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

Heavily Ho3+-doped lead silicate glass fiber for ~2 μm fiber lasers

Open Access Open Access

Abstract

A heavily Ho3+-doped lead silicate glass with low content of hydroxyls was prepared in this paper. Luminescent properties of this glass were characterized. Peak emission cross section reached 3.07 × 10−21 cm2 around 2055 nm and lifetime of Ho3+: 5I7 was fitted to be 1.31 ms. A single-cald fiber was prepared using this glass as the core. An all-fiber configuration was constructed. Under in-band core-pump of a silica fiber laser, a single-mode maximum 60 mW was realized in a 6 cm-long fiber, which shows high potential for single-frequency fiber lasers applications.

© 2016 Optical Society of America

1. Introduction

Two micrometer fiber lasers have received increasing attention recently due to its potential applications in remote sensing, ranging, medical surgery, and as pumping sources for mid-infrared lasers [1–4]. For ~2 μm fiber lasers, thulium or holmium have been widely used as doping rare earth ions. Till now, Tm3+-doped fiber laser have drawn wide attention due to available commercial diode laser pump source and high quantum efficiency [5, 6]. However, one favorable LIDAR wavelength of 2053 nm is far away from the gain peak of thulium doped glass fiber [7]. Compared to thulium, the emission peak of Ho3+ is 200 nm longer. Hence, holmium ions have a much higher gain at 2050 nm. Though there is no commercial diode pump laser available, great progress have been made in pumping Ho3+ with Tm3+-doped fiber laser or Yb3+-doped fiber laser [8, 9].

Silica fiber laser have gained great success in power scaling owing to its high strength and low loss. Until now, kilowatt-level laser output has been achieved in thulium-doped silica fiber lasers [10], and 407 W laser output has also been obtained in Ho3+-doped silica fibers pumped by Tm3+-doped fiber lasers [11]. However, doping concentration of rare earth ions is relatively low in silica, and thus hinders its gain ability [12]. Till now, to the best of our knowledge, single frequency fiber lasers have not been reported in Ho3+-doped silica fiber. Multi-component glass fiber gained wide attention for single frequency glass fiber laser [13, 14]. Ho3+-doped fiber laser have been achieved in various kinds of multi-component glass materials. In 2009, Wu et al demonstrated a single frequency fiber laser operating near 2 μm with over 50 mW output power using a short piece of holmium-doped germanate glass fiber [7]. In 2012, single frequency fiber laser has also been obtained with gain-switch method in a Ho3+-doped silicate fiber by Geng et al [15]. Recently, laser output has been realized in a Ho3+-doped flurotellurite fiber by Qin et al [16].

Rare earth ion doped lead silicate glass fibers have been fewly investigated, mainly ascribed to hardness of melting corresponding glass. Earlier, Tm3+-doped lead silicate glass fiber laser has been realized [17]. Here we investigated a Ho3+-doped lead silicate glass for ~2 μm fiber lasers. A single-clad fiber was prepared successfully. An all-fiber configuration was constructed. With in-band core-pumped by Tm3+-doped fiber laser, a maximum 60 mW single-mode signal laser was realized from a 6 cm-long single-clad fiber. This is the first report of ~2 μm laser output from a Ho3+: SPANK fiber.

2. Experiments

2.1 Material synthesis and fiber drawing

Ho3+-doped glass system SiO2-PbO-Al2O3-Na2O-K2O was prepared, which was named SPANK afterwards. Na2O and K2O were introduced by corresponding carbonates. The glass was melted in a 500 mL crucible made of high purity silica. The glass melt was cleared at 1400°C for 1 h and then stirred at 1200°C for 3 h. After that, the glass melt were poured into a preheated stainless iron mould and placed in a muffle furnace for annealing. The annealing temperature was 10°C above Tg of corresponding glass sample. The whole melting process was protected by high purity oxygen. Finally, the glass sample was machined to a size of 10 × 10 × 2 mm for analysis.

2.2 Measurements

The refractive index was measured by the prism coupler method, while the density was measured by the Archimedes method using distilled water as the immersion liquid. The transmittance spectra in the 2000-4000 cm–1 range were measured on a Nicolet 6700 series Fourier transform infrared (FTIR) spectrometer. In addition, the absorption spectra were recorded using a Perkin-Elmer Lambda 900 UV/VIS/NIR spectrophotometer in the 300-2200 nm range. Doping concentration of Ho3+ in glass sample was measured using an ICP-AES 6300 apparatus. The emission spectrum of the core glass was measured upon excitation at 640 nm using a Triax 320 type spectrometer (Jobin-Yvon Co, France). Lifetime measurement of the Ho3+: 5I7 state was conducted in a FLSP 920 (Edinburgh instruments Ltd, UK) instrument. Laser experimental setup will be given in section 3.2. All the experiments were carried out at room temperature.

