Open Access
29 March 2016 Fiber-integrated tungsten disulfide saturable absorber (mirror) for pulsed fiber lasers
Author Affiliations +
Abstract
We propose two schemes for achieving tungsten disulfide (WS2)-based saturable absorber (SA) and saturable absorber mirror (SAM). By utilizing the pulsed laser deposition method, we grow the WS2 film on microfiber to form an evanescent field interaction SA device. Incorporating this SA device into a common ring-cavity erbium-doped fiber (EDF) laser, stably passive mode-locking can be achieved with pulse duration of 395 fs and signal-to-noise ratio of 64 dB. We also produce a fiber tip integrated WS2 -SAM by utilizing the magnetron sputtering technique (MST). This new type of SAM combines the WS2 layer as SA and gold mirror as high reflective mirror. By employing the WS2 -SAM, we construct the linear-cavity EDF lasers, and achieve passive mode-locking operation with pulse duration of ∼1  ns and SNR of ∼61  dB. We further achieve stably passive Q-switching operation with pulse duration of ∼160  ns and pulse energy of 54.4 nJ. These fiber-integrated SAs and SAMs have merits of compactness and reliability, paving the way for the development of new photonic devices such as SAs for pulsed laser technology.

1.

Introduction

Pulsed fiber lasers have attracted much attention owing to many potential applications in such areas as environmental sensing, biomedical diagnostics, and nonlinear frequency generation.1,2 Mainly two schemes are selected for generating pulsed laser emission—active mode-locking or Q-switching by employing costly and complex electrically driven modulators,35 and passive mode-locking or Q-switching by incorporating an artificial or real saturable absorber (SA) as an intensity-dependent optical switch. Compared with active scheme, passive scheme has the advantages of compactness, simplicity, and flexibility. In the past, the nonlinear polarization rotation (NPR), nonlinear optical loop mirror (NOLM),6 and semiconductor saturable absorber mirror (SESAM)7 were the dominant techniques with wide application in commercial laser system. Both NPR and NOLM are the fiber nonlinearity-based artificial SAs with higher damage threshold, but are also highly sensitive to environmental fluctuations. SESAM requires complicated fabrication and packaging process, and has limited bandwidth, few picoseconds response times. From the last decade, carbon nanotube (CNT)815 has emerged as the first case of nanomaterial SA for mode-locking, which has one-dimensional (1-D) nanostructure and owns advantages such as ultrafast photoresponse, easy fabrication, and low cost. Gold nanorod (GNR)16,17 is another familiar 1-D nanomaterial SA, which has two surface plasmon resonance absorption bands and has been utilized for Q-switching or mode-locking. These 1-D nanomaterial SAs have absorption peaks related with the tube diameter (for CNT) or the aspect ratio (for GNR), thus leading to relatively narrowband operation restricted by the optimal absorption band. For wideband operation, the 1-D nanomaterial SAs require the combination of their nanotubes with a broad diameters distribution or nanorods with different aspect ratio. Unlike CNT or GNR, graphene18,19 has a Dirac-like electronic band structure with unique zero bandgap, which endows it with remarkable optical properties (i.e., ultrafast photoresponse, ultrabroadband absorption) and electrical properties (i.e., remarkable electron mobility). Extensively, research2047 has employed graphene as a candidate material for the development of functional photonic technologies, e.g., ultrafast mode-locker, broadband polarizer, and phase shifter. But graphene also holds two main disadvantages, the weak modulation depth (typically 1.3% per layer20) and the difficulty of creating an optical bandgap. Therefore, substantial endeavors have been focused on developing new SA beyond graphene from other layered crystal,48 such as topological insulators (TIs),4982 transition mental dichalcogenides (TMDs),83114 including of molybdenum disulfide (MoS2) or tungsten disulfide (WS2), as well as their diselenide analogues (MoSe2, MoTe2, WSe2, and WTe2) and black phosphorus.115118 The series of TIs (i.e., Bi2Te3, Sb2Te3, Bi2Se3, and so on) demonstrate the characteristics of a small bandgap in the bulk state and a gapless metallic state in the edge/surface. Both the bulk- and nanostructured-TIs have been applied in pulsed laser cavity as the SA device. The TMDs possesses a thickness-dependent electronic and optical property—their bulk states have indirect bandgap with weak light–matter interaction while monolayer or few-layer structures are direct bandgap semiconductor with enhanced light activity. For example, a bulk WS2 is a semiconductor with bandgap of 1.4 eV (0.886  μm), while its monolayer has a direct bandgap of 2.1 eV (0.59  μm). Its thickness-dependent bandgap and electronic band structure endow it with many potential optical applications in optical fields that require strong light–matter interaction. These newly emerged two-dimensional (2-D) materials possess remarkable abilities in the field of optoelectronics and nonlinear photonics, such as broadband SA, high third-order nonlinear susceptibility, and ultrafast carrier dynamics.

Several methods have been used to obtain the TMD SAs, such as solution processing method,84 evanescent field-interaction method,39,52 composite films composed of nanomaterial flakes in a polyvinyl alcohol (PVA) host,43,53 and the bottom-up growth techniques such as chemical vapor deposition (CVD),92,99 pulsed laser deposition (PLD),6364 and the magnetron sputtering technique (MST).6061 In the laser cavity, these SAs are usually pasted on fiber ferrules, or embedded in the air channels of photonic crystal fiber (PCF),30,36,67,77,108,119124 or deposited on microfiber or side polished fiber (SPF).56,58,59 The fiber ferrule-type SAs have inherently short nonlinear interaction length. The composite films SAs could maintain the thermal stability of SA materials, but is vulnerable to destruction by high power operation. Moreover, it should be guaranteed that light must transmit through the SA materials. The PCF-based SAs can supply strong interaction effect and large nonlinear effect, but their fabrication processes are complicated according to the previous report.108 The microfiber-based SAs are attractive for long light–matter interaction length, making such SAs as light modulator and high nonlinear device simultaneously. As in Ref. 52, Luo et al. demonstrated the combination of high nonlinearity induced by the real part of third-order nonlinear optical property in passively mode-locked fiber lasers with high-repetition rates. The SPF-based SAs have the merits of high power tolerance and longer light–matter interaction length, but the SPF requires accurate polishing technique and usually causes unwanted polarization-dependent insertion loss.56,58 CVD offers a scalable method for the production of monolayer or few-layer TMD (i.e., MoS2 or WS2), but the film growth is limited by the low nucleation rate on bare substrates, and pretreatment of the substrate is often necessary to seed the TMD growth. The PLD and MST can produce the film by irradiating the target placed under vacuum chamber, and the film can be directly deposited on the substrate (i.e., quartz glass substrate,85 microfiber,63 SPF,61 or fiber-tip80).

In this paper, we propose two schemes for achieving new WS2-based SA and SAM devices as shown in Fig. 1. By utilizing the PLD method, we grow the WS2 film on microfiber to form an evanescent field interaction SA device as in Fig. 1(a), which would have the combined advantages from the strong nonlinear optical response in material together with the sufficiently long-range interaction length in fiber taper. Incorporating this SA device into a common ring-cavity erbium-doped fiber (EDF) laser, stably passive mode-locking can be achieved with pulse duration of 395 fs and signal-to-noise ratio of 64 dB. On the other hand, we produce a fiber tip integrated WS2-SAM by utilizing the MST method. This new type of SAM combines the WS2 layer as SA and gold mirror as high reflective mirror, as shown in Fig. 1(b). By employing the WS2-SAM, we construct the linear-cavity EDF lasers, and achieve passive mode-locking operation with pulse duration of 1  ns and SNR of 61  dB. We further achieve stably passive Q-switching operation with pulse duration of 160  ns and pulse energy of 54.4 nJ. These fiber-integrated SAs and SAMs have merits of compactness and reliability. Our research, accompanied by studies from Bogusławski et al.,60,61 paves the way for the development of new photonic devices such as SAs for pulsed laser technology.

