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

Vertically standing layered MoS2 nanosheets on TiO2 nanofibers for enhanced nonlinear optical property

Open Access Open Access

Abstract

Vertical layered MoS2 nanosheets rooting into TiO2 nanofibers were successfully prepared by a facile two-step method: prefabrication of porous TiO2 nanofibers based on an electrospinning technique, and assembly of MoS2 ultrathin nanosheets through a simple hydrothermal reaction. Significant enhancement of nonlinear optical response of the MoS2/TiO2 nanocomposite was confirmed by an open-aperture z-scan measurement. The nanocomposite displayed strong optical limiting (OL) effects to ultrafast laser pulses with a low OL threshold of ~22.3 mJ/cm2, which is lower than that of pristine TiO2 nanofibers and MoS2 nanosheets. In addition to the contribution of the strong nonlinear absorption of MoS2 nanosheets and TiO2 nanofibers, such phenomenon is also attributed to the unique structure of vertically standing layered MoS2 nanosheets on TiO2 nanofibers with a large amount of exposed edge states, large surface areas and fast electron transfer between TiO2 and MoS2. This work broadens our vision to engineering novel hierarchical MoS2-based nanocomposite for efficiently enhanced nonlinear light-matter interaction.

© 2016 Optical Society of America

1. Introduction

With increasing use of laser for various fields such as medicine, electronics, scientific research, industry and military, the development of nonlinear optical (NLO) material (saturable absorber and optical limiting (OL) material) which can modulate the light intensity is crucial. Saturable absorber, which allows high intensity light pass, can be employed for Q-switching, mode locking and pulse compression [1–3]. OL material, in contrast, which effectively forbids high intensity light but allows low intensity pass, can be applied to protect human eyes and sensitive instruments from laser induced damage [4]. Despite a variety of optical nanomaterials have been extensively studied for their NLO behavior, for instance, noble metal nanoparticles, quantum dots, organic nanomaterials, carbon-based nanomaterials and graphene [5–9], there is still in great request for highly NLO responsive components.

Layered molybdenum disulfide (MoS2), an emerging graphene-like two-dimensional (2D) nanomaterial, has aroused the interest of researchers because of its exotic electronic and optical properties induced by the indirect-to-direct transition with decreasing the number of layer [10–21]. Researches demonstrated that layered MoS2 exhibits strong nonlinear light-matter interaction due to the Van Hove singularities in the density of states [1,22]. Due to the unique performance, layered MoS2 can be a promising NLO material used in optoelectronics [23–25]. However, the inherent drawbacks including the basic layered building blocks limit its practical applications. Recently, the fabrication of 2D nanocomposites assembled by MoS2 and other semiconductor materials has been proved as an attractive strategy to obtain novel 2D nanomaterials with enhanced electronic and optoelectronic properties [1,26–30]. The greatest motivation to study the nanocomposites is to determine their unique structures and synergistic coupling effects between the constituents. MoS2/carbon [29], MoS2/graphene [1], MoS2/CdS [28] and MoS2/ZnO nanocomposites [27] were successfully synthesized and they cannot only boost photocatalytic activity and electrochemical property, but also increase NLO response. Most recently, MoS2/TiO2 composite was also prepared, and it displayed excellent lithium storage property with high specific capacity and outstanding rate capability due to the integrated smart architecture [31]. In addition, owing to the more exposed active edge sites compared to the pristine MoS2 nanosheets, MoS2/TiO2 composite exhibited improved hydrogen evolution reaction [32]. However, the NLO response of the MoS2/TiO2 nanocomposite remains un-explored so far. Interestingly, Zhang et al reported that the passive Q-switching behavior for the hierarchical MoS2 thin film was better than that of exfoliated MoS2 and comparable to that of chemical vapor deposition MoS2 nanosheets. They inferred that the superior ultrafast pulse emission may be associated with the large NLO absorption which induced by the hierarchical structure. This structure was composed of orthogonally oriented vertical/horizontal layers which had a crowd of exposed edge states [33]. Strong second-order NLO susceptibilities of the 2D MoS2 crystal were observed by Yin et al most recently, due to the abundant edge states [34]. In addition, TiO2, as a direct band gap semiconductor, also has been confirmed experimentally that it has excellent NLO property [35]. These factors were believed that investigation NLO property of the MoS2/TiO2 nanocomposite is of tremendous significance because the combination of layered MoS2 and TiO2 will be likely to enhance the NLO property.