3. Results and discussions

3.1 Luminescent properties of Ho3+-doped SPANK

Doping concentration of Ho3+ was confirmed by ICP-AES to be 5.91 × 1020 cm−3, much higher than that of Ho3+-doped silica fiber [18]. Figure 1 shows absorption spectra of Ho3+ in SPANK in the range of 500-2200 nm. Absorption peaks around 536, 640, 880, 1150, and 1938 nm are ascribed to transitions from ground state 5I8 to excited states 5I7, 5I6, 5I5, 5I4, and 5F5, respectively. Integrated areas of 5I6, 5I5, 5I4, and 5F5 are calculated for J-O parameters simulation. J-O parameters Ω2, Ω4, and Ω6 of Ho3+ for SPANK is calculated to be 5.52, 2.19, and 0.59, respectively. The room-mean-square error deviation of intensity parameters is 0.6 × 10−6, which indicates the validity of the J-O theory for predicting the spectral intensities of Ho3+ and the reliable calculations. As listed in Table 1, compared to other glass matrices, Ω2 is much higher owing to high polarizability of PbO in Ho3+: SPANK. As a result, spontaneous transition rate A10 is higher than silicate and ZBYA glass though smaller than that of germanate glass. However, A10 of Ho3+ is still much lower than that of Tm3+ in the same matrix [17].

 figure: Fig. 1

Fig. 1 Absorption spectrum of Ho3+-doped SPANK glass.

Download Full Size | PDF

Tables Icon

Table 1. J-O parametersΩ2, Ω4, Ω6, spontaneous transition rate (A10), radiative lifetime τrad in various glass matrices

OH is one of the dominant quenching centers in Ho3+-doped glasses [22], especially for populations in 5I7 level, since only a few hydroxyl groups (2700 cm–1-3700 cm–1) are required for non-radiative de-excitation through multi-phonon relaxation. In this work, this glass was melted in a low hydroxyl atmosphere with 8% humidity. Bubbling with liquid oxygen was used to eliminate OH from the glass melt. The OH content in the glass could be expressed by the absorption coefficient of OH vibration band at 2880 cm–1, which is given by [23]:

αOH=ln(T0/T)/l
where l is the sample thickness, whereas T0 and T are the transmitted and incident intensities, respectively.

Figure 2(a) shows transmittance spectra of 2 mm-thick Ho3+: SPANK glass sample in the range of 2000-4000 cm−1. According to Eq. (1), absorption of hydroxyls was calculated to be 0.4 cm−1. The relatively low content of hydroxyls contribute to a long lifetime of Ho3+: 5I7 energy level, which is fitted to be 1.41ms.The lifetime of Ho3+: 5I7 in SPANK is much longer than that in Ho3+: silica though doping concentration of Ho3+ is much higher, indicating fine distribution of Ho3+ in the glass [18]. The relatively long lifetime contributes to population inversion during laser operation.

 figure: Fig. 2

Fig. 2 (a) Transmittance spectra of Ho3+: SPANK glass in the range of 2000-4000 cm−1. Inset is decay curve of Ho3+: 5I7 . (b) Normalized absorption spectrum and emission spectrum of Ho3+-doped SPANK in the range of 1700-2200 nm.

Download Full Size | PDF

The absorption cross section is calculated according to σabs = 2.303*OD/N0*l and emission cross section is calculated according to Füchtbauer–Ladenburg (F-L) equation [24]. Figure 2(b) shows absorption cross section and emission cross section of Ho3+: SPANK glass in the range of 1700-2200 nm. The peak absorption cross section reaches 3.94 × 10−21 cm2 around 1938 nm while peak emission cross section arrives at 3.07 × 10−21 cm2 around 2055 nm. Thus, pump laser source at 1940 nm was picked up to pump single-clad Ho3+: SPANK fiber in the laser experiment. A FBG with central wavelength at 2040 nm is chosen owing to relatively high emission cross section here.

