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

1.06μm Q-switched ytterbium-doped fiber laser using few-layer topological insulator Bi2Se3 as a saturable absorber

Open Access Open Access

Abstract

Passive Q-switching of an ytterbium-doped fiber (YDF) laser with few-layer topological insulator (TI) is, to the best of our knowledge, experimentally demonstrated for the first time. The few-layer TI: Bi2Se3 (2–4 layer thickness) is firstly fabricated by the liquid-phase exfoliation method, and has a low saturable optical intensity of 53 MW/cm2 measured by the Z-scan technique. The optical deposition technique is used to induce the few-layer TI in the solution onto a fiber ferrule for successfully constructing the fiber-integrated TI-based saturable absorber (SA). By inserting this SA into the YDF laser cavity, stable Q-switching operation at 1.06 μm is achieved. The Q-switched pulses have the shortest pulse duration of 1.95 μs, the maximum pulse energy of 17.9 nJ and a tunable pulse-repetition-rate from 8.3 to 29.1 kHz. Our results indicate that the TI as a SA is also available at 1 μm waveband, revealing its potential as another broadband SA (like graphene).

© 2013 Optical Society of America

1. Introduction

Passively Q-switched fiber lasers with the advantages of compactness and flexibility have attracted much attention, and have been opening up some important applications in laser material processing, remote sensing, telecommunications and medicine. One of the most effective methods for passive Q-switching in fiber lasers is to use the saturable absorption of optical materials (i.e. saturable absorbers, SAs). In the past two decades, several kinds of SAs (e.g. semiconductor saturable absorber mirrors SESAMs [1], carbon nanotubes CNTs [2], graphene [3,4]) have been successively discovered, and successfully used to passively Q-switch fiber lasers. Although SESAMs have been well developed to be very mature and commercial for the past two decades, they have also some limitations, including 1) relatively narrowband wavelength operation range, 2) complex fabrication and expensive packaging. The carbon-based nanomaterials (e.g. CNTs, graphene) are easy-fabrication and cost-effective, but the bandgap engineering or diameter/charity control of the CNT is often demanded for obtaining a broad wavelength range of saturable absorption. Although graphene possesses the wavelength-independent saturable absorption, it has a small absorption of only 2.3% at 1550 nm and a relatively low modulation depth (less than 66.5% (normalized)) [4]. Therefor, researchers are still making some efforts even now to seek for new SAs which are expected to have the ideal characteristics of wavelength-independent saturable absorption, low saturable optical intensity, large modulation depth, high damage threshold and low cost. Most recently, similar to the graphene SA [59], the topological insulators (TIs) have been found to be another potential and ideal SAs for mode-locking or Q-switching [1013]. Especially, TIs have impressively shown the normalized modulation depth as high as 98% [11].

TIs are states of quantum matter with the metallic states on the surface and have the narrow topological non-trivial energy gaps (e.g. TI: Bi2Se3, ΔE∼0.3 eV) [14], exhibiting the Dirac-like linear dispersion. When the narrow-bandgap TI is excited by strong light with the single-photon energy more than the TI bandgap ΔE, the saturable absorption could happen due to the Pauli-blocking principle [1012], like graphene. Therefore, TIs could also possess the ultra-broadband saturable absorption. For example, the saturable-absorption wavelength range of TI: Bi2Se3 is calculated to be from ultraviolet to mid-infrared <4.1 μm. Recently, Zhao et al. have measured that the saturable optical intensity of multilayer (>50 layers) TI:Bi2Se3 is 10.12 GW/cm2 at 800 nm [15] and 0.49 GW/cm2 at 1550 nm [11]. By further inserting such TI-based SAs into the fiber laser cavities, their group has successfully mode-locked the Er-doped fiber lasers at 1.55 μm [10,11]. However, it should be noted that: 1) the broadband saturable-absorption characteristic of the TIs has not been fully exploited, 2) the TIs were not used previously to Q-switch fiber laser at 1 μm, and 3) if using few-layer TIs instead of multilayer TIs, its saturable optical intensity is expected to be lower and the corresponding mode-locking/Q-switching might be started more easily. In this letter, we report a passively Q-switched ytterbium-doped fiber (YDF) laser with few-layer TI: Bi2Se3 as a SA. The Q-switched YDF laser operates in the 1.06 μm wavelength which is sufficiently away from the 1.55 μm waveband of TI-based Er-doped lasers reported previously [1013], further demonstrating the broadband saturable-absorption nature of the TI.

