ArticlePDF Available

Lead Sulfide Saturable Absorber Based Passively Mode-Locked Tm-Doped Fiber Laser

Authors:

Abstract and Figures

In this paper, a passively mode-locked Tm-doped fiber laser by employing lead sulfide (PbS) nanoparticles as the saturable absorber (SA) is successfully demonstrated in the 2 μm region for the first time, to the best of our knowledge. Measured by a home-made balanced twin-detector setup at 2 μm, the PbS SA was characterized by the modulation depth of 10.69%, the non-saturable loss of 74.27%, and the saturable peak intensity of 8.62 MW/cm2, respectively. The laser delivered stable conventional soliton with a pulse duration of 1.24 ps and a repetition rate of 21.93 MHz. The center wavelength and 3 dB bandwidth are 1957.37 nm and 3.43 nm, respectively. Additionally, the second-order harmonic soliton pulses with a repetition rate of 43.86 MHz were also observed by further increasing the pump power. Our work reveals that PbS material is a reliable SA applied for pulse generation in the 2 μm spectral region.
Content may be subject to copyright.
This work is licensed under a Creative Commons Attribution 4.0 License. For more information, see https://creativecommons.org/licenses/by/4.0/.
This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/JPHOT.2020.2964981, IEEE
Photonics Journal
Lead Sulfide Saturable Absorber Based Passively Mode-locked
Tm-doped Fiber Laser
Fei Liu1, Ying Zhang2, Xiaodong Wu1, Jianfeng Li1, Senior Member, IEEE, Fei Yan1,
Xiaohui Li2, Abdul Qyyum2, Zhu Hu1, Chen Zhu3, and Yong Liu1, Senior Member, IEEE
1State Key Laboratory of Electronic Thin Films and Integrated Devices, School of Optoelectronic Science and Engineering, University of
Electronic Science and Technology of China (UESTC), Chengdu 610054, China
2School of Physics and Information Technology, Shaanxi Normal University, Xi’an 710119, China
3Science and Technology on Solid-State Laser Laboratory, Beijing 100015, China
This work is supported by National Natural Science Foundation of China under Grant 61435003, Grant 61722503, and Grant 61421002,
the Fundamental Research Funds for the Central Universities under Grant ZYGX2019Z012, the Science and Technology on
Solid-State Laser Laboratory, the Joint Fund of Ministry of Education for Equipment Pre-research under Grant 6141A02033411, and
the Field Funding for Equipment Pre-research under Grant 61404140106. Corresponding author: Jian-feng Li (e-mail:
lijianfeng@uestc.edu.cn).
Abstract: In this paper, a passively mode-locked Tm-doped fiber laser by employing lead sulfide (PbS) nanoparticles as the saturable
absorber (SA) is successfully demonstrated in the 2 μm region for the first time, to the best of our knowledge. Measured by a
home-made balanced twin-detector setup at 2 μm, the PbS SA was characterized by the modulation depth of 10.69%, the
non-saturable loss of 74.27%, and the saturable peak intensity of 8.62 MW/cm2, respectively. The laser delivered stable conventional
soliton with a pulse duration of 1.24 ps and a repetition rate of 21.93 MHz. The center wavelength and 3 dB bandwidth are 1957.37 nm
and 3.43 nm, respectively. Additionally, the second-order harmonic soliton pulses with a repetition rate of 43.86 MHz were also
observed by further increasing the pump power. Our work reveals that PbS material is a reliable SA applied for pulse generation in the
2 μm spectral region.
Index Terms: Tm-doped fiber laser, lead sulfide, saturable absorber, mode-locking.
1. Introduction
In recent years, 2 μm mode-locked Tm-doped fiber lasers have attracted much attention owing to its unique applied
advantages in mid-infrared wavelength conversion, medical diagnosis, laser LIDAR, nanoscale imaging, special material
processing and so on [1-7]. A key point for the formation of mode-locking is the saturable absorption effect, which can be
achieved using either artificial SAs based on fiber’s nonlinear Kerr effect or real SAs based on the nonlinear absorption
effect [8-10]. However, the former usually needs strong pump strength and long cavity length due to the high nonlinear
threshold of silica fiber, which inevitably leads to bulky construction and is not conducive to realize the self-starting of
mode-locking [11,12]. Therefore, the real SAs with a low response threshold are alternatively adopted to achieve
mode-locking. In the past few decades, semiconductor saturable absorber mirrors (SESAMs) [13-15] has been regarded as
one of the most important elements for mode-locking. However, it usually requires complex fabrication process and costly
systems. Nowadays, a series of low dimensional SAs including CNTs [16-18], graphene [12, 19-21], transition metal
dichalcogenides (TMDCs) (e.g., WS2, MoS2, WSe2, MoSe2 and ReS2) [22-27] and black phosphorus [28-31] have been
regarded as good substitutes for SESAMs. These SAs have been widely used to achieve mode-locking at different optical
bands. Among them, CNTs are widely adopted because of its good stability and fast response time. Especially, Liu et al.
revealed the dynamics buildup process of soliton molecules, harmonic mode-locking states, and the transition from
Q-Switching to mode-locking in a CNT SA based ultrafast fiber laser [32-34], providing the theoretical and experimental
basis for the researchers to better understand the soliton dynamics in the fiber laser. Nevertheless, the accurate control of
CNT’s diameter is required to match the desired laser wavelengths with the absorption, which restricts its broadband
tunability. Although graphene possesses an excellent broadband absorption performance as a result of its zero bandgap,
suitable modulation depth for pulse generation cannot be ensured due to the weak absorption of approximately 2.3% per
layer. Owing to the unique absorption, TMDCs have been also employed to achieve pulsed fiber laser in the visible spectral
range. But large direct bandgaps of TMDCs result in that suitable defects should be introduced to extend laser wavelengths,
which increases their fabricating complexity and limits their application in mid-infrared region. In contrast, black phosphorus
as an emerging SA in recent years, offers a layer-dependent bandgap from 0.3 (bulk) to 2 eV (monolayer), widely covering
the bandgap range of graphene and TMDCs. Unfortunately, it is easily oxidized and has a bad stability.
This work is licensed under a Creative Commons Attribution 4.0 License. For more information, see https://creativecommons.org/licenses/by/4.0/.
This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/JPHOT.2020.2964981, IEEE
Photonics Journal
Most recently, semiconductor PbS nanoparticles as the oldest light-harvesting materials for infrared detectors [35],
biomarkers [36, 37] and solar cells [38, 39], presented its great optical potential from near-infrared to mid-infrared region (1
~3.2 μm) [40, 41]. Owing to its narrow bandgap (0.41 eV), bandgap tunability and good long-term stability, PbS
nanoparticles have attracted people’s intensive attention and been regarded as an effective SA material to overcome the
shortcomings of the aforementioned materials. Lee et al. demonstrated a 1.55 μm passively Q-switched erbium-doped fiber
laser by using the film-type PbS quantum dots (QDs) material as the SA [42]. After that, based on the PbS/CdS core/shell
QDs SA, Ming et al. obtained the mode-locked Er-doped fiber laser with 54 ps pulse duration and 2.71 mW maximal output
power [43]. Then, at the same waveband, the PbS QDs in polystyrene films was fabricated to realize the Q-switched
Er-doped fiber laser by Sun et al. This fiber laser yielded stable Q-switched pulses with maximum output power of 40.19 mW
and pulse duration of 3.9 μs [44]. Due to the polymer films structure, however, these PbS QDs as SAs have low damage
threshold and consequently limits the performance of mode-locking. By employing the PbS nanoparticles without any
polymer as the SA, Zhang et al. have obtained a 1533 nm mode-locked fiber laser with good stability for more than 1 month,
which proved the potential of PbS nanomaterials in ultra-short pulse generation [45]. Furthermore, using the same PbS
nanoparticles, our research group has successfully demonstrated the Q-switched Dy3+-doped fiber laser that is tunable from
2.71 to 3.08 μm [46]. Although these works indicate that PbS material as the SA is a good candidate for the fiber lasers at
1.5 μm and 3 μm wavelength range, it has never been reported in the 2 μm region.