3.2 Laser properties of Ho3+: SPANK single-clad fiber

The Ho3+: SPANK single-clad fiber was prepared with stack-and-draw method [17]. The core glass was Ho3+: SPANK while the cladding glass was ZF-2, provided by CDGM GLASS CO.,LTD. Inset of Fig. 3 shows cross section of Ho3+: SPANK single-clad fiber with a diameter of 125 μm. This fiber has a core of 13 μm and a NA of 0.30 at 2040 nm. Loss of the fiber was confirmed to be 0.05dB/cm at 1310 nm using cut-back method, which is much lower than our previous fiber [17]. Absorption around 1940 nm of this fiber was theoretically estimated to be 9.94 dB/cm. However, the practical absorption at 1940 nm was measured to be 2.34 dB/cm. The large discrepancy between the calculated and measured is ascribed to off-center distribution of the fiber core. That is to say, part of light was injected into the cladding. A length of 6 cm could make the best use of the pump light due to high absorption of the fiber.

 figure: Fig. 3

Fig. 3 Setup of test platform for Ho3+-doped single-clad SPANK fiber. Cross section of Ho3+: SPANK single-clad fiber is shown inset.

Download Full Size | PDF

Ho3+: SPANK single-clad fiber was in-band pumped by a Tm3+-doped silica fiber laser. The Tm3+: silica fiber was provided by FUTONG GROUP. A schematic of built-in-house in-band pump test platform is depicted in Fig. 3. In the Tm3+-doped silica fiber laser, a 793 nm LD was chosen as pump source. The pump light was collimated and focused into the fiber with two aspheric mirrors. A FBG with high reflectivity at 1940 nm on the front end and a FBG with 55% reflectivity at 1940nm on the back end were used to form a resonant cavity. A 4 m-long Tm3+-doped silica fiber could make the best of pump laser. An isolator was placed after the output end in case of back-reflection of signal fiber laser. For Ho3+: SPANK fiber laser, a FBG with high reflectivity at 2040 nm on the front end and a FBG with 40% reflectivity at 2040nm on the back end were used to form a resonant cavity. All FBGs used here are written on commercial single-mode silica fibers. A 6 cm-long Ho3+: SPANK fiber was butted to the fiber gratings to form an all-fiber configuration. Here, a reflector with high reflection (99%) at 1940nm was chosen to filter the 1940 nm pump laser for power and luminescence measurement.

As is shown in Fig. 4(a), the black curve is spectra detected from the output fiber end at a LD pump current of 2.8 A. ASE of Ho3+-doped fiber covers from 1981 nm to 2086 nm. The wide ASE indicates feasibility of a short pulse for mode-locking [25]. It should be noticed that signal of Tm3+: silica fiber laser could not be entirely blocked by the reflector. With LD pump power increasing to 3.5 A, a stable signal laser peaking at 2040 nm arises. The signal laser has a FWHM of 6 nm, which was ascribed to narrow reflection band of FBG (Δλ - 0.766 nm). Furthermore, at this point, signal of Tm3+: silica fiber laser was very weak and could hardly be detected by the fiber spectrograph.

 figure: Fig. 4

Fig. 4 (a) Normalized spectra of pump laser, ASE, and signal laser; (b) Efficiency of signal power with respect to LD pump current.

Download Full Size | PDF

Figure 4(b) shows efficiency of signal power with respect to LD pump power. It is hard to determine how much pump light was coupled into core of the fiber owing to off-center distribution of the fiber core in the Ho3+: SPANK fiber. Thus, pump current of 793 nm LD laser was chosen as x axis. The signal laser has a threshold of 3.4 A. With increasing pump current, a maximum signal power of 60 mW was generated from Ho3+: SPANK single-clad fiber without saturation. It is believed that signal out-power could be further scaled. Though large NA of this single-clad fiber does not supports single-mode operation, the signal laser was single-mode outputted owing to the single-mode FBG was pigtailed. The fiber shows high potential for single-frequency fiber lasers if a more complete fiber structure was picked up.

4. Conclusion

A heavily Ho3+-doped lead silicate glass with low content of hydroxlys was prepared. Doping concentration of Ho3+ was confirmed to be 5.91 × 1020 cm−3. Luminescent properties of this glass were characterized. Peak emission cross section reaches 3.07 × 10−21 cm2 around 2055 nm and lifetime of Ho3+: 5I7 was fitted to be 1.31 ms. A single-cald fiber was prepared using this glass as the core. Under in-band core-pump of a silica fiber laser, a single-mode maximum 60 mW was realized in a 6 cm-long fiber. This is the first report of ~2 μm laser output from a Ho3+: SPANK fiber. All these results show that Ho3+: SPANK fiber has high potential for single-frequency fiber lasers applications.

