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

Nonlinear optical properties of MoS2-WS2 heterostructure in fiber lasers

Open Access Open Access

Abstract

As a saturable absorption material, the heterostructure with the van der Waals structure has been paid much attention in material science. In general, the heterogeneous combination is able to neutralize, or even exceed, the individual material’s advantages in some aspects. In this paper, which describes the magnetron sputtering deposition method, the tapered fiber is coated by the MoS2-WS2 heterostructure, and the MoS2-WS2 heterostructure saturable absorber (SA) is fabricated. The modulation depth of the prepared MoS2-WS2 heterostructure SA is measured to be 19.12%. Besides, the theoretical calculations for the band gap and carrier mobility of the MoS2-WS2 heterostructure are provided. By employing the prepared SA, a stable and passively erbium-doped fiber laser is implemented. The generated pulse duration of 154 fs is certified to be the shortest among all fiber lasers based on transition mental dichalcogenides. Results in this paper provide the new direction for the fabrication of ultrafast photon modulation devices.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Since the emergence of graphene [1–4], a variety of two-dimensional (2D) materials, such as topological insulators [5–7], black phosphorus (BP) [8–10] and some new materials [11–13], have sprung up rapidly and occupied an important position in the field of optoelectronics. As effective optical modulation materials, 2D materials can be employed in fiber lasers to generate ultrashort pulses [14,15]. As a result, they have attracted increasing attention in recent years.

Transition mental dichalcogenides (TMDs), which are considered to be the supplements or even the substitutes of graphene in some cases, have also been investigated in the field effect transistors, photocatalysts, photodetectors and optical modulators [16–29]. WS2 and MoS2, as two representative materials of TMDs, have been fully explored from the physical to optoelectronic properties. The nonlinear optical response of MoS2 which is stronger than graphene at 800 nm has been investigated by using Z-scan technique [30]. Then, the broadband saturable absorption of MoS2 has been illustrated by theoretical arithmetic and further successfully demonstrated at 1.06 μm, 1.42 μm and 2.1 μm [31]. As the analogue of MoS2, WS2 not only has good performance in ultra-wide response, but also exhibits a large second-order nonlinear susceptibility [32]. Therefore, WS2 has great potential in realizing ultrashort pulses. Both MoS2 and WS2 perform well as optical modulation materials in fiber lasers.

As far as we know, molecular layers of TMDs are connected by van der Waals (vdW) forces. The weak vdW forces between two adjacent TMD layers not only enable them easier to be stripped into monolayer nanosheets, but also provide a chance to manufacture optoelectronic devices by stacking different TMDs without considering the problem of mismatch [33,34]. The structure of heterostructure opens up a different avenue for the fabrication of better photoelectric devices. From previous reports, the MoS2-WS2 heterostructure exhibits remarkable performances. The type-II semiconductor heterostructures came from stacked MoS2-WS2 heterostructure facilitate the transfer of holes [35,36]. Moreover, the absorption of MoS2-WS2 heterostructure is larger than the simple superposition of the respective absorptions of MoS2 and WS2 [37]. Therefore, the heterostructure consisting of two different materials with various geometric composition and electronic energy exhibits unique electrical and optical properties [38–40]. However, the MoS2-WS2 heterostructure SA is rarely used in the fiber lasers for ultrafast photonics. As far as we know, only Chen et al. has made an attempt at the application of the WS2-MoS2-WS2 heterostructure in the fiber laser so far [41].

Among the various preparation methods, the magnetron sputtering deposition (MSD) method with the simple operation is considered to be the most suitable method for the preparation of MoS2-WS2 heterostructure with high quality in this paper. MoS2 and WS2 are coated on the tapered fibers to manufacture the MoS2-WS2 heterostructure SA. Herein, the fiber laser is implemented to investigate the related nonlinear optical performance of the proposed MoS2-WS2 heterostructure SA. After the application of SA in fiber laser, a stable mode-locking system is obtained. The obtained pulse duration of 154 fs is proved to be the shortest in the congeneric fiber lasers. The theoretical and experimental results illustrate that the MoS2-WS2 heterostructure provides reference value for the innovation and development of fiber lasers.

2. Result and discussion

2.1 Theoretical calculation of MoS2-WS2 heterostructure SA

By the density functional theory (DFT) implemented in the Vienna ab initio simulation package (VASP), the theoretical calculations were performed. The interactions between valence electrons and core electrons were described with the projector augmented wave (PAW) pseudopotential. A cutoff energy was chosen to be 420 eV for the plane-wave expansion of wave functions. For the integration over the first Brillouin zone, the Monkhorst-Pack scheme of the k-point sampling was adopted. In order to avoid the interactions between two slabs in the nearest-neighbor unit cells, the periodic boundary conditions and a vacuum space of 30 Å along the z direction were applied. A 7 × 7 × 1 grid for the k-point sampling was used for the geometry optimization, while 37 × 37 × 1 was used for the static total energy calculations. When the change of the total energy was less than 10−4 eV, the forces became smaller than 0.01 eV/Ao, the structure began to relax. The calculated lattice constant of monolayer MoS2, monolayer WS2 and MoS2-WS2 heterostructures are 3.160 Ao, 3.153 Ao and 3.160 Ao, respectively.