Herein, the MoS2/TiO2 composite was successfully synthesized by an electrospinning technique and a hydrothermal process. SEM, TEM, XRD, Raman spectroscopy and XPS were utilized to characterize the MoS2/TiO2 nanocomposite. Vertically standing layered MoS2 nanosheets on TiO2 nanofibers not only endows structure stability, large surface areas and intimate contact between TiO2 and MoS2 for fast electrons transfer, but also possesses maximally exposed edge states [20,21]. The NLO response of the MoS2/TiO2/polyvinyl alcohol (PVA) composite was investigated applying an open-aperture (OA) z-scan measurement. The composite film exhibited improved OL property, which demonstrates that the MoS2/TiO2/PVA composite is a new promising OL material for laser.

2. Experimental

Typically, 0.5 g polyvinylpyrrolidone (PVP, Mw = 1,300,000, Sigma-Aldrich) power was dissolved in 9.5 mL absolute ethanol which contained 2 mL acetic acid. After stirring for 2 h, 2 mL tetrabutyl titanate (Ti(OBu)4, Sigma-Aldrich) was added to the homogeneous solution and stirred for another 30 min. The resultant precursor solution was loaded into a 20 mL plastic syringe attached with a blunt metal needle of 0.6 mm inner diameter. The flow rate was 1 mL/h and a grounded stainless steel plate was placed 15 cm below the spinneret to collect the nanofibers. A high voltage of 13 kV was employed. The as-collected electrospun fibers were calcined in a covered corundum crucible in a tube furnace at 500 °C for 2 h under air atmosphere. At the same time, 0.0450 g sodium molybdate (Na2MoO4·2H2O, Sigma-Aldrich) and 0.0090 g thioacetamide (C2H5NS, Sigma-Aldrich) were dispersed in 20 mL deionized water under vigorous stirring to achieve a homogeneous solution. Subsequently, 0.0020 g calcined TiO2 nanofibers were added into the above solution and continuously stirred to get a suspension. The mixture was transferred to a 50 mL Teflon-lined stainless steel autoclave and heated at 220 °C for 24 h. The reaction system was allowed to cool down to room temperature naturally. Finally, the resulting precipitate was rinsed with ethanol by several times and finally dried at 60 °C for 12 h in vacuum. Then, the products were dispersed in PVA to fabricate the MoS2/TiO2/PVA composite film according to the previous work [3]. For comparison, the nanoscale TiO2/PVA and MoS2/PVA composite films were also prepared.

Morphologies and sizes of the products were observed by field-emission scanning electron microscope (FESEM, Nova NanoSEM430). Transmission electron microscopy (TEM) and high resolution TEM (HRTEM) characterizations were conducted on a JEOL-2100F field emission microscope operating at 200 kV, the sample was prepared via dispersing in ethanol and dropping into a copper grid. X-ray diffraction (XRD) patterns were recorded with a Bruker diffractometer utilizing Cu Kα radiation (λ = 1.5418 Å). Raman spectra were obtained by a Raman spectrometer (Renishaw inVia, Gloucestershire, UK) with a 785 nm laser as the excitation source. Chemical compositions of the products were characterized by X-ray photoelectron spectra (XPS, K-Alpha, Thermo Scientific, UK) with achromatic 200W Al Kα as the X-ray source and the resolution of 0.05 eV. NLO characteristics of the different samples were measured applying the femtosecond (fs) OA z-scan technique. Laser pulses with a pulse width of 130 fs were generated from a commercial Ti: sapphire laser at the wavelength of 800 nm with a repetition rate of 1 kHz. The normalized transmittance of the product was measured with respect to the incident pump power, which was changed via translating the product along the propagation direction of a focused laser beam with a linear motorized stage controlled by the step motor controller. To more precisely confirm the measured data, CS2 solution contained in a cuvette (1 mm in thick) was employed to calibrate. All the measurements were performed at room temperature.

