• Opto-Electronic Advances
  • Vol. 6, Issue 10, 220162 (2023)
Yiduo Wang1, Yingwei Wang1、*, Yulan Dong2, Li Zhou1, Jianlong Kang1, Ning Wang1, Yejun Li1, Xiaoming Yuan1, Zhengwei Zhang1, Han Huang1, Mengqiu Long1, Si Xiao1, and Jun He1、**
Author Affiliations
  • 1Hunan Key Laboratory of Nanophotonics and Devices, School of Physics and Electronics, Central South University, Changsha 410083, China
  • 2Key Laboratory of Hunan Province for Statistical Learning and Intelligent Computation, School of Mathematics and Statistics, Hunan University of Technology and Business, Changsha 410205, China
  • show less
    DOI: 10.29026/oea.2023.220162 Cite this Article
    Yiduo Wang, Yingwei Wang, Yulan Dong, Li Zhou, Jianlong Kang, Ning Wang, Yejun Li, Xiaoming Yuan, Zhengwei Zhang, Han Huang, Mengqiu Long, Si Xiao, Jun He. 2D Nb2CTx MXene/MoS2 heterostructure construction for nonlinear optical absorption modulation[J]. Opto-Electronic Advances, 2023, 6(10): 220162 Copy Citation Text show less

    Abstract

    Two-dimensional (2D) nonlinear optical mediums with high and tunable light modulation capability can significantly stimulate the development of ultrathin, compact, and integrated optoelectronics devices and photonic elements. 2D carbides and nitrides of transition metals (MXenes) are a new class of 2D materials with excellent intrinsic and strong light-matter interaction characteristics. However, the current understanding of their photo-physical properties and strategies for improving optical performance is insufficient. To address this issue, we rationally designed and in situ synthesized a 2D Nb2C/MoS2 heterostructure that outperforms pristine Nb2C in both linear and nonlinear optical performance. Excellent agreement between experimental and theoretical results demonstrated that the Nb2C/MoS2 inherited the preponderance of Nb2C and MoS2 in absorption at different wavelengths, resulting in the broadband enhanced optical absorption characteristics. In addition to linear optical modulation, we also achieved stronger near infrared nonlinear optical modulation, with a nonlinear absorption coefficient of Nb2C/MoS2 being more than two times that of the pristine Nb2C. These results were supported by the band alinement model which was determined by the X-ray photoelectron spectroscopy (XPS) experiment and first-principal theory calculation. The presented facile synthesis approach and robust light modulation strategy pave the way for broadband optoelectronic devices and optical modulators.

    Introduction

    Since their discovery in 20041, two dimensional (2D) layered materials have been recognized as the foundation of the next generation optoelectronics devices2 and photonic elements3 due to their strong and unique light-matter interaction, such as ultrafast and broad optical response in graphene4, strong excitonic optical properties in single layer transition-metal dichalcogenides (TMDs)5-7, and tunable direct optical band gap in black phosphorus (BP)8, 9, etc. Recently, a series of newly emerged 2D materials10-13 have shown interesting nonlinear optical response. In light of remarkable optical properties, 2D layered materials endow new optoelectronics applications that are radically different from their bulk counterparts14-17.

    MXenes are a new family of 2D materials. It can be produced by selective etching methods. Since the first reported Ti3C2 MXenes in 201118, over 100 kinds of MXenes have been theoretically predicted19. They shared a universal formula of Mn+1XnTx, where M is a transitional metal, X represents C/N, and T is the surface termination. Due to the diversity of compositions, MXenes have exhibited tunable and fascinating optical, chemical, and electronic properties, prompting the proposal of the concept of MXetronics20. The strong plasmonic feature peaks of different MXenes covering the visible to near-infrared (NIR) spectral range21, 22, resulting in the diversity of photoelectronic23, photothermal24, and photovoltaic applications25. In terms of their nonlinear optical response, Ti3C2Tx thin films were recently reported to have thickness-dependent saturable absorption26, which was attributed to the plasmon-induced increased ground-state absorption at high optical intensity. In addition, the nonlinear optical absorption response of MXenes has been demonstrated to be modulated by the excitation wavelength27 and surface group28, promoting the development of MXenes-based nonlinear optical devices. However, most nonlinear optical research about MXenes is currently focused on reporting the performance or intrinsic properties of a specific MXene. The fundamental understanding of MXenes photo-physics and the strategies for regulating the optical performance of MXenes is still very rudimentary29.