The mobility μ of the MoS2-WS2 heterstructure was calculated using the deformation potential (DP) theory on the basis of the effective mass approximation:

μ=2e3C3kBT|m*|2E2,
where T is the temperature, and C is the elastic modulus. For the 2D system, the in-plane value is defined asC2D=[2E/δ2]/S0, where E, δ, and S0 are the total energy, applied uniaxial strain and area of the investigated system, respectively. The DP constant E along a certain direction is obtained byE=dEedge/dδ, where Eedge is the energy of the band edges (valence-band maximum for holes, and conduction-band minimum for electrons).

The heterostructure was modeled by a MoS2-WS2 bilayer structure in Fig. 1(a). To get the most stable structure, the heterostructure of the strain WS2 monolayer was used, and the corresponding lattice mismatch is 0.15%. The band gap of the MoS2-WS2 heterostructure is 1.26 eV, which is smaller than the band gap of the monolayer MoS2 (1.77 eV) shown in Fig. 1. The effective mass of electrons and the effective mass of holes in K are 0.46 m0 and 0. 72 m0, respectively. The effective mass of holes is smaller than that of the monolayer MoS2, and the effective mass of electrons is almost equal to the monolayer MoS2.

 figure: Fig. 1

Fig. 1 Atomic and electronic structure of the MoS2-WS2 heterostructure. (a) Top and side views of the MoS2-WS2 heterostructure, the dashed rectangle denotes the primitive cell. (b), (c) and (d) are band structures of MoS2-WS2 heterostructure, monolayer MoS2 and monolayer WS2, respectively. Here, the fermi level is set to be zero, and the orange line denotes valence-band maximum for the holes and conduction-band minimum for the electrons. Besides, red points project the contribution from the MoS2 in (b). (e) is the band alignment of the MoS2-WS2 heterostructure. The energy levels of MoS2 and monolayer WS2 slabs are shown in both sides. Here, the monolayer MoS2, monolayer WS2, and MoS2-WS2 heterostructure are considered.

Download Full Size | PDF

The effective mass and carrier mobility of the MoS2-WS2 heterostructure, monolayer MoS2 and WS2 are shown in Table 1. The carrier mobility of the MoS2-WS2 heterostructure is larger than that of the monolayer MoS2. These results may be related to the type-II heterostructure from the stacked MoS2-WS2 heterostructure. Specifically, upon optical excitation, the electron tends to stay in the MoS2 layer which reside with the minimum conduction band, and the holes prefer to stay in the WS2 layer which reside with the maximum valence band. Therefore, the type-II heterostructure effectively assist electrons and holes separate quickly. The relationship between relaxation time (<τ>) and electron mobility (μ2D) can be calculated asμ2D=e<τ>m, where m is the effective mass. As far as we know, the large relaxation time can bring about greater modulation depth. Therefore, in the same case, the greater the product of effective mass (m) and electron mobility (μ2D), the higher the modulation depth.

Tables Icon

Table 1. Band Gap (Eg), Effective Mass and Carrier Mobility (µ) of Monolayer and Heterostructure Materials

2.2 Fabrication and characterization of MoS2-WS2 heterostructure SA

The MoS2-WS2 heterostructure SA was prepared by the MSD method. Before deposition, the surface impurities of the WS2 and MoS2 raw materials were removed, which guaranteed the purity of materials produced. During deposition, the MoS2 target and taper fiber were placed in the vacuum chamber at the same time, and the pressure of the cavity was set to be 1.7 × 10−3 pa. The Ar gas was continuously excited under the 0.4 A alternating current (AC) with voltage of 6 V for 70 s. Following, WS2 nanosheets were deposited on the surface of MoS2 in the same way under the 0.5 A with voltage of 6 V for 180 s. Finally, a dense layer of gold film, which prevents the material from being damaged oxidized, was deposited on the surface of the MoS2-WS2 heterostructure.

To characterize the surface and lateral properties of the MoS2-WS2 heterostructure SA, the scanning electron microscope (SEM) was employed. Figure 2(a) exhibits the surface morphology of the SA. We can observe that the particles are arranged in a compact and uniform manner. As shown in Fig. 2(b), the lateral surface illustrates the thickness of the heterostructure. To ensure accuracy, the thickness of the sample is measured in three different places. The average value of three different places which is calculated as 63 nm illustrates the thickness of MoS2-WS2 heterostructure. According to measuring results, the thickness of MoS2 is 21 nm and the thickness of WS2 is 42 nm. The thickness of the SiO2 substrate is measured to be 59 nm. Those results prove that the obtained MoS2-WS2 has layered structure.

 figure: Fig. 2

Fig. 2 SEM of the MoS2-WS2 heterostructure SA. (a) The surface morphology; (b) The thickness of heterostructure indicated by the lateral.

Download Full Size | PDF

Raman analysis is a common method to distinguish the types of materials through different vibration modes. In order to confirm the successful manufacture of the MoS2-WS2 heterostructure SA, the corresponding Raman spectra are measured in Fig. 3(a). The measurements show that the Raman spectra of the pure MoS2 sample have peaks at 378 cm−1 and 404 cm−1, which correspond to the E2g and A1g, respectively [42]. The peaks at 355 cm−1 and 419 cm−1 in Raman spectra of WS2 is corresponding to E21g and A1g modes [34]. The four peaks shown in the Raman spectra of the MoS2-WS2 heterostructure are in good agreement with the four vibration modes of MoS2 and WS2. The nonlinear absorption properties of the MoS2-WS2 heterostructure SA are investigated by the balanced twin-detector method. In the measurement, the mode-locked fiber laser with the repetition rate of 135 MHz and pulse duration of 100 fs is used as the exciting source. The measurement results in Fig. 3(b) demonstrate that the prepared MoS2-WS2 heterostructure SA has the modulation depth of 19.12%. Moreover, the saturation absorption intensity of 1.361 MW/cm2 is relatively small, which is beneficial to the low self-starting threshold for the fiber laser. The optical damage threshold of the current absorber device is calculated to be 4.17 mJ/cm2, which is higher than the commercial available semiconductor saturable absorber mirrors (SESAMs) (500 μJ/cm2).

 figure: Fig. 3

Fig. 3 (a) The Raman spectra of the MoS2, WS2 and MoS2-WS2 heterostructure; (b) The nonlinear absorption of the MoS2-WS2 heterostructure SA.

