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

Enhanced 2.7 μm mid-infrared emissions of Er3+ via Pr3+ deactivation and Yb3+ sensitization in LiNbO3 crystal

Open Access Open Access

Abstract

The use of Pr3+ codoping for enhancement of the transition of Er3+: 4I11/24I13/2 2.7 μm emissions was investigated in the Er/Yb codoped LiNbO3 crystal for the first time. It is found that the codoped of Pr3+ ion in Er3+, Yb3+ and Pr3+ triply doped LiNbO3 crystal (Er/Yb/Pr: LN) greatly enhances Er3+: 2.7 μm emission under excitation of a common 970 nm laser diode, depopulates the lower laser level of Er3+:4I13/2, and has little influence on the higher laser level of Er3+:4I11/2 at the same time for population inversion. The 2.7 μm emission characteristics and energy transfer were investigated in detail. The energy transition efficiency from lower laser level of Er3+:4I13/2 to Pr3+:3F4 level is as high as 0.42, indicating that the Pr3+ ion is an effective deactivation ion for Er3+ ion in LiNbO3 crystal. These results suggest that Er/Yb/Pr: LiNbO3 crystal may become an attractive host for developing solid state lasers at around 2.7 μm under a conventional 970 nm LD pump.

© 2016 Optical Society of America

1. Introduction

Over the past several decades, mid-IR (MIR) lasers operating around 2.7–3 μm have received extensive attention for numerous applications in medicine, remote sensing, military countermeasures, and atmosphere pollution monitoring due to the strong water absorption and short penetration depths of a few micrometers into biological tissue around this spectral region [1–5]. In addition, 2.7–3 μm lasers can be used as efficient and high-quality pump sources for longer wavelength mid-IR lasers and optical parametric oscillators [6–8].

Erbium ion is an ideal luminescent center for 2.7 μm emission with the transition of 4I11/24I13/2, and Er3+ doped garnets [9–12] and fluorides [13] crystalline materials have been widely studied and demonstrated as 2.7 μm laser sources. However, the 2.7 μm laser operation cannot be obtained efficiently due to (i) the absorption coefficient of Er3+ ion around 970 nm is significantly low, limiting the pump efficiency, and (ii) the self-terminating “bottleneck” effect, which may result from the fluorescence lifetime of the upper 4I11/2 level is considerably shorter than that of the lower level 4I13/2 [14]. In order to conquer these problems, on one hand, because the strong overlap of the luminescence spectrum of Yb3+ ions and the absorption spectrum of acceptor Er3+ ions, Yb3+ has can be as an efficient sensitizer ion to transfer the accumulated excitation energy to the Er3+ ions to improve the absorption efficiencies with commercially available 970 nm pumping sources [15–17]. On the other hand, high doping concentrations (>30 at. %) of Er3+ can mitigate the self-termination bottleneck by well known concentration dependent up-conversion process [4I13/2+4I13/24I11/2+4I9/2]. However, high doping concentration of Er3+ may give rise to a decline in the quality of crystal, limiting the laser output efficiency and beam quality. Another method of mitigating the self-termination bottleneck is the co-doping of deactivation ion, which can efficiently quench the lower level 4I13/2 of Er3+, leading to population inversion. Up to now, with the codoped of deactivation ion (such as Pr3+, Ho3+, Tm3+, etc.), efficient “quenchers” of luminescence from the level 4I13/2 have been achieved in luminescent materials [18–20]. We are now focusing our scientific program on Pr3+ ion, because its energy level 4F4 is adjacent to level 4I13/2 of Er3+. Predictably, it is believed that co-doping of both Yb3+ and Pr3+ with Er3+ without heavily doping of Er3+ may turn on the possibility of excellent 2.7 μm lasers performance from Er3+ under a conventional 970 nm LD pump.

Figure 1 shows the simplified energy level diagram of Er3+, Yb3+ and Pr3+ triply doped system. When the crystal is pumped by a 970 nm laser diode, ions of Yb3+: 2F7/2 state is excited to Yb3+:2F5/2. Then transferring the energy to the Er3+: 4I11/2 level thanks to the effective resonance energy transfer from Yb3+ to Er3+. Then, on one hand, some Er3+ ions in 4I11/2 level would decay radiatively to the ground state 4I15/2 with 1.0 μm emission, on the other hand, other ions in 4I11/2 level would decay to 4I13/2 with 2.7 μm emission, or nonradiatively to 4I13/2 without emission. Subsequently, ions in the 4I13/2 level would undergo energy transition (ET) process to the 3F4 level of Pr3+, depopulating the Er3+: 4I13/2 level, which may induce the population inversion and facilitate laser operation.

 figure: Fig. 1

Fig. 1 Simplified energy level diagram of Er3+, Yb3+ and Pr3+ triply doped system.