    The strategy for the construction of 2D heterostructure is an important tool for improving the photoelectronic performance of 2D materials-based devices30-32. The advantage of each component in 2D heterostructure can be preserved by careful design, and novel properties such as charge transfer or energy transfer between constituents of the 2D heterostructure may appear due to the interfacial effect. Study on the optical response of metal–semiconductor 2D heterostructures is essential for its photodetection and photocatalysis applications33. Recently, Nb2CTx MXene has been shown to have the highest optical extinction coefficient in the near-infrared region21. The few-layer MoS2 generally demonstrated remarkable optical properties in its excitonic resonance region (~1.8 eV)34. Moreover, 2D MXene/MoS2 heterostructure have shown the great potential for batteries35, 36 and supercapacitor37, which was attributed to its long-term stability.

    Inspired by these advantages, we demonstrated an enhanced linear and nonlinear optical performance of an Nb2C/MoS2 heterostructure by in situ growing MoS2 on the surface of Nb2C nanosheets. Interestingly, the species of surface group in Nb2C can modulate the work function of Nb2C/MoS2, which has been confirmed by X-ray photoelectron spectroscopy (XPS) measurements and density functional theory (DFT) calculation. After comparing the experimental and theoretical results, we determined that the surface group of Nb2C/MoS2 was dominated by O termination, leading to the decrease in the work function of Nb2C after the in situ growth of MoS2. The Nb2C/MoS2 inherited the preponderance at a different wavelength of Nb2C and MoS2 in absorption and exhibited enhanced broadband optical absorption, which was confirmed by measurement of UV-vis spectrum and DFT calculation. In addition, the saturable absorption of Nb2C/MoS2 and pristine Nb2C was investigated by the Z-scan technique. The nonlinear absorption coefficient and modulation depth of Nb2C/MoS2 are greater than those of Nb2C, but the saturated intensity comparison shows the opposite result. This demonstrates the superior nonlinear optical performance of Nb2C/MoS2 to that of Nb2C. The improved NLO performance can be attributed to the hole transfer from Nb2C to MoS2, which caused nonlinear optical response modulation in the heterostructure when combined with the determined energy level alignment. The current findings demonstrated that the Nb2C/MoS2 is a promising candidate for high-performance optoelectronic devices and provided an effective method for regulating the nonlinear optical response of MXenes.

    Methods

    Synthesis of few-layer Nb2C MXenes

    The 2 g Nb2AlC powder (11 Technology Co., Ltd., China) was added to a 40 mL of 49% HF solution (Macklin Inc.) and stirred for 60 h at room temperature. Then, the excrescent HF solution was washed with water by centrifugation until the pH was close to 6. Subsequently, the washed dispersion was added to the 5% 25 mL TMAOH solution (Macklin Inc.) with stirring for 12 h. The few-layer Nb2C nanosheets dispersed in water were obtained after washing the excrescent TMAOH.

    Preparation of the Nb2C/MoS2 heterostructure

    The ammonium thiomolybdate (13.3 mg) was added to as prepared Nb2C nanosheets dispersion (2 mL, 5 mg/mL) and stirred for 12 h until the MoS42− was fully inserted into the surface of the Nb2C nanosheets in dispersion. Then the precipitate was collected with quick freezing and freeze-drying. After annealing (2 h, 500 °C) in an H2 (10%)/Ar (90%) atmosphere, we obtained the heterostructure with MoS2 nanocrystals in situ grown on the surface of Nb2C nanosheets.

    Characterization

    The TEM, HRTEM, STEM, and element mapping scans were acquired using a Talos™F200X S/TEM (Thermo Fisher Scientific). A BRUKER D8 ADVANCE XRD system was employed for X-ray diffraction (XRD) phase characterizing. A Renishaw InVia Qontor confocal Raman microscope system was employed for collecting Raman spectra. The XPS measurement was performed with the help of a Thermo Scientific ESCALAB Xi+. The optical absorption spectra were characterized with a UV−vis spectrophotometer (UV-2600, Shimadzu).

    OA Z-scan system

    The nonlinear optical (NLO) absorption was characterized by an open aperture (OA) Z-scan system. A mode-locked Ti: sapphire system operating at 800 nm with 35 fs pulses at a 2 kHz repetition rate and a TOPAS (Light-Conversion) optical parametric amplifier was employed as a laser source. The Z-scan system measures the transmittance of the sample as a function of optical intensity, where the focal length of lens is 175 mm and beam waist is determined to be 24 μm. The signal was collected with the average power of the optical detectors (OPHIR, PD300 IR).