Acknowledgments

This research was supported by the Chinese National Natural Science Foundation (No. 51272262 61308084 and 60573172) and Shanghai Natural Science Foundation (No.15ZR1444800).

References and links

1. S. D. Jackson, “Towards high-power mid-infrared emission from a fibre laser,” Nat. Photonics 6(7), 423–431 (2012). [CrossRef]  

2. B. M. Walsh, “Review of Tm and Ho materials: spectroscopy and lasers,” Laser Phys. 19(4), 855–866 (2009). [CrossRef]  

3. J. Boguslawski, J. Sotor, G. Sobon, R. Kozinski, K. Librant, M. Aksienionek, L. Lipinska, and K. M. Abramski, “Graphene oxide paper as a saturable absorber for Er- and Tm-doped fiber lasers,” Photon. Res. 3(4), 119–124 (2015). [CrossRef]  

4. N. M. Fried and K. E. Murray, “High-power thulium fiber laser ablation of urinary tissues at 1.94 um,” J. Endourol. 19(1), 25–31 (2005). [CrossRef]   [PubMed]  

5. S. D. Jackson, A. Sabella, and D. G. Lancaster, “Application and Development of High-Power and Highly Efficient Silica-Based Fiber Lasers Operating at 2μm,” IEEE J. Sel. Top. Quant. 13(3), 567–572 (2007). [CrossRef]  

6. S. D. Jackson and S. Mossman, “Efficiency dependence on the Tm3+ and Al3+ concentrations for Tm3+-doped silica double-clad fiber lasers,” Appl. Opt. 42(15), 2702–2707 (2003). [CrossRef]   [PubMed]  

7. J. Wu, Z. Yao, J. Zong, A. Chavez-Pirson, N. Peyghambarian, and J. Yu, “Single frequency fiber laser at 2.05 μm based on Ho-doped germanate glass fiber,” in SPIE LASE: Lasers and Applications in Science and Engineering, (International Society for Optics and Photonics, 2009), 71951K–71951K–71957.

8. X. Fan, M. Chen, A. Shirakawa, K. Ueda, C. B. Olausson, J. K. Lyngsø, and J. Broeng, “High power Yb-doped photonic bandgap fiber oscillator at 1178 nm,” Opt. Express 20(13), 14471–14476 (2012). [CrossRef]   [PubMed]  

9. A. Hemming, S. Bennetts, N. Simakov, A. Davidson, J. Haub, and A. Carter, “High power operation of cladding pumped holmium-doped silica fibre lasers,” Opt. Express 21(4), 4560–4566 (2013). [CrossRef]   [PubMed]  

10. T. Ehrenreich, R. Leveille, I. Majid, K. Tankala, G. Rines, and P. Moulton, “1-kW, all-glass Tm: fiber laser,” Proc. SPIE 7580, 758016 (2010).

11. A. Hemming, N. Simakov, A. Davidson, S. Bennetts, M. Hughes, N. Carmody, P. Davies, L. Corena, D. Stepanov, and J. Haub, “A monolithic cladding pumped holmium-doped fibre laser,” in CLEO: Science and Innovations, (Optical Society of America, 2013), CW1M. 1.

12. F. Lou, P.-W. Kuan, L. Zhang, S. Wang, Q. Zhou, M. Wang, S. Feng, K. Li, C. Yu, and L. Hu, “2 μm laser properties of Tm3+-doped large core sol-gel silica fiber,” Opt. Mater. Express 4(6), 1267–1275 (2014). [CrossRef]  

13. S. H. Xu, Z. M. Yang, T. Liu, W. N. Zhang, Z. M. Feng, Q. Y. Zhang, and Z. H. Jiang, “An efficient compact 300 mW narrow-linewidth single frequency fiber laser at 1.5 µm,” Opt. Express 18(2), 1249–1254 (2010). [CrossRef]   [PubMed]  

14. S. Xu, Z. Yang, W. Zhang, X. Wei, Q. Qian, D. Chen, Q. Zhang, S. Shen, M. Peng, and J. Qiu, “400 mW ultrashort cavity low-noise single-frequency Yb³⁺-doped phosphate fiber laser,” Opt. Lett. 36(18), 3708–3710 (2011). [CrossRef]   [PubMed]  