In our work, we experimentally demonstrated a stable 2 μm passively mode-locked Tm-doped fiber laser using PbS
nanomaterials as the SA. The nonlinear characteristics of PbS were measured by a 2 μm home-made mode-locked fiber
laser. The fiber laser mode-locked by PbS SA delivered stable conventional soliton pulses operating at fundamental
repetition rate state. When further increasing pump power, a second-order harmonic mode-locking was also obtained.
2. Preparation and Characterization of PbS SA
In our experiment, the PbS sample was prepared by using the sol-gel method and mixing PbS powder with acetone solvent
in a volume ratio of 1:3. Stable PbS nanoparticles dispersion was collected from the supernatant after centrifugation to
separating large agglomerations for 30 minutes at 2000 rpm in a vortex mixer (Langyue SK-1, China). For the PbS
nanoparticles dispersion, the scanning electron microscopy (SEM: Hitachi SU8220, Japan) images with different scales (2
μm and 500 nm) are shown in Figs. 1(a) and 1(b). It is seen that PbS nanoparticles obviously present irregular
microstructure, mainly existing in the forms of stereo-structure and polygon. Feng et al. explained the reason why the
irregular shapes of PbS nanoparticles formed [47]. In Fig. 1(c), the transmission electron microscopy (TEM: JEOL JEM-2100,
Japan) image provides that the lateral size of PbS nanoparticles is in the range of 75-200 nm. The measured average size
of individual PbS nanoparticle is approximately 80 nm. Note that some of the larger and the darker nanoparticles were
generated by the agglomeration of smaller particles [48, 49]. Thus, the size distribution of PbS nanoparticles is not as broad
as we can see. Figure 1(d) presents a high resolution TEM (HRTEM: JEOL JEM-2100, Japan) image with 10 nm lateral size.
The observed spacing of lattice fringes is approximately 0.33 nm. A clear cladding was formed by acetone attached solution
during the fabrication process of PbS nanoparticles dispersion, which reveals a signal of good dispersibility. As illustrated in
Fig. 1(e), the energy dispersive spectroscopy (EDS) analysis of PbS nanoparticles was also executed. It is seen that the
chemical composition of the sample corresponds well to the atomic structure of PbS, except for Si and Al derived from the
sample placement stations. To further confirm the pulse generation performance of PbS nanoparticles dispersion, it was
deposited on the end face of fiber jumper (shown in the inset of Fig. 1(f)) and then was fabricated as a fiber-compatible SA.
Fig. 1. Characterization of PbS nanoparticles dispersion. (a) The SEM image with 2 μm scale. (b) The SEM image
with 500 nm scale. (c) The TEM image with 200 nm scale. (d) The HRTEM (high-resolution TEM) image with 10 nm
scale. (e) The EDS analysis. (f) Photograph of PbS nanoparticles dispersion and the fiber-compatible PbS-SA device.
This work is licensed under a Creative Commons Attribution 4.0 License. For more information, see https://creativecommons.org/licenses/by/4.0/.
This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/JPHOT.2020.2964981, IEEE
Photonics Journal
Additionally, we designed a typical balanced twin-detector setup to investigate the nonlinear transmission characteristics
of PbS SA in the experiment, as depicted in Fig. 2(a). The home-made 2 μm all-fiber pulse laser source consists of a
picosecond Tm-doped fiber laser seed and a Tm-doped fiber amplifier. For the amplified picosecond pulse with a maximum
average power of 39.6 mW, its repetition rate, pulse duration, and signal-to-noise ratio (SNR) are 10.27 MHz, 21 ps, and
60.5 dB, respectively. The laser output was divided into a reference path and a measurement path using a 3-dB optical
coupler. To minimize the measuring errors, the average output powers of two optical paths were detected using the same
model power meters with the measurement accuracy of microwatt level. Figure 2(b) presents the nonlinear transmission
curve of PbS SA by fitting the measured data with the following formula [23]:
T (I) =1-ΔT×exp (-I/Isat)-αns, (1)
where T (I) is the transmittance and I is the incident peak intensity. ΔT, αns and Isat represent the modulation depth, the
non-saturable loss and the saturable peak intensity of PbS SA, which are calculated to be 10.69%, 74.27%, and 8.62
MW/cm2, respectively.
Fig. 2. (a) The experimental setup for the nonlinear transmission measurement of PbS SA. (b) Measured nonlinear transmission curves.
3. Experiment Setup
The experimental ring configuration of the Tm-doped fiber laser is presented in Fig. 3. With a (2+1) ×1 pump combiner (ITF,
Canada), a commercial multimode 793 nm laser diode (BWT, China) was utilized as pump source to supply outside energy
for the laser oscillator. A 1.1 m Tm-doped double-cladding fiber (Coractive-DCF-TM-10/128) was served as the gain fiber.
Tm-doped fiber is characterized by an octagonal shaped inner cladding with 0.45 numerical aperture (NA) and the circular
fiber core with 0.22 NA. A polarization-independent isolator (Advanced Photonics, USA) ensured laser’s unidirectional
operation. 10% of a 9:1 optical coupler (OC) (AdValue Photonics, USA) was utilized to output the laser. A polarization
controller (PC) was used to optimize the performance of mode-locking. PbS SA was positioned between the OC and PC.
The total laser cavity length is 9.43 m, including 1.1 m TDF and 8.33 m SMF28e tail fibers of the intra-cavity fiber
components. The anomalous dispersion values of these two fibers at 1.993 µm ware about −84 ps2/km and −80 ps2/km,
respectively [50]. The estimated value of net cavity dispersion is -0.75 ps2. In order to obtain the laser signal, an optical
spectrum analyzer (Yokogawa AQ6375, Japan) with a high resolution of 0.05 nm and an interference autocorrelator (APE
Pulsecheck, Germany) were respectively applied for the real-time measurement of optical spectrum and pulse duration. In
addition, a 2-μm InGaAs photodetector (EOT ET-5000F, USA) with a bandwidth of 12.5 GHz and a response time of 28 ps
was employed to detect the radio-frequency (RF) spectrum and oscilloscope trace.
Fig. 3. Setup of passively mode-locked Tm-doped fiber laser based on PbS SA.
This work is licensed under a Creative Commons Attribution 4.0 License. For more information, see https://creativecommons.org/licenses/by/4.0/.
This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/JPHOT.2020.2964981, IEEE
Photonics Journal
4. Experiment Results and Discussion
In the experiment, once the PC’s position was properly set, stable mode-locking was self-started by simply increasing the
pump power. Figure 4(a) presents the relationship between pump power and output power, as well as the operation state of
the passively mode-locked Tm-doped fiber oscillator. Continuous wave (CW) with the output power of 0.66 mW firstly
appeared at the pump power of 0.88 W. When the pump power was increased to 1.01 W, the laser switched into the
mode-locking regime. Conventional soliton with a maximum output power of 6.34 mW was obtained at the pump power of
1.54 W. A segment of double-cladding single-mode Thulium-doped fiber (DC-SM-TDF) was adopted in our work is the
reason why the pump threshold of the mode-locking is up to ~1 W. We can further improve the output power by optimizing
the parameter performance of PbS SA (such as the deposition thickness and the size of nanoparticles), and decreasing the
intra-cavity loss. Additionally, using a linear cavity mode-locking structure based on PbS SA is also a good method to realize
the higher power ultrashort pulse laser output. Figure 4(b) shows the corresponding optical spectrum of conventional soliton.