Download Full Size | PDF

3. Nonlinear optical performance

To investigate the nonlinear absorption characteristics of the MoS2-WS2 heterostructure SA, it has been integrated into the erbium-doped fiber (EDF) laser in Fig. 4. The wavelength division multiplexer (WDM) (980/1550 nm) coupled the pump light into the ring cavity. With the excitation of the pump which operated at 976 nm, a piece of EDF of 42 cm amplified the pulse through the energy level transition. The polarization controller (PC) was used to fine tune the polarization state and birefringence in the cavity. In order to avoid reflecting light damage to the device, isolator (ISO) enforced light to transmit in a fixed direction. The starting threshold of the laser was 180 mW. When the pump power was greater than this value, the laser maintained the stable mode-locking operation. Through the 20:80 optical coupler (OC), the experimental results were displayed and recorded. The main test instruments we adopted were a 250 MHz oscilloscope and spectrum analyzer.

 figure: Fig. 4

Fig. 4 Experimental device diagram of the passively mode-locked EDF laser employed with the MoS2-WS2 heterostructure SA.

Download Full Size | PDF

Figure 5 summarizes the laser performance at the maximum pump power of 630 mW. As indicated in Fig. 5(a), the central optical spectrum is located at 1560 nm with 3 dB spectral width of 24.4 nm. From the symmetrical Kelly sidebands of the spectrum, it can be inferred that this is the soliton mode locking system. The regular array of the pulse train in Fig. 5(b) indicates that the mode-locked system is in a stable operative condition. The time interval of two neighboring pulses is 13.4 ns, which is corresponding to the fundamental repetition rate 74.6 MHz of mode-locked pulses. The resolution bandwidth (RBW) and span adopted in the measurement of the fundamental frequency are 10 Hz and 10 kHz, respectively. Under high resolution and small span, there is no obvious frequency interference signals appear, and the signal-to-noise ratio (SNR) is 91.2 dB in Fig. 5(c), which further illustrate that the mode-locked operation is relatively stable. Moreover, the pulses in the illustration of Fig. 5(c) are in the uniform arrangement. The symmetrical autocorrelation trace is fitted by the Sech2 function in Fig. 5(d), which indicates that pulse duration of the mode-locked fiber laser is 154 fs. The corresponding time-bandwidth product (TBP) is 0.4403, indicating that the output mode-locked pulses are slightly chirped. The maximum output power of the laser is 19.8 mW. Results demonstrate that the MoS2-WS2 heterostructure SA showns the saturable absorption property around 1.5 μm. According to previous reports, the direct bandgap of the monolayer MoS2 is 1.8 eV, and the indirect bandgap of the bulk MoS2 is 1.29 eV. Obviously, the photon energy is below the optical (excitonic) bandgap of MoS2 at 1560 nm in our work. Although the governing mechanism is still unclear, there have been several convincing theories put forward by researchers, such as multiphoton absorption, edge state saturable absorption and defect state saturable absorption [43]. In the theory of defect state, the imperfection of the 2D material is inevitable in the production process, which has an impact both on its electronic and optical properties [44]. Wang et al. has demonstrated that the MoS2 bandgap can be reduced from 1.08 to 0.08 eV by introducing the defects in a suitable range. By the introduction of S defects, the MoS2 has been successfully applied in fiber lasers at the operating wavelength of 1.06, 1.42, and 2.1 μm [45]. Therefore, it is justified to believe that there are unavoidable defects in the material, which result in the decrease of the band gap and broadband absorption beyond expectation.

 figure: Fig. 5

Fig. 5 The performance of the passively mode-locked EDF laser employed with the MoS2-WS2 heterostructure SA. (a) The optical spectrum located at 1560 nm with 3 dB spectral width of 24.4 nm; (b) The mode-locked pulse train; (c) The radio-frequency spectrum; (d) The symmetrical autocorrelation trace of mode-locked pulses.

Download Full Size | PDF

To prove the long-term stability of new SA device, the continuous monitoring for output power of the fiber laser is implemented in Fig. 6. The total monitoring time was up to 16 hours, and the output power of the fiber laser was recorded every second. As shown in Fig. 6, the standard deviation of the output power is only 0.123. This result indicates that the system has good stability. Limited by the experimental conditions, we only monitored the output power within 16 hours, but it can be observed that the trend of data remained stable in the later period, which indicated that the system could maintain stability for a longer time. Therefore, the new SA device can withstand long-term illumination.

 figure: Fig. 6

Fig. 6 The continuous monitoring of output power of the fiber laser.