3. Results and discussion

Figure 1(a) presents the morphology of the calcined electrospun TiO2 nanofibers with a diameter of approximately 200 nm. It notes that the surface of the TiO2 nanofibers which consisting of interconnected TiO2 particles is highly rough and porous, which is attributed to the removal of PVP porogenic agents by calcination. Figure 1(b) displays that MoS2 nanosheets with the absence of TiO2 nanofibers are clustered to generate micron-sized agglomerates. After undergoing an electrospinning technique and a hydrothermal coating reaction, MoS2/TiO2 nanocomposite was successfully prepared and exhibited in Fig. 1(c). The TEM image in Fig. 1(d) reveals that the MoS2 sheets are deeply rooted into the TiO2 nanofibers. The vertical growth of MoS2 nanosheets on TiO2 nanofibers endows a highly structural stability, larger surface areas, intimate contact between TiO2 and MoS2, and more exposed MoS2 edge states. The coated MoS2 nanosheets are flexible, curly, and transparent, indicative of the ultrathin 2D nanosheets features. The freely suspended MoS2 sheets show apparent distortion edges and random elastic deformations which facilitate the stability of nanomaterials. Figure 1(e) displays a HRTEM image taken from the edge of the MoS2/TiO2 nanocomposite. The uniform d interlayer spacing of about 0.642 nm is observed, corresponding to the (002) plane of hexagonal-phase MoS2. As directly evidenced from the folded edge in Fig. 1(e), the MoS2 nanosheets mainly contain 3 sandwiched S-Mo-S layers, that is, the thickness is below 3 nm. In addition, Fig. 1(e) also apparently presents the lattice fringes with a spacing of 0.374 nm, which is in good agreement with the (101) plane of anatase TiO2. Selected area electron diffraction (SAED) pattern in Fig. 1(f) further manifests the existence of TiO2 and MoS2 crystal phases in the MoS2/TiO2 nanocomposite.

 figure: Fig. 1

Fig. 1 (a) SEM images of TiO2 nanofibers, (b) TEM images of pure MoS2 nanosheets, (c) SEM, (d) TEM, (e) HRTEM images and SAED pattern of MoS2/TiO2 nanocomposite.

Download Full Size | PDF

Optical absorption spectra of the TiO2 nanofibers, pure MoS2 nanosheets and MoS2/TiO2 nanocomposite were depicted in Fig. 2(a). Two absorption bands at 380-450 and 600-700 nm regions are also observed in the MoS2/TiO2 nanocomposite, indicative of the features of 2D layered MoS2 nanosheets with hexagonal symmetry. The dual peaks located at 619 and 667 nm are assigned to the inter-band excitonic transitions at the K point of the 2D Brillouin zone of MoS2, identified as the B and A transitions [23]. In addition, according to a distribution for 2D MoS2 nanosheets with different thickness reported by Wang et al [16], we can also deduce that the average thickness of the MoS2 nanosheets in MoS2/TiO2 nanocomposite is less than 3 nm from the A exciton position (667 nm).

 figure: Fig. 2

Fig. 2 (a) Optical absorption spectra, (b) XRD patterns and (c) Raman spectra of the TiO2 nanofibers, pure MoS2 nanosheets and MoS2/TiO2 nanocomposite, respectively.

Download Full Size | PDF

XRD technique is applied to investigate the crystal structure of the resultant material. As shown in Fig. 2(b), all of the diffraction peaks can be well-indexed to anatase-phase TiO2 (JCPDS card no. 21-1272) and hexagonal-phase MoS2 (JCPDS card no. 24-0513), respectively. No obvious peaks from the impurity phase are observed in the XRD patterns, indicating the MoS2/TiO2 nanocomposite containing only TiO2 and MoS2 components. It is worth mentioning that the (002) peak of MoS2, corresponding to the periodicity in c-axis of MoS2 plane, is very weak in the MoS2/TiO2 nanocomposite compared to the pristine MoS2, which declared that the c-axis stacking of MoS2 happening in pure MoS2 is effectively restrained in MoS2/TiO2 nanocomposite.