Its center wavelength and 3 dB bandwidth are 1957.37 nm and 3.43 nm, respectively. A series of Kelly sidebands
symmetrically distributed around the main peak is a good indication of the conventional soliton operation, which is caused by
the spectral interference of dispersive waves. Some sub-sidebands caused by the four-wave mixing (FWM) between the
orthogonal soliton components in the fiber laser [51], also can be observed on the mode-locked spectrum. Figure 4(c) shows
the measured interference and intensity autocorrelation trace with a 7-ps scanning range. Autocorrelation signal with a full
width at half maximum (FWHM) of 1.91 ps was fitted by using the sech2 function, corresponding to a pulse duration of 1.24
ps. Thus, the calculated time-bandwidth product (TBP) of 0.335 indicates that the mode-locking is almost transform-limited.
In Fig. 4(d), the RF spectrum with a resolution bandwidth of 3 kHz and a scanning range of 2 MHz shows a high signal-noise
ratio of 61.2 dB, which suggests the good uniformity of the pulse train. The measured repetition rate of 21.93 MHz is in
agreement with the theoretical cavity length dependent value, implying that single pulse was obtained per round trip. The
pulse train with the amplitude fluctuation of less than 0.8% is illustrated in the inset of Fig. 4(d), which exhibits the good
stability of the passive mode-locked fiber laser based on PbS SA.
Fig. 4. (a) Output power as a function of pump power. (b) Soliton optical spectrum. (c) Pulse autocorrelation with sech2 fitting.
(d) RF spectrum with the resolution bandwidth of 3 kHz and the scanning range of 2 MHz, inset: Oscilloscope trace.
The above single soliton state maintained at the pump range from 1.01 W to 1.54 W. When pump power exceeded 1.54
W, the multi-pulses appeared. By appropriately adjusting the pump power and the PC’s position, the multi-pulses operated
at the 2nd-order harmonic state. Figure 5 shows an example of this state at the pump power of 1.77 W. The spectrum is
presented in Fig. 5(a), showing a 3-dB bandwidth of 4.22 nm visibly wider than previous fundamental frequency soliton. It is
seen from the inset of Fig. 5(b) that the spectrum peak intensity monotonously decreases with increased order of the RF
spectrum component, and the separation between two adjacent spectrum components always keeps constant, suggesting
the high uniformity of the 2nd harmonic pulse train. The high signal noise ratio of 60.26 dB also indicates its good stability for
This work is licensed under a Creative Commons Attribution 4.0 License. For more information, see https://creativecommons.org/licenses/by/4.0/.
This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/JPHOT.2020.2964981, IEEE
Photonics Journal
pulse generation. Moreover, the RF spectrum with a scanning range of 1 MHz and a resolution bandwidth of 3 kHz was
measured to investigate the quality of the 2nd harmonic soliton pulses. In Fig. 5(c), the measured oscilloscope trace shows
that a stable pulse train uniformly exists with constant pulse separation of 22.8 ns, corresponding to the pulse repetition rate
of 43.86 MHz. This is twice of the fundamental repetition rate of the laser cavity, indicating the 2nd-order harmonic operation.
Figure 5(d) shows the measured autocorrelation trace with sech2 fitting in a 5-ps scan range. According to the FWHM of
1.682 ps when sech2-pulse fit is assumed, the pulse duration of harmonic mode-locking (HML) pulse is estimated to be 1.09
ps. The calculated TBP of 0.36 implies that the pulse is slightly chirped. The average output power is 9.06 mW
corresponding to the pulse energy of 206.8 pJ. Note that it is difficult to further increase the harmonic order in our
experiment, due to the high soliton splitting threshold in the large anomalous dispersion region determined by the soliton
area theorem [52, 53]. By reducing the net anomalous cavity dispersion or inducing high-nonlinear component to reduce the
soliton splitting threshold inside the laser cavity, the higher order harmonic or noise-like pulse could be realized. Until the
maximum pump power of 1.77 W, the mode-locking is still observed, suggesting that PbS SA is undamaged. According to
the intra-cavity laser intensity of 90.6 mW at stable mode-locking operation, the peak power density in the fiber laser is
estimated to be 360 MW/cm2, revealing that the PbS SA has a damage threshold higher than this value. In addition, the
mode-locked fiber laser could stably operate for more than half a month without any deterioration, which shows its good
long-term stability. Note that no Q-switching was observed due to low intra-cavity energy loss and a modulation depth
(10.69%) capable of achieving mode-locking. Many similar phenomena have been reported in graphene, carbon-nanotube,
or alcohol based pulse fiber lasers [54-58]. A key factor to realize the Q-switching operation is controlling the loss of the
oscillator. Thus, the Q-switched pulse can be achieved by adjusting the parameters of PbS SA or changing the operating
condition of this oscillator. For the SA material, we can prepare the PbS SA with a higher modulation depth and saturable
peak intensity to increase the pump threshold of fiber laser. Meanwhile, by changing the splitting ratio of optical coupler from
9/1 to 5/5, the Q-switched state can also be generated in the PbS SA based Tm-doped fiber laser due to the higher
intra-cavity loss. To indicate the unique advantages of PbS SA based mode-locking fiber laser, the mode-locked fiber lasers
with PbS SA and other different SAs were compared in Table 1. It can be seen that PbS SA has a higher modulation depth
appropriate for the mode-locking pulse generation. Moreover, the temporal pulse width of our fiber laser is almost
comparable to those from the lasers using graphene and TMDCs, and it is shorter than the laser output using CNT or
SESAM. The results suggest that PbS nanoparticles dispersion can be regarded as an effective saturable absorber for
ultra-short pulse generation in the Tm-doped fiber laser. This also provides us for a reliable ultra-short pulse source to
realize the nonlinear mid-infrared light generation in the future work.
Fig. 5. The second-order harmonic mode-locking pulse at the pump power of 1.77 W. (a) The optical spectrum. (b) RF spectrum
with the 3 kHz resolution bandwidth and 2 MHz scanning range, inset: RF spectrum with a scanning range of 1.0 GHz. (c)
Oscilloscope trace. (d) Pulse autocorrelation with sech2 fitting in a 5-ps scan range.
This work is licensed under a Creative Commons Attribution 4.0 License. For more information, see https://creativecommons.org/licenses/by/4.0/.