Download Full Size | PDF

The well-known versatile LiNbO3 (LN) crystal has been widely used in optoelectronic devices because of its electro-optic, piezo-electric, and nonlinear optical physical properties [21–23]. LN crystal has been extensively doped with rare-earth ions, in which laser oscillation, electro-optic and nonlinear optics effects can be integrated within just one crystal [24–26]. Er3+ doped LN crystal has been widely investigated for it operates at the eye-safe 1.54 μm laser, and the miniaturization of solar cells [27, 28]. However, to our knowledge, the observation of emission around 2.7 μm in Er3+, Yb3+, and Pr3+ triply doped LN crystal has not been successfully obtained. In this letter, Er3+, Yb3+, and Pr3+ triply doped LN crystal (Er/Yb/Pr: LN) was successfully prepared. Yb3+ ion was demonstrated to be an effective sensitizer for Er3+ ion in Er/Yb/Pr: LN crystal. Pr3+ was demonstrated to enhance the Er3+: 4I11/24I13/2 2.7 μm emission in Er/Yb/Pr: LN crystal by depopulating the Er3+: 4I13/2 level while having little influence on the Er3+: 4I11/2 level. The spectroscopic investigation of 2.7 μm emission has also been made for future applications in MIR lasers.

2. Experimental section

The 0.6 at.% Er3+, 1.0 at.% Yb3+, and 0.3 at.% Pr3+ triply doped and 0.6 at.% Er3+, 1.0 at.% Yb3+ codoped LN (Er/Yb: LN) crystals were grown by the Czochralski method with intermediate frequency induction heating. Li2CO3 (99.999%), Nb2O5 (99.999%), Yb2O3 (99.999%), Er2O3 (99.999%), and Pr2O3 (99.999%) were used as raw material for crystal growth. the mixture was pressed into disks and heated in air at 1150 °C for 12 h to remove CO2 and form polycrystalline powers. Then, the polycrystalline powers were followed loaded into platinum crucible for crystal growth along the ferroelectric c axis. The pulling rate and rotation rate were 0.6–0.8 mm/h, and 8–12 rpm, respectively. Before being cut into wafers for experiment, the crystals were polarized at 1200 °C with a current density of 6–8 mA/cm2 to produce a single-domain crystal.

The inductively coupled plasma-atomic emission spectrometry (ICP-AES) was used to measure the concentrations of Er3+, Yb3+, and Pr3+ ions in the as grown crystals. The concentrations of Er3+ and Yb3+ in double doped crystal were 0.75 at.%, and 1.28 at.%, respectively. The corresponding segregator coefficients of Er3+ and Yb3+ of Er/Yb: LN crystal were 1.25, and 1.28, respectively. The concentrations of Er3+, Yb3+, and Pr3+ in triply doped crystal were 0.73 at.%, 1.26 at.%, and 0.27 at.%, respectively. The corresponding segregator coefficients of Er3+, Yb3+, and Pr3+ were 1.21, 1.26, and 0.91, respectively. It is clear to see that although the concentration of Er3+ ions in the melt is constant (0.6 at.%), the concentration of Er3+ ions in the crystal decreases (from 0.75 at.% to 0.73 at.%) when the total impurity concentration increases (after the codoped of Pr3+ ions), which is consistent with the results of Er3+/Yb3+ codoped LN crystal published by E. Cantelar et al. [29].

The analysis of Er3+, Yb3+, and Pr3+ ion absorption was carried out by a UV-VIS-NIR spectrophotometer (ModelV-570, JASCO Co.). The fluorescence spectra in range of 2700 nm to 3000 nm and fluorescence decay profiles of the double-doped and triply-doped crystals were acquired by Edinburgh Instruments FLS920 and FSP920 spectrophotometers with a laser diode (LD) as the pump source (excited at 970 nm), and an optical parametric oscillator pulse laser. All measurements were done at room temperature.