Download Full Size | PDF

To illustrate the relative advantages and improvements of the proposed mode-locked EDF laser based on the MoS2-WS2 heterostructure SA, the performance of mode-locked fiber lasers employed with pure TMDs materials are compared in Table 2. Although there have been some researches on the electronic properties and nonlinear characteristics of the monolayer MoS2 (or WS2), almost all MoS2 (or WS2) that applied in mode-locked fiber lasers are multilayered according to the previous reports. Because the preparation of the highly efficient SA based on monolayer material is complicated and difficult. The pulse duration obtained is 154 fs, which is almost the shortest among the similar fiber lasers. Considering the structural particularity of the MoS2-WS2 heterostructure, we think there are two reasons for the remarkable performance of the fiber laser: Firstly, the ultrafast transfer time of carriers from the type-II semiconductor heterostructures may beneficial to the generation of ultra-short pulses. For the monolayer MoS2, the intra-layer carrier recombination time is 2 ps [41]. However, it has reported that the electronic transfer in the MoS2-WS2 heterostructure occur within 50 fs upon photo-excitation [35]. Compared with MoS2 itself, the carrier recombination of the MoS2-WS2 heterostructure is much faster. Moreover, we can see that no matter what the thickness of the material is, the pulse duration of the corresponding laser is not as short as that of the fiber laser based on the MoS2-WS2 heterostructure in Table 2. The above results give us reason to believe that the vertically stacking of WS2 and MoS2 may cause ultrafast carrier recombination, which further benefit to the generation of ultrashort pulses. Secondly, the large modulation depth of the MoS2-WS2 heterostructure SA is also propitious to the generation of ultrashort pulses. On the one hand, compared with the pure TMDs materials, the combination of the nonlinear absorption properties of two materials make it more advantageous in the modulation of light. It has been reported that the absorption of the MoS2-WS2 heterostructure is larger than the simple superposition of the respective absorptions of MoS2 and WS2 in previous works [40]. On the other hand, the tapered fiber structure of the MoS2-WS2 heterostructure SA enhance its nonlinearity. The tapered fiber allows a very long interaction length. In commonly used sandwich structures, the interaction length of the material and light is limited by the thickness of the material, often in the nanoscale. However, in the tapered fiber, the interaction length can be extended to centimeter magnitude order by means of the evanescent field effect. This sufficient reaction of the material and light enables the material to fully exhibit its nonlinearity. Besides, SAs based on tapered fibers with different specifications show differences in the modulation depth, and SAs owning small waist diameter tend to show larger modulation depth. Therefore, we believe that the tapered fiber structure of the MoS2-WS2 heterostructure SA can enhance its nonlinearity.

Tables Icon

Table 2. Comparisons of Mode-Locked Fiber Lasers Employed with Different SAs

4. Conclusions

In this paper, we have prepared the MoS2-WS2 heterostructure SA using the MSD method. The modulation depth and saturation absorption intensity of the SA is 19.12% and 1.361 MW/cm2, respectively. The corresponding band gap and electron mobility have been theoretically calculated. In order to investigate the nonlinear absorption characteristics of the MoS2-WS2 heterostructure SA, it has been integrated into the EDF laser. In addition, the stable mode-locked fiber laser operating at 1560 nm with SNR of 91.2 dB and output power of 19.8 mW has been implemented. The obtained pulse duration of 154 fs has been proved to be the shortest in the congeneric fiber lasers. Results in this paper not only reveal the impressive optical nonlinearity of the MoS2-WS2 heterostructure SA, but also provide reference value for the application and development of TMDs heterostructures.

Funding

National Natural Science Foundation of China (11674036, 11875008); Beijing Youth Top-Notch Talent Support Program (2017000026833ZK08); State Key Laboratory of Information Photonics and Optical Communications (IPOC2017ZZ05).

References

1. A. K. Geim and K. S. Novoselov, “The rise of graphene,” Nat. Mater. 6(3), 183–191 (2007). [CrossRef]   [PubMed]  

2. F. Bonaccorso, Z. Sun, T. Hasan, and A. C. Ferrari, “Graphene photonics and optoelectronics,” Nat. Photonics 4(9), 611–622 (2010). [CrossRef]  

3. Q. Bao, H. Zhang, B. Wang, Z. Ni, C. H. Y. X. Lim, Y. Wang, D. Y. Tang, and K. P. Loh, “Broadband graphene polarizer,” Nat. Photonics 5(7), 411–415 (2011). [CrossRef]  

4. Z. T. Wang, Y. Chen, C. J. Zhao, H. Zhang, and S. C. Wen, “Switchable dual-wavelength synchronously Q-switched erbium-doped fiber laser based on graphene saturable absorber,” IEEE Photonics J. 4(3), 869–876 (2012). [CrossRef]  

5. W. Liu, L. Pang, H. Han, W. Tian, H. Chen, M. Lei, P. Yan, and Z. Wei, “70-fs mode-locked erbium-doped fiber laser with topological insulator,” Sci. Rep. 6(1), 19997 (2016). [CrossRef]   [PubMed]  

6. Y. H. Lin, S. F. Lin, Y. C. Chi, C. L. Wu, C. H. Cheng, W. H. Tseng, J. H. He, C. I. Wu, C. K. Lee, and G. R. Lin, “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). [CrossRef]  

7. J. Ma, S. Lu, Z. Guo, X. Xu, H. Zhang, D. Tang, and D. Fan, “Few-layer black phosphorus based saturable absorber mirror for pulsed solid-state lasers,” Opt. Express 23(17), 22643–22648 (2015). [CrossRef]   [PubMed]  