This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/JPHOT.2020.2964981, IEEE
Photonics Journal
Table 1. The comparison of 2 μm Tm-doped fiber lasers mode-locked with different SA materials
Saturable
Absoption
Materials
Modulation
Depth
@ 2 μm
Non-saturable
Loss
@ 2 μm
Pulse
Wavelength
(nm)
Output Pulse
Width
(ps)
References
SWCNT
-
-
1944
10
[59]
graphene
-
-
1940
3.6
[20]
graphene
-
-
1901.6
0.37
[60]
graphene
-
-
1944
0.933
[61]
graphene
0.8%
-
1884
1.2
[62]
graphene
~2.7%
-
1912-1918
6500
[63]
graphene
-
-
1953
2.1
[64]
SESAM
-
-
1980
1.5
[13]
SESAM
20%
16%
1962
18
[14]
SESAM
26%
-
1978
815
[15]
SESAM
20%
16%
1945
0.7
[65]
WS2
~10.9%
38.4%
1925
~1.3
[66]
MoS2
13.6
16.7%
1940
843
[23]
MoS2
12.5%
27.2%
1927
1.51
[67]
MoSe2
4.4%
45.5%
1912.6
0.92
[68]
WSe2
1.83%
87%
1863.96
1.16
[69]
MoTe2
5.7%
70%
1930.22
0.952
[70]
PbS
10.69%
74.27%
1957.37
1961.2
1.24
1.09
This work
5. Conclusions
In summary, PbS nanoparticles dispersion was successfully employed as SA to achieve a passively mode-locked Tm-doped
fiber laser. Nonlinear saturable absorption properties of the PbS nanoparticles dispersion were also measured in the 2 μm
spectral region, by using the homemade measurement setup in our experiment. PbS SA has the measured modulation
depth of 10.69%, non-saturable loss of 74.27% and saturation peak intensity of 8.62 MW/cm2. When the laser operated at
the stable fundamental frequency mode-locked state, the pulse duration was measured to be 1.24 ps. In addition, by further
increasing the pump power, the mode-locking switches into the second-order harmonic regime with the repetition rate of
43.86 MHz and the SNR of 60.26 dB. These results suggest that PbS nanoparticles dispersion could be developed as an
effective SA material for 2 μm mode-locking generation.
Acknowledgements to author contribution
F. Liu designed the experiment, prepared the paper and discussed with J. F. Li. Y. Zhang, Abdul Qyyum and X. H. Li
fabricated PbS nanoparticles dispersion and provided the characterizations for PbS samples. X. D. Wu, F. Yan and Z. Hu
built up the system and finished the measurements with F. Liu. C. Zhu presented some good suggestion for the paper
writing. Y. Liu supervised the project.
References
[1] I. T. Sorokina, V. V. Dvoyrin, N. Tolstik, and E. Sorokin, “Mid-IR ultrashort pulsed fiber-based lasers, IEEE J. Sel. Top. Quant. Electron., vol. 20, no. 5,
pp. 99-110, 2014.
[2] K. Scholle, S. Lamrini, P. Koopmann, and P. Fuhrberg, “2 µm laser sources and their possible applications,” in Frontiers in Guided Wave Optics and
Optoelectronics, B. Pal, ed., 2010.
[3] K. Yin, B. Zhang, G. Xue, L. Li, J. Hou, “High-power all-fiber wavelength-tunable thulium doped fiber laser at 2 μm,” Opt. Express, vol. 22, no.17, pp.
19947-19952, 2014.
[4] M. Jung, J. Lee, J. Koo, J. Park, Y. Song, K. Lee, S. Lee, and J. H. Lee, “A femtosecond pulse fiber laser at 1935 nm using a bulk-structured Bi2Te3
topological insulator,” Opt. Express, vol. 22, no. 1935, pp. 7865-7874, 2014.
[5] D. Mao, B. Jiang, X. Gan, C. Ma, Y. Chen, C. J. Zhao, H. Zhang, J. B. Zheng, and J. L. Zhao, “Soliton fiber laser mode locked with two types of
film-based Bi2Te3 saturable absorbers,” Photon. Res., vol. 3, pp. A43-A46, 2015.
[6] W. Yang, B. Zhang, J. Hou, K. Yin, Z. Liu, “A novel 2-μm pulsed fiber laser based on a supercontinuum source and its application to mid-infrared
supercontinuum generation,” Chin. Phys. B, vol. 23, pp. 054208, 2014.
[7] K. Yin, B. Zhang, L. Y. Yang, and J. Hou, “15.2 W spectrally flat all-fiber supercontinuum laser source with >1 W power beyond 3.8 μm,” Opt. Lett., vol.
42, no. 12, pp 2334-2337, 2017.
[8] X. Wang, P. Zhou, X. Wang, H. Xiao, and Z. Liu, “Pulse bundles and passive harmonic mode-locked pulses in Tm-doped fiber laser based on nonlinear
polarization rotation,” Opt. Express, vol. 22, no. 5, pp. 6147-6153, 2014.
[9] C. W. Rudy, K. E. Urbanek, M. J. F. Digonnet, and R. L. Byer, “Amplified 2-μm thulium-doped all-fiber mode-locked figure-eight laser,” J. Lightwave
Technol., vol. 31, no. 11, pp. 18091812, 2013.
[10] J. Li, Z. Zhang, Z. Sun, H. Luo, Y. Liu, Z. Yan, C. Mou, L. Zhang, and S. K. Turitsyn, “All-fiber passively mode-locked Tm-doped NOLM-based oscillator
operating at 2-μm in both soliton and noisy-pulse regimes,” Opt. Express, vol. 22, no. 7, pp. 7875-7882, 2014.
[11] R. C. Sharp, D. E. Spock, N. Pan, and J. Elliot, “190-fs passively mode-locked thulium fiber laser with a low threshold,” Opt. Lett., vol. 21, no. 12, pp.
881-883, 1996.
[12] J. Sotor, M. Pawliszewska, G. Sobon, P. Kaczmarek, A. Przewolka, I. Pasternak, J. Cajzl, P. Peterka, P. Honzatko, “All-fiber Ho-doped mode-locked
oscillator based on a graphene saturable absorber,” Opt. Lett., vol. 41, no. 11, pp. 2592-2505, 2016.
This work is licensed under a Creative Commons Attribution 4.0 License. For more information, see https://creativecommons.org/licenses/by/4.0/.
This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/JPHOT.2020.2964981, IEEE
Photonics Journal
[13] Q. Wang, J. Geng, T. Luo, and S. Jiang, “Mode-locked 2 µm laser with highly thulium-doped silicate fiber,” Opt. Lett., vol. 34, no. 23, pp. 3616-3618,
2009.
[14] J. Liu, Q. Wang, and P. Wang, “High average power picosecond pulse generation from a thulium-doped all-fiber MOPA system,” Opt. Express, vol. 20,
no. 20, pp. 22442-22447, 2012.
[15] W. Zhou, D. Y. Shen, Y. S. Wang, J. Y. Long, and Y. An, “Mode-locked thulium-doped fiber laser with a narrow bandwidth and high pulse energy,”
Laser Phys. Lett., vol. 9, no. 8, pp. 587-590, 2012.
[16] M. A. Chernysheva, A. A. Krylov, N. R. Arutyunyan, A. S. Pozharov, E. D. Obraztsova, and E. M. Dianov, SESAM and SWCNT mode-locked all-fiber
thulium-doped lasers based on the nonlinear amplifying loop mirror,” IEEE J. Sel. Top. Quant. Electron., vol. 20, no. 5, pp. 1101208, 2014.
[17] K. Keiu, F. W. Wise, “Soliton thulium-doped fiber laser with carbon nanotube saturable absorber,” IEEE Photon. Technol. Lett., vol. 21, no. 3, pp.
128-130, 2009.
[18] G. Sobon, A. Duzynska, M. Świniarski, J. Judek, J. Sotor, and M. Zdrojek, “CNT-based saturable absorbers with scalable modulation depth for
Thulium-doped fiber lasers operating at 1.9 μm,” Sci. Rep., vol. 7, pp. 45491, 2017.