In order to reduce an artificial lifetime lengthening due to radiation trapping [30], the following test technologies have been taken: (i) the thickness of crystals for experiment measuring was as thin as 1.0 mm; (ii) the excited laser beam was focused near the edge of crystals; (iii) the measurements were performed by the pin-hole method [31], the crystals were placed behind a pin-hold with a variable aperture from 0.5 mm to 1.5 mm at an angle of 45 °C with respect to the excited laser beam and the axis of the collection optics.

3. Experimental results and discussion

The visible-near-IR absorption spectra of Er/Yb/Pr: LN crystal was shown in Fig. 2. It is clear to see that in the range of 400–2100 nm, there are 10 typical absorption bands of Er3+ ion, centered at approximately 380, 406, 454, 492, 524, 548, 658, 804, 978, and 1528 nm, which correspond to the transitions from Er3+: 4I15/2 to 4G11/2, 2H9/2, 4F5/2+4F3/2, 4F7/2, 2H11/2, 4S3/2, 4F9/2, 4I9/2, 4I11/2, and 4I13/2, respectively.

 figure: Fig. 2

Fig. 2 Absorption spectrum of Er/Yb/Pr: LN crystal in the range of 400–2100 nm. The inset shows absorption spectrum of Er: LN, Er/Yb: LN, and Er/Yb/Pr: LN crystals in the range of 850–1700 nm.

Download Full Size | PDF

On one side, as L. E. Bausa [32] and E. Montoya [33] have pointed out that there is a wide and strong absorption band corresponds to the Yb3+ optical transition 2F7/22F5/2 centered at 970 nm in LN crystal. It is clear to see from the inset of Fig. 2, in the range of 850–1050 nm, there is a strong overlap between the transition of Yb3+ ions and Er3+ ions. By contrast, the absorption of the crystal without Yb3+ ion (Er: LN) is significantly weak. Therefore, Yb3+ ion can act as a sensitizer to Er3+ ion by nonradiative energy transfer, resulting in an increased absorption efficiency, which is beneficial to effectively enhance the pump efficiency.

On the other side, I. Baumann [34] has reported that in the range of 1310–1700 nm, there is an absorption band due to the transition of Pr3+: 3H43F4 in Ti/Pr: LN waveguide. As the inset of Fig. 2 shows, in the range of 1400–1700 nm, a strong overlap between Er3+: 4I15/24I13/2 transition and Pr3+: 3H43F4 transition can be seen. It is expected that efficient energy transfer will occur from the lower 4I13/2 level of Er3+ to the 3F4 states of Pr3+ in the Er/Yb/Pr: LN crystal, leading in depopulating the Er3+: 4I13/2 level.

L. Nunez et al. [35] have made a systematic study on the polarized optical absorption of Er3+ doped LN crystal, and they found that compared with the unpolarized theory, the Judd-Ofelt parameters Ω4 and Ω6 were in good agreement, while the Ω2 was different. Therefore, it is best to apply a fully polarized Judd-Ofelt treatment for optical performance characterization. However, in some cases, it is supposed that a unpolarized absorption spectra approximates to some extent an average of π and σ polarizations and may be used within in the framework of the unpolarized Judd-Ofelt theory [36], and limited to the existing test-bed condition, here we only made a fully unpolarized Judd-Ofelt treatment for optical performance characterization. In order to separate the Pr3+ and Yb3+ ions absorption peaks, only the following 8 absorption bands centered at approximately 380, 406, 454, 492, 524, 548, 658, and 804 nm, which correspond to the transitions from Er3+: 4I15/2 to 4G11/2, 2H9/2, 4F5/2+4F3/2, 4F7/2, 2H11/2, 4S3/2, 4F9/2, and 4I9/2, respectively, were used to calculated the intensity parameters Ω2,4,6 of Er3+ based on the Judd-Ofelt theory [37,38]. Moreover, the absorption coefficient for measured concentrations was corrected. The intensity parameters Ω2,4,6 of Er3+ are shown in Table 1. To justify our fitting quality, the root-mean-square (RMS) deviation between the experimental and the calculated line strengths was calculated to be as low as 0.08×10−20 cm2, which confirms the creditability of calculations and the validity of the J-O theory for predicting the spectral intensities of Er3+. It is clear to see that the Ω2 of Er3+ in Er/Yb/Pr: LN crystal is higher than that of Er/Yb: LN crystal. It is well known that the value of Ω2 drops with the improved symmetry of rare-earth ions site [39]. The larger Ω2 of Er3+ in Er/Yb/Pr: LN crystal indicates that the introduction of Pr3+ ions leads to a lower symmetry surrounding Er3+ ions. The fluorescence branching ratio (β) and the radiative lifetime (τr) were calculated by using the intensity parameters Ω2,4,6 of Er3+ in our crystal and are also shown in Table 1. It is clear to see that the fluorescence branching ratio β of 4I11/24I13/2 transition for 2.7 μm fluorescence emission in Er/Yb/Pr: LN crystal is as high as 16.32%, which is higher than that (12.21%) of Er/Yb: LN crystal, indicating that the codoping of Pr3+ ions makes the Er/Yb: LN crystal more easily induces the 2.7 μm fluorescence emission and facilitates laser operation.