    DFT calculations

    The first-principles calculations of Nb2C/MoS2 heterostructure were performed by density functional theory as implemented in the Vienna ab initio Simulation Package (VASP)38 with the projector augmented wave (PAW) pseudopotentials. The generalized gradient approximation (GGA) of Perdew-Burke-Ernzerhof (PBE) functional was used. The plane-wave energy cutoff was set to 500 eV with a precision force of 0.01 eV/Å (1Å=10−10 m). The Brillouin zone was sampled with a 9×9×1 k-mesh point setting, and a denser k-mesh 15×15×1 was used for optical properties computations. The vdW interaction was treated with the semi-empirical D2 method (Grimme method). The absorbance A(ω) was determined by the 2D optical conductivity for low-dimensional systems39, 40. For the calculations of the Nb atom, the Hubbard U method was employed to treat strong correlation effects, and the U value of the Nb d orbital was set to 6.5 eV41.

    Results and discussion

    At first, the few-layer Nb2C nanosheets are synthesized with selective HF etching and intercalation methods27. The synthesis route of Nb2C/MoS2 heterostructure is shown in Fig. 1(a). Briefly, the MoS42− was inserted into the surface of Nb2C nanosheets during dispersion. MoS2 nanocrystals were grown in situ on the surface of Nb2C nanosheets after freeze-drying and annealing in an H2 atmosphere. Fig. 1(b) and Fig. S1(a) illustrate the transmission electron microscopy (TEM) image of Nb2C/MoS2, indicating a uniform, transparent flake structure of Nb2C/MoS2. The elemental mapping images were shown in Fig. 1(c). The Nb, Mo, and S elements overlap very well, indicating that the MoS2 and Nb2C are integrated very well in a single flake. To further characterize the morphology of MoS2 on the Nb2C sheets, high-resolution transmission electron microscopy (HRTEM) was performed (Fig. 1(d) and Fig. S1(b)). The surface of the Nb2C sheet has a clear lattice fringe. After careful measurement, the interplanar crystal spacing was determined to be 0.62 nm, which agrees well with the d-spacing of the (002) plane in MoS2. The selected area electron diffraction (SAED) pattern of Nb2C/MoS2 (Fig. 1(e)) matched well with (002) and (103) planes for Nb2C, (100) and (110) planes for MoS2, indicating a good overlap of the Nb2C and MoS2.

    Characterization of the Nb2C/MoS2 heterostructure. (a) Schematic illustration of the in situ synthesis of the Nb2C/MoS2. (b) Transmission electron microscope (TEM) image of Nb2C/MoS2. (c) Elemental mapping images of Nb2C/MoS2, upper left plane: Nb (yellow), upper right plane: Mo (cyan), left lower plane: S (blue), and right lower plane Merge. (d) High-resolution transmission electron microscopy (HRTEM) of Nb2C/MoS2. (e) Selected area electron diffraction (SAED) pattern of Nb2C/MoS2. (f) Raman spectra of Nb2C/MoS2. (g) XRD pattern of Nb2C/MoS2, Nb2C, and Nb2AlC.

    Figure 1.Characterization of the Nb2C/MoS2 heterostructure. (a) Schematic illustration of the in situ synthesis of the Nb2C/MoS2. (b) Transmission electron microscope (TEM) image of Nb2C/MoS2. (c) Elemental mapping images of Nb2C/MoS2, upper left plane: Nb (yellow), upper right plane: Mo (cyan), left lower plane: S (blue), and right lower plane Merge. (d) High-resolution transmission electron microscopy (HRTEM) of Nb2C/MoS2. (e) Selected area electron diffraction (SAED) pattern of Nb2C/MoS2. (f) Raman spectra of Nb2C/MoS2. (g) XRD pattern of Nb2C/MoS2, Nb2C, and Nb2AlC.

    The Raman spectroscopy analysis of Nb2C/MoS2 heterostructure is illustrated in Fig. 1(f). The prominent peaks appeared at 218 cm−1 and 255 cm−1 are attributed to the vibration mode of E22g and A1g of Nb2C42, respectively, and the peaks located at 378 cm−1 and 402 cm−1 were assigned to the vibration mode of E12g and A1g in MoS243, indicating the successful combination of Nb2C and MoS2. X-ray diffraction (XRD) was performed to further characterize the structure of Nb2C/MoS2. When compared to the XRD pattern of Nb2AlC (JCDS PDF#30-0033), the (002) peak of Nb2C downshifts and broadens to 2θ=7.1, demonstrating a successful exfoliation of 2D Nb2C27. For Nb2C/MoS2 heterostructure, the peaks located at 14.2°, 33.4°, and 58.7° were attributed to the (002), (101), and (110) planes of MoS2 (JCDS PDF#37-1492), and the signal at 8.7° can be ascribed to the (002) plane of Nb2C. Because the (002) peak of MXene represents the interlayer spacing, the downshifting of the (002) peak in Nb2C indicates decreased layer spacing after combing with MoS2.