Raman spectrum is conducted to identify the atomic structure arrangement of sample. As presented in Fig. 2(c), the phonon vibrational mode of the pure TiO2 nanofibers at 143.9, 398.2, 516.2 and 638.9 cm−1 coincide with the Eg1, B1g1, A1g+B1g2 and Eg2 modes of anatase TiO2, respectively [36]. While two characteristic peaks of the pristine MoS2 at 383.2 and 406.6 cm−1 can be assigned to the in-plane E2g1 and out-of-plane A1g vibrational modes of hexagonal MoS2, respectively [23]. In the Raman spectrum of the MoS2/TiO2 nanocomposite, there is a blue shift of ~3.5 cm−1 for Eg1 mode compared to the individual TiO2 nanofibers, which is mostly owing to the surface strain, it is possibly caused by the TiO2 surface. Also, a surface strain mostly formed in the coated MoS2 nanosheets, which interprets the red shift of ~1.3 cm−1 for E2g1 mode and ~1.2 cm−1 forA1g mode compared with pristine MoS2 [36].

The XPS was studied to further understand the chemical composition and bonding configuration of the as-prepared samples and the XPS measurements were presented in Fig. 3(a)-3(c). For pristine MoS2 nanosheets, the binding energies of Mo 3d3/2, Mo 3d5/2, S 2p1/2 and S 2p3/2 peaks [Fig. 3(a)-3(b)] were located at 232.3, 229.1, 163.1 and 161.8 eV, respectively [37]. After MoS2 nanosheets vertically rooting into the TiO2 nanofibers, these peaks shift to 232.0, 228.7, 162.7 and 161.5 eV, respectively, which are lower than the homologous values of pristine MoS2 nanosheets. Such shifts imply electronic interaction between TiO2 nanofibers and MoS2 nanosheets [38]. For O 1s peaks of the MoS2/TiO2 nanocomposite in Fig. 3(c), apart from the peak at 529.2 eV ascribed to the Ti-O-Ti bond, 532.3 eV corresponding to O-H bonds of the surface-adsorbed water, the peak at 530.9 eV emerged which cannot be observed in pristine TiO2 nanofibers, it might be associated with the formation of the Ti-O-Mo bonds between MoS2 nanosheets and TiO2 nanofibers [32,39].

 figure: Fig. 3

Fig. 3 XPS spectra pure MoS2 and MoS2/TiO2 nanocomposites: (a) Mo 3d peaks and (b) S 2p peaks of MoS2 and MoS2/TiO2, respectively. (c) O1s peaks of TiO2 and MoS2/TiO2, respectively.

Download Full Size | PDF

Figure 4(a) illustrates the typical OA z-scan curves for the MoS2/TiO2 nanocomposite, obvious reverse sarturable absorption (RSA) is observed. The MoS2/TiO2 nanocomposite exhibits increased RSA, characterized by the significantly reducing valley value of transmittance with increasing the incident pump power. In other words, the MoS2/TiO2 nanocomposite can effectively inhibit high intensity light but allow low intensity light to pass, which clearly suggest the fs OL effects. Furthermore, the MoS2/TiO2 nanocomposite shows much stronger nonlinear absorption than that of the pristine TiO2 nanofibers and pure MoS2 nanosheets in the same input pump power [Fig. 4(b)]. Owing to the lack of the any pronounced nonlinear behavior for the pure PVA film, we can suggest that the significant RSA mainly originate from the MoS2/TiO2 nanocomposite. Fitting the OA z-scan curves [Fig. 4(b)] using formula given in the previous reports [3], the nonlinear absorption coefficient β ~23.68 cm/GW of MoS2/TiO2 nanocomposite was calculated, which is larger than the sum of that of TiO2 nanofibers (β ~5.90 cm/GW) and MoS2 nanosheets (β ~11.03 cm/GW). The imaginary part of the third-order NLO susceptibility Im χ(3) is directly associated with β, so that Im χ(3) ~3.32 x 10−12 esu of the MoS2/TiO2 nanocomposite can be deduced. To eliminate the influence of the linear absorption α0, the nonlinear figure of merit (FOM) is used to evaluate the NLO response of the materials. The larger FOM signifies the better NLO property of the sample. FOM for the third-order optical nonlinearity is defined as FOM = |Im χ(3)/α0| [23]. Hence, FOM of the MoS2/TiO2 nanocomposite is determined to be 3.10 x 10−14 esu cm−1, which is much larger than the sum of that of TiO2 nanofibers (FOM ~0.77 x 10−14 esu cm−1) and MoS2 nanosheets (FOM ~1.44 x 10−14 esu cm−1). In addition, The OL threshold (FOL, deifind as the input fluence at which the normalized transmittance falls to 50%) is a crucial parameter to evaluate the OL effects of a given material [24]. The fs OL responses of the samples in different input fluences were presented in Fig. 4(c), it is clear that the FOL value of the MoS2/TiO2 nanocomposite is 22.3 mJ/cm2, significantly lower than that of the TiO2 nanofibers and MoS2 nanosheets. All these results indicate that the MoS2/TiO2 nanocomposite has great potential as a type of excellent NLO material for OL utilization.