8. F. Xia, H. Wang, and Y. Jia, “Rediscovering black phosphorus as an anisotropic layered material for optoelectronics and electronics,” Nat. Commun. 5(1), 4458 (2014). [CrossRef]   [PubMed]  

9. H. R. Mu, S. H. Lin, Z. C. Wang, S. Xiao, P. F. Li, Y. Chen, H. Zhang, H. F. Bao, S. P. Lau, C. X. Pan, D. Y. Fan, and Q. L. Bao, “Pulsed lasers: black phosphorus-polymer composites for pulsed lasers,” Adv. Opt. Mater. 3(10), 1447–1453 (2015). [CrossRef]  

10. J. Li, H. Luo, B. Zhai, R. Lu, Z. Guo, H. Zhang, and Y. Liu, “Black phosphorus: a two-dimension saturable absorption material for mid-infrared Q-switched and mode-locked fiber lasers,” Sci. Rep. 6(1), 30361 (2016). [CrossRef]   [PubMed]  

11. Y. H. Xu, Z. T. Wang, Z. N. Guo, H. Huang, Q. L. Xiao, H. Zhang, and X.-F. Yu, “Solvothermal synthesis and ultrafast photonics of black phosphorus quantum dots,” Adv. Opt. Mater. 4(8), 1223–1229 (2016). [CrossRef]  

12. X. T. Jiang, S. X. Liu, W. Y. Liang, S. J. Luo, Z. L. He, Y. Q. Ge, H. D. Wang, R. Cao, F. Zhang, Q. Wen, J. Q. Li, Q. L. Bao, D. Y. Fan, and H. Zhang, “Broadband nonlinear photonics in few-layer MXene Ti3C2Tx (T= F, O, or OH),” Laser Photonics Rev. 12(2), 1700229 (2018). [CrossRef]  

13. L. Lu, Z. M. Liang, L. M. Wu, Y. X. Chen, Y. F. Song, S. C. Dhanabalan, J. S. Ponraj, B. Q. Dong, Y. J. Xiang, F. Xing, D. Y. Fan, and H. Zhang, “Few‐layer bismuthene: sonochemical exfoliation, nonlinear optics and applications for ultrafast photonics with enhanced stability,” Laser Photonics Rev. 12(1), 1700221 (2018). [CrossRef]  

14. Q. H. Wang, K. Kalantar-Zadeh, A. Kis, J. N. Coleman, and M. S. Strano, “Electronics and optoelectronics of two-dimensional transition metal dichalcogenides,” Nat. Nanotechnol. 7(11), 699–712 (2012). [CrossRef]   [PubMed]  

15. F. N. Xia, H. Wang, D. Xiao, M. Dubey, and A. Ramasubramaniam, “Two-dimensional material nanophotonics,” Nat. Photonics 8(12), 899–907 (2014). [CrossRef]  

16. J. Baek, D. M. Yin, N. Liu, I. Omkaram, C. Jung, H. Im, S. Hong, S. M. Kim, Y. K. Hong, J. Hur, Y. Yoon, and S. Kim, “A highly sensitive chemical gas detecting transistor based on highly crystalline CVD-grown MoSe2 films,” Nano Res. 10(6), 1861–1871 (2017). [CrossRef]  

17. W. J. Liu, M. L. Liu, M. Lei, S. B. Fang, and Z. Y. Wei, “Titanium selenide saturable absorber mirror for passive Q-switched Er-doped fiber laser,” IEEE J. Sel. Top. Quant. 24(3), 1 (2018). [CrossRef]  

18. J. A. Wilson and A. D. Yoffe, “The transition metal dichalcogenides discussion and interpretation of the observed optical, electrical and structural properties,” Adv. Phys. 18(73), 193–335 (1969). [CrossRef]  

19. W. Liu, M. Liu, Y. OuYang, H. Hou, G. Ma, M. Lei, and Z. Wei, “Tungsten diselenide for mode-locked erbium-doped fiber lasers with short pulse duration,” Nanotechnology 29(17), 174002 (2018). [CrossRef]   [PubMed]  

20. F. Schwierz, J. Pezoldt, and R. Granzner, “Two-dimensional materials and their prospects in transistor electronics,” Nanoscale 7(18), 8261–8283 (2015). [CrossRef]   [PubMed]  

21. W. Liu, M. Liu, J. Yin, H. Chen, W. Lu, S. Fang, H. Teng, M. Lei, P. Yan, and Z. Wei, “Tungsten diselenide for all-fiber lasers with the chemical vapor deposition method,” Nanoscale 10(17), 7971–7977 (2018). [CrossRef]   [PubMed]  

22. J. Yin, H. Chen, W. Lu, M. Liu, I. Ling Li, M. Zhang, W. Zhang, J. Wang, Z. Xu, P. Yan, W. Liu, and S. Ruan, “Large-area and highly crystalline MoSe2 for optical modulator,” Nanotechnology 28(48), 484001 (2017). [CrossRef]   [PubMed]  

23. M. L. Liu, Y.-Y. OuYang, H.-R. Hou, M. Lei, W.-J. Liu, and Z.-Y. Wei, “MoS2 saturable absorber prepared by chemical vapor deposition method for nonlinear control in Q-switching fiber laser,” Chin. Phys. B 27(8), 084211 (2018). [CrossRef]  