    Sampleλ (nm)α0 (cm−1)αNL (cm/GW)Modulation depth α2Is (GW/cm2)
    Nb2C/MoS213003.68−0.180.3961.3
    15507.23−0.090.45107.8
    Nb2C13003.68−0.080.37136.9
    15507.23−0.050.43203.2

    Table 1. Linear and nonlinear optical parameters of Nb2C and Nb2C/MoS2.

    To analyze the bonding and chemical composition, X-ray photoelectron spectroscopy (XPS) of Nb2C and Nb2C/MoS2 was performed. In the XPS survey pattern spectrum (Fig. S1), the peaks of Nb2C matched well with Nb, C, O, and F elements, which were presented in the XPS spectrum of Nb2C/MoS2. Furthermore, the additional peaks can be found in the Nb2C/MoS2 spectrum, which were identified as the feature of Mo and S elements. The C 1s peaks of Nb2C/MoS2 (Fig. 2(a) upper plane) consist of three components. The peaks at 284.8 eV, 286.2 eV, and 288.9 eV were designated to sp3 C-C bond44, C-O bond45, and O-C=O bond45, respectively. As compared to the results shown in Fig. 2(a) (lower plane), the C bonding properties are nearly unchanged after combing with MoS2 because the position and intensity of the three peaks do not change. In the Nb 3d spectrum (Fig. 3(b) upper plane), the peaks at 207.5 eV and 210.2 eV are related to the 3d5/2 and 3d3/2 orbits of Nb atom46, respectively, which could be assigned to the Nb2O5 component47, 48. In comparison, the peaks of the same components in pristine Nb2C are downshifted to 206.8 eV and 209.5 eV, and the ratio of the two components does not change (Fig. 2(b) lower plane). These results indicate that the chemical environment of Nb atoms has changed due to the growth of MoS2, but the valence state of Nb atoms has not changed. The upper plane of Fig. 2(c) shows the O 1s high-resolution spectrum, which revealed three peaks in Nb2C/MoS2: one resulted from Nb-O contaminations at 531.3 eV49, one from C-O contaminations at 532.5 eV44, and one from OH surface termination groups at 533.7 eV44. These three components shifted to 530.0 eV, 531.4 eV, and 532.5 eV for pristine Nb2C. It is worth noting that the ratio of the C-O component in Nb2C/MoS2 is significantly enhanced as compared to the pristine Nb2C, demonstrating an increase of oxygen content at the surface of Nb2C after the in situ growth of MoS2 (Fig. 2(c) lower plane).

    XPS spectra of the Nb2C/MoS2 and Nb2C: (a) C 1s, (b) Nb 3d, (c) O 1s, (d) S 2p, (e) F 1s, and (f) valence band.

    Figure 2.XPS spectra of the Nb2C/MoS2 and Nb2C: (a) C 1s, (b) Nb 3d, (c) O 1s, (d) S 2p, (e) F 1s, and (f) valence band.

    The calculated work functions of (a) Nb2CF2/MoS2, (b) Nb2CO2/MoS2, (c) Nb2C(OH)2/MoS2. The charge-transfer and band alinement diagram of (d) Nb2CO2/MoS2 and (e) Nb2C(OH)2/MoS2.

    Figure 3.The calculated work functions of (a) Nb2CF2/MoS2, (b) Nb2CO2/MoS2, (c) Nb2C(OH)2/MoS2. The charge-transfer and band alinement diagram of (d) Nb2CO2/MoS2 and (e) Nb2C(OH)2/MoS2.

    The Mo 3d peaks shown in Fig. S2 consist of four components. The main peaks located at 229.0 eV (3d5/2) and 232.1 eV (3d3/2) corresponds to the Mo-S bonding50. For the S element (Fig. 2(d)), the peaks located at 226.3 eV, 162.4 eV, and 161.2 eV were assigned to S 2s, S 2p1/2, and S 2p3/2, respectively36. Moreover, the peaks located at 168.4 eV were ascribed to the Nb-S bonding46, indicating a strong combination of Nb2C and MoS2. Figure 2(e) shows the high-resolution XPS spectrum of F 1s, it can be seen that the -F surface termination was removed from the Nb2C/MoS2 heterostructure, which could be attributed to the F desorption during the heat treatment51.