[19] W. D. Tan, C. Y. Su, R. J. Knize, G. Q. Xie, L. J. Li, and D. Y. Tang, “Mode locking of ceramic Nd: yttrium aluminum garnet with graphene as a
saturable absorber,” Appl. Phys. Lett., vol. 96, no. 3, pp. 031106, 2010.
[20] M. Zhang, E. J. R. Kelleher, F. Torrisi, Z. Sun, T. Hasan, et al., “Tm-doped fiber laser mode-locked by graphene-polymer composite,” Opt. Express, vol.
20, no. 22, pp. 25077-25084, 2012.
[21] J. Sotor, I. Pasternak, A. Krajewska, W. Strupinski, and G. Sobon, “Sub-90 fs a stretched-pulse mode-locked fiber laser based on a graphene saturable
absorber,” Opt. Express, vol. 23, no. 21, pp. 27503-27508, 2015.
[22] B. Guo, Q. Lyu, Y. Yao, and P. F. Wang, “Direct generation of dip-type sidebands from WS2 mode-locked fiber laser,” Opt. Mater. Express, vol. 6, no.8,
pp. 2475-2486, 2016.
[23] T. Zhen, W. Kan, K. Lingchen, Y. Nan, W. Yao, C. Rong, H. Weisheng, X. Jianqiu, T. Yulong, “Mode-locked thulium fiber laser with MoS2,” Laser Phys.
Lett., vol. 12, no. 6, pp. 065104, 2015.
[24] J. Zhang, H. Ouyang, X. Zheng, J. You, R. Z. Chen, T. Zhou, Y. Z. Sui, Y. Liu, X. A. Cheng, and T. Jiang, “Ultrafast saturable absorption of MoS2
nanosheets under different pulse-width excitation conditions,” Opt. Lett., vol. 43, no. 2, pp. 243-246, 2018.
[25] J. Wang, W. Lu, J. Li, H. Chen, Z. Jiang, J. Wang, W. Zhang, M. Zhang, I.L. Li, Z. Xu, W. Liu, P. Yan, “Ultrafast thulium-doped fiber laser mode locked
by monolayer WSe2,” IEEE J. Sel. Top. Quant. Electron., vol. 24, no. 3, pp. 1100706, 2018.
[26] R.I. Woodward, R.C.T. Howe, T.H. Runcorn, G. Hu, F. Torrisi, E.J.R. Kelleher, T. Hasan, “Wideband saturable absorption in few-layer molybdenum
diselenide (MoSe2) for Q-switching Yb-, Er- and Tm-doped fiber lasers,” Opt. Express, vol. 23, no. 15, pp. 20051-20061, 2015.
[27] D. Mao, X. Cui, X. Gan, M. Li, W. Zhang, H. Lu, and J. Zhao, “Passively Q-switched and mode-locked fiber laser based on an ReS2 saturable
absorber,” IEEE J. Sel. Top. Quant. Electron., vol. 24, no. 3, pp. 1100406, 2018.
[28] Y. Chen, G. Jiang, S. Chen, Z. Guo, X. Yu, C. Zhao, H. Zhang, Q. Bao, S. Wen, D. Tang, and D. Fan, “Mechanically exfoliated black phosphorus as a
new saturable absorber for both Q-switching and mode-locking laser operation,” Opt. Express, vol. 23,no. 10, pp. 12823-12833, 2015.
[29] J. Sotor, G. Sobon, M. Kowalczyk, W. Macherzynski, P. Paletko, and K. M. Abramski, “Ultrafast thulium-doped fiber laser mode locked with black
phosphorus,” Opt. Lett., vol. 40, no.16, pp. 3885-3888, 2015.
[30] K. Wu, B. Chen, X. Zhang, S. Zhang, C. Guo, C. Li, P. Xiao, J. Wang, L. Zhou, W. Zou, and J. Chen, “High performance mode-locked and Q-switched
fiber lasers based on novel 2D materials of topological insulators, transition metal dichalcogenides and black phosphorus: review and perspective
(invited),” Opt. Commun., vol. 406, pp. 214-229, 2018.
[31] H. Yu, X. Zheng, K. Yin, X. Cheng, and T. Jiang, “Thulium/holmium-doped fiber laser passively mode locked by black phosphorus nanoplatelets-based
saturable absorber,” Appl. Optics, vol. 54, no. 34, pp. 10290-10294, 2015.
[32] X. M. Liu, X. K. Yao, and Y. D. Cui, “Real-time observation of the buildup of soliton molecules,” Phys. Rev. Lett., vol. 121, pp. 023905, 2018.
[33] X. M. Liu, D. Popa, and N. Akhmediev, “Revealing the transition dynamics from Q-Switching to mode-locking in a soliton laser,” Phys. Rev. Lett., vol.
123, pp. 093901, 2019.
[34] X. M. Liu, and M. Pang, “Revealing the buildup dynamics of harmonic mode-locking states in ultrafast lasers,” Laser Photonics Rev., vol. 13, pp.
1800333, 2019.
[35] M. Li, Q. Wang, X. Shi, L. A. Hornak, and N. Wu, “Detection of mercury (II) by quantum dot/DNA/goldnanoparticle ensemble based nanosensor via
nanometal surface energy transfer,” Anal. Chem., vol. 83, no.18, pp. 7061-7065, 2011.
[36] D. Y. Godovsky, “Device applications of polymer-nanocomposites,” Adv. Polym. Sci., vol. 165, pp. 153, 2000.
[37] N. N. Zhu, A. P. Zhang, Q. J. Wang, P. He, Y. Z. Fang, “Lead sulfide nanoparticle as oligonucleotides labels for electrochemical stripping detection of
DNA hybridization,” Eletroanalysis, vol. 16, no. 7, pp. 577-582, 2004.
[38] J. J. Qiu, B. B. Weng, W. Y. Ge, L. L. McDowell, Z. H. Cai, and Z. S. Shi, “A broadband Pb-chalcogenide/CdS solar cells with tandem quantum-dots
embedded in the bulk matrix (QDiM) absorption layers by using chemical bath deposition,” Sol. Energ. Mater. Sol. C., vol. 172, pp. 117-123, 2017.
[39] M. J. Speirs, D. M. Balazs, H. H. Fang, L. H. Lai, L. Protesescu, M. V. Kovalenko, and M. A. Loi, “Origin of the increased open circuit voltage in
PbS-CdS core-shell quantum dot solar cells,” J. Mater. Chem. A, vol. 3, no. 4, pp. 1450-1457, 2015.
[40] I. Kang and F. W. Wise, “Electronic structure and optical properties of PbS and PbSe quantum dots,” J. Opt. Soc. Am. B, vol. 14, no. 7, pp. 1632-1646,
1997.
[41] J. Kuljanin, M. I. Čomor, V. Djoković, J. M. Nedeljković, “Synthesis and characterization of nanocomposite of polyvinyl alcohol and lead sulfide
nanoparticles,” Mater. Chem. Phys., vol. 95, no. 1, pp. 67-71, 2006.
[42] Y. W. Lee, C. M. Chen, C. W. Huang, S. K. Chen, and J. R. Jiang, “Passively Q-switched Er3+-doped fiber lasers using colloidal PbS quantum dot
saturable absorber,” Opt. Express, vol. 24, no. 10, pp. 10675-10681, 2016.
[43] N. Ming, S. Tao, W. Q. Yang, Q. Y. Chen, R. Y Sun, C. Wang, S.Y. Wang, B. Y. Man, and H. N. Zhang, “Mode-locked Er-doped fiber laser based on
PbS/CdS core/shell quantum dots as saturable absorber,” Opt. Express, vol. 26, no. 7, pp. 9017-9026, 2018.