24. W. J. Liu, M. L. Liu, H. N. Han, S. B. Fang, H. Teng, M. Lei, and Z. Y. Wei, “Nonlinear optical properties of WSe2 and MoSe2 films and their applications in passively Q-switched erbium doped fiber lasers,” Photon. Res. 6(10), C15–C21 (2018). [CrossRef]  

25. W. Liu, M. Liu, Y. OuYang, H. Hou, M. Lei, and Z. Wei, “CVD-grown MoSe2 with high modulation depth for ultrafast mode-locked erbium-doped fiber laser,” Nanotechnology 29(39), 394002 (2018). [CrossRef]   [PubMed]  

26. W. Y. Li, Y. OuYang, G. Ma, M. Liu, and W. Liu, “Q-switched all-fiber laser with short pulse duration based on tungsten diselenide,” Laser Phys. 28(5), 055104 (2018). [CrossRef]  

27. K. Wang, J. Wang, J. Fan, M. Lotya, A. O’Neill, D. Fox, Y. Feng, X. Zhang, B. Jiang, Q. Zhao, H. Zhang, J. N. Coleman, L. Zhang, and W. J. Blau, “Ultrafast saturable absorption of two-dimensional MoS2 nanosheets,” ACS Nano 7(10), 9260–9267 (2013). [CrossRef]   [PubMed]  

28. Y. Li, N. Dong, S. Zhang, X. Zhang, Y. Feng, K. Wang, L. Zhang, and J. Wang, “Giant two-photon absorption in monolayer MoS2,” Laser Photonics Rev. 9(4), 427–434 (2015). [CrossRef]  

29. X. Zhang, S. Zhang, B. Chen, H. Wang, K. Wu, Y. Chen, J. Fan, S. Qi, X. Cui, L. Zhang, and J. Wang, “Direct synthesis of large-scale hierarchical MoS2 films nanostructured with orthogonally oriented vertically and horizontally aligned layers,” Nanoscale 8(1), 431–439 (2016). [CrossRef]   [PubMed]  

30. K. Wang, J. Wang, J. Fan, M. Lotya, A. O’Neill, D. Fox, Y. Feng, X. Zhang, B. Jiang, Q. Zhao, H. Zhang, J. N. Coleman, L. Zhang, and W. J. Blau, “Ultrafast saturable absorption of two-dimensional MoS2 nanosheets,” ACS Nano 7(10), 9260–9267 (2013). [CrossRef]   [PubMed]  

31. S. Wang, H. Yu, H. Zhang, A. Wang, M. Zhao, Y. Chen, L. Mei, and J. Wang, “Broadband few-layer MoS2 saturable absorbers,” Adv. Mater. 26(21), 3538–3544 (2014). [CrossRef]   [PubMed]  

32. C. Janisch, Y. Wang, D. Ma, N. Mehta, A. L. Elías, N. Perea-López, M. Terrones, V. Crespi, and Z. Liu, “Extraordinary second harmonic generation in tungsten disulfide monolayers,” Sci. Rep. 4(1), 5530 (2015). [CrossRef]   [PubMed]  

33. J. Kang, H. Sahin, and F. M. Peeters, “Tuning carrier confinement in MoS2 /WS2 lateral heterostructure,” J. Phys. Chem. C 119(17), 9580–9586 (2015). [CrossRef]  

34. J. Zhang, J. Wang, P. Chen, Y. Sun, S. Wu, Z. Jia, X. Lu, H. Yu, W. Chen, J. Zhu, G. Xie, R. Yang, D. Shi, X. Xu, J. Xiang, K. Liu, and G. Zhang, “Observation of strong interlayer coupling in MoS2 /WS2 heterostructures,” Adv. Mater. 28(10), 1950–1956 (2016). [CrossRef]   [PubMed]  

35. X. Hong, J. Kim, S. F. Shi, Y. Zhang, C. Jin, Y. Sun, S. Tongay, J. Wu, Y. Zhang, and F. Wang, “Ultrafast charge transfer in atomically thin MoS₂/WS₂ heterostructures,” Nat. Nanotechnol. 9(9), 682–686 (2014). [CrossRef]   [PubMed]  

36. N. J. Huo, J. Kang, Z. M. Wei, S. S. Li, J. B. Li, and S. H. Wei, “Novel and enhanced optoelectronic performances of multilayer MoS2 -WS2 heterostructure transistors,” Adv. Funct. Mater. 24(44), 7025–7031 (2014). [CrossRef]  

37. Y. Yu, S. Hu, L. Su, L. Huang, Y. Liu, Z. Jin, A. A. Purezky, D. B. Geohegan, K. W. Kim, Y. Zhang, and L. Cao, “Equally efficient interlayer exciton relaxation and improved absorption in epitaxial and nonepitaxial MoS2/WS2 heterostructures,” Nano Lett. 15(1), 486–491 (2015). [CrossRef]   [PubMed]  

38. G. H. Han, J. A. Rodríguez-Manzo, C. W. Lee, N. J. Kybert, M. B. Lerner, Z. J. Qi, E. N. Dattoli, A. M. Rappe, M. Drndic, and A. T. Johnson, “Continuous growth of hexagonal graphene and boron nitride in-plane heterostructures by atmospheric pressure chemical vapor deposition,” ACS Nano 7(11), 10129–10138 (2013). [CrossRef]   [PubMed]  