2. Preparation and characterization of few-layer TI:Bi2Se3

The few-layer TI: Bi2Se3 used in our experiment was prepared by the liquid-phase exfoliation method as follows. Initially, we synthesized the Bi2Se3 crystal from bismuth oxide powder (Bi2O3)and selenium powder (Se) through the hydrothermal method. Then, the as-synthesized Bi2Se3 were added into the N-methyl-2-pyrrolidone (NMP) solution and sonicated for 24 h to produce the few-layer TI: Bi2Se3 suspension [the inset of Fig. 1(b)]. The thickness of the as-prepared TI nanosheets was characterized by atomic force microscopy (AFM), as shown in Fig. 1. The average thickness from the height profile diagram [Fig. 1(b)] was measured to be ∼2–4 nm. This indicates that the TI nanosheets are around 2–4 layers, because the single-layer thickness of Bi2Se3 is 0.96 nm [16]. It is interestingly noticed that the layer number (2–4 layer) of our TI: Bi2Se3 is much less than that (>50 layer) of TI: Bi2Se3 [11,15] previously reported by the hydrothermal intercalation/exfoliation method.

 figure: Fig. 1

Fig. 1 (a) The AFM image and (b) the height profile diagram of the few-layer TI: Bi2Se3. Inset: the as-prepared TI: Bi2Se3 suspension.

Download Full Size | PDF

In order to conveniently measure the optical absorption properties of the few-layer TI: Bi2Se3, we sprayed the TI suspension onto a glass substrate by the spin-coating method, and then evaporating to dryness in an oven. As shown in Fig. 2(a), we measured its transmission spectrum (i.e. linear-absorption spectrum) from 300 to 2000 nm using an optical spectrometer (Perkinelmer Lambda 7500). The linear-absorption curve is flat from ultraviolet to infrared region with 79.8% transmittance at 1.06 μm, indicating the potential of the few-layer TI as a broadband optical material. Using the Z-scan technique, we also characterized the saturable absorption of the TI sample. As described in Fig. 2(b), the open-aperture Z-scan curve shows the real modulation depth (δT) of 3.8%, comparable to that of graphene [5,17] and CNTs [2]. The saturable optical intensity (Isa) of the few-layer TI is 53 MW/cm2 at 800 nm wavelength, which is about two-orders of magnitude less than the reported one (10.12 GW/cm2) of the multilayer (>50 layer) TI [15], mainly benefiting from the few-layer structure. It is noted that the Isa (53 MW/cm2 @800 nm) is also less than that (2.77 GW/cm2 @800 nm [18]) of single-layer graphene. With such low saturable intensity, it can be predicted that the threshold of mode-locking or Q-switching using the few-layer TI can be significantly reduced. In addition, the larger modulation depth could be enabled by increasing the thickness of the TI SA (i.e. superimposing more nanosheets of few-layer TI), but correspondingly the insertion loss will also increase, leading to a high threshold of Q-switching/mode-locking.

 figure: Fig. 2

Fig. 2 (a) The transmission spectrum of the few-layer TI: Bi2Se3 sample, and (b) its saturable absorption characteristic at 800 nm wavelength by the Z-scan technique.

Download Full Size | PDF

3. Experimental setup of TI-based passively Q-switched Yb-doped fiber laser

To further exploit the broadband saturable absorption of the TI, we designed a 1.06 μm passively Q-switched YDF laser with the few-layer TI: Bi2Se3 as a Q-switcher. The experimental setup of the proposed Q-switched YDF laser is schematically shown in Fig. 3. The compact linear cavity consists of a 20-cm high-concentration YDF, two fiber Bragg gratings (FBG1 & FBG2) and the TI: Bi2Se3 Q-switcher. The total cavity length is ∼3.2 m. The pump laser from a 190 mW/974 nm laser diode (LD) was directly launched into the YDF for providing the laser gain. The FBG1 and FBG2 were used to form the linear resonant cavity at the 1067 nm wavelength. The chirped FBG1 has a high reflectivity of >99.9% in the range of 1067.01∼1068.20 nm, and as given in Fig. 3(a), the FBG2 has two reflection wavelengths of 1067.66 and 1067.92 nm with the reflectivity of 98.0% and 97.8%, respectively. Therefore, ∼2% intracavity light can be extracted as laser output from the FBG2. The output laser is measured by an optical spectrum analyzer (OSA, Advantest Q8384), an optical powermeter and a high-speed photodetector together with a digital oscilloscope (Tektronix TDS2024) or radio-frequency (RF) spectrum analyzer (Gwinstek GSP-930).

