∕2 → Ho3+: 5I6 as marked in Fig. 2, Yb3+ ions can prolong the higher laser level of Ho3+: 5I6 and contribute to the ~2.8μm emission in YAG crystal.

 figure: Fig. 3

Fig. 3 Fluorescence decay curves of the Yb:Ho:YAG and Yb:Ho:Pr:YAG crystals for the Ho: 5I6 and 5I7 mainfolds.

Download Full Size | PDF

To further investigate the Yb3+-Ho3+ and Ho3+-Pr3+ energy interaction mechanism, a simplified diagram of energy-level [22] and energy-transfer of Yb3+, Ho3+ and Pr3+ co-doped system is shown in Fig. 4. As the energy gap between Yb3+: 2F5∕2 (~10000 cm−1) and Ho3+: 5I6 (~8500 cm−1) is small, efficient ET process from Yb3+: 2F5∕2 to Ho3+: 5I6 is expected to occur with the phonon energy of the host bridging this energy difference (∼1500 cm−1), which populates the Ho3+: 5I6 level. On the other hand, it is evident that ions in the Ho3+: 5I7 level will undergo ET process to the Pr3+: 3F2 level due to rather small energy gap (∼30 cm−1) between these two energy levels, which depopulates the Ho3+: 5I7 level. Both ET processes make population inversion for Ho3+: 5I65I7 possible and enhance the MIR ~2.8 μm emission. Moreover, efficiency of the energy transfer within Ho3+ and Pr3+ ions can be estimated by: η=1τYb/Ho/PrτYb/Ho, where τYb/Ho/Pr and τYb/Ho are the lifetimes of Ho3+: 5I7 for the Yb:Ho:Pr:YAG and Yb:Ho:YAG crystals, respectively. Consequently, the value of η is calculated to be 82.8%, indicating that Pr3+ ions can efficiently quench the excited-state population of Ho3+: 5I7 by ET process and enhance the ~2.8 μm emission.

 figure: Fig. 4

Fig. 4 Schematic of energy-level and energy-transfer of Yb3+, Ho3+ and Pr3+ co-doped system.

Download Full Size | PDF

4. Conclusion

In conclusion, high-quality Yb:Ho:YAG and Yb:Ho:Pr:YAG crystals were successfully grown using Czochralski method. Absorption spectra show that co-doping with Yb enhances the absorption intensity within 850-1050nm and makes both crystals suitable to be pumped by commercially available high-power InGaAs LDs. Fluorescence spectra exhibit that output of ~2.8μm lasers may be realized on both crystals at different wavelengths with the maximum emission cross section of 1.69 × 10−20 cm2 at 2780 nm for Yb:Ho:Pr:YAG and 1.23 × 10−20 cm2 at 2848 nm for Yb:Ho:YAG. Lifetime results and energy transfer analyses indicate that Yb3+ ions can prolong the higher laser level of Ho3+: 5I6 from 0.045ms of Ho:YAG to 0.112ms of Yb:Ho:YAG and Pr3+ ions are able to depopulate the lower laser level of Ho3+: 5I7 from 10.63ms of Yb:Ho:YAG to 1.83ms of Pr:Yb:Ho:YAG, both beneficial to population inversion and contributed to facilitate Ho3+: 5I65I7 ~2.8μm laser operations. It is suggested that the Yb:Ho:Pr:YAG is a favorable LD pumping active laser media for ~2.8μm solid-state lasers.

Acknowledgments

This work has been supported by the National Natural Science Foundation of China (Grant No.51302283 and Grant No. 51502321), and the Shanghai Natural Science Foundation under Projects (No.13ZR1463400).

References and links

1. F. Moser, D. Bunimovich, A. DeRowe, O. Eyal, A. German, Y. Gotshal, A. Levite, L. Nagli, A. Ravid, V. Scharf, S. Shalem, D. Shemesh, R. Simchi, I. Vasserman, and A. Katzir, “Medical Applications of Infrared Transmitting Silver Halide Fibers,” IEEE J. Sel. Top Quant. 2(4), 872–879 (1996). [CrossRef]  

2. J. S. Sanghera, L. Brandon Shaw, and I. D. Aggarwal, “Chalcogenide Glass-Fiber-Based Mid-IR Sources and Applications,” IEEE J. Sel. Top Quant. 15(1), 114–119 (2009). [CrossRef]  

3. M. Tacke, “New Developments and Applications of Tunable IR Lead Salt Lasers,” Infrared Phys. Technol. 36(1), 447–463 (1995). [CrossRef]  

4. K. L. Vodopyanov, F. Ganikhanov, J. P. Maffetone, I. Zwieback, and W. Ruderman, “ZnGeP2 optical parametric oscillator with 3.8-12.4 µm tunability,” Opt. Lett. 25(11), 841–843 (2000). [CrossRef]   [PubMed]  

5. W. S. Rabinovich, S. R. Bowman, B. J. Feldman, and M. J. Winings, “Tunable Laser Pumped 3 μm Ho:YA1O3 Laser,” IEEE J. Quantum Electron. 27(4), 895–897 (1991). [CrossRef]  