39. A. Mishchenko, J. S. Tu, Y. Cao, R. V. Gorbachev, J. R. Wallbank, M. T. Greenaway, V. E. Morozov, S. V. Morozov, M. J. Zhu, S. L. Wong, F. Withers, C. R. Woods, Y. J. Kim, K. Watanabe, T. Taniguchi, E. E. Vdovin, O. Makarovsky, T. M. Fromhold, V. I. Fal’ko, A. K. Geim, L. Eaves, and K. S. Novoselov, “Twist-controlled resonant tunnelling in graphene/boron nitride/graphene heterostructures,” Nat. Nanotechnol. 9(10), 808–813 (2014). [CrossRef]   [PubMed]  

40. T. Gao, X. Song, H. Du, Y. Nie, Y. Chen, Q. Ji, J. Sun, Y. Yang, Y. Zhang, and Z. Liu, “Temperature-triggered chemical switching growth of in-plane and vertically stacked graphene-boron nitride heterostructures,” Nat. Commun. 6(1), 6835 (2015). [CrossRef]   [PubMed]  

41. H. Chen, J. Yin, J. Yang, X. Zhang, M. Liu, Z. Jiang, J. Wang, Z. Sun, T. Guo, W. Liu, and P. Yan, “Transition-metal dichalcogenides heterostructure saturable absorbers for ultrafast photonics,” Opt. Lett. 42(21), 4279–4282 (2017). [CrossRef]   [PubMed]  

42. K. Chen, X. Wan, J. Wen, W. Xie, Z. Kang, X. Zeng, H. Chen, and J. B. Xu, “Electronic properties of MoS2-WS2 heterostructures synthesized with two-step lateral epitaxial strategy,” ACS Nano 9(10), 9868–9876 (2015). [CrossRef]   [PubMed]  

43. D. Mao, B. Du, D. Yang, S. Zhang, Y. Wang, W. Zhang, X. She, H. Cheng, H. Zeng, and J. Zhao, “Nonlinear saturable absorption of liquid-exfoliated molybdenum/tungsten ditelluride nanosheets,” Small 12(11), 1489–1497 (2016). [CrossRef]   [PubMed]  

44. X. Zhang, S. Zhang, Y. Xie, J. Huang, L. Wang, Y. Cui, and J. Wang, “Tailoring the nonlinear optical performance of two-dimensional MoS2 nanofilms via defect engineering,” Nanoscale 10(37), 17924–17932 (2018). [CrossRef]   [PubMed]  

45. S. Wang, H. Yu, H. Zhang, A. Wang, M. Zhao, Y. Chen, L. Mei, and J. Wang, “Broadband few-layer MoS2 saturable absorbers,” Adv. Mater. 26(21), 3538–3544 (2014). [CrossRef]   [PubMed]  

46. 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 23(21), 27503–27508 (2015). [CrossRef]   [PubMed]  

47. H. Liu, X. W. Zheng, M. Liu, N. Zhao, A. P. Luo, Z. C. Luo, W. C. Xu, H. Zhang, C. J. Zhao, and S. C. Wen, “Femtosecond pulse generation from a topological insulator mode-locked fiber laser,” Opt. Express 22(6), 6868–6873 (2014). [CrossRef]   [PubMed]  

48. J. Sotor, G. Sobon, W. Macherzynski, P. Paletko, K. Grodecki, and K. M. Abramski, “Mode-locking in Er-doped fiber laser based on mechanically exfoliated Sb2Te3 saturable absorber,” Opt. Mater. Express 4(1), 1–6 (2014). [CrossRef]  

49. 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 23(10), 12823–12833 (2015). [CrossRef]   [PubMed]  

50. H. Chen, L. Li, S. C. Ruan, T. Guo, and P. G. Yan, “Fiber-integrated tungsten disulfide saturable absorber (mirror) for pulsed fiber lasers,” Opt. Eng. 55(8), 081318 (2016). [CrossRef]  

51. K. Wu, X. Zhang, J. Wang, X. Li, and J. Chen, “WS2 as a saturable absorber for ultrafast photonic applications of mode-locked and Q-switched lasers,” Opt. Express 23(9), 11453–11461 (2015). [CrossRef]   [PubMed]  

52. R. Khazaeinezhad, S. H. Kassani, H. Jeong, K. J. Park, B. Y. Kim, D. Yeom, and K. Oh, “Ultrafast pulsed all-fiber laser based on tapered fiber enclosed by few-layer WS2 nanosheets,” IEEE Photonic. Tech. L. 27(15), 1581–1584 (2015). [CrossRef]  

53. B. Guo, Y. Yao, P. G. Yan, K. Xu, J. J. Liu, S. G. Wang, and Y. Li, “Dual-wavelength soliton mode-locked fiber laser with a WS2-based fiber taper,” IEEE Photonic. Tech. L. 28(3), 323–326 (2016). [CrossRef]  

54. P. G. Yan, A. J. Liu, Y. S. Chen, H. Chen, S. C. Ruan, C. Y. Guo, S. F. Chen, I. L. Li, H. P. Yang, J. G. Hu, and G. Z. Cao, “Microfiber-based WS2-film saturable absorber for ultra-fast photonics,” Opt. Mater. Express 5(3), 479–489 (2015). [CrossRef]  

55. D. Mao, S. Zhang, Y. Wang, X. Gan, W. Zhang, T. Mei, Y. Wang, Y. Wang, H. Zeng, and J. Zhao, “WS₂ saturable absorber for dissipative soliton mode locking at 1.06 and 1.55 µm,” Opt. Express 23(21), 27509–27519 (2015). [CrossRef]   [PubMed]  