    To determine the change of work function, the comparison of the XPS secondary electron cut-off (SEC) between Nb2C and Nb2C/MoS2 was shown in Fig. 2(f). The SEC correspond to the position where the detected electron has the highest binding energy, which is usually combined with the position of the Fermi edge to determine the escape work of the material: ϕm=hvEK,maxmeas, where ϕm is the work function of sample, hv and EK,maxmeas are the photon energy and the detected highest binding energy in the SEC. The SEC energy of Nb2C/MoS2 is larger than that of Nb2C. According to the basic principle of photoelectron spectroscopy52, 53, the work function of Nb2C/MoS2 is smaller than that of Nb2C. In other words, in situ MoS2 construction on the surface of Nb2C can be used to tailor the work function of Nb2C.

    Figure 3 shows crystal models of Nb2C with surface groups of -F, -O, and - OH27, monolayer MoS2, and Nb2C/MoS2 heterostructure used to investigate interface carrier-transfer and work function variations. The electrostatic potential of MoS2 and Nb2C monolayer with surface groups of -F, -O, and -OH, along the z direction (vertical plane), were shown in Fig. S3(a–d), respectively. The work function is defined as the energy required to extract an electron from the Fermi level to the surface of a solid54, which is equal to the energy difference between the vacuum level and Fermi level used in DFT calculation. Hence, the work function was determined to be 5.82 eV, 4.58 eV, 6.53 eV, and 2.33 eV for MoS2, Nb2CF2, Nb2CO2, Nb2C(OH)2, respectively. Interestingly, the work function of Nb2C MXene is very sensitive to its surface terminations: changing from O termination to OH termination causes the work function to decrease over 4 eV, which could be attributed to the variation of dipole moment density of MXene with different terminations55. The surface termination-dependent work function properties of MXene may provide a golden opportunity to modulate carrier behavior at the 2D heterostructure interface.

    Figure 3(a–c) illustrated the work function of the interface of Nb2CF2/MoS2, Nb2CO2/MoS2, and Nb2C(OH)2/MoS2, which give the work function of 4.56 eV, 6.43 eV, and 3.95 eV, respectively. The work function of heterostructures is different from its pristine compound due to charge redistribution and interface interaction. Moreover, the variation of surface terminations in Nb2C also changes the work function of the heterostructure. As schematically shown in Fig. 3(d), the electron could transfer from MoS2 to Nb2CO2 after the contact because the Fermi level of MoS2 is higher than that of Nb2CO2, resulting in a decrease in work function in Nb2CO2 after contact. In contrast, the electron could transfer from Nb2C(OH)2 to MoS2, resulting in the increase of work function in Nb2C(OH)2 after contact (Fig. 3(e)). When the XPS results in the valence band region are combined, we can conclude that the O termination dominated the surface group after the synthesis of heterostructures. We did not consider the Nb2CF2/MoS2 situation because the XPS experiment results indicated F termination desorption during the synthesis process.

    To analyze the effects of in situ growth of MoS2 and surface group changes on linear optical properties of Nb2C, the optical absorption properties of Nb2C and Nb2C/MoS2 were investigated with UV-visible spectrometry combined with DFT calculation. As shown in Fig. 4(a), the strong broadband optical extinction was observed in the near-infrared (NIR) region having a peak at about 850 nm, which was recognized as the optical feature of Nb2C MXenes21. Due to its good dispersibility in water, the absorbance of Nb2C dispersion is proportional to its concentration. The extinction coefficient was linearly fitted to be 16.13 L cm−1g−1 (Fig. S4) using the Beer-Lambert law. The DFT calculated optical absorption spectra with different terminations were shown in Fig. 4(b). Because the incident directions are random in dispersion during the UV-visible spectrometry, the calculated values of absorption are the average of three directions (x, y, and z). The difference between experimental and calculated results can be attributed to the termination and environment deviations. Even though the calculated results are not consistent with the experimental results, the average of three calculated spectrums (Nb2CF2, Nb2C(OH)2, and Nb2CO2) shows a similar tendency to the experimental results, which indicates that the experimentally synthesized Nb2C naturally terminated with F, O, and OH. The enhancement in the near-infrared region can be observed in the average curve, which is inherited from Nb2CF2, indicating that the broadband optical extinction in the near-infrared region of Nb2C in Fig. 4(a) was due to the F termination.

    (a) UV-vis spectrum of Nb2C solution with different concentrations. Inset: absorbance as a function of concentration. (b) DFT calculated absorption spectrum of Nb2C with different terminations. (c) UV-vis spectrum of Nb2C/MoS2 solution. (d) DFT calculated absorption spectrum of Nb2C/MoS2 with different terminations.