Tables Icon

Table 1. Judd-Ofelt parameters Ω2,4,6, branching ratio β, lifetime of 4I11/2 and 4I13/2 levels for Er3+ ions (τm and τr are the measured and calculated radiative lifetime) of Er/Yb/Pr: LN and Er/Yb: LN crystals.

In order to understand the energy transfer between Yb3+ and Er3+ ions, the decay curve of the Yb3+:2F5/2 energy level for the luminescence at 1060 nm of Er/Yb/Pr: LN crystal was measured and shows in Fig. 3. The inset of Fig. 3 shows the Ln of fluorescence intensity I versus time t, which is unable to be linear fitting, indicating that the decay curve can not be fitted to single exponential function. It can be well fitted to dual exponential decaying, and the fitted curve is also shown in Fig. 3. The measured lifetime of the Yb3+:2F5/2 manifold in the Er/Yb/Pr: LN crystal is 320 μs, which is much shorter than that of Er: LN crystal (580 μs) [40]. This reduced lifetime indicates that an additional relaxation channel (energy transfer from Yb3+ to Er3+) has been activated [41]. The energy transition efficiency from Yb3+ to Er3+ can be estimated by the following equation: ηET1=1−τEr/Yb/Pr@1060nm/τYb@1060nm, where τEr/Yb/Pr@1060nm and τYb@1060nm are the Yb3+ lifetimes of Er/Yb/Pr: LN crystal, and Yb: LN crystal, respectively. The calculated energy transfer efficiency ηET1 from Yb3+ to Er3+ is as high as 44.8 %. This confirms that Yb3+ ion is a suitable sensitizer in Er/Yb/Pr: LN crystal and can efficiently transfer energy to Er3+ ion, making this crystal propitious to be pumped by commercialized InGaAs LD. It is worth mentioning that based on the results of Er3+/Yb3+ codoped LN crystal published by E. Cantelar [41], the energy transfer efficiency ηET1 can be further improved by reducing the concentration of Yb3+ in Er/Yb/Pr: LN crystal.

 figure: Fig. 3

Fig. 3 Fluorescence decay curves of the Yb3+:2F5/2 energy level of Er/Yb/Pr: LN crystal. The inset show the Ln of fluorescence intensity I versus time t.

Download Full Size | PDF

Figure 4 shows the emission spectra of Er/Yb/Pr: LN and Er/Yb: LN crystals in the range of 1400–1700 nm, and 2500–3000 nm. It can be see that the emission intensity of the Er/Yb/Pr: LN crystal is at least two times that of the Er/Yb: LN crystal at 2.7 μm, whereas emission at 1.5 μm is reverse. This enhanced 2.7 μm emission and reduced 1.5 μm emission in the Er/Yb/Pr: LN crystal is mainly attributed to the quenching effect of the Er3+: 4I13/2 by energy transition to Pr3+: 3F4. The emission cross sections are subsequently calculated by the Fuchtbauer-Ladenburg equation [42]:

σem=Aβλ5I(λ)8πcn2λI(λ)dλ
where I(λ)/∫ λI(λ)dλ is the normalized line shape function of the experimental emission spectrum, β is the fluorescence branching ratio, c is the speed of light, n is the refractive index, and A is the spontaneous emission probability. It is worth noting that the maximum emission cross section of the crystal codoped with Pr3+ is 2.35×10−20 cm2 at 2745 nm, which is larger than that of the crystal without Pr3+ codoping (1.37×10−20 cm2). It may be ascribed to the higher fluorescence branching ratio β of 4I11/24I13/2 transition (from 12.21% to 16.32%), and the shorter radiative lifetime (from 1.48 ms to 1.05 ms) with the codoped of Pr3+ ions (shown in Table 1).

 figure: Fig. 4

Fig. 4 (a) 2.7 μm emission spectrum of Er/Yb/Pr: LN and Er/Yb: LN crystals; (b) 1.5 μm emission spectrum of Er/Yb/Pr: LN and Er/Yb: LN crystals.

Download Full Size | PDF

The ability of a deactivation ion (Pr3+) receiving energy from a donor (Er3+) is an important factor to evaluate the acceptor (Pr3+) as a deactivation ion. In order to investigate the energy interaction mechanism between Er3+ and Pr3+, the time-resolved decays of the Er3+: 4I11/2, and 4I13/2 multiplets for the Er/Yb/Pr: LN and Er/Yb: LN crystals were measured, and the results are shown in Fig. 5, and the inset of (a)(b)(c)(d) show the Ln of fluorescence intensity I versus time t. It is clear to see that the curves of the inset of (a)(b)(c)(d) are unable to be linear fitting, indicating that the these four decay curves can not be fitted to single exponential functions. They can be well fitted to dual exponential decaying, and the fitted curves are also shown in Fig. 5.

 figure: Fig. 5

Fig. 5 (a)(b)Fluorescence decay curves of the Er3+:4I11/2 and 4I13/2 energy levels of Er/Yb: LN crystal. (c)(d)Fluorescence decay curves of the Er3+:4I11/2 and 4I13/2 multiplets of Er/Yb/Pr: LN crystal. The inset of (a)(b)(c)(d) show the Ln of fluorescence intensity I versus time t.

Download Full Size | PDF

On the one hand, the decay curves of the Er3+:4I11/2 energy level of Er/Yb: LN and Er/Yb/Pr: LN crystals are shown in Fig. 5(a) and 5(c). And the measured lifetime of Er3+: 4I11/2 level in Er/Yb/Pr: LN crystal is 234 μs, which is only 16% shorter than that of Er/Yb: LN crystal (280 μs). This unconspicuous shortening of the measured lifetime indicates that the codoping of Pr3+ ions has little influence on the higher laser 4I11/2 level of Er3+. On the other hand, the decay curves of the Er3+:4I13/2 energy level of Er/Yb: LN and Er/Yb/Pr: LN crystals are shown in Fig. 5(b) and 5(d). The measured lifetime of the lower laser 4I13/2 level in the Er/Yb/Pr: LN crystal is 2.18 ms, which is 42% shorter when compared with the date for the Er/Yb: LN crystal (3.78 ms, which is consistent with the results of Er3+ codoped LN crystal published by J. A. Munaoz [30]). This shortening of the measured lifetime confirms that the codoped of Pr3+ ions are able to effectively depopulate the Er3+: 4I13/2 for 2.7 μm emission in LN crystal by energy transition from Er3+: 4I13/2 to Pr3+: 3F4, which may induce the population inversion and facilitate laser operation.

To further investigate the energy interaction mechanism between Er3+ and Pr3+, the energy transition efficiency can be estimated by the following equation: ηET =1−τEr/Yb/Pr/τEr/Yb, where τEr/Yb/Pr and τEr/Yb are the Er3+ lifetimes monitored with and without Pr3+ ions, respectively. Therefore, the energy transition efficiency from Er3+: 4I11/2 level to Pr3+: 1G4 level was calculated to be only 0.16. However, the energy transition efficiency from Er3+: 4I13/2 level to Pr3+: 3F4 level was calculated to be as high as 0.42. It is demonstrated again that Pr3+ can efficiently quench the lower laser level of Er3+ by energy transition, which may induce the population inversion and enhance the emission of 2.7 μm.