[44] X. L. Sun, B. Zhou, C. H. Zou, W. Zhao, Q. Q. Huang, N. N. Li, T. X. Wang, C. B. Mou, T. Y. Wang, and A. R. Kost, “Stable passively Q-switched
erbium-doped fiber laser incorporating a PbS quantum dots polystyrene composite film based saturable absorber,” Appl. Optics, vol. 57, no. 12, pp.
3231-3236, 2018.
[45] Y. Zhang, X. H. Li, A. Qyyum, T. C. Feng, P. L. Guo, J. Jiang, and H. R. Zheng, “PbS nanoparticles for ultrashort pulse generation in optical
communication region,” Part. Part. Syst. Char., vol. 35, no. 11, pp. 1800341, 2018.
[46] H. Y. Luo, J. F. Li, Y. Gao, Y. Xu, X. H. Li, and Y. Liu, “Tunable passively Q-switched Dy3+-doped fiber laser from 2.71 to 3.08 μm using PbS
nanoparticles,” Opt. Lett., vol. 44, no. 9, pp. 2322-2325, 2019.
[47] R. K Joshi, A. Kanjilal and H. K. Sehgal, “Size dependence of optical properties in solution grown Pb1−xFexS nanoparticle films,” Nanotechnology, vol.
14, pp. 809-812, 2003.
[48] Y. Y. Feng, J. Zhang, P. Zhou, G. F. Luc, J.C. Bao, W. Wang, Z. Xu, “A facile method to prepare PbS nanorods,” Mater. Res. Bull., vol. 39, pp.
1999-2005, 2004.
[49] M. S. Dhlamini, J. J. Terblans, O. M. Ntwaeaborwa, J. M. Ngaruiya, K. T. Hillie, J. R. Botha, and H. C. Swart, “Photoluminescence properties of powder
and pulsed laser-deposited PbS nanoparticles in SiO2,” J. Lumin., vol. 128, pp. 1997-2003, 2008.
[50] J. Li, Z. Yan, Z. Sun, H. Luo, Y. He, Z. Li, Y. Liu, and L. Zhang, “Thulium-doped all-fiber mode-locked laser based on NPR and 45°-tilted fiber grating,”
Opt. Express, vol. 22, no. 25, pp. 3102031028, 2014.
[51] H. Zhang, D. Y. Tang, L. M. Zhao, and N. Xiang, “Coherent energy exchange between components of a vector soliton in fiber lasers,” Opt. Express, vol.
16, no. 17, pp. 12618-12623, 2018.
[52] M. Pang, W. He, and P. S. J. Russell, “Gigahertz-repetition-rate Tm-doped fiber laser passively mode-locked by optoacoustic effects in nanobore
photonic crystal fiber,” Opt. Lett., vol. 41, no. 19, pp. 4601-4604, 2016.
This work is licensed under a Creative Commons Attribution 4.0 License. For more information, see https://creativecommons.org/licenses/by/4.0/.
This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/JPHOT.2020.2964981, IEEE
Photonics Journal
[53] M. Pang, X. Jiang, W. He, G. K. L. Wong, G. Onishchukov, N. Y. Joly, G. Ahmed, C. R. Menyuk, and P. St. J. Russell, “Stable s ubpicosecond soliton
fiber laser passively mode-locked by gigahertz acoustic resonance in photonic crystal fiber core,” Optica, vol. 2, no. 4, pp. 339-342, 2015.
[54] C.Hönninger, R. Paschotta, F. Morier-Genoud, M. Moser,and U. Keller, Q-switching stability limites of continuous-wave passive mode locking, J. Opt.
Soc. Am. B, vol. 16, pp. 4656, 1999.
[55] Q. Bao, H. Zhang, Z. Ni, Y. Wang, L. Polavarapu, Z. Shen, et al., Monolayer graphene as a saturable absorber in a mode-locked laser, Nano
Research, vol. 4, pp. 297-307, 2011.
[56] J. Z. Wang, X. Y. Liang, G. H. Hu, Z. J. Zheng, S. H. Lin, et al. 152 fs nanotube-mode-locked thulium-doped all-fiber laser, Sci. Rep., vol. 6, pp.
28885, 2016.
[57] Z. Q. Wang, L. Zhan, M. L. Qin, J. Wu, L. Zhang, Z. Zou, K. Qian, Passively Q-switched Er-doped fiber Lasers using alcohol, J. Lightwave Technol.,
vol. 33, pp. 4857-4861, 2015.
[58] Z. Q. Wang, L. Zhan, J. Wu, Z. X. Zou, L. Zhang, K. Qian, L. He, X. Fang, Self-starting ultrafast fiber lasers mode-locked with alcohol, Opt. Lett., vol.
40, no. 16, pp. 3699-3672, 2015.
[59] W. B. Cho, A. Schmidt, Jong H. Yim, S. Y. Choi, et al., “Passive mode-locking of a Tm-doped bulk laser near 2 µm using a carbon nanotube saturable
absorber,” Opt. Express, vol. 17, no. 13, pp. 11007-11012, 2009.
[60] D. I. M. Zen, N. Saidin, S. S. A. Damanhuri, S. W. Harun, et al., “Mode-locked thulium-bismuth co-doped fiber laser using graphene saturable absorber
in ring cavity,” Appl. Optics, vol. 52, no. 6, pp. 1226-1229, 2013.
[61] J. Sotor, G. Sobon, I. Pasternak, A. Krajewska, W. Strupinski, and K. M. Abramski, “Simultaneous mode-locking at 1565 nm and 1944 nm in fiber laser
based on common graphene saturable absorber,” Opt. Express, vol. 21, no. 16, pp. 18994-19002, 2013.
[62] G. Sobon, J. Sotor, I. Pasternak, A. Krajewska, W. Strupinski, and K. M. Abramski, “Thulium-doped all-fiber laser mode-locked by
CVD-graphene/PMMA saturable absorber,” Opt. Express, vol. 21, no. 10, pp. 127971-127976, 2013.
[63] B. Fu, Yi Hua, X. S. Xiao, H. W. Zhu, Z. P. Sun, and C. X. Yang, “Broadband graphene saturable absorber for pulsed fiber lasers at 1, 1.5, and 2 μm,”
IEEE J. Sel. Top. Quant. Electron., vol. 20, no. 5, pp. 411-415, 2014.
[64] Q. Wang, T. Chen, B. Zhang, M. Li, Y. Lu, and K. P. Chen, “All-fiber passively mode-locked thulium-doped fiber ring laser using optically deposited
graphene saturable absorbers,” Appl. Phys. Lett., vol. 102, pp. 131117, 2013.
[65] H. Li, J. Liu, Z. Cheng, J. Xu, F. Tan, and P. Wang, “Pulse-shaping mechanisms in passively mode-locked thulium-doped fiber lasers,” Opt. Express,
vol. 23, no. 5, pp. 6292-6303, 2015.
[66] M. Jung, J. Lee, J. Park, J. Koo, Y. M. Jhon, J. H. Lee, “Mode-locked, 1.94μm, all-fiberized laser using WS2-based evanescent field interaction,” Opt.
Express, vol. 23, no. 15, pp. 19996-20006, 2015.
[67] L. M. Cao, X. Li, R. Zhang, D. D. Wu, S. X. Dai, “Tm-doped fiber laser mode-locking with MoS2-polyvinyl alcohol saturable absorber,” Opt. Fiber
Technol., vol. 48, pp. 187-192, 2018.