56. J. Lee, J. Park, J. Koo, Y. M. Jhon, and J. H. Lee, “Harmonically mode-locked femtosecond fiber laser using non-uniform, WS2-particle deposited side-polished fiber,” J. Opt. 18(3), 035502 (2016). [CrossRef]  

57. M. Liu, X. W. Zheng, Y. L. Qi, H. Liu, A. P. Luo, Z. C. Luo, W. C. Xu, C. J. Zhao, and H. Zhang, “Microfiber-based few-layer MoS2 saturable absorber for 2.5 GHz passively harmonic mode-locked fiber laser,” Opt. Express 22(19), 22841–22846 (2014). [CrossRef]   [PubMed]  

58. H. Zhang, S. B. Lu, J. Zheng, J. Du, S. C. Wen, D. Y. Tang, and K. P. Loh, “Molybdenum disulfide (MoS2) as a broadband saturable absorber for ultra-fast photonics,” Opt. Express 22(6), 7249–7260 (2014). [CrossRef]   [PubMed]  

59. J. Du, Q. Wang, G. Jiang, C. Xu, C. Zhao, Y. Xiang, Y. Chen, S. Wen, and H. Zhang, “Ytterbium-doped fiber laser passively mode locked by few-layer Molybdenum Disulfide (MoS2) saturable absorber functioned with evanescent field interaction,” Sci. Rep. 4(1), 6346 (2015). [CrossRef]   [PubMed]  

60. R. Khazaeizhad, S. H. Kassani, H. Jeong, D. I. Yeom, and K. Oh, “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(19), 23732–23742 (2014). [CrossRef]   [PubMed]  

61. K. Wu, X. Zhang, J. Wang, and J. Chen, “463-MHz fundamental mode-locked fiber laser based on few-layer MoS(2) saturable absorber,” Opt. Lett. 40(7), 1374–1377 (2015). [CrossRef]   [PubMed]  

62. D. Mao, X. She, B. Du, D. Yang, W. Zhang, K. Song, X. Cui, B. Jiang, T. Peng, and J. Zhao, “Erbium-doped fiber laser passively mode locked with few-layer WSe2/MoSe2 nanosheets,” Sci. Rep. 6(1), 23583 (2016). [CrossRef]   [PubMed]  

63. Z. Q. Luo, Y. Y. Li, M. Zhong, Y. Z. Huang, X. J. Wan, J. Peng, and J. Weng, “Nonlinear optical absorption of few-layer molybdenum diselenide (MoSe2) for passively mode-locked soliton fiber laser,” Photon. Res. 3(3), A79–A86 (2015). [CrossRef]  

64. J. Koo, J. Park, J. Lee, Y. M. Jhon, and J. H. Lee, “Femtosecond harmonic mode-locking of a fiber laser at 3.27 GHz using a bulk-like, MoSe2-based saturable absorber,” Opt. Express 24(10), 10575–10589 (2016). [CrossRef]   [PubMed]  

65. J. Koo, Y. I. Jhon, J. Park, J. Lee, Y. M. Jhon, and J. H. Lee, “Near-infrared saturable absorption of defective bulk-structured WTe2 for femtosecond laser mode-locking,” Adv. Funct. Mater. 26(41), 7454–7461 (2016). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (6)

Fig. 1
Fig. 1 Atomic and electronic structure of the MoS2-WS2 heterostructure. (a) Top and side views of the MoS2-WS2 heterostructure, the dashed rectangle denotes the primitive cell. (b), (c) and (d) are band structures of MoS2-WS2 heterostructure, monolayer MoS2 and monolayer WS2, respectively. Here, the fermi level is set to be zero, and the orange line denotes valence-band maximum for the holes and conduction-band minimum for the electrons. Besides, red points project the contribution from the MoS2 in (b). (e) is the band alignment of the MoS2-WS2 heterostructure. The energy levels of MoS2 and monolayer WS2 slabs are shown in both sides. Here, the monolayer MoS2, monolayer WS2, and MoS2-WS2 heterostructure are considered.
Fig. 2
Fig. 2 SEM of the MoS2-WS2 heterostructure SA. (a) The surface morphology; (b) The thickness of heterostructure indicated by the lateral.
Fig. 3
Fig. 3 (a) The Raman spectra of the MoS2, WS2 and MoS2-WS2 heterostructure; (b) The nonlinear absorption of the MoS2-WS2 heterostructure SA.
Fig. 4
Fig. 4 Experimental device diagram of the passively mode-locked EDF laser employed with the MoS2-WS2 heterostructure SA.
Fig. 5
Fig. 5 The performance of the passively mode-locked EDF laser employed with the MoS2-WS2 heterostructure SA. (a) The optical spectrum located at 1560 nm with 3 dB spectral width of 24.4 nm; (b) The mode-locked pulse train; (c) The radio-frequency spectrum; (d) The symmetrical autocorrelation trace of mode-locked pulses.
Fig. 6
Fig. 6 The continuous monitoring of output power of the fiber laser.

Tables (2)

Tables Icon

Table 1 Band Gap (Eg), Effective Mass and Carrier Mobility (µ) of Monolayer and Heterostructure Materials

Tables Icon

Table 2 Comparisons of Mode-Locked Fiber Lasers Employed with Different SAs

Equations (1)

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

μ= 2e 3 C 3 k B T | m* | 2 E 2 ,
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.