 figure: Fig. 3

Fig. 3 The experimental setup of the TI-based passively Q-switched YDF laser. (a) the transmission spectrum of the FBG2, (b) the optical deposition setup, (c) the image of the fiber-ferrule facet after depositing the TI nanosheets.

Download Full Size | PDF

The saturable absorption of TI: Bi2Se3 plays a important role in the experiment, but the key problem is how to integrate the TI nanosheets into the fiber laser for guaranteeing the compact all-fiber structure. In this work, we used the optical deposition technique [19,20] to induce the few-layer TI nanosheets in the suspension onto a fiber ferrule. The optical-deposition setup is given in Fig. 3(b). When the 20 mW laser at 974 nm was injected into the fiber, the fiber ferrule in the TI suspension started the optical deposition for about 30 minutes, and was then moved out the suspension to evaporate for 3 h at 60 °C. After the optical-deposition process, it is clearly seen from the image [Fig. 3(c)] of the fiber-ferrule facet that some TI nanosheets were successfully deposited on the fiber core. Finally, the TI-deposited fiber ferrule was connected with another ferrule by an optical adaptor to construct the fiber-compatible TI: Bi2Se3 Q-switcher. The insertion loss of this device is 1.9 dB at 1067 nm.

4. Experimental results and discussions

In the experiment, the YDF laser started the continuous-wave lasing at the pump power of 36.7 mW, and then transited to the Q-switching operation at ∼42.5 mW. When the pump power was gradually increased from 42.7 to 106.2 mW, as shown in Fig. 4, the stable pulse trains with the different repetition rates were observed, which is the typical feature of passive Q-switching. However, once the pump power is over 107 mW, we found that the Q-switched pulses became unstable, as usually observed in some passively Q-switched fiber lasers reported previously [8,19,21]. For excluding the thermal damage of the TI-based SA, we gradually decreased the pump power from 150 mW to 0, and the stable Q-switching operation was observed again at the pump power of <107 mW, indicating that the TI-based SA is not damaged. We think that the possible reason for the unstable Q-switching is the over-saturation of the SA at higher pump intensity. At the pump power of 77.7 mW, Fig. 5 summarizes the typical characteristics of the Q-switching operation. The output optical spectrum in Fig. 5(a) has the two lasing wavelengths of 1067.66 and 1067.91 nm, corresponding to the reflection wavelengths of the FBG2. One can see that the intensity of the weak peak (1067.91 nm) is much less (<10 dB) than that of the strong peak (1067.66 nm), and therefore the effect of the weak peak on the Q-switching can be neglected. The 3-dB bandwidth of the 1067.66-nm lasing channel is 0.10 nm. Fig. 5(b) shows the oscilloscope trace of the Q-switched pulse trains. The pulse train has the period of 51.3 μs and the pulse intensity fluctuation of less than 6%. Fig. 5(c) plots the typical pulse envelope, having the symmetrical Gaussian-like shape. The pulse duration is 2.71 μs, comparable to those reported fiber lasers Q-switched with other SAs (e.g. CNTs [21], SESAM [22], graphene [19,2325]). The RF output spectrum in Fig. 5(d) shows that the Q-switched pulses have the repetition rate of 19.5 kHz which agrees with the pulse period of 51.3 μs measured in Fig. 5(b). Moreover, the RF signal-to-noise ratio is as high as 48 dB (∼105 contrast), and the broadband RF spectrum in the inset of Fig. 5(d) shows no spectral modulation, indicating that the repetition rate of the passively Q-switching was very stable.

 figure: Fig. 4

Fig. 4 The Q-switched pulse trains under different pump powers Pp, (a) Pp=42.7 mW, (b) Pp=49.2 mW, (c) Pp=81.7 mW, (d) Pp=106.2 mW.