4. Conclusion

In conclusion, Er/Yb: LN crystals codoped with and without Pr3+ ion were successfully grown. Compared with the Er/Yb: LN crystal, the Er/Yb/Pr: LN crystal presents an enhanced 2.7 μm luminescence emission under 970 nm excitation, and an higher fluorescence branching ratio (16.32%) corresponding to the stimulated emission of Er3+: 4I11/24I13/2 transition. It was also demonstrated that the introduced Pr3+ depopulates the lower laser level of Er3+: 4I13/2, and has little influence on the higher laser level of Er3+: 4I11/2 at the same time, which is benefited the possible population inversion for the transition of Er3+: 4I11/24I13/2. Moreover, the energy transition efficiency from Er3+: 4I11/2 level to Pr3+: 1G4 level, and from Er3+: 4I13/2 level to Pr3+: 3F4 level were calculated to be 0.16, and 0.42, respectively. The results and analyzed energy transfer indicate codoping of Pr3+ with Yb3+ in Er:LN crystal plays an important role in 2.7 μm emission when the crystal is pumped by a conventional 970 nm LD.

Funding

National Natural Science Foundation of China (NSFC) (614750671, 1404332, 51302283, 51472257); Guangdong Science and Technology Department (2014B090903014, 2014B010131004, 2014B010124002, 2015B090901014, 2016B090917002).

References and links

1. I. D. Aggarwal, L. B. Shaw, and J. S. Sanghera, “Chalcogenide glass fiber-based mid-IR sources and applications,” Proc. SPIE. 6453, 645312 (2007). [CrossRef]  

2. H. Guo, L. Liu, Y. Wang, C. Hou, W. Li, M. Lu, K. Zou, and B. Peng, “Host dependence of spectroscopic properties of Dy3+–doped and Dy3+, Tm3+–codped Ge-Ga-S-CdI2 chalcohalide glasses,” Opt. Express 17, 15350 (2009). [CrossRef]   [PubMed]  

3. J. S. Sanghera, L. B. Shaw, and I. D. Aggarwal, “Chalcogenide Glass-Fiber-Based Mid-IR Sources and Applications,” IEEE J. Sel. Top. Quantum Electron. 15, 114 (2009). [CrossRef]  

4. M. D. Alessandro, U. W. Niklaus, L. B. Sonia, and M. R. Izilda, “Pulse–Energy–Enhanced, Strongly Modulated Er:YLF Laser for Medical Applications,” AIP Conf. Proc. 992, 415 (2008). [CrossRef]  

5. S. D. Jackson and A. Lauto, “Diode-pumped fiber lasers: A new clinical tool?” Laser Surg. Med. 30, 184 (2002). [CrossRef]  

6. K. L. Vodopyanov, “Mid-infrared optical parametric generator with extra-wide (3–19 μm) tunability: applications for spectroscopy of two-dimensional electrons in quantum wells,” J. Opt. Soc. Am. B 16(9), 1579–1586 (1999). [CrossRef]  

7. 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]  

8. B. C. Dickinson, P. S. Golding, M. Pollnau, T. A. King, and S. D. Jackson, “Investigation of a 791 nm pulsed-pumped 2.7 μm Er-doped ZBLAN fibre laser,” Opt. Commun. 191, 315–321 (2001). [CrossRef]  

9. E. Arbabzadah, S. Chard, H. Amrania, C. Phillips, and M. Damzen, “Comparison of a diode pumped Er:YSGG and Er:YAG laser in the bounce geometry at the 3 μm transition,” Opt. Express 19(27), 25860–25865 (2011). [CrossRef]  

10. J. Chen, D. Sun, J. Luo, H. Zhang, R. Dou, J. Xiao, Q. Zhang, and S. Yin, “Spectroscopic properties and diode end-pumped 2.79 μm laser performance of Er,Pr:GYSGG crystal,” Opt. Express 21(20), 23425–23432 (2013). [CrossRef]   [PubMed]  

11. D. W. Chen, C. L. Fincher, T. S. Rose, F. L. Vernon, and R. A. Fields, “Diode-pumped 1-W continuous-wave Er:YAG 3 μm laser,” Opt. Lett. 24(6),385–387 (1999). [CrossRef]  

12. C. Ziolek, H. Ernst, G. F. Will, H. Lubatschowski, H. Welling, and W. Ertmer, “High-repetition-rate, high-average-power, diode-pumped 2.94 μm Er:YAG laser,” Opt. Lett. 26(9),599–601 (2001). [CrossRef]  