Fig. 1

(a) WS2-SA integrated on microfiber and (b) WS2-SAM integrated on fiber-tip.

OE_55_8_081318_f001.png

2.

Fabrication, Characterization of Fiber-Integrated SA/SAM and Their Application in Pulsed Fiber Laser

2.1.

Microfiber-Based WS2-SA and its Application in Passively Mode-Locked EDF Laser

We employed the PLD method to fabricate the SA device. In this process, the WS2 target was placed into a vacuum chamber where the vacuum degree was set at 5×104  Pa. A high energy Nd:YAG laser (SL II-10, Surelite) could emit 2  mJ/pulse laser beam, which was delivered into the chamber and focused on the target to inspire out of the plasma plume. When arrived to the microfiber, the W and S elements would grow on the side surface of microfiber. In the experiment, the deposition time was 2 h, and the deposition temperature was fixed at room temperature. To verify that the film was really deposited on microfiber, we executed a scanning electron microscope (SEM) on the WS2 film morphology, as shown in Fig. 2. The waist region of microfiber had a diameter of 16  μm. The guided light in taper region would effectively penetrate into the film and would be modulated along the microfiber. Figure 2(a) shows that a layer of WS2 film clearly stuck tightly on the side of microfiber. The thickness of WS2 film was measured to be 1  μm. Figure 2(b) shows the film morphology. It illustrates that the film was quite different from the previously reported SA with relatively uniform size and thickness by LPE method.

Fig. 2

SEM characteristic of the microfiber-based WS2-SA. (a) WS2 layer on fiber and (b) film morphology.

OE_55_8_081318_f002.png

For checking the Raman shift property of the as-prepared materials, the Raman spectra were measured by using a Raman spectrometer (LabRAM HR Evolution) with a laser at 514 nm. Figure 3(a) shows the measured Raman spectra of WS2 film on quartz glass together with that of bare quartz glass. Notably, these peaks were also observed for the film with locations of LA(M) at 174  cm1, 2LA(M) at 350  cm1, E2g1 at 356  cm1 and A1g at 420.7  cm1, where the LA(M) and 2LA(M) are the longitudinal acoustic modes, E2g1 is an in-plane optical mode, and A1g corresponds to the out-of-plane vibrations along the c-axis direction of the S atoms.

Fig. 3

(a) Raman spectra of bare quartz glass and WS2 film on quartz glass. (b) Measured linear transmission (inset) and nonlinear transmission.

OE_55_8_081318_f003.png

We also measured the linear and nonlinear transmission of the SA device, as shown in Fig. 3(b). The linear transmission was at the level of 65.6% at 1560 nm by using an ASE source (Glight, 1250 to 1650 nm) and optical spectrum analyzer (OSA). The power-dependent nonlinear transmission is the key parameter to evaluate the mode-locking ability of SA. The SA of our fiber-integrated WS2-SA was investigated by standard two-arm experiment. A homemade fs laser (central wavelength: 1562 nm, repetition rate: 22.5 MHz, pulse duration: 650 fs, average power: 12 mW) was utilized as test source, a variable optical attenuator was applied to continuously change the input optical intensity into the sample. A 50:50 optical coupler (OC) was used to split the laser into two arms with the 50% arm for power-dependent transmission measurement of SA device and the 50% arm for reference. A two-channel power meter with measuring range from 10  μW to 10 mW was used to measure the power. The modulation depth, saturation intensity, and nonsaturable loss were 7.8%, 189  MW/cm2, and 25.7%, respectively. Furthermore, the SA might suffer from two-photon absorption as the transmittance declined when the intensity of input light exceeds around 450  MW/cm2.

Figure 4 shows the schematic of mode-locked fiber laser with our WS2-SA device. The pump source was a laser diode (LD) with emission centered at 974.5 nm. A piece of 2.4 m EDF was used as the laser gain medium with absorption coefficient of 25  dB/m at 980 nm (IsoGainTM I-25, Fibercore). The pump was delivered into EDF via a 980/1550 fused wavelength-division multiplexer (WDM) coupler. A polarization independent isolator (PI-ISO), placed after the EDF, was used to ensure unidirectional operation and eliminate undesired feedback from the output end facet. A fused fiber OC was used to extract 30% energy from the cavity. A polarization controller (PC), consisting of three spools of SMF-28 fiber, was placed in the ring cavity after the ISO. The WS2-SA was inserted between the PC and the WDM coupler. Apart from the gain fiber, all the fiber devices in cavity were made by SMF-28 fiber. The laser performance was observed using an OSA (Yokogawa, AQ6370B), 1 GHz digital oscilloscope (Tektronix, DPO7104C), 3 GHz RF spectrum analyzer (Agilent, N9320A) coupled with a 15 GHz photodetector (EOT, ET-3500FEXT), and an optical autocorrelator (APE, PulseCheck).

Fig. 4

Schematic of mode-locked fiber laser with WS2-SA.

OE_55_8_081318_f004.png

Figure 5(a) shows the typical spectrum of mode-locked pulses at the pump power of 65 mW. The generated optical soliton was centered at 1560 nm with a 3 dB spectral width of 6.75 nm. The appearance of pronounced Kelly sidebands indicated that the laser was in soliton operation. The radio frequency (RF) spectrum of the laser was shown in Fig. 5(b). The fundamental repetition frequency was 19.57 MHz with an SNR of 64 dB measured with a 1 kHz resolution bandwidth (RBW). Figure 5(c) shows the autocorrelation trace. It had a full width at half maximum (TFWHM) of 609 fs, corresponding to pulse duration (Δt) of 395  fs if a sech2 pulse profile was assumed. The time-band product (TBP) was calculated to be 0.328 and deviated from the transform-limited value of 0.315, indicating that some chirp was included in the output pulse that had a broader bottom. The average output power was 1.5 mW, thus the output pulse energy was about 76.6 pJ.

Fig. 5

Performance of mode-locked fiber laser. (a) Spectrum, (b) RF spectrum measured with a 1 kHz RBW, and (c) measured pulse duration.

OE_55_8_081318_f005.png

2.2.

Fiber-Tip-Integrated WS2-SAM and its Application in Passively Mode-Locked and Q-Switched EDF Lasers

The MST was employed to fabricate the integrated WS2-SAM. The WS2 target was with the diameter of 49 mm, thickness of 3 mm, and purity of more than 99.5%. All the coating processes occurred in vacuum with the sputtering insert argon gas at pressure of 103  Pa. During the deposition process, the AC voltage was applied to be 100 W in the chamber. The sputtered tungsten and sulfur atoms were ejected out from the target and then condensed on the fiber end. The synthesis lasted around 1 h at room temperature, forming a thin WS2 film as the SA layer. Subsequently, the Au target was excited under the DC voltage for 3 min until the Au atoms formed a tightly thin film with thickness of 150  nm on the SA layer. The gold film (GF) here not only functioned as a high reflective mirror, but also as a protective barrier that isolated the inner SA layer from the contamination, corrosion, and oxidation. Figure 6(a) shows the three-dimensional (3-D) view of the fabricated sample by confocal scanning microscopy. It indicated that the device was compact. To testify whether the SA was grown on the fiber end, we took out several samples from the chamber before the Au-film fabrication, which was direct deposited WS2 film on pollution-free quartz glasses in the same conditions as production WS2-SAM. Figure 6(b) shows the SEMs of the as-prepared sample at different scales. It illustrates that a layer of WS2 film was successively deposited on the quartz glass. The film was composited by thin layer of nanoparticles with diameters from 10 nm to around 40 nm, thus we can deduce that the WS2 film were also deposited on the fiber tip.

Fig. 6

(a) 3-D image of WS2-SAM and (b) SEM of deposited WS2 film.