Download Full Size | PDF

 figure: Fig. 5

Fig. 5 The typical Q-switching characteristics at the pump power of 77.7 mW. (a) output optical spectrum, (b) the oscilloscope trace of the Q-switched pulse train, (c) single pulse envelope, (d) RF output spectrum, and Inset: broadband RF output spectrum.

Download Full Size | PDF

As shown in Fig. 6(a), we recorded the pulse repetition rate and the pulse duration as a function of the pump power. As increasing the pump power from 42.5 to 106.2 mW, the repetition rate of the Q-switching pulses can be tuned from 8.3 to 29.1 kHz, and meanwhile the pulse duration was significantly narrowed from 8.3 to 1.95 μs. The pulse duration could be further narrowed by optimizing the parameters, including 1) shortening the cavity length and 2) improving the modulation depth of the TI: Bi2Se3 Q-switcher [26]. In addition, as plotted in Fig. 6(b), we also measured the average output power and correspondingly calculated the single-pulse energy. The average output power almost linearly increased with the input pump power, and at the pump power of 106.2 mW, the maximum average output power is 0.46 mW. One can see from Fig. 6(b) that the pulse energy linearly grew in the initial stage, but after the pump power over 90 mW, the pulse energy became to saturate obviously. The maximum pulse energy is 17.9 nJ, comparable to those Q-switched fiber lasers using graphene [19,2325], CNTs [21] and SESAMs as passive Q-switchers. For the purpose of enlarging the pulse energy, we replaced the FBG2 with a fiber loop mirror (15% output) and increased the Yb-doped gain fiber to be 50-cm length. The pulse energy was indeed enlarged up to 41.3 nJ (corresponding to the average output power of 1.1 mW), but the pump threshold of Q-switching is significantly increased to 66.3 mW and the minimum pulse duration is broadened to 4.9 μs. The higher pulse energy could be enabled by using high-gain fiber (e.g. double-clad fiber) with high-performance TI-based SA, and further optimizing the cavity designs (e.g. output coupling ratio, cavity loss) [26].

 figure: Fig. 6

Fig. 6 (a) The pulse duration and the pulse repetition rate as a function of pump power, (b) the average output power and the pulse energy as a function of pump power.

Download Full Size | PDF

5. Conclusion

In summary, we have proposed and experimentally demonstrated the TI-based passively Q-switched all-fiber laser operating at 1 μm waveband. Using the optical deposition method, the few-layer TI: Bi2Se3 in the suspension was well deposited onto a fiber ferrule to fabricate the fiber-compatible TI-based Q-switcher. The stable Q-switching operation at 1.067 μm was successfully achieved with the maximum pulse energy of 17.9 nJ, the shortest pulse duration of 1.95 μs and the pulse repetition rate from 8.3 to 29.1 kHz. Our results show that: 1) the Q-switching performances based on the few-layer TI are comparable to those obtained with other SAs (e.g. CNTs [21], graphene [19,2325]) and 2) the TI nanomaterials could operate as the ultra-broadband SA.

Acknowledgments

This work is supported partially by the National Natural Science Foundation of China (NSFC) (No. 61177044 and 61107038), and partially supported by the Natural Science Foundation of Fujian Province of China (No. 2011J01370).

References and links

1. U. Keller, K. J. Weingarten, F. X. Kartner, D. Kopf, B. Braun, I. D. Jung, R. Fluck, C. Honninger, N. Matuschek, and J. Aus der Au, “Semiconductor saturable absorber mirrors (SESAM’s) for femtosecond to nanosecond pulse generation in solid-state laser,” IEEE J. Sel. Top. Quantum Electron. 2, 435–453 (1996). [CrossRef]  

2. S. Yamashita, Y. Inoue, S. Maruyama, Y. Murakami, H. Yaguchi, M. Jablonski, and S. Set, “Saturable absorbers incorporating carbon nanotubes directly synthesized onto substrates and fibers and their application to mode-locked fiber lasers,” Opt. Lett. 29, 1581–1583 (2004). [CrossRef]   [PubMed]  