13. T. Jensen, A. Diening, G. Huber, and B. H. Chai, “Investigation of diode-pumped 2.8 μm Er:LiYF4 lasers with various doping levels,” Opt. Lett. 21(8),585–587 (1996). [CrossRef]   [PubMed]  

14. T. Sandrock, A. Diening, and G. Huber, “Laser emission of erbium-doped fluoride bulk glasses in the spectral range from 2.7 to 2.8 μm,” Opt. Lett. 24, 382 (1999). [CrossRef]  

15. J. Silver, M.I. Martinez-Rubio, T.G. Ireland, G.R. Fern, and R. Withnall, “Yttrium Oxide Upconverting Phosphors. Part 2: Temperature Dependent Upconversion Luminescence Properties of Erbium in Yttrium Oxide,” J. Phys. Chem. B 105, 948 (2001). [CrossRef]  

16. S. Xu, Z. Yang, J. Zhang, G. Wang, S. Dai, L. Hu, and Z. Jiang, “Upconversion fluorescence spectroscopy of Er3+/Yb3+-codoped lead oxyfluorosilicate glass,” Chem. Phys. Lett. 385, 263 (2004). [CrossRef]  

17. A. S. Oliveira, M. T. de Araujo, A. S. Gouveia-Neto, J. A. Medeiros Neto, A. S. B. Sombra, and Y. Messaddeq, “Frequency upconversion in Er3+/Yb3+-codoped chalcogenide glass,” Appl. Phys. Lett. 72, 753 (1998). [CrossRef]  

18. Z. Zhu, J. Li, Z. You, Y. Wang, S. Lv, E. Ma, J. Xu, H. Wang, and C. Tu, “Beneft of Pr3+ ions to the spectral properties of Pr3+/Er3+: CaGdAlO4 crystal for a 2.7 μm laser,” Opt. Lett. 37, 4838–4840 (2012). [CrossRef]   [PubMed]  

19. Y. Choi, K. Kim, B. Lee, Y. Shin, Y. Kim, and J. Heo, “Emission properties of the Er3+: 4I11/24I13/2 transition in Er3+- and Er3+/Tm3+-doped Ge-Ga-As-S glasses,” J. Non-Cryst. Solids 278, 137–144 (2000). [CrossRef]  

20. Y. Huang, F. Hu, and L. Zhang, “Er3+/Ho3+-codoped fuorotellurite glasses for 2.7 μm fber laser materials,” Fibers 1, 11–20 (2013). [CrossRef]  

21. R. S. Weis and T. K. Gaylord, “Lithium niobate: Summary of physical properties and crystal structure,” Appl. Phys. A. 37, 191–203 (1985). [CrossRef]  

22. X. Li, D. Qu, W. Wang, X. Zhao, L. Zhang, and X. Meng, “Improved nonvolatile holographic storage sensitivity of near-stoichiometric LiNbO3:Fe:Mn crystals,” Chin. Opt. Lett. 10(12), 122101 (2012). [CrossRef]  

23. J. G. Sole, L. Bausa, D. Jaque, E. Montoya, H. Murrieta, and F. Jaque, “Rare earth and transition metal ion centers in LiNbO3,” Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 54, 1571(1998). [CrossRef]  

24. R. Zhang, H. Li, P. Zhang, Y. Hang, and J. Xu, “Efficient 1856 nm emission from Tm,Mg:LiNbO3 laser,” Opt. Express 21, 20990–20998 (2013). [CrossRef]   [PubMed]  

25. P. Zhang, Y. Hang, J. Gong, C. Zhao, J. Yin, and L. Zhang, “Growth, optical characterization and evaluation of laser properties of Yb3+, Mg2+:LiTaO3 crystal,” J. Cryst. Growth. 364, 57–61(2013). [CrossRef]  

26. H. Qi, Z. Song, G. Peng, S. Li, and C. Wang, “Short distributed feedback fiber laser with unidirectional output for sensing applications,” Chin. Opt. Lett. 11(4), 041407 (2013). [CrossRef]  

27. J. Yang, Y. Tang, and J. Xu, “Development and applications of gain-switched fiber lasers [Invited],” Photon. Res. 1, 52–57(2013). [CrossRef]  