OE_55_8_081318_f006.png

The Raman spectrum of the as-grown WS2 was measured by a Raman spectrometer (LabRAM HR Evolution) with laser at 488 nm. Figure 7(a) shows the characteristic Raman bands, e.g., two optical phonon modes (E2g1 at 356  cm1 and A1g at 417.5  cm1) and typical longitudinal acoustic modes 2LA(M) at 349.8  cm1, where the E2g1 is an in-plane optical mode and A1g corresponds to the out-of-plane vibrations along the c-axis direction of the S atoms. The main Raman bands agreed with the earlier reports. The linear transmission of the sample was measured in the range from 1000 to 2000 nm by using an ASE source and OSA. It was at the level of 92.8±0.6%, and the transmittance at 1560 nm was 93.1%, as shown in Fig. 7(b). The nonlinear absorption curve gave a modulation depth of 4.48%, saturation intensity of 138  MW/cm2, and nonsaturable loss of 2%, as shown in Fig. 7(b). The nonsaturable loss here might be the smallest value when compared with other WS2-SA reported in Refs. 103 and 105. We tend to believed that this remarkable improvement comes from the compactness of our device—no extra insertion loss is imported in this SAM. This design can allow the reflectivity up to 98% after the SA layer is saturable and only Au film is remained as a high reflective mirror.

Fig. 7

(a) Raman spectra of WS2 film. (b) Measured linear transmission (inset) and nonlinear transmission.

OE_55_8_081318_f007.png

A simplest linear cavity was implemented to obtain self-starting mode-locking operation, as shown in Fig. 8. The pump source was a 976 nm LD with maximum output power of 410 mW. A 980/1550 nm WDM was utilized to deliver the 976 nm pump light into the laser cavity and extract the laser output. The cavity was constructed by fiber Bragg grating (FBG), EDF, and WS2-SAM. An FBG centered at 1560 nm was utilized as output coupler with 3-dB bandwidth of 0.2 nm and reflectivity of 88.52%. A 13 cm EDF (Liekki 110-4/125) was used as active media with absorption coefficient of 250  dB/m at 980 nm. The WS2-SAM, serving as a light modulator and high reflective mirror, was spliced directly with the EDF. The other fibers in cavity were single mode fiber (Corning, SMF-28) with 16.2-cm length. At the output end, an isolator was added to eliminate undesired feedback from the output end facet.

Fig. 8

Schematic of mode-locked fiber laser with WS2-SAM.

OE_55_8_081318_f008.png

The CW lasing state started to occur around the pump power of 6 mW, and the mode-locking operation would self-start and quickly stabilize when the pump power was beyond 30 mW. Figure 9 shows the measured spectrum, RF spectra, and pulse traces of the laser at the pump power of 410 mW. Figure 9(a) shows the spectrum of mode-locked pulses. The generated pulses are centered at 1560 nm with a 3-dB bandwidth of 0.031 nm. The mode-locking operated at a fundamental frequency of 352 MHz, corresponding well to the cavity round trip time. The RF spectrum in a 2 GHz span was also measured with RBW of 10 kHz and SNR of 61  dB, as displayed in Fig. 9(b), certifying the stability of this mode-locking EDFL. Restricted by the bandwidth of our FBG, the output pulse had a width to be 1.04 ns, as shown in Fig. 9(c). Its long range stability had been characterized in a span of 1  μs by the inset. In the full pump range, the fundamental mode-locking state remained stable and no pulse-breaking and harmonic waves were observed on the oscilloscope. The power characteristic curve was shown in Fig. 9(d). It can be seen that the maximum output was 5.28 mW at 1560 nm.

Fig. 9

Performance of mode-locked fiber laser. (a) Spectrum, (b) RF spectrum measured with a 10 kHz RBW, (c) measured pulse duration and pulse trace, and (d) output power versus input power.

OE_55_8_081318_f009.png

It was noted that no PC was inserted into the cavity for stabilizing the pulse, unlike that in Ref. 100. Also, the pump threshold was much lower than the mode-locking EDFL based on MoS2-PVA SA.100 This is because that the high reflection of narrowband FBG and WS2-SAM was beneficial to the rapid growth of cavity energy. As a fiber-integrated photonic device, this SAM also demonstrated well-thermal stability as the EDFLs worked at the maximum pump power. It was interesting that the generated pulse was restricted in the ns regime in our experiment, which might come from the inherent property of WS2 nanoparticles. It was expected that the shorter pulse could be generated when the quality of WS2 film was improved by optimizing the deposition condition and applying the postprocessing.

The deposition condition and integrated fiber type would have great impact on the laser performance. Apart from the above WS2-SAM sample, we also integrated the WS2-SAM on a polarization-maintaining fiber (PMF, PM980). Before the deposition, the vacuum pressure specification in chamber was settled to 103  Pa to remove various impurity gases. During the deposition, the RF power was fixed at 100 W. The SMF tips were coated with a thin WS2 nanomaterial functioned as SA layer during 1.5-h deposition process. Then the gold film was deposited under direct-current magnetron sputtering at the power of 80 W with 300-nm thickness.

In this case, the same linear-cavity structure was implemented to obtain the pulse operation by the WS2-SAM on PMF, whereas the EDF was replaced by 23-cm length of Liekki 110-4/125 and the total cavity was 44-cm long. It was very interesting that the fiber laser could easily achieve the stable Q-switched pulse output at such a simple structure. The laser performance was measured and shown in Fig. 10. The Q-switched pulse started to occur around the pump power of 50 mW, and then the Q-switched operation would maintain stably even the pump power up to 600 mW. Figures 10(a) and 10(b) show the measured optical spectrum and RF spectrum at pump power of 450 mW. It can be seen that the central wavelength was at 1560 nm with a 3-dB bandwidth of 0.025 nm, while the fundamental frequency was at 296.7 kHz with an RF SNR of 40 dB under RBW of 1 kHz. Figure 10(c) plotted the typical Q-switched pulse trains. At pump of 50 mW, the pulses had an FWHM of 750  ns. The duration of Q-switched pulses decreased when the pump power increased, which was agreed with Q-switching properties. At the maximum pump of 600 mW, the minimum pulse width was measured to be 160  ns. Figure 10(d) shows the repetition rate and output power versus pump power. The repetition rate varied with a wide range from 91 to 318 kHz as a function of pump power. It was interesting to see that the output power almost linearly increased even when the pump power was up to 600 mW, indicating that the SAM functioned well without any extra protection at this pump level. The maximum output power was 17.3 mW, corresponding to the pulse energy of 54.4 nJ.

Fig. 10

Performance of Q-switched fiber laser. (a) Spectrum, (b) RF spectrum measured with a 1 kHz RBW, (c) measured pulse traces, and (d) output power versus input power.

OE_55_8_081318_f010.png

For the direct comparison with the previous works using CNT/graphene/TI as Q-switcher in EDF laser, we summarized their Q-switched pulse parameters, as listed in Table 1. It is notable that our work has the largest output power or pulse energy; meanwhile, the pulse width of 160 ns is the shortest for the Q-switched EDFL with linear-cavity structure. It is expected that the pulse duration can be further shortened down to 100 ns by decreasing the length of the laser cavity and employing the high absorption gain fiber pumped by high power LD at 980 nm.

Table 1

Parameters of Q-switched EDFLs using CNT/graphene/TI and WS2-SAM.

SA typeOperating wavelength (nm)Cavity length (cm)Frequency range (kHz)Pulse width (ns)Pulse energy (nJ)References
CNT1534.110.555 to 143.56620.9113
1534.111014.5 to 141.43307.312
Graphene1538.33731.7 to 236.320633.2125
TI1543.26512.6 to 177.72177.555
WS2-SAM15604491 to 31816054.4This work

3.