    Figure 4.(a) UV-vis spectrum of Nb2C solution with different concentrations. Inset: absorbance as a function of concentration. (b) DFT calculated absorption spectrum of Nb2C with different terminations. (c) UV-vis spectrum of Nb2C/MoS2 solution. (d) DFT calculated absorption spectrum of Nb2C/MoS2 with different terminations.

    The experimentally obtained absorption spectra of Nb2C, MoS2, Nb2C/MoS2 were shown in Fig. 4(c) for comparison. We normalized the absorption coefficients of three spectra at 300 nm in Fig. 4(c) to highlight the significance of the three groups. The absorption peak of MoS2 located in the range of 600–700 nm was attributed to the general features of MoS256. Furthermore, the absorption of MoS2 increases sharply as the wavelength decreases below 800 nm, whereas the absorption of Nb2C decreases. It appears that MoS2 replenishes the Nb2C absorption in the visible region, resulting in the increased absorption of Nb2C/MoS2 heterostructure in the visible region (450 nm – 800 nm). While the increased absorption of Nb2C/MoS2 in the NIR region (>800 nm) is inherited from Nb2C. In other words, Nb2C/MoS2 inherited the preponderance of Nb2C and MoS2 in absorption, exhibiting broadband enhanced optical absorption. The DFT calculated spectra of MoS2, Nb2CO2/MoS2, and Nb2C(OH)2/MoS2 shown in Fig. 4(d) are used to explore the mechanism of their outstanding absorption. The calculated MoS2 spectra exhibited a similar absorption trend to the experimental results, but the peaks in the calculation spectrum show a blue shift when compared to the experimental spectrum. The blue shift could be ascribed to an underestimation of the band gap in the generalized gradient approximation (GGA) level during DFT calculation. Due to the F desorption during the heat treatment, we averaged the calculated spectra of Nb2CO2/MoS2 and Nb2C(OH)2/MoS2. The averaged line is more consistent with the measured Nb2C/MoS2 spectrum than the pristine Nb2CO2/MoS2 and Nb2C(OH)2/MoS2 absorption curves, manifesting that the O and OH groups are interacting with the surface of Nb2C.

    In order to further study the nonlinear optical properties of Nb2C/MoS2, the OA Z-scan technique was utilized to determine the NLO response of Nb2C/MoS2 and Nb2C in the near-infrared region. The details of the measurement setup can be found in the experimental section. The OA Z-scan results of Nb2C/MoS2 and Nb2C with the excitation wavelength of 1300 nm and 1550 nm are shown in Fig. 5(a) and 5(b), respectively. Both Nb2C/MoS2 and Nb2C exhibit typical saturable absorption (SA), that is, the normalized transmittance increases as the sample approaches z=z0 point. For the control of the NLO experiment, the linear transmittance of Nb2C/MoS2 and Nb2C were set to be approximate consistency. Given this premise, the Nb2C/MoS2 exhibited a stronger SA response than that of the Nb2C at the wavelength of 1300 nm and 1550 nm.

    OA Z-scan results of Nb2C/MoS2 and Nb2C with an excitation wavelength of (a) 1300 nm and (b) 1550 nm. The corresponding nonlinear transmittance curves under excitation optical intensity are shown in (c) and (d). (e) Histogram of Is and αNL. (f) Schematic diagram of the transfer process of photogenerated carriers and enhanced nonlinear absorption process in Nb2C/MoS2.

    Figure 5.OA Z-scan results of Nb2C/MoS2 and Nb2C with an excitation wavelength of (a) 1300 nm and (b) 1550 nm. The corresponding nonlinear transmittance curves under excitation optical intensity are shown in (c) and (d). (e) Histogram of Is and αNL. (f) Schematic diagram of the transfer process of photogenerated carriers and enhanced nonlinear absorption process in Nb2C/MoS2.

    The intrinsic NLO absorption coefficient can be calculated by fitting the OA Z-scan results to the NLO theory. The NLO propagation equation can be written as56: dIdz"=(α0+αNLI)I, where z' is the transmission distance in the sample, α0 and αNL are the linear part and nonlinear part of the absorption coefficient, respectively. Giving the optical intensity Iz=I0/(1+z2/z02), where z is the position along the laser propagation direction, z0 is the Rayleigh length and I0 is the peak intensity on the axis, the NLO propagation can be solved as:

    T=1πq0ln[1+q0exp(x2)]dx.