3. T. Hasan, Z. Sun, F. Wang, F. Bonaccorso, P. H. Tan, A. G. Rozhin, and A. C. Ferrari, “Nanotube-polymer composites for ultrafast photonics,” Adv. Mater. 21, 3874–3899 (2009). [CrossRef]  

4. Q. Bao, H. Zhang, Y. Wang, Z. Ni, Y. Yan, Z. X. Shen, K. P. Loh, and D. Y. Tang, “Atomic-layer graphene as a saturable absorber for ultrafast pulsed lasers,” Adv. Funct. Mater. 19, 3077–3083 (2009). [CrossRef]  

5. Z. P. Sun, T. Hasan, F. Torrisi, D. Popa, G. Privitera, F. Q. Wang, F. Bonaccorso, D. M. Basko, and A. C. Ferrari, “Graphene mode-locked ultrafast laser,” ACS Nano 4, 803–810 (2010). [CrossRef]   [PubMed]  

6. H. Zhang, D. Y. Tang, L. M. Zhao, Q. L. Bao, and K. P. Loh, “Large energy mode locking of an erbium-doped fiber laser with atomic layer graphene,” Opt. Express 17, 17630–17635 (2009). [CrossRef]   [PubMed]  

7. Y. W. Song, S. Y. Jang, W. S. Han, and M. K. Bae, “Graphene mode-lockers for fiber lasers functioned with evanescent field interaction,” Appl. Phys. Lett. 96, 051122 (2010). [CrossRef]  

8. J. Liu, S. Wu, Q. H. Yang, and P. Wang, “Stable nanosecond pulse generation from a graphene-based passively Q-switched Yb-doped fiber laser,” Opt. Lett. 36, 4008–4010 (2011). [CrossRef]   [PubMed]  

9. G. Sobon, J. Sotor, and K. M. Abramski, “Passive harmonic mode-locking in Er-doped fiber laser based on graphene saturable absorber with repetition rates scalable to 2.22 GHz,” Appl. Phys. Lett. 100, 161109 (2012). [CrossRef]  

10. C. J. Zhao, H. Zhang, X. Qi, Y. Chen, Z. Wang, S. C. Wen, and D. Y. Tang, “Ultra-short pulse generation by a topological insulator based saturable absorber,” Appl. Phys. Lett. 101, 211106 (2012). [CrossRef]  

11. C. J. Zhao, Y. Zou, Y. Chen, Z. Wang, S. Lu, H. Zhang, S. C. Wen, and D. Y. Tang, “Wavelength-tunable picosecond soliton fiber laser with topological insulator: Bi2Se3 as a mode locker,” Opt. Express 20, 27888–27895 (2012). [CrossRef]   [PubMed]  

12. F. Bernard, H. Zhang, S. P. Gorza, and P. Emplit, “Towards mode-locked fiber laser using topological insulators,” in Nonlinear Photonics (Optical Society of America, 2012), NTh1A.5.

13. P. Tang, X. Zhang, C. J. Zhao, Y. Wang, H. Zhang, D. Y. Shen, S. C. Wen, D. Y. Tang, and D. Y. Fan, “Topological insulator: Bi2Te3 saturable absorber for the passive Q-switching operation of an in-band pumped 1645-nm Er: YAG ceramic laser,” IEEE Photonics J. 5, 1500707 (2013). [CrossRef]  

14. H. Zhang, C. X. Liu, X. L. Qi, X. Dai, Z. Fang, and S. C. Zhang, “Topological insulators in Bi2Se3, Bi2Te3, Sb2Te3 with a single dirac cone on the surface,” Nat. Phys. 5, 438–442 (2009). [CrossRef]  

15. S. B. Lu, C. J. Zhao, Y. Zou, S. Chen, Y. Chen, Y. Li, H. Zhang, S. C. Wen, and D. Y. Tang, “Third order nonlinear optical property of Bi2Se3,” Opt. Express 21, 2072–2082 (2013). [CrossRef]   [PubMed]  

16. N. Bansal, Y. S. Kim, E. Edrey, M. Brahlek, Y. Horibe, K. Iida, M. Tanimura, G. H. Li, T. Feng, and H. D. Lee, “Epitaxial growth of topological insulator Bi2Se3 film on si (111) with atomically sharp interface,” Thin Solid Films 520, 224–229 (2011). [CrossRef]  