Conclusions

In summary, the fiber-integrated WS2-based SA and SAM were fabricated with the interesting merits of compactness and reliability. The compactness of these devices can effectively reduce their insertion loss, thus lowering the pump threshold for mode-locking or Q-switching. For the case of WS2-SAM, the nonsaturable loss can be regarded as nearly negligible after the SA layer was saturable. The suppression of unwanted loss (both the insertion loss and nonsaturable loss) allows these devices to possess more reliable performance than those CNT/graphene/TIs sandwiched between fiber connecters. For the WS2-SAM in experiment, no scheme was applied to protect them from the thermal damage. Once the Q-switching state self-started, it would remain stable even at the maximum power of 600 mW, and occur repeatedly with the booting of pump LD. It is believed that the fiber-integrated SAM can serve as a candidate practical SAM operating at watt-level pump.

As the library of 2-D materials grows, a number of studies emerge to produce and characterize SAs using heterostructures126 or graphene and TMD–polymer composite.127 New exotic properties are likely to emerge when different materials are combined. Unlike the inkjet printing method, the fabrication technique of WS2-based SA or SAM can be applied to materials of other TMDs and the TI series. Both the PLD and PVD technologies allow the control over the ratio of different elements in the fabricated 2-D material film, and also allow the grown of different 2-D materials to build superlattice that can be artificially engineered their bandgap and thickness. However, substantial endeavors still need to improve the fiber-integrated WS2-SA or SAM in the development of pulsed fiber lasers, especially to find the optimized deposition condition in the fabrication procedure, e.g., the deposition temperature and time, the lattice match of different SA materials.

Acknowledgments

This work was supported by the National Natural Science Foundation of China (Nos. 61205080 and 61308049), the Natural Science Fund of Guangdong Province (2014A030313387), the Science and Technology Project of Shenzhen City (JCYJ20140418091413568, 20140418095735582, JCYJ20150324140036862, JCYJ20130329142116637, JCYJ20150324140036870, and JCYJ20140418091413577), and the Youth Science and Technology Innovation Talents of Guangdong (No. 2014TQ01X539).

References

1. 

X. Liu et al., “Distributed ultrafast fibre laser,” Sci. Rep., 5 9101 (2015). http://dx.doi.org/10.1038/srep09101 SRCEC3 2045-2322 Google Scholar

2. 

X. Liu and Y. Cui, “Flexible pulse-controlled fiber laser,” Sci. Rep., 5 9399 (2015). http://dx.doi.org/10.1038/srep09399 SRCEC3 2045-2322 Google Scholar

3. 

H. Chen et al., “Flexible rectangular wave-breaking-free pulse generation in actively mode-locked ytterbium-doped fiber laser,” Opt. Express, 22 (22), 26449 –26456 (2014). http://dx.doi.org/10.1364/OE.22.026449 OPEXFF 1094-4087 Google Scholar

4. 

H. Chen et al., “All fiber actively mode-locked ytterbium-doped laser with large range temporal tunability,” IEEE Photonics Technol. Lett., 26 (17), 1786 –1789 (2014). http://dx.doi.org/10.1109/LPT.2014.2336371 Google Scholar

5. 

H. Chen et al., “Versatile long cavity widely tunable pulsed Yb-doped fiber laser with up to 27655th harmonic mode locking order,” Opt. Express, 23 (2), 1308 –1318 (2015). http://dx.doi.org/10.1364/OE.23.001308 OPEXFF 1094-4087 Google Scholar

6. 

H. Chen et al., “0.4  μJ, 7 kW ultrabroadband noise-like pulse direct generation from an all-fiber dumbbell-shaped laser,” Opt. Lett., 40 (23), 5490 –5493 (2015). http://dx.doi.org/10.1364/OL.40.005490 OPLEDP 0146-9592 Google Scholar

7. 

U. Keller et al., “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). http://dx.doi.org/10.1109/2944.571743 IJSQEN 1077-260X Google Scholar

8. 

S. Yamashita et al., “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). http://dx.doi.org/10.1364/OL.29.001581 OPLEDP 0146-9592 Google Scholar

9. 

F. Wang et al., “Wideband tuneable, nanotube mode-locked, fibre laser,” Nat. Nanotechnol., 3 738 –742 (2008). http://dx.doi.org/10.1038/nnano.2008.312 NNAABX 1748-3387 Google Scholar

10. 

X. M. Liu et al., “Versatile multiwavelength ultrafast fiber laser mode-locked by carbon nanotubes,” Sci. Rep., 3 2718 (2013). http://dx.doi.org/10.1038/srep02718 Google Scholar

11. 

M. Zhang et al., “Mid-infrared Raman-soliton continuum pumped by a nanotube-mode-locked sub-picosecond Tm-doped MOPFA,” Opt. Express, 21 23261 (2013). http://dx.doi.org/10.1364/OE.21.023261 Google Scholar

12. 

B. Dong et al., “Wide pulse-repetition rate range tunable nanotube Q-switched low threshold erbium-doped fiber laser,” IEEE Photonics Technol. Lett., 22 (24), 1853 –1855 (2010). http://dx.doi.org/10.1109/LPT.2010.2089507 Google Scholar

13. 

B. Dong et al., “Short linear-cavity Q-switched fiber laser with a compact short carbon nanotube based saturable absorber,” Opt. Fiber Technol., 17 (2), 105 –107 (2011). http://dx.doi.org/10.1016/j.yofte.2010.12.001 1068-5200 Google Scholar

14. 

X. Li et al., “Nonlinear absorption of SWNT film and its effects to the operation state of pulsed fiber laser,” Opt. Express, 22 17227 –17235 (2014). http://dx.doi.org/10.1364/OE.22.017227 OPEXFF 1094-4087 Google Scholar

15. 

J. Z. Wang et al., “Pulse dynamics in carbon nanotube mode-locked fiber lasers near zero cavity dispersion,” Opt. Express, 23 9947 –9958 (2015). http://dx.doi.org/10.1364/OE.23.009947 OPEXFF 1094-4087 Google Scholar

16. 

Z. Kang et al., “Passively mode-locking induced by gold nanorods in erbium-doped fiber lasers,” Appl. Phys. Lett., 103 041105 (2013). http://dx.doi.org/10.1063/1.4816516 APPLAB 0003-6951 Google Scholar

17. 

Z. Kang et al., “Passively mode-locked fiber lasers at 1039 and 1560 nm based on a common gold nanorod saturable absorber,” Opt. Mater. Express, 5 794 –801 (2015). http://dx.doi.org/10.1364/OME.5.000794 Google Scholar

18. 

Q. L. Bao et al., “Atomic layer graphene as saturable absorber for ultrafast pulsed laser,” Adv. Funct. Mater., 19 (19), 3077 –3083 (2009). http://dx.doi.org/10.1002/adfm.200901007 AFMDC6 1616-301X Google Scholar

19. 

T. Hasan et al., “Nanotube–polymer composites for ultrafast photonics,” Adv. Mater., 21 (38–39), 3874 –3899 (2009). http://dx.doi.org/10.1002/adma.200901122 Google Scholar

20. 

A. Martinez and Z. Sun, “Nanotube and graphene saturable absorbers for fibre lasers,” Nat. Photonics, 7 842 –845 (2013). http://dx.doi.org/10.1038/nphoton.2013.304 NPAHBY 1749-4885 Google Scholar

21. 

H. Zhang et al., “Large energy mode locking of an erbium-doped fiber laser with atomic layer graphene,” Opt. Express, 17 7630 –17635 (2009). http://dx.doi.org/10.1364/OE.17.007630 OPEXFF 1094-4087 Google Scholar

22. 

H. Zhang et al., “Large energy soliton erbium-doped fiber laser with a graphene–polymer composite mode locker,” Appl. Phys. Lett., 95 (14), 141103 (2009). http://dx.doi.org/10.1063/1.3244206 APPLAB 0003-6951 Google Scholar

23. 

Z. Q. Luo et al., “Graphene-based passively Q-switched dual-wavelength erbium-doped fiber laser,” Opt. Lett., 35 3709 –3711 (2010). http://dx.doi.org/10.1364/OL.35.003709 OPLEDP 0146-9592 Google Scholar

24. 