[68] J. Lee, J. Koo, J. Lee, Y. M. Jhon, J. H. Lee, “All-fiberized, femtosecond laser at 1912 nm using a bulk-like MoSe2 saturable absorber,” Opt. Mater.
Express, vol. 7, no. 1912, pp. 2968-2979, 2017.
[69] J. Wang, W. Lu, J. Li, H. Chen, Z. Jiang, J. Wang, W. Zhang, M. Zhang, I.L. Li, Z. Xu, W. Liu, P. Yan, “Ultrafast thulium-doped fiber laser mode locked
by monolayer WSe2,” IEEE J. Sel. Top. Quant. Electron., vol. 24, no. 3, pp. 1100706, 2018.
[70] J. Wang, H. Chen, Z. Jiang, J. Yin, J. Wang, M. Zhang, T. He, J. Li, P. Yan and S. Ruan, “Mode-locked thulium-doped fiber laser with chemical vapor
deposited molybdenum ditelluride,Opt. Lett., vol. 43, no. 9, pp. 1998-2001, 2018.
... Besides that, the fabrication method of SA was also different for both cases that is by dropcasting the desired material directly onto the surface of fiber ferrule itself, thus creating a condition similar to thin-film SA, which may not allow high power operation. More recent studies of chalcogenides as SA for mode-locking operations have been carried out by Liu et al. (2020) and He et al. (2022). Liu et al. demonstrated the capabilities of PbS 2 as an SA in a thulium system. ...
Article
Full-text available
The performance of FePS3 as a saturable absorber (SA) for generating mode-locked pulses in a Thulium-Holmium co-doped fiber (THDF) laser via the evanescent field interaction was demonstrated in this work. FePS3 solution was drop-casted onto an arc-shaped fiber prepared using the polishing method. The fabricated FePS3/Arc-shaped SA exhibits nonlinear optical absorption behavior with modulation depth, saturation intensity, and non-saturable loss of 15.12%, 0.015 MW/cm², and 84.88%, respectively. Upon integrating the FePS3/Arc-shaped SA into the THDF ring cavity, mode-locking pulses were produced at a central wavelength of 1910.06 nm with a 3-dB bandwidth of 2.46 nm and a pulse width of 1.66 ps. At a pump power of 274.3 mW, the generated pulsed laser demonstrates a signal-to-noise ratio of 50 dB at a fundamental frequency of 11.37 MHz. The maximum average output power produced was 3.16 mW at pump power 449.9 mW, with the highest peak power and pulse energy of 0.17 kW and 277.92 pJ, respectively. The interaction between evanescent wave and FePS3 enhances the optical interaction and provides a damage threshold of higher than 4.65 GW/cm², giving a better laser performance. The finding of this work shed light on the applicability of a new group of 2-dimensional material as reliable SA material for pulsed laser generation and other potential photonics applications.
... Recently, PbS quantum dot (QD), a kind of quasi zerodimensional nanomaterial exhibiting high optical damage threshold, easy fabrication, and the possibility of mass production, has been emerged to be an effective SA to realized high-power pulse. Many approaches have been developed to fabricate PbS-SAs such as evanescent field interaction [19], fiber end deposition [20][21][22][23][24][25], and composite film [26][27][28][29]. For this type of evanescent field interaction, it is necessary to synthesize the PbS QDs onto tapered [19,30] or D-shaped [31] fibers, which requires excessive dependency on the particular fiber structure. ...
Article
Full-text available
A wavelength-tunable noise-like pulse (NLP) erbium-doped fiber laser incorporating PbS quantum dot (QD) polystyrene (PS) composite film as a saturable absorber (SA) is experimentally demonstrated. The wavelength tuning is implemented via a Lyot filter consisting of a segment of polarization-maintaining fiber (PMF) and a 45° tilted fiber grating. By adjusting the polarization state of the ring cavity, the laser can deliver NLP with a continuous wavelength-tunable range from 1550.21 to 1560.64 nm. During continuous wavelength tuning, the output power varies between a range of 30.88–48.8 mW. Worthwhile noting is that the output power of 48.8 mW is the reported highest output power for wavelength-tunable NLP operation in an erbium-doped fiber laser using composite film as a saturable absorber.
Article
Ultrafast photonics has become an interdisciplinary topic of great significance owing to the spectacular development of compact and efficient ultrafast pulse generation. Saturable absorbers (SAs) are the kernel for ultrafast pulse generation. However, the conventional SAs performance is plateauing owing to the complex preparation process, the diversification of demand for pulse width, and the relatively low laser output at mill watt level. Herein, high-quality comprehensive vertical Lead Sulfide (PbS) nanoflakes were synthesized by the physical vapor deposition (PVD), and a home-build handy dry-transfer system was used to simplify the fabrication process of PbS saturable absorbers device (PbS-SAs). Total-energy calculations by the density functional theory (DFT) were performed to further understand the saturable absorption characteristic of PbS. A time-domain flexible erbium-doped fiber pulse laser has obtained Q-switched, mode-locked, and highest 14th order harmonic mode-locked pulse. Based on a selected mode-locked operation, an all-fiber master oscillator power amplifier (MOPA) pulse laser system was proposed and demonstrated. The integrated system based on PbS-SAs achieved a time-domain flexible fiber pulse laser, high order harmonic mode-locked pulse, and a picosecond pulse (∼287ps) with high average power of ∼57.5 W and peak power of ∼13.7 kW at the central wavelength of 1562 nm. This research not only provides a simple manufacturing process for high uniformity SAs devices but also greatly improves output power, which would attract interest in exploring the unique properties of nanocomposite SAs for photonic applications.
Article
As a new two-dimensional (2D) material, zirconium telluride (ZrTe3) is a typical charge density wave (CDW) material with strong electrical and optical anisotropy compared with graphene, transition metal disulfides (TMDs), topological insulators (TIs), and black phosphorus (BPs). To date, the nonlinear optical properties of ZrTe3 and its applications in information photonics aspect have not been reported. In this work, the nonlinear optical saturable absorption properties of ZrTe3 nanosheets and their application in passively mode-locked fiber laser are systematically investigated for the first time. The results show that the modulation depth and saturation intensity are 6.5% and 89.87 MW/cm² at the optical fiber communication band (C band), respectively. In the erbium-doped fiber (EDF) ring laser with ZrTe3 nanosheets as saturable absorber (SA), a stable mode-locked pulse train is obtained with the central wavelength of 1562.12 nm, the 3 dB optical spectral width of 2.07 nm, the basic repetition rate of 3.377 MHz, and the pulse duration of 1.469 ps when the pump power is 65 mW. Moreover, the dynamic evolution process of the optical spectrum and pulse sequence under different pump power are illustrated. All results reveal that ZrTe3 has excellent potential for applying in many fields, including ultra-fast all-optical sampling, optical frequency comb, optical fiber communication, supercontinuum, etc, and provides an intuition of opening new avenues toward various advanced information photonic devices.
Article
Full-text available
We present the first widely tunable passively Q -switched Dy 3 + -doped ZBLAN fiber laser around 3 μm pumped at 1.1 μm using PbS nanoparticles as the saturable absorber (SA) to our knowledge. At 2.87 μm, the modulation depth and saturation intensity of the SA were measured to be 12.5% and 1.10 MW / cm 2 , respectively. Stable Q -switching was achieved over a wavelength range of 2.71–3.08 μm ( ∼ 370 nm ), which is a record tuning range from a pulsed rare-earth-doped fiber laser, to our knowledge. A maximum output power of 252.7 mW was obtained, with a pulse energy of 1.51 μJ, a pulse width of 795 ns, and a repetition rate of 166.8 kHz. This demonstration implies that Dy 3 + is a promising gain medium for tunable pulsed sources in the 3 μm band and shows for the first time, to the best of our knowledge, the potential of PbS as a mid-infrared SA.