 figure: Fig. 4

Fig. 4 (a) OA z-scan curves of the MoS2/TiO2 nanocomposite with incident pump powers from 10.7 to 68.0 μW, (b) OA z-scan curves of the TiO2 nanofibers, pure MoS2 nanosheets and MoS2/TiO2 nanocomposite with an incident pump power of 21.0 μW, (c) OL responses of the TiO2 nanofibers, pure MoS2 nanosheets and MoS2/TiO2 nanocomposite under different input fluences and (d) schematic diagram of the nonlinear absorption in MoS2/TiO2 nanocomposite.

Download Full Size | PDF

The improved nonlinear light-matter interaction between the MoS2/TiO2 nanocomposite and fs laser, other than the contribution of the strong nonlinear absorption of MoS2 nanosheets and TiO2 nanofibers, can be further understood as follows: the hierarchical structure with vertical layered MoS2 nanosheets rooting into TiO2 nanofibers can afford a crowd of exposed edge states and larger surface areas to absorb light, leading to the significant enhancement of NLO property [3,33]. In addition, such unique structure also can enable the efficiently intimate contact between TiO2 and MoS2. When excited by the fs laser, the highly efficient interface electron transfer between TiO2 and MoS2, and excited state absorption (ESA) may occur, as shown in Fig. 4(d). The electrons are excited to either the excited state (ES) of MoS2 or TiO2 via multiphoton absorption (MA). The upward band bending and the interfacial band energy build up a strong internal electric filed. Under this driving force, the photoexcited electrons flow into TiO2. Hence, owing to interface charge transfer, charge accumulates in the ES of TiO2, which would improved the ESA process of TiO2, and developed a strong potential between the ES and the ground state (GS), which also promotes MA for TiO2 and MoS2 [40].

4. Conclusions

We have successfully fabricated the MoS2/TiO2 nanocomposite with vertical layered MoS2 nanosheets rooting into highly porous electrospum TiO2 nanofibers through a hydrothermal approach. The MoS2/TiO2 nanocomposite exhibited enhanced OL effects with very low threshold FOL ~22.3 mJ/cm2, identified by an OA z-scan measurement. Apart from the contribution of the strong OL responses of MoS2 nanosheets and TiO2 nanofibers, the hierarchical structure with a crowd of exposed edge states, large surface areas and fast electron transfer from TiO2 and MoS2 also contribute to the enhancement of OL property of MoS2/TiO2 nanocomposite. This work advances the development of the novel 2D nanocomposite for efficiently improved NLO performance.

Funding

The National Natural Science Foundation of China (Grant No. 51132004, 51102096, 11404114), Guangdong Natural Science Foundation (Grant No. S2011030001349, 1045106410104887) and Open Fund from the State Key Laboratory of Precision Spectroscopy (East China Normal University).