H. Zhang et al., “Graphene mode locked, wavelength-tunable, dissipative soliton fiber laser,” Appl. Phys. Lett., 96 (11), 111 –112 (2010). http://dx.doi.org/10.1063/1.3367743 APPLAB 0003-6951 Google Scholar

25. 

Q. Bao et al., “Graphene–polymer nanofiber membrane for ultrafast photonics,” Adv. Funct. Mater., 20 (5), 782 –791 (2010). http://dx.doi.org/10.1002/adfm.200901658 AFMDC6 1616-301X Google Scholar

26. 

Q. L. Bao and K. P. Loh, “Graphene photonics, plasmonics, and broadband optoelectronic devices,” ACS Nano, 6 3677 –3694 (2012). http://dx.doi.org/10.1021/nn300989g ANCAC3 1936-0851 Google Scholar

27. 

H. Zhang et al., “Z-scan measurement of the nonlinear refractive index of graphene,” Opt. Lett., 37 (11), 1856 –1858 (2012). http://dx.doi.org/10.1364/OL.37.001856 Google Scholar

28. 

M. Zhang et al., “Tm-doped fiber laser mode-locked by graphene–polymer composite,” Opt. Express, 20 25077 –25084 (2012). http://dx.doi.org/10.1364/OE.20.025077 OPEXFF 1094-4087 Google Scholar

29. 

J. Ma et al., “Graphene mode-locked femtosecond laser at 2  μm wavelength,” Opt. Lett., 37 2085 –2087 (2012). http://dx.doi.org/10.1364/OL.37.002085 OPLEDP 0146-9592 Google Scholar

30. 

Y. H. Lin et al., “Using graphene nano-particle embedded in photonic crystal fiber for evanescent wave mode-locking of fiber laser,” Opt. Express, 21 16763 –16776 (2013). http://dx.doi.org/10.1364/OE.21.016763 OPEXFF 1094-4087 Google Scholar

31. 

S. Jaroslaw et al., “Simultaneous mode-locking at 1565 nm and 1944 nm in fiber laser based on common graphene saturable absorber,” Opt. Express, 21 18994 –19002 (2013). http://dx.doi.org/10.1364/OE.21.018994 OPEXFF 1094-4087 Google Scholar

32. 

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). http://dx.doi.org/10.1063/1.4704913 APPLAB 0003-6951 Google Scholar

33. 

J. Q. Zhao et al., “Graphene-oxide-based q-switched fiber laser with stable five-wavelength operation,” Chin. Phys. Lett., 29 (11), 114206 (2012). http://dx.doi.org/10.1088/0256-307X/29/11/114206 CPLEEU 0256-307X Google Scholar

34. 

S. S. Huang et al., “Tunable and switchable multi-wavelength dissipative soliton generation in a graphene oxide mode-locked Yb-doped fiber laser,” Opt. Express, 22 11417 –11426 (2014). http://dx.doi.org/10.1364/OE.22.011417 OPEXFF 1094-4087 Google Scholar

35. 

Y. Meng et al., “Multiple-soliton dynamic patterns in a graphene mode-locked fiber laser,” Opt. Express, 20 (6), 6685 –6692 (2012). http://dx.doi.org/10.1364/OE.20.006685 OPEXFF 1094-4087 Google Scholar

36. 

J. Q. Zhao et al., “Cladding-filled graphene in a photonic crystal fiber as a saturable absorber and its first application for ultrafast all-fiber laser,” Opt. Eng., 52 (10), 106105 (2013). http://dx.doi.org/10.1117/1.OE.52.10.106105 Google Scholar

37. 

H. Ahmad et al., “Photonic crystal fiber based dual-wavelength Q-switched fiber laser using graphene oxide as a saturable absorber,” Appl. Opt., 53 (16), 3581 –3586 (2014). http://dx.doi.org/10.1364/AO.53.003581 Google Scholar

38. 

S. S Huang et al., “Soliton rains in a graphene-oxide passively mode-locked ytterbium-doped fiber laser with all-normal dispersion,” Laser Phys. Lett., 11 025102 (2014). http://dx.doi.org/10.1088/1612-2011/11/2/025102 1612-2011 Google Scholar

39. 

A. P. Luo et al., “Microfiber-based, highly nonlinear graphene saturable absorber for formation of versatile structural soliton molecules in a fiber laser,” Opt. Express, 22 27019 –27025 (2014). http://dx.doi.org/10.1364/OE.22.027019 OPEXFF 1094-4087 Google Scholar

40. 

S. S. Huang, Y. G. Wang and P. G. Yan, “High-order harmonic mode locking in an all-normal dispersion Yb-doped fiber laser with a graphene-oxide saturable absorber,” Laser Phys., 24 (1), 015001 (2014). http://dx.doi.org/10.1088/1054-660X/24/1/015001 Google Scholar

41. 

R. Y. Lin et al., “Bright- and dark-square-pulse generated from a graphene-oxide mode-locked ytterbium-doped fiber laser,” IEEE Photonics J., 6 (3), 1 (2014). http://dx.doi.org/10.1109/JPHOT.2014.2319099 Google Scholar

42. 

Y. Meng et al., “16.1  μm high-order passive harmonic mode locking in a fiber laser based on graphene saturable absorber,” Opt. Express, 22 (24), 29921 –29926 (2014). http://dx.doi.org/10.1364/OE.22.029921 OPEXFF 1094-4087 Google Scholar

43. 

J. Q. Zhao et al., “An L-band graphene-oxide mode-locked fiber laser delivering bright and dark pulses,” Laser Phys., 23 (7), 075105 (2013). http://dx.doi.org/10.1088/1054-660X/23/7/075105 Google Scholar

44. 

J. Q. Zhao, P. G. Yan and S. C. Ruan, “Observations of three types of pulses in an erbium-doped fiber laser by incorporating a graphene saturable absorber,” Appl. Opt., 52 8465 –8470 (2013). http://dx.doi.org/10.1364/AO.52.008465 Google Scholar

45. 

K. Wu et al., “Towards low timing phase noise operation in fiber lasers mode locked by graphene oxide and carbon nanotubes at 1.5  μm,” Opt. Express, 23 (1), 501 (2015). http://dx.doi.org/10.1364/OE.23.000501 OPEXFF 1094-4087 Google Scholar

46. 

X. T. Gan et al., “Graphene-assisted all-fiber phase shifter and switching,” Optica, 2 468 –471 (2015). http://dx.doi.org/10.1364/OPTICA.2.000468 Google Scholar

47. 

X. Li et al., “High-power graphene mode-locked Tm/Ho co-doped fiber laser with evanescent field interaction,” Sci. Rep., 5 16624 (2015). http://dx.doi.org/10.1038/srep16624 Google Scholar

48. 

A. K. Geim and I. V. Grigorieva, “Van der Waals heterostructures,” Nature, 499 419 –425 (2013). http://dx.doi.org/10.1038/nature12385 Google Scholar

49. 

C. Zhao et al., “Wavelength-tunable picosecond soliton fiber laser with topological insulator: Bi2Se3 as a mode locker,” Opt. Express, 20 (25), 27888 –27895 (2012). http://dx.doi.org/10.1364/OE.20.027888 OPEXFF 1094-4087 Google Scholar

50. 

S. Lu et al., “Third order nonlinear optical property of Bi2Se3,” Opt. Express, 21 (2), 2072 –2082 (2013). http://dx.doi.org/10.1364/OE.21.002072 Google Scholar

51. 

Z. Q. Luo et al., “1.06  μm Q-switched ytterbium-doped fiber laser using few-layer topological insulator Bi2Se3 as a saturable absorber,” Opt. Express, 21 29516 –29522 (2013). http://dx.doi.org/10.1364/OE.21.029516 OPEXFF 1094-4087 Google Scholar

52. 

Z. C. Luo et al., “2 GHz passively harmonic mode-locked fiber laser by a microfiber-based topological insulator saturable absorber,” Opt. Lett., 38 5216 –5219 (2013). http://dx.doi.org/10.1364/OL.38.005216 OPLEDP 0146-9592 Google Scholar

53. 