6. S. A. Payne, L. K. Smith, and W. F. Krupke, “Cross sections and quantum yields of the 3 μm emission for Er3+ and Ho3+ dopants in crystals,” J. Appl. Phys. 77(9), 4274–4279 (1995). [CrossRef]  

7. P. Zhang, J. Yin, B. Zhang, L. Zhang, J. Hong, J. He, and Y. Hang, “Intense 2.8 μm emission of Ho3+ doped PbF2single crystal,” Opt. Lett. 39(13), 3942–3945 (2014). [CrossRef]   [PubMed]  

8. A. Diening and S. Kück, “Spectroscopy and diode-pumped laser oscillation of Yb3+, Ho3+ -doped yttrium scandium gallium garnet,” J. Appl. Phys. 87(9), 4063–4068 (2000). [CrossRef]  

9. H. Chen, F. Chen, T. Wei, Q. Liu, R. Shen, and Y. Tian, “Ho3+ doped fluorophosphate glasses sensitized by Yb3+ for efficient 2 μm laser applications,” Opt. Commun. 321, 183–188 (2014). [CrossRef]  

10. Y. Wang, J. Li, Z. Zhu, Z. You, J. Xu, and C. Tu, “Activation effect of Ho3+ at 2.84 μm MIR luminescence by Yb3+ ions in GGG crystal,” Opt. Lett. 38(20), 3988–3990 (2013). [CrossRef]   [PubMed]  

11. T. Hu, D. D. Hudson, and S. D. Jackson, “Actively Q-switched 2.9 μm Ho3+Pr3+-doped fluoride fiber laser,” Opt. Lett. 37(11), 2145–2147 (2012). [CrossRef]   [PubMed]  

12. P. Zhang, Y. Hang, and L. Zhang, “Deactivation effects of the lowest excited state of Ho3+ at 2.9 μm emission introduced by Pr3+ ions in LiLuF4 crystal,” Opt. Lett. 37(24), 5241–5243 (2012). [CrossRef]   [PubMed]  

13. H. Zhang, D. Sun, J. Luo, S. Cao, M. Cheng, Q. Zhang, and S. Yin, “Growth and spectroscopic investigations of Yb,Ho: YAP and Yb,Ho,Pr:YAP laser crystals,” J. Lumin. 158, 215–219 (2015). [CrossRef]  

14. T. Zhao, F. Wang, and D. Y. Shen, “High-power Ho:YAG laser wing-pumped by a Tm:fiber laser at 1933 nm,” Appl. Opt. 54(7), 1594–1597 (2015). [CrossRef]  

15. Th. Rothacher, W. Luthy, and H. P. Weber, “Diode pumping and laser properties of Yb:Ho:YAG,” Opt. Commun. 155(1), 68–72 (1998). [CrossRef]  

16. A. A. Nikitichev and V. A. Pis’mennyi, “1 W CW 2.12 μm Lamp Pumped Room Temperature YAG:Yb-Ho Laser,” Proc. SPIE 2772, 35–36 (1996). [CrossRef]  

17. G. S. Ofelt, “Intensities of Crystal Spectra of RareEarth Ions,” J. Chem. Phys. 37(3), 511–520 (1962). [CrossRef]  

18. B. R. Judd, “Optical Absorption Intensities of Rare-Earth Ions,” Phys. Rev. 127(3), 750–761 (1962). [CrossRef]  

19. E. Preda, M. Stef, G. Buse, A. Pruna, and I. Nicoara, “Concentration dependence of the Judd–Ofelt parameters of Er3+ ions in CaF2 crystals,” Phys. Scr. 79(3), 035304 (2009). [CrossRef]  

20. B. M. Walsh, G. W. Grew, and N. P. Barnes, “Energy levels and intensity parameters of Ho3+ ions in Y3Al5O12 and Lu3Al5O12,” J. Phys. Chem. Solids 67(7), 1567–1582 (2006). [CrossRef]  

21. 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 Electron. 18(5), 925–930 (1982). [CrossRef]  

22. G. H. Dieke, Spectra and Energy Levels of Rare Earth Ions in Crystals (Interscience Publishers, 1968).

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 (4)

Fig. 1
Fig. 1 Absorption spectra of the Yb:Ho:YAG and Yb:Ho:Pr:YAG crystals.
Fig. 2
Fig. 2 Fluorescence spectra of the Yb:Ho:YAG and Yb:Ho:Pr:YAG crystals.
Fig. 3
Fig. 3 Fluorescence decay curves of the Yb:Ho:YAG and Yb:Ho:Pr:YAG crystals for the Ho: 5I6 and 5I7 mainfolds.
Fig. 4
Fig. 4 Schematic of energy-level and energy-transfer of Yb3+, Ho3+ and Pr3+ co-doped system.

Tables (1)

Tables Icon

Table 1 Judd–Ofelt parameters, fluorescence branching ratio and lifetimes of Ho-doped materials.

Equations (2)

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

σ a b s ( λ ) = O D ( λ ) L × N 0 × log e ,
σ e m ( λ ) = β λ 5 I ( λ ) 8 π c n 2 τ r 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.