References and links

1. Y. Jiang, L. Miao, G. Jiang, Y. Chen, X. Qi, X. F. Jiang, H. Zhang, and S. Wen, “Broadband and enhanced nonlinear optical response of MoS2/graphene nanocomposites for ultrafast photonics applications,” Sci. Rep. 5(282), 16372 (2015). [CrossRef]   [PubMed]  

2. 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]  

3. R. Wei, H. Zhang, X. Tian, T. Qiao, Z. Hu, Z. Chen, X. He, Y. Yu, and J. Qiu, “MoS2 nanoflowers as high performance saturable absorbers for an all-fiber passively Q-switched erbium-doped fiber laser,” Nanoscale 8(14), 7704–7710 (2016). [CrossRef]   [PubMed]  

4. X. Zheng, M. Feng, and H. Zhan, “Giant optical limiting effect in ormosil gel glasses doped with graphene oxide materials,” J. Mater. Chem. C Mater. Opt. Electron. Devices 1(41), 6759–6766 (2013). [CrossRef]  

5. B. Anand, S. R. Krishnan, R. Podila, S. S. S. Sai, A. M. Rao, and R. Philip, “The role of defects in the nonlinear optical absorption behavior of carbon and ZnO nanostructures,” Phys. Chem. Chem. Phys. 16(18), 8168–8177 (2014). [CrossRef]   [PubMed]  

6. X. F. Jiang, L. Polavarapu, S. T. Neo, T. Venkatesan, and Q. H. Xu, “Graphene oxides as tunable broadband nonlinear optical materials for femtosecond laser pulses,” J. Phys. Chem. Lett. 3(6), 785–790 (2012). [CrossRef]   [PubMed]  

7. M. O. Senge, M. Fazekas, E. G. A. Notaras, W. J. Blau, M. Zawadzka, O. B. Locos, and E. M. N. Mhuircheartaigh, “Nonlinear optical properties of porphyrins,” Adv. Mater. 19(19), 2737–2774 (2007). [CrossRef]  

8. D. Tan, Y. Yamada, S. Zhou, Y. Shimotsuma, K. Miura, and J. Qiu, “Carbon nanodots with strong nonlinear optical response,” Carbon 69(69), 638–640 (2014). [CrossRef]  

9. H. Zhang, Z. Hu, Z. Ma, M. Gecevičius, G. Dong, S. Zhou, and J. Qiu, “Anisotropically enhanced nonlinear optical properties of ensembles of gold nanorods electrospun in polymer nanofiber film,” ACS Appl. Mater. Interfaces 8(3), 2048–2053 (2016). [CrossRef]   [PubMed]  

10. M. Amani, D. H. Lien, D. Kiriya, J. Xiao, A. Azcatl, J. Noh, S. R. Madhvapathy, R. Addou, S. Kc, M. Dubey, K. Cho, R. M. Wallace, S. C. Lee, J. H. He, J. W. Ager 3rd, X. Zhang, E. Yablonovitch, and A. Javey, “Near-unity photoluminescence quantum yield in MoS2,” Science 350(6264), 1065–1068 (2015). [CrossRef]   [PubMed]  

11. M. Chhowalla, H. S. Shin, G. Eda, L. J. Li, K. P. Loh, and H. Zhang, “The chemistry of two-dimensional layered transition metal dichalcogenide nanosheets,” Nat. Chem. 5(4), 263–275 (2013). [CrossRef]   [PubMed]  

12. X. Huang, Z. Zeng, S. Bao, M. Wang, X. Qi, Z. Fan, and H. Zhang, “Solution-phase epitaxial growth of noble metal nanostructures on dispersible single-layer molybdenum disulfide nanosheets,” Mol. Ther. 4(1444), 1444 (2013). [PubMed]  

13. J. Kibsgaard, Z. Chen, B. N. Reinecke, and T. F. Jaramillo, “Engineering the surface structure of MoS2 to preferentially expose active edge sites for electrocatalysis,” Nat. Mater. 11(11), 963–969 (2012). [CrossRef]   [PubMed]  

14. S. Najmaei, Z. Liu, W. Zhou, X. Zou, G. Shi, S. Lei, B. I. Yakobson, J. C. Idrobo, P. M. Ajayan, and J. Lou, “Vapour phase growth and grain boundary structure of molybdenum disulphide atomic layers,” Nat. Mater. 12(8), 754–759 (2013). [CrossRef]   [PubMed]  

15. B. Radisavljevic, A. Radenovic, J. Brivio, V. Giacometti, and A. Kis, “Single-layer MoS2 transistors,” Nat. Nanotechnol. 6(3), 147–150 (2011). [CrossRef]   [PubMed]  

16. 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]  

17. L. Zhi, H. Zhang, Z. Yang, W. Liu, and B. Wang, “Interface coassembly of mesoporous MoS2 based-frameworks for enhanced near-infrared light driven photocatalysis,” Chem. Commun. (Camb.) 52(38), 6431–6434 (2016). [CrossRef]   [PubMed]  

18. G. X. Bai, M. K. Tsang, and J. H. Hao, “Luminescent ions in advanced composite materials for multifunctional applications,” Adv. Mater. 26(35), 6330–6350 (2016).