H. Liu et al., “Femtosecond pulse generation from a topological insulator mode-locked fiber laser,” Opt. Express, 22 6868 –6873 (2014). http://dx.doi.org/10.1364/OE.22.006868 OPEXFF 1094-4087 Google Scholar

54. 

M. Liu et al., “Dual-wavelength harmonically mode-locked fiber laser with topological insulator saturable absorber,” IEEE Photonics Technol. Lett., 26 983 –986 (2014). http://dx.doi.org/10.1109/LPT.2014.2311101 IPTLEL 1041-1135 Google Scholar

55. 

M. Wu et al., “Nanosecond-switched erbium-doped fiber laser with wide pulse-repetition-rate range based on topological insulator,” IEEE J. Quantum Electron., 50 (6), 393 –396 (2014). http://dx.doi.org/10.1109/JQE.2014.2314774 Google Scholar

56. 

J. Koo et al., “Passively Q-switched 1.56  μm all-fiberized laser based on evanescent field interaction with bulk-structured bismuth telluride topological insulator,” J. Opt. Soc. Am. B, 31 (9), 2157 –2162 (2014). http://dx.doi.org/10.1364/JOSAB.31.002157 JOBPDE 0740-3224 Google Scholar

57. 

Y. Chen et al., “Large energy, wavelength widely tunable, topological insulator Q-switched erbium-doped fiber laser,” IEEE J. Sel. Top. Quantum Electron., 20 (5), 315 –322 (2014). http://dx.doi.org/10.1109/JSTQE.2013.2295196 IJSQEN 1077-260X Google Scholar

58. 

J. Lee et al., “A femtosecond pulse erbium fiber laser incorporating a saturable absorber based on bulk-structured Bi2Te3 topological insulator,” Opt. Express, 22 (5), 6165 –6173 (2014). http://dx.doi.org/10.1364/OE.22.006165 OPEXFF 1094-4087 Google Scholar

59. 

J. Sotor et al., “Mode-locked erbium-doped fiber laser based on evanescent field interaction with Sb2Te3 topological insulator,” Appl. Phys. Lett., 104 (25), 251112 (2014). http://dx.doi.org/10.1063/1.4885371 APPLAB 0003-6951 Google Scholar

60. 

J. Bogusławski et al., “Investigation on pulse shaping in fiber laser hybrid mode-locked by Sb2Te3 saturable absorber,” Opt. Express, 23 29014 –29023 (2015). http://dx.doi.org/10.1364/OE.23.029014 OPEXFF 1094-4087 Google Scholar

61. 

J. Bogusławski et al., “Dissipative soliton generation in Er-doped fiber laser mode-locked by Sb2Te3 topological insulator,” Opt. Lett., 40 (12), 2786 –2789 (2015). http://dx.doi.org/10.1364/OL.40.002786 OPLEDP 0146-9592 Google Scholar

62. 

Y. H. Lin et al., “Using n- and p-type Bi2Te3 topological insulator nanoparticles to enable controlled femtosecond mode-locking of fiber lasers,” ACS Photonics, 2 (4), 481 –490 (2015). http://dx.doi.org/10.1021/acsphotonics.5b00031 Google Scholar

63. 

P. G. Yan et al., “A 2.95 GHz, femtosecond passive harmonic mode-locked fiber laser based on evanescent field interaction with topological insulator film,” Opt. Express, 23 (1), 154 –164 (2015). http://dx.doi.org/10.1364/OE.23.000154 OPEXFF 1094-4087 Google Scholar

64. 

P. G. Yan et al., “A practical topological insulator saturable absorber for mode-locked fiber laser,” Sci. Rep., 5 8690 (2015). http://dx.doi.org/10.1038/srep08690 SRCEC3 2045-2322 Google Scholar

65. 

W. J. Liu et al., “Generation of dark solitons in erbium-doped fiber lasers based Sb2Te3 saturable absorbers,” Opt. Express, 23 26023 –26031 (2015). http://dx.doi.org/10.1364/OE.23.026023 OPEXFF 1094-4087 Google Scholar

66. 

L. N. Duan et al., “Passively harmonic mode-locked fiber laser with a high signal-to-noise ratio via evanescent-light deposition of bismuth telluride topological insulator based saturable absorber,” IEEE Photonics J., 7 1 –7 (2015). http://dx.doi.org/10.1109/JPHOT.2015.2404315 Google Scholar

67. 

P. G. Yan et al., “Topological insulator solution filled in photonic crystal fiber for passive mode-locked fiber laser,” IEEE Photonics Technol. Lett., 27 (3), 264 –267 (2015). http://dx.doi.org/10.1109/LPT.2014.2361915 Google Scholar

68. 

Y. C. Meng et al., “High power L-band mode-locked fiber laser based on topological insulator saturable absorber,” Opt. Express, 23 23053 –23058 (2015). http://dx.doi.org/10.1364/OE.23.023053 OPEXFF 1094-4087 Google Scholar

69. 

M. Liu et al., “Microfiber-based, highly-nonlinear topological insulator photonic device for the formation of versatile multi-soliton patterns in a fiber laser,” J. Lightwave Technol., 33 2056 –2061 (2015). http://dx.doi.org/10.1109/JLT.2015.2396939 JLTEDG 0733-8724 Google Scholar

70. 

D. Mao et al., “Soliton fiber laser mode locked with two types of film-based Bi2Te3 saturable absorbers,” Photonics Res., 3 (2), A43 –A46 (2015). http://dx.doi.org/10.1364/PRJ.3.000A43 Google Scholar

71. 

B. Guo et al., “Observation of bright-dark soliton pair in a mode-locked fiber laser with topological insulator,” IEEE Photonics Technol. Lett., 27 (7), 701 –704 (2015). http://dx.doi.org/10.1109/LPT.2015.2390212 Google Scholar

72. 

M. Liu et al., “Dissipative rogue waves induced by long-range chaotic multi-pulse interactions in a fiber laser with a topological insulator-deposited microfiber photonics device,” Opt. Lett., 40 2667 –2770 (2015). http://dx.doi.org/10.1364/OL.40.004767 OPLEDP 0146-9592 Google Scholar

73. 

J. L. Xu et al., “Ultrasensitive nonlinear absorption response of large-size topological insulator and application in low-threshold bulk pulsed lasers,” Sci. Rep., 5 14856 (2015). http://dx.doi.org/10.1038/srep14856 SRCEC3 2045-2322 Google Scholar

74. 

Y. J. Sun et al., “Passively Q-switched tri-wavelength Yb3+:GdAl3(BO3)4 solid-state laser with topological insulator Bi2Te3 as saturable absorber,” Photonics Res., 3 A97 –A101 (2015). http://dx.doi.org/10.1364/PRJ.3.000296 Google Scholar

75. 

H. Ahmad et al., “Tunable S-Band Q-switched fiber laser using Bi2Se3 as the saturable absorber,” IEEE Photonics J., 7 (3), 1 (2015). http://dx.doi.org/10.1109/JPHOT.2015.2433020 Google Scholar

76. 

Q. Wang et al., “Wide spectral and wavelength-tunable dissipative soliton fiber laser with topological insulator nano-sheets self-assembly films sandwiched by PMMA polymer,” Opt. Express, 23 (6), 7681 –7693 (2015). http://dx.doi.org/10.1364/OE.23.007681 OPEXFF 1094-4087 Google Scholar

77. 

L. Gao et al., “Ultrafast pulse mode-locked by topological insulator nanosheets interacting with photonic crystal fiber: from anomalous dispersion to normal dispersion,” IEEE Photonics J., 7 1 –8 (2015). http://dx.doi.org/10.1109/JPHOT.2015.2402594 Google Scholar

78. 