Article
Full-text available
We demonstrated a passively Q -switched erbium-doped fiber laser (EDFL) using PbS quantum dots in polystyrene films (QDPFs) as saturable absorbers (SAs). Compared to other SAs, PbS QDPFs have advantages of broad absorption range, high quantum yield, low cost, and facile preparation. We have successfully generated stable Q -switched pulses with an average output power of 40.19 mW, a single pulse energy of 586.1 nJ, a repetition rate of 68.04 kHz, a pulse width of 3.9 μs, and a signal-to-noise ratio of 50.5 dB under 660 mW pump power. The output of the EDFL has been monitored for 5 consecutive hours under laboratory conditions to show stable operation of the laser system.
Article
Full-text available
A passively mode-locked thulium-doped fiber (TDF) laser was realized by employing chemical vapor deposited few-layer molybdenum ditelluride ( MoTe 2 ) as a saturable absorber (SA). The few-layer MoTe 2 film was transferred onto the waist of a microfiber and then incorporated into a TDF laser with a typical all-fiber ring cavity configuration. Stable soliton pulses emitting at 1930.22 nm were obtained with a 3 dB bandwidth of 4.45 nm, a pulse duration of 952 fs, and an average power of 36.7 mW.
Article
Full-text available
Previously, PbS/CdS core/shell quantum dots with excellent optical properties have been widely used as light-harvesting materials in solar cell and biomarkers in bio-medicine. However, the nonlinear absorption characteristics of PbS/CdS core/shell quantum dots have been rarely investigated. In this work, PbS/CdS core/shell quantum dots were successfully employed as nonlinear saturable absorber (SA) for demonstrating a mode-locked Er-doped fiber laser. Based on a film-type SA, which was prepared by incorporating the quantum dots with the polyvinyl alcohol (PVA), mode-locked Er-doped operation with a pulse width of 54 ps and a maximum average output power of 2.71 mW at the repetition rate of 3.302 MHz was obtained. Our long-time stable results indicate that the CdS shell can effectively protect the PbS core from the effect of photo-oxidation and PbS/CdS core/shell quantum dots were efficient SA candidates for demonstrating pulse fiber lasers due to its tunable absorption peak and excellent saturable absorption properties.
Article
Full-text available
The newly raised two-dimensional material MoS 2 is regarded as an ideal candidate for saturated absorbers. Here, the open-aperture Z-scan method is used to study the saturation absorption (SA) response of monolayer and multilayer MoS 2 , considering laser irradiation with different pulse widths. Specifically, in cases of 10 ns and 10 ps laser pulses, the accumulative nonlinearity [e.g., free carrier absorption (FCA)] coupled with SA is found in both monolayer and multilayer MoS 2 . However, under a 65 fs pulse laser, the instantaneous nonlinearity [e.g., two-photon absorption (TPA)] and the SA effect turn to play a significant role. Additionally, the saturation of both TPA and FCA is observed in MoS 2 . Importantly, the modulation depth of MoS 2 shows different change trends by adjusting the laser pulse width.
Article
Q switching (QS) and mode locking (ML) are the two main techniques enabling generation of ultrashort pulses. Here, we report the first observation of pulse evolution and dynamics in the QS-ML transition stage, where the ML soliton formation evolves from the QS pulses instead of relaxation oscillations (or quasi-continuous-wave oscillations) reported in previous studies. We discover a new way of soliton buildup in an ultrafast laser, passing through four stages: initial spontaneous noise, QS, beating dynamics, and ML. We reveal that multiple subnanosecond pulses coexist within the laser cavity during the QS, with one dominant pulse transforming into a soliton when reaching the ML stage. We propose a theoretical model to simulate the spectrotemporal beating dynamics (a critical process of QS-ML transition) and the Kelly sidebands of the as-formed solitons. Numerical results show that beating dynamics is induced by the interference between a dominant pulse and multiple subordinate pulses with varying temporal delays, in agreement with experimental observations. Our results allow a better understanding of soliton formation in ultrafast lasers, which have widespread applications in science and technology.
Article
Harmonic mode‐locking (HML) is an important technique enabling the generation of high‐repetition‐rate ultrashort pulses. Using an emerging time‐stretch dispersive Fourier transform technique, the experimental observation of the entire buildup process of the passive HML state in an ultrafast fiber laser is reported here. It is unveiled that the whole process of HML buildup successively undergoes seven different ultrafast phases: raised relaxation oscillation, spectral beating behavior, birth of a giant pulse, self‐phase‐modulation‐induced instability, pulse splitting, repulsion and separation of multiple pulses, and a stable HML state. It is observed that the multiple HML pulses originate from a single‐pulse splitting phenomenon and a remarkable breathing behavior occurs at an early stage of the HML buildup process. The numerical results confirm that the effects of dispersive wave, gain depletion and recovery, and acoustic wave play key roles in the earlier, middle, and later stages of this HML buildup process, respectively; as well, the acoustic resonance in the single‐mode fiber stabilizes the final HML state of lasers. The experimental observation of the entire buildup process of the passive harmonic mode‐locking (HML) state, which undergoes seven different ultrafast phases, is demonstrated. Multiple HML pulses originate from the single‐pulse splitting phenomenon. The acoustic resonance in the single‐mode fiber stabilizes the final HML state of lasers.
Article
Real-time spectroscopy access to ultrafast fiber lasers opens new opportunities for exploring complex soliton interaction dynamics. Here, we have reported the first observation, to the best of our knowledge, of the entire buildup process of soliton molecules (SMs) in a mode-locked laser. We have observed that the birth dynamics of a stable SM experiences five different stages, i.e., the raised relaxation oscillation (RO) stage, beating dynamics stage, transient single pulse stage, transient bound state, and finally the stable bound state. We have discovered that the evolution of pulses in the raised RO stage follows a law that only the strongest one can ultimately survive and, meanwhile, the pulses periodically appear at the same temporal positions for all lasing spikes during the same RO stage (named as memory ability) but they lose such ability between different RO stages. Moreover, we have found that the buildup dynamics of SMs is quite sensitive to both the polarization state of intracavity light and the fluctuation of pump power. These results provide new perspectives into the ultrafast transient process in mode-locked lasers and the dynamics of complex nonlinear systems.
Article
We have designed an all-fiber passive mode-locking thulium-doped fiber laser that uses molybdenum disulfide (MoS2) as a saturable absorber (SA) material. A free-standing few-layer MoS2-polyvinyl alcohol (PVA) film is fabricated by liquid phase exfoliation (LPE) and is then transferred onto the end face of a fiber connector. The excellent saturable absorption of the fabricated MoS2-based SA allows the laser to output soliton pulses at a pump power of 500 mW. Fundamental frequency mode-locking is realized at a repetition frequency of 13.9 MHz. The central wavelength is 1926 nm, the 3 dB spectral bandwidth is 2.86 nm and the pulse duration is 1.51 ps. Additionally, third-order harmonic mode-locking of the laser is also achieved. The pulse duration is 1.33 ps, which is slightly narrower than the fundamental frequency mode-locking bandwidth. The experimental results demonstrate that the few-layer MoS2-PVA SA is promising for use in 2 μm laser systems.