19. G. Bai, S. Yuan, Y. Zhao, Z. Yang, S. Y. Choi, Y. Chai, S. F. Yu, S. P. Lau, and J. Hao, “2D layered materials of rare-earth Er-doped MoS2 with NIR-to-NIR down- and up-conversion photoluminescence,” Adv. Mater. 28(34), 7472–7477 (2016). [CrossRef]   [PubMed]  

20. Y. Jung, J. Shen, Y. Liu, J. M. Woods, Y. Sun, and J. J. Cha, “Metal seed layer thickness-induced transition from vertical to horizontal growth of MoS2 and WS2.,” Nano Lett. 14(12), 6842–6849 (2014). [CrossRef]   [PubMed]  

21. D. Kong, H. Wang, J. J. Cha, M. Pasta, K. J. Koski, J. Yao, and Y. Cui, “Synthesis of MoS2 and MoSe2 films with vertically aligned layers,” Nano Lett. 13(3), 1341–1347 (2013). [CrossRef]   [PubMed]  

22. K. G. Zhou, M. Zhao, M. J. Chang, Q. Wang, X. Z. Wu, Y. Song, and H. L. Zhang, “Size-dependent nonlinear optical properties of atomically thin transition metal dichalcogenide nanosheets,” Small 11(6), 694–701 (2015). [CrossRef]   [PubMed]  

23. 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]  

24. R. Wei, H. Zhang, X. He, Z. Hu, X. Tian, Q. Xiao, Z. Chen, and J. Qiu, “Versatile preparation of ultrathin MoS2 nanosheets with reverse saturable absorption response,” Opt. Mater. Express 5(8), 1807–1814 (2015). [CrossRef]  

25. 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]  

26. F. Liu, Y. Jiang, J. Yang, M. Hao, Z. Tong, L. Jiang, and Z. Wu, “MoS2 nanodot decorated In2S3 nanoplates: a novel heterojunction with enhanced photoelectrochemical performance,” Chem. Commun. (Camb.) 52(9), 1867–1870 (2016). [CrossRef]   [PubMed]  

27. B. Qu, Q. Ouyang, X. Yu, W. Luo, L. Qi, and Y. Chen, “Nonlinear absorption, nonlinear scattering, and optical limiting properties of MoS2-ZnO composite-based organic glasses,” Phys. Chem. Chem. Phys. 17(8), 6036–6043 (2015). [CrossRef]   [PubMed]  

28. C. Wang, H. Lin, Z. Xu, H. Cheng, and C. Zhang, “One-step hydrothermal synthesis of flowerlike MoS2/CdS heterostructures for enhanced visible-light photocatalytic activities,” RSC Advances 5(20), 15621–15626 (2015). [CrossRef]  

29. X. Xie, T. Makaryan, M. Zhao, K. L. Van Aken, Y. Gogotsi, and G. Wang, “MoS2 nanosheets vertically aligned on carbon paper: A freestanding electrode for highly reversible sodium-ion batteries,” Adv. Energy Mater. 6(5), 1502161 (2016). [CrossRef]  

30. M. Zhao, M. J. Chang, Q. Wang, Z. T. Zhu, X. P. Zhai, M. Zirak, A. Z. Moshfegh, Y. L. Song, and H. L. Zhang, “Unexpected optical limiting properties from MoS2 nanosheets modified by a semiconductive polymer,” Chem. Commun. (Camb.) 51(61), 12262–12265 (2015). [CrossRef]   [PubMed]  

31. B. Chen, N. Zhao, L. Guo, F. He, C. Shi, C. He, J. Li, and E. Liu, “Facile synthesis of 3D few-layered MoS2 coated TiO2 nanosheet core-shell nanostructures for stable and high-performance lithium-ion batteries,” Nanoscale 7(30), 12895–12905 (2015). [CrossRef]   [PubMed]  

32. C. Liu, L. Wang, Y. Tang, S. Luo, Y. Liu, S. Zhang, Y. Zeng, and Y. Xu, “Vertical single or few-layer MoS2 nanosheets rooting into TiO2 nanofibers for highly efficient photocatalytic hydrogen evolution,” Appl. Catal. B 164(8), 1–9 (2015).