17. Y.-H. Lin, C.-Y. Yang, J.-H. Liou, C.-P. Yu, and G.-R. Lin, “Using graphene nano-particle embedded in photonic crystal fiber for evanescent wave modelocking of fiber laser,” Opt. Express 21, 16763–16777 (2013). [CrossRef]   [PubMed]  

18. S. Lu, S. Chen, Z. Zheng, H. Zhang, C. Zhao, and S. Wen, “Saturable absorption in graphene at 800 nm band,” Proc. SPIE 8555, 855512 (2012). [CrossRef]  

19. Z. Q. Luo, M. Zhou, J. Weng, G. Huang, H. Xu, C. Ye, and Z. P. Cai, “Graphene-based passively Q-switched dual-wavelength erbium-doped fiber laser,” Opt. Lett. 35, 3709–3711 (2010). [CrossRef]   [PubMed]  

20. A. Martinez, K. Fuse, B. Xu, and S. Yamashita, “Optical deposition of graphene and carbon nanotubes in a fiber ferrule for passive mode-locked lasing,” Opt. Express 18, 23054–23061 (2010). [CrossRef]   [PubMed]  

21. D. P. Zhou, L. Wei, B. Dong, and W. K. Liu, “Tunable passively Q-switched erbium-doped fiber laser with carbon nanotubes as a saturable absorber,” IEEE Photonics Technol. Lett. 22, 9–11 (2010). [CrossRef]  

22. R. Paschotta, R. Häring, E. Gini, H. Melchior, U. Keller, H. Offerhaus, and D. Richardson, “Passively Q-switched 0.1-mJ fiber laser system at 1.53 μ m,” Opt. Lett. 24, 388–390 (1999). [CrossRef]  

23. D. Popa, Z. P. Sun, T. Hasan, F. Torrisi, F. Wang, and A. C. Ferrari, “Graphene Q-switched, tunable fiber laser,” Appl. Phys. Lett. 98, 073106 (2011). [CrossRef]  

24. W. Cao, H. Wang, A. P. Luo, Z. C Luo, and W. C. Xu, “Graphene-based, 50 nm wide-band tunable passively Q-switched fiber laser,” Laser Phys. Lett. , 9, 54–57 (2012). [CrossRef]  

25. Z. Q. Luo, M. Zhou, D. Wu, C. Ye, J. Weng, J. Dong, H. Xu, Z. Cai, and L. Chen, “Graphene-induced nonlinear four-wave-mixing and its application to multiwavelength Q-switched rare-earth-doped fiber lasers,” J. Lightwave Technol. 29, 2732–2739 (2011). [CrossRef]  

26. J. J. Zayhowski and P. L. Kelley, “Optimization of Q-switched lasers,” IEEE J. Quantum Electron. 27, 2220–2225 (1991). [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) The AFM image and (b) the height profile diagram of the few-layer TI: Bi2Se3. Inset: the as-prepared TI: Bi2Se3 suspension.
Fig. 2
Fig. 2 (a) The transmission spectrum of the few-layer TI: Bi2Se3 sample, and (b) its saturable absorption characteristic at 800 nm wavelength by the Z-scan technique.
Fig. 3
Fig. 3 The experimental setup of the TI-based passively Q-switched YDF laser. (a) the transmission spectrum of the FBG2, (b) the optical deposition setup, (c) the image of the fiber-ferrule facet after depositing the TI nanosheets.
Fig. 4
Fig. 4 The Q-switched pulse trains under different pump powers Pp, (a) Pp=42.7 mW, (b) Pp=49.2 mW, (c) Pp=81.7 mW, (d) Pp=106.2 mW.
Fig. 5
Fig. 5 The typical Q-switching characteristics at the pump power of 77.7 mW. (a) output optical spectrum, (b) the oscilloscope trace of the Q-switched pulse train, (c) single pulse envelope, (d) RF output spectrum, and Inset: broadband RF output spectrum.
Fig. 6
Fig. 6 (a) The pulse duration and the pulse repetition rate as a function of pump power, (b) the average output power and the pulse energy as a function of pump power.
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.