    To further assess the saturable absorption properties, a simplified saturable absorption model with a two-level system can be expressed as57:

    T=exp[(α1+α21+I/Is]),

    where T is transmission, α1 is the non-saturable loss component, α2 is modulation depth, and Is is saturated intensity. Here, the Z-scan curves were converted into a function of transmission with optical intensity for fitting using the Eq. (2). The stand OA Z-scan model with Eq. (1) works very well for the curves in Fig. 5(a) and 5(b). The two-level system model with Eq. (2) also fits the converted traces well in Fig. 5(c) and 5(d). All the NLO fitting parameters are listed in Table 1.

    As shown in Fig. 5(e), the NLO absorption coefficient αNL of Nb2C/MoS2 was determined to be −0.18 cm/GW, and −0.09 cm/GW at 1300, and 1550 nm. For Nb2C, these values are −0.08 cm/GW, and −0.05 cm/GW at 1300, and 1550 nm. As a comparison, αNL of different 2D materials are shown in Fig. S6, indicating a comparable NLO performance of Nb2C/MoS2 with classical 2D materials. In addition, the saturated intensity Is of Nb2C/MoS2 was fitted as 61.3 GW/cm2, and 107.8 GW/cm2 at 1300, and 1550 nm, respectively. While Is of Nb2C was estimated to be 136.9 GW/cm2, and 203.2 GW/cm2 at 1300, and 1550 nm, respectively. As expected, the αNL of Nb2C/MoS2 is larger than that of Nb2C at both 1300 and 1550 nm wavelengths, that is, Nb2C/MoS2 possesses a greater ability to alter absorption coefficient than the Nb2C at the given optical intensity. In comparison, the fitted Is shows opposite trend with αNL: Nb2C/MoS2 had a lower saturated intensity than Nb2C, and a shorter wavelength corresponding to a lower saturated intensity. Because Is represents half of the optical intensity required for a complete bleaching of materials58, implying that the Nb2C/MoS2 is more easily saturated than Nb2C.

    Key insight into the enhanced saturable absorption process of Nb2C/MoS2 was schematically shown in Fig. 5(f). The saturable absorption in the NIR region of Nb2C/MoS2 and Nb2C can be attributed to the Pauli blocking induced by the intense single-photon transition. That is, the excited electrons gradually occupied possible states in the conduction band until all available states are full. After further increasing the incident light intensity, optical bleaching occurs according to Pauli blocking theory. Hence, the intensity of saturable absorption is naturally linked with the number of excited states58, 59. In addition, the single photon transition mainly originated from Nb2C instead of MoS2, because the excitation photon energy (< 1 eV) is smaller than the band gap of MoS2 (~1.8 eV)34. Moreover, the previous characterization, and calculation indicated that the surface group of Nb2C/MoS2 was dominated by O elements, implying that the Fermi level of Nb2C/MoS2 is close to the valence band of MoS2 as presented in Fig. 3(d). Because the electron and hole can spontaneously transfer to energies closer to the Fermi level, the photon-excited holes in Nb2C could transfer to MoS2, whereas this process is prohibited for the electrons. On the one hand, saturable absorption is the process by which states gradually exhaust. On the other hand, MoS2 offers the additional occupying state for carriers through the transfer process. Hence, Nb2C/MoS2 can offer more states to be exhausted, leading to a stronger and easier saturable absorption than the pristine Nb2C.

    Conclusions

    In summary, we have observed the improved linear and nonlinear optical performance of an Nb2C/MoS2 heterostructure designed in this study. The XPS characterizing and work function calculation revealed that the dominated surface group of Nb2C/MoS2 was O termination, leading to the decrease in the work function of Nb2C after the in situ growth of MoS2. Experiment and theoretical calculations revealed that the Nb2C/MoS2 heterostructure has broadband-enhanced optical absorption. Furthermore, the OA Z-scan results showed that Nb2C/MoS2 has better NLO performance than Nb2C. The enhanced NLO performance is attributed to the hole transfer from Nb2C to MoS2. The 2D Nb2C/MoS2 heterostructure was proved to possess excellent nonlinear optical response, indicating that Nb2C/MoS2 can be applicated in the field of optoelectronics and ultrafast photonics, such as saturable absorbers, optical switches, and optical diodes. This facile strategy of in situ construction of the 2D Nb2C/MoS2 heterostructure provides guidance for achieving robust light modulation and paves the way for the development of broadband optoelectronic devices and optical modulators.

    References

    [1] KS Novoselov, AK Geim, SV Morozov, D Jiang, Y Zhang et al. Electric field effect in atomically thin carbon films. Science, 666-669(2004).