B. Guo et al., “Dual-wavelength soliton mode-locked fiber laser with a WS2-based fiber taper,” IEEE Photonics Technol. Lett., 28 (3), 323 –326 (2016). http://dx.doi.org/10.1109/LPT.2015.2495330 Google Scholar

79. 

W. J. Liu et al., “70-fs mode-locked erbium-doped fiber laser with topological insulator,” Sci. Rep., 5 19997 (2016). http://dx.doi.org/10.1038/srep19997 SRCEC3 2045-2322 Google Scholar

80. 

P. G. Yan et al., “Multi-pulses dynamic patterns in a topological insulator mode-locked ytterbium-doped fiber laser,” Opt. Commun., 335 65 –72 (2015). http://dx.doi.org/10.1016/j.optcom.2014.09.009 OPCOB8 0030-4018 Google Scholar

81. 

L. Lu et al., “All-normal dispersion passively mode-locked Yb-doped fiber laser with Bi2Te3 absorber,” Opt. Eng., 54 046101 (2015). http://dx.doi.org/10.1117/1.OE.54.4.046101 Google Scholar

82. 

P. G. Yan et al., “Q-switched fiber laser using a fiber-tip-integrated TI saturable absorption mirror,” IEEE Photonics J., 8 (1), 1 –6 (2016). http://dx.doi.org/10.1109/JPHOT.2015.2509864 IPTLEL 1041-1135 Google Scholar

83. 

K. P. Wang et al., “Ultrafast saturable absorption of two-dimensional MoS2 nanosheets,” ACS Nano, 7 9260 –9267 (2013). http://dx.doi.org/10.1021/nn403886t ANCAC3 1936-0851 Google Scholar

84. 

Q. Cui et al., “Transient absorption microscopy of monolayer and bulk WSe2,” ACS Nano, 8 2970 –2976 (2014). http://dx.doi.org/10.1021/nn500277y ANCAC3 1936-0851 Google Scholar

85. 

S. Wang et al., “Broadband few-layer MoS2 saturable absorbers,” Adv. Mater., 26 3538 –3544 (2014). http://dx.doi.org/10.1002/adma.201306322 ADVMEW 0935-9648 Google Scholar

86. 

H. Zhang et al., “Molybdenum disulfide (MoS2) as a broadband saturable absorber for ultra-fast photonics,” Opt. Express, 22 7249 –7260 (2014). http://dx.doi.org/10.1364/OE.22.007249 OPEXFF 1094-4087 Google Scholar

87. 

F. Bonaccorso and Z. P. Sun, “Solution processing of graphene, topological insulators and other 2D crystals for ultrafast photonics,” Opt. Mater. Express, 4 63 –78 (2014). http://dx.doi.org/10.1364/OME.4.000063 Google Scholar

88. 

R. I. Woodward et al., “Tunable Q-switched fiber laser based on saturable edge-state absorption in few-layer molybdenum disulfide (MoS2),” Opt. Express, 22 (25), 31113 –31122 (2014). http://dx.doi.org/10.1364/OE.22.031113 OPEXFF 1094-4087 Google Scholar

89. 

R. I. Woodward et al., “Wideband saturable absorption in few-layer molybdenum diselenide (MoSe2) for Q-switching Yb-, Er- and Tm-doped fiber lasers,” Opt. Express, 23 20051 –20061 (2015). http://dx.doi.org/10.1364/OE.23.020051 OPEXFF 1094-4087 Google Scholar

90. 

Y. J. Sun et al., “Comparison of MoS2 nanosheets and hierarchical nanospheres in the application of pulsed solid-state lasers,” Opt. Mater. Express, 5 2924 –2932 (2015). http://dx.doi.org/10.1364/OME.5.002924 Google Scholar

91. 

Y. J. Sun et al., “Wavelength-tunable, passively Q-switched Yb3+:Ca3Y2(BO3)4 solid state laser using MoS2 saturable absorber,” Mater. Lett., 160 268 –270 (2015). http://dx.doi.org/10.1016/j.matlet.2015.07.128 MLETDJ 0167-577X Google Scholar

92. 

H. Xia et al., “Ultrafast erbium-doped fiber laser mode-locked by a CVD-grown molybdenum disulfide (MoS2) saturable absorber,” Opt. Express, 22 17341 –17348 (2014). http://dx.doi.org/10.1364/OE.22.017341 OPEXFF 1094-4087 Google Scholar

93. 

J. Du et al., “Ytterbium-doped fiber laser passively mode locked by few-layer molybdenum disulfide (MoS2) saturable absorber functioned with evanescent field interaction,” Sci. Rep., 4 6346 (2014). http://dx.doi.org/10.1038/srep06346 SRCEC3 2045-2322 Google Scholar

94. 

R. Khazaeizhad et al., “Mode-locking of Er-doped fiber laser using a multilayer MoS2 thin film as a saturable absorber in both anomalous and normal dispersion regimes,” Opt. Express, 22 23732 –23742 (2014). http://dx.doi.org/10.1364/OE.22.023732 OPEXFF 1094-4087 Google Scholar

95. 

Y. Z. Huang et al., “Widely-tunable, passively Q-switched erbium-doped fiber laser with few-layer MoS2 saturable absorber,” Opt. Express, 22 25258 –25266 (2014). http://dx.doi.org/10.1364/OE.22.025258 OPEXFF 1094-4087 Google Scholar

96. 

Z. Q. Luo et al., “1-, 1.5-, and 2-μm fiber lasers Q-switched by a broadband few-layer MoS2 saturable absorber,” J. Lightwave Technol., 32 (24), 4077 –4084 (2014). http://dx.doi.org/10.1109/JLT.2014.2362147 JLTEDG 0733-8724 Google Scholar

97. 

H. Liu et al., “Femtosecond pulse erbium-doped fiber laser by a few-layer MoS2 saturable absorber,” Opt. Lett., 39 4591 –4594 (2014). http://dx.doi.org/10.1364/OL.39.004591 OPLEDP 0146-9592 Google Scholar

98. 

M. Liu et al., “Microfiber-based few-layer MoS2 saturable absorber for 2.5 GHz passively harmonic mode-locked fiber laser,” Opt. Express, 22 22841 –22846 (2014). http://dx.doi.org/10.1364/OE.22.022841 OPEXFF 1094-4087 Google Scholar

99. 

R. Khazaeizhad et al., “Passively mode-locked fiber laser based on CVD WS2,” in 2015 Conf. on Lasers and Electro-Optics (CLEO), (2015). Google Scholar

100. 

K. Wu et al., “463-MHz fundamental mode-locked fiber laser based on few-layer MoS2 saturable absorber,” Opt. Lett., 40 (7), 1374 –1377 (2015). http://dx.doi.org/10.1364/OL.40.001374 OPLEDP 0146-9592 Google Scholar

101. 

M. Zhang et al., “Solution processed MoS2-PVA composite for sub-bandgap mode-locking of a wideband tunable ultrafast Er:fiber laser,” Nano Res., 8 1522 –1534 (2015). http://dx.doi.org/10.1007/s12274-014-0637-2 1998-0124 Google Scholar

102. 

H. Wang et al., “Ethanol catalytic deposition of MoS2 on tapered fiber,” Photonics Res., 3 (3), A102 (2015). http://dx.doi.org/10.1364/PRJ.3.00A102 Google Scholar

103. 

D. Mao et al., “WS2 mode-locked ultrafast fiber laser,” Sci. Rep., 5 7965 (2015). http://dx.doi.org/10.1038/srep07965 SRCEC3 2045-2322 Google Scholar

104. 

P. G. Yan et al., “Microfiber-based WS2-film saturable absorber for ultra-fast photonics,” Opt. Mater. Express, 5 (3), 479 –489 (2015). http://dx.doi.org/10.1364/OME.5.000479 Google Scholar

105. 

K. Wu et al., “WS2 as a saturable absorber for ultrafast photonic applications of mode-locked and Q-switched lasers,” Opt. Express, 23 (9), 11453 –11461 (2015). http://dx.doi.org/10.1364/OE.23.011453 OPEXFF 1094-4087 Google Scholar

106. 