15. J. Geng, Q. Wang, T. Luo, B. Case, S. Jiang, F. Amzajerdian, and J. Yu, “Single-frequency gain-switched Ho-doped fiber laser,” Opt. Lett. 37(18), 3795–3797 (2012). [CrossRef]   [PubMed]  

16. C. Yao, C. He, Z. Jia, S. Wang, G. Qin, Y. Ohishi, and W. Qin, “Holmium-doped fluorotellurite microstructured fibers for 2.1 μm lasing,” Opt. Lett. 40(20), 4695–4698 (2015). [CrossRef]   [PubMed]  

17. X. Liu, X. Wang, L. Wang, P. Kuan, M. Li, W. Li, X. Fan, K. Li, L. Hu, and D. Chen, “Realization of 2 µm laser output in Tm3+-doped lead silicate double cladding fiber,” Mater. Lett. 125, 12–14 (2014). [CrossRef]  

18. X. Wang, L. Hu, W. Xu, S. Wang, L. Zhang, C. Yu, and D. Chen, “Spectroscopic properties of Ho3+ and Al3+

19. F. Huang, J. Cheng, X. Liu, L. Hu, and D. Chen, “Ho3+/Er3+ doped fluoride glass sensitized by Ce3+ pumped by 1550 nm LD for efficient 2.0 μm laser applications,” Opt. Express 22(17), 20924–20935 (2014). [CrossRef]   [PubMed]  

20. B. Peng and T. Izumitani, “Optical properties, fluorescence mechanisms and energy transfer in Tm3+, Ho3+ and Tm3+ -Ho3+ doped near-infrared laser glasses, sensitized by Yb3+,” Opt. Mater. 4(6), 797–810 (1995). [CrossRef]  

21. Q. Zhang, J. Ding, Y. Shen, G. Zhang, G. Lin, J. Qiu, and D. Chen, “Infrared emission properties and energy transfer between Tm3+ and Ho3+ in lanthanum aluminum germanate glasses,” J. Opt. Soc. Am. B 27(5), 975–980 (2010). [CrossRef]  

22. E. J. Friebele, C. G. Askins, J. R. Peele, B. M. Wright, N. J. Condon, S. O’Connor, C. G. Brown, and S. R. Bowman, “Ho-doped fiber for high energy laser applications,” in SPIE LASE, (International Society for Optics and Photonics, 2014), 896120–896120–896129.

23. F. Huang, X. Liu, W. Li, L. Hu, and D. Chen, “Energy transfer mechanism in Er3+ doped fluoride glass sensitized by Tm3+ or Ho3+ for 2.7μm emission,” Chin. Opt. Lett. 12(5), 051601 (2014). [CrossRef]  

24. S. Li, P. Wang, H. Xia, J. Peng, L. Tang, Y. Zhang, and H. Jiang, “Tm3+ and Nd3+ singly doped LiYF4 single crystals with 3-5μm mid-infrared luminescence,” Chin. Opt. Lett. 12(2), 021601 (2014). [CrossRef]  

25. Q. Wang, J. Geng, T. Luo, and S. Jiang, “Mode-locked 2 µm laser with highly thulium-doped silicate fiber,” Opt. Lett. 34(23), 3616–3618 (2009). [CrossRef]   [PubMed]  

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 spectrum of Ho3+-doped SPANK glass.
Fig. 2
Fig. 2 (a) Transmittance spectra of Ho3+: SPANK glass in the range of 2000-4000 cm−1. Inset is decay curve of Ho3+: 5I7 . (b) Normalized absorption spectrum and emission spectrum of Ho3+-doped SPANK in the range of 1700-2200 nm.
Fig. 3
Fig. 3 Setup of test platform for Ho3+-doped single-clad SPANK fiber. Cross section of Ho3+: SPANK single-clad fiber is shown inset.
Fig. 4
Fig. 4 (a) Normalized spectra of pump laser, ASE, and signal laser; (b) Efficiency of signal power with respect to LD pump current.

Tables (1)

Tables Icon

Table 1 J-O parametersΩ2, Ω4, Ω6, spontaneous transition rate (A10), radiative lifetime τrad in various glass matrices

Equations (1)

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

α O H =ln( T 0 /T)/l
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.