28. Y. Liao, J. Xu, Y. Cheng, Z. H. Zhou, F. He, H. Y. Sun, J. Song, X. S. Wang, Z. Z. Xu, K. Sugioka, and K. Midorikawa, “Electro-optic integration of embedded electrodes and waveguides in LiNbO3 using a femtosecond laser,” Opt. Lett. 33, 2281 (2008). [CrossRef]   [PubMed]  

29. E. Cantelar, J. A. Sanz-Garcasa, and F. Cusso, “Growth of LiNbO3 co-doped with Er3+/Yb3+,” J. Crystal Growth , 205, 196–201 (1999). [CrossRef]  

30. J. A. Munoz, B. Herreros, G. Lifante, and F. Cusso, “Concentration Dependence of the 1.5 μmEmission Lifetime of Er3+ in LiNbO3 by Radiation Trapping,” Phys. Stat. Sol. (a) 168, 525–530 (1998). [CrossRef]  

31. H. Kuhn, S. T. Fredrich-Thornton, C. Krankel, R. Peters, and K. Petermann, “Model for the calculation of radiation trapping and description of the pinhole method,” Opt. Lett. 32(13),, 1908–1910 (2007). [CrossRef]   [PubMed]  

32. L. E. Bausa, M.O. Ramirez, and E. Montoya, “Optical performance of Yb3+ in LiNbO3 laser crystal,” Phys. Stat. Sol (a) 201, 289–297 (2004). [CrossRef]  

33. E. Montoya, A. Lorenzo, and L. E. Bausa, “Optical characterization of LiNbO3: Yb3+ crystals,” J. Phys. Condens. Matter 11, 311–320 (1999). [CrossRef]  

34. I. Baumann, F. Cusso, B. Herreros, H. Holzbrecher, H. Paulus, K. Schafer, and W. Sohler, “Praseodymium-doped Ti:LiNbO3 waveguides,” Appl. Phys. A 68, 321–324 (1999). [CrossRef]  

35. L. Nunez, G. Lifante, and F. Cusso, “Polarization effects on the line-strength calculations of Er3+-doped LiNbO3,” Appl. Phys. B. 62, 485–491 (1996). [CrossRef]  

36. T. S. Lomheim and L. G. DeShazer, “Optical-absorption intensities of trivalent neodymium in the uniaxial crystal yttrium orthovanadate,” J. Appl. Phys. 49, 5517 (1978). [CrossRef]  

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

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

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

40. E. Cantelar and F. Cusso, “Dynamics of the Yb3+ to Er3+ energy transfer in LiNbO3,” Appl. Phys. B 69, 29–33 (1999). [CrossRef]  

41. E. Cantelar and F. Cusso, “Analytical solution of the transfer rate equations in LiNbO3:Er3+/Yb3+,” J. Phys. Condensed Matter 12, 521–527 (2000). [CrossRef]  

42. 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, 925 (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 Er3+, Yb3+ and Pr3+ triply doped system.
Fig. 2
Fig. 2 Absorption spectrum of Er/Yb/Pr: LN crystal in the range of 400–2100 nm. The inset shows absorption spectrum of Er: LN, Er/Yb: LN, and Er/Yb/Pr: LN crystals in the range of 850–1700 nm.
Fig. 3
Fig. 3 Fluorescence decay curves of the Yb3+:2F5/2 energy level of Er/Yb/Pr: LN crystal. The inset show the Ln of fluorescence intensity I versus time t.
Fig. 4
Fig. 4 (a) 2.7 μm emission spectrum of Er/Yb/Pr: LN and Er/Yb: LN crystals; (b) 1.5 μm emission spectrum of Er/Yb/Pr: LN and Er/Yb: LN crystals.
Fig. 5
Fig. 5 (a)(b)Fluorescence decay curves of the Er3+:4I11/2 and 4I13/2 energy levels of Er/Yb: LN crystal. (c)(d)Fluorescence decay curves of the Er3+:4I11/2 and 4I13/2 multiplets of Er/Yb/Pr: LN crystal. The inset of (a)(b)(c)(d) show the Ln of fluorescence intensity I versus time t.

Tables (1)

Tables Icon

Table 1 Judd-Ofelt parameters Ω2,4,6, branching ratio β, lifetime of 4I11/2 and 4I13/2 levels for Er3+ ions (τm and τr are the measured and calculated radiative lifetime) of Er/Yb/Pr: LN and Er/Yb: LN crystals.

Equations (1)

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

σ em = 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.