33. 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]  

34. X. Yin, Z. Ye, D. A. Chenet, Y. Ye, K. O’Brien, J. C. Hone, and X. Zhang, “Edge nonlinear optics on a MoS2 atomic monolayer,” Science 344(6183), 488–490 (2014). [CrossRef]   [PubMed]  

35. S. Divya, V. P. N. Nampoori, P. Radhakrishnan, and A. Mujeeb, “Morphology dependent dispersion of third-order optical nonlinear susceptibility in TiO2,” Appl. Phys. Adv. Mater. 114(4), 1079–1084 (2014).

36. M. Shen, Z. Yan, L. Yang, P. Du, J. Zhang, and B. Xiang, “MoS2 nanosheet/TiO2 nanowire hybrid nanostructures for enhanced visible-light photocatalytic activities,” Chem. Commun. (Camb.) 50(97), 15447–15449 (2014). [CrossRef]   [PubMed]  

37. J. Xie, J. Zhang, S. Li, F. Grote, X. Zhang, H. Zhang, R. Wang, Y. Lei, B. Pan, and Y. Xie, “Controllable disorder engineering in oxygen-incorporated MoS2 ultrathin nanosheets for efficient hydrogen evolution,” J. Am. Chem. Soc. 135(47), 17881–17888 (2013). [CrossRef]   [PubMed]  

38. W. Zhou, Z. Yin, Y. Du, X. Huang, Z. Zeng, Z. Fan, H. Liu, J. Wang, and H. Zhang, “Synthesis of few-layer MoS2 nanosheet-coated TiO2 nanobelt heterostructures for enhanced photocatalytic activities,” Small 9(1), 140–147 (2013). [CrossRef]   [PubMed]  

39. C. Liao, Z. Ma, G. Dong, and J. Qiu, “BiOI nanosheets decorated TiO2 nanofiber: Tailoring water purification performance of photocatalyst in structural and photo-responsivity aspects,” Appl. Surf. Sci. 314(30), 481–489 (2014). [CrossRef]  

40. J. G. Tao, J. W. Chai, L. X. Guan, J. S. Pan, and S. J. Wang, “Effect of interfacial coupling on photocatalytic performance of large scale MoS2/TiO2 hetero-thin films,” Appl. Phys. Lett. 106(8), 081602 (2015). [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 (4)

Fig. 1
Fig. 1 (a) SEM images of TiO2 nanofibers, (b) TEM images of pure MoS2 nanosheets, (c) SEM, (d) TEM, (e) HRTEM images and SAED pattern of MoS2/TiO2 nanocomposite.
Fig. 2
Fig. 2 (a) Optical absorption spectra, (b) XRD patterns and (c) Raman spectra of the TiO2 nanofibers, pure MoS2 nanosheets and MoS2/TiO2 nanocomposite, respectively.
Fig. 3
Fig. 3 XPS spectra pure MoS2 and MoS2/TiO2 nanocomposites: (a) Mo 3d peaks and (b) S 2p peaks of MoS2 and MoS2/TiO2, respectively. (c) O1s peaks of TiO2 and MoS2/TiO2, respectively.
Fig. 4
Fig. 4 (a) OA z-scan curves of the MoS2/TiO2 nanocomposite with incident pump powers from 10.7 to 68.0 μW, (b) OA z-scan curves of the TiO2 nanofibers, pure MoS2 nanosheets and MoS2/TiO2 nanocomposite with an incident pump power of 21.0 μW, (c) OL responses of the TiO2 nanofibers, pure MoS2 nanosheets and MoS2/TiO2 nanocomposite under different input fluences and (d) schematic diagram of the nonlinear absorption in MoS2/TiO2 nanocomposite.
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.