M. W. Jung et al., “Mode-locked, 1.94-μm, all-fiberized laser using WS2-based evanescent field interaction,” Opt. Express, 23 19996 –20006 (2015). http://dx.doi.org/10.1364/OE.23.019996 OPEXFF 1094-4087 Google Scholar

107. 

J. Lin et al., “Wavelength-tunable Yb-doped passively Q-switching fiber laser based on WS2 saturable absorber,” Opt. Express, 23 29059 –29064 (2015). http://dx.doi.org/10.1364/OE.23.029059 OPEXFF 1094-4087 Google Scholar

108. 

P. G. Yan et al., “Passively mode-locked fiber laser by a cell-type WS2 nanosheets saturable absorber,” Sci. Rep., 5 12587 (2015). http://dx.doi.org/10.1038/srep12587 SRCEC3 2045-2322 Google Scholar

109. 

A. P. Luo et al., “Few-layer MoS2-deposited microfiber as highly nonlinear photonic device for pulse shaping in a fiber laser [invited],” Photonics Res., 3 A69 –A78 (2015). http://dx.doi.org/10.1364/PRJ.3.000A69 Google Scholar

110. 

M. Zhang et al., “Yb- and Er-doped fiber laser Q-switched with an optically uniform, broadband WS2 saturable absorber,” Sci. Rep., 5 17482 (2015). http://dx.doi.org/10.1038/srep17482 SRCEC3 2045-2322 Google Scholar

111. 

D. Mao et al., “WS2 saturable absorber for dissipative soliton mode locking at 1.06 and 1.55  μm,” Opt. Express, 23 27509 –27519 (2015). http://dx.doi.org/10.1364/OE.23.027509 OPEXFF 1094-4087 Google Scholar

112. 

L. Li et al., “WS2/fluorine mica (FM) saturable absorbers for all-normal-dispersion mode-locked fiber laser,” Opt. Express, 23 28698 –28706 (2015). http://dx.doi.org/10.1364/OE.23.028698 OPEXFF 1094-4087 Google Scholar

113. 

P. G. Yan et al., “Self-starting mode-locking by fiber-integrated WS2 saturable absorber mirror [invited],” IEEE J. Sel. Top. Quantum Electron., (2016). http://dx.doi.org/10.1109/JSTQE.2016.2515546 Google Scholar

114. 

B. H. Chen et al., “Q-switched fiber laser based on transition metal dichalcogenides MoS2, MoSe2, WS2, and WSe2,” Opt. Express, 23 (20), 26723 –26737 (2015). http://dx.doi.org/10.1364/OE.23.026723 OPEXFF 1094-4087 Google Scholar

115. 

S. B. Lu et al., “Broadband nonlinear optical response in multi-layer black phosphorus: an emerging infrared and mid-infrared optical material,” Opt. Express, 23 (9), 11183 –11194 (2015). http://dx.doi.org/10.1364/OE.23.011183 OPEXFF 1094-4087 Google Scholar

116. 

Y. Chen et al., “Mechanically exfoliated black phosphorus as a new saturable absorber for both Q-switching and mode-locking laser operation,” Opt. Express, 23 (10), 12823 –12833 (2015). http://dx.doi.org/10.1364/OE.23.012823 OPEXFF 1094-4087 Google Scholar

117. 

Z. C. Luo et al., “Microfiber-based few-layer black phosphorus saturable absorber for ultra-fast fiber laser,” Opt. Express, 23 20030 –20039 (2015). http://dx.doi.org/10.1364/OE.23.020030 OPEXFF 1094-4087 Google Scholar

118. 

J. Sotor et al., “Ultrafast thulium-doped fiber laser mode locked with black phosphorus,” Opt. Lett., 40 (16), 3885 –3888 (2015). http://dx.doi.org/10.1364/OL.40.003885 OPLEDP 0146-9592 Google Scholar

119. 

J. C. Knight et al., “All-silica single-mode optical fiber with photonic crystal cladding,” Opt. Lett., 21 1547 –1549 (1996). http://dx.doi.org/10.1364/OL.21.001547 OPLEDP 0146-9592 Google Scholar

120. 

P. G. Yan et al., “Double cladding seven-core photonic crystal fibers with different GVD properties and fundamental supermode output,” J. Lightwave Technol., 31 3658 –3662 (2013). http://dx.doi.org/10.1109/JLT.2013.2286210 JLTEDG 0733-8724 Google Scholar

121. 

H. F. Wei et al., “A compact seven-core photonic crystal fiber supercontinuum source with 42.3 W output power,” Laser Phys. Lett., 10 045101 (2013). http://dx.doi.org/10.1088/1612-2011/10/4/045101 1612-2011 Google Scholar

122. 

P. G. Yan et al., “Polarization dependent visible supercontinuum generation in the nanoweb fiber,” Opt. Express, 19 4985 –4990 (2011). http://dx.doi.org/10.1364/OE.19.004985 OPEXFF 1094-4087 Google Scholar

123. 

P. G. Yan et al., “Numerical simulation on the coherent time-critical 2-5  μm supercontinuum generation in an As2S3 microstructured optical fiber with all-normal flat-top dispersion profile,” Opt. Commun., 293 133 –138 (2013). http://dx.doi.org/10.1016/j.optcom.2012.11.093 OPCOB8 0030-4018 Google Scholar

124. 

P. G. Yan et al., “Supercontinuum generation in a photonic crystal fibre,” Chin. Phys. Lett., 21 1093 –1095 (2004). http://dx.doi.org/10.1088/0256-307X/21/6/032 CPLEEU 0256-307X Google Scholar

125. 

L. Wei et al., “Graphene-based Q-switched erbium-doped fiber laser with wide pulse-repetition-rate range,” IEEE Photonics Technol. Lett., 24 (4), 1853 –1855 (2012). http://dx.doi.org/10.1109/LPT.2011.2178821 Google Scholar

126. 

H. Mu et al., “Graphene-Bi2Te3 heterostructure as saturable absorber for short pulse generation,” ACS Photonics, 2 (7), 832 –841 (2015). http://dx.doi.org/10.1021/acsphotonics.5b00193 Google Scholar

127. 

Y. Q. Jiang et al., “Broadband and enhanced nonlinear optical response of MoS2/graphene nanocomposites for ultrafast photonics applications,” Sci. Rep., 5 16372 (2015). http://dx.doi.org/10.1038/srep16372 SRCEC3 2045-2322 Google Scholar

Biography

Hao Chen is studying for a master’s degree at the Shenzhen Key Lab of Laser Technology, Key Laboratory of Advanced Optical Precision Manufacturing Technology of Guangdong Higher Education Institutes, Shenzhen University, China. His current research interests concentrate on passively mode-locked fiber lasers based on topological insulator/few-layer transition-mental dichalcogenides and black phosphorus.

Peiguang Yan received his PhD in the College of Physics Science from Nankai University, China, in 2005. He joined Shenzhen University (China) in 2005 as a lecturer and became an associate professor (2007–2012) and promoted to professor in 2013. His research interests include PCF-based nonlinear fiber optics and passively mode-lock fiber laser by 2D materials. He has authored/coauthored over 50 technical publications.

Biographies for the other authors are not available.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 Unported License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Hao Chen, Ling Li, Shuangchen Ruan, Tuan Guo, and Peiguang Yan "Fiber-integrated tungsten disulfide saturable absorber (mirror) for pulsed fiber lasers," Optical Engineering 55(8), 081318 (29 March 2016). https://doi.org/10.1117/1.OE.55.8.081318
Published: 29 March 2016
Lens.org Logo
CITATIONS
Cited by 28 scholarly publications.
Advertisement
Advertisement
KEYWORDS
Mode locking

Mirrors

Fiber lasers

Pulsed fiber lasers

Pulsed laser operation

Tungsten

Raman spectroscopy

Back to Top