    [2] C Zeng, H Lu, D Mao, YQ Du, H Hua et al. Graphene-empowered dynamic metasurfaces and metadevices. Opto-Electron Adv, 200098(2022).

    [3] ZP Sun, A Martinez, F Wang. Optical modulators with 2D layered materials. Nat Photonics, 227-238(2016).

    [4] F Bonaccorso, Z Sun, T Hasan, AC Ferrari. Graphene photonics and optoelectronics. Nat Photonics, 611-622(2010).

    [5] A Chernikov, TC Berkelbach, HM Hill, A Rigosi, YL Li et al. Exciton binding energy and nonhydrogenic Rydberg series in monolayer WS2. Phys Rev Lett, 076802(2014).

    [6] YW Wang, ZL Deng, DJ Hu, J Yuan, QD Ou et al. Atomically thin noble metal dichalcogenides for phase-regulated meta-optics. Nano Lett, 7811-7818(2020).

    [7] A Elbanna, K Chaykun, Y Lekina, YD Liu, B Febriansyah et al. Perovskite-transition metal dichalcogenides heterostructures: recent advances and future perspectives. Opto-Electron Sci, 220006(2022).

    [8] GW Zhang, SY Huang, FJ Wang, QX Xing, CY Song et al. The optical conductivity of few-layer black phosphorus by infrared spectroscopy. Nat Commun, 1847(2020).

    [9] ZT Wang, YH Xu, SC Dhanabalan, J Sophia, CJ Zhao et al. Black phosphorus quantum dots as an efficient saturable absorber for bound soliton operation in an erbium doped fiber laser. IEEE Photonics J, 1503310(2016).

    [10] LM Wu, WC Huang, YZ Wang, JL Zhao, DT Ma et al. 2D tellurium based high-performance all-optical nonlinear photonic devices. Adv Funct Mater, 1806346(2019).

    [11] LM Wu, YZ Dong, JL Zhao, DT Ma, WC Huang et al. Kerr nonlinearity in 2D graphdiyne for passive photonic diodes. Adv Mater, 1807981(2019).

    [12] ZJ Xie, F Zhang, ZM Liang, TJ Fan, ZJ Li et al. Revealing of the ultrafast third-order nonlinear optical response and enabled photonic application in two-dimensional tin sulfide. Photonics Res, 494-502(2019).

    [13] J Ren, H Lin, XR Zheng, WW Lei, D Liu et al. Giant and light modifiable third-order optical nonlinearity in a free-standing h-BN film. Opto-Electron Sci, 210013(2022).

    [14] YW Wang, L Zhou, MZ Zhong, YP Liu, S Xiao et al. Two-dimensional noble transition-metal dichalcogenides for nanophotonics and optoelectronics: Status and prospects. Nano Res, 3675-3694(2022).

    [15] YW Wang, S Liu, BW Zeng, H Huang, J Xiao et al. Ultraviolet saturable absorption and ultrafast carrier dynamics in ultrasmall black phosphorus quantum dots. Nanoscale, 4683-4690(2017).

    [16] H Wei, YD Wang, YW Wang, WX Fan, L Zhou et al. Giant two-photon absorption in MXene quantum dots. Opt Express, 8482-8493(2022).

    [17] LM Wu, TJ Fan, SR Wei, YJ Xu, Y Zhang et al. All-optical logic devices based on black arsenic–phosphorus with strong nonlinear optical response and high stability. Opto-Electron Adv, 200046(2022).

    [18] M Naguib, M Kurtoglu, V Presser, J Lu, JJ Niu et al. Two-dimensional nanocrystals produced by exfoliation of Ti3AlC2. Adv Mater, 4248-4253(2011).

    [19] NC Frey, J Wang, GIV Bellido, B Anasori, Y Gogotsi et al. Prediction of synthesis of 2D metal carbides and nitrides (MXenes) and their precursors with positive and unlabeled machine learning. ACS Nano, 3031-3041(2019).

    [20] A VahidMohammadi, J Rosen, Y Gogotsi. The world of two-dimensional carbides and nitrides (MXenes). Science, eabf1581(2021).

    [21] K Maleski, CE Shuck, AT Fafarman, Y Gogotsi. The broad chromatic range of two-dimensional transition metal carbides. Adv Opt Mater, 2001563(2021).

    [22] MK Han, K Maleski, CE Shuck, YZ Yang, JT Glazar et al. Tailoring electronic and optical properties of MXenes through forming solid solutions. J Am Chem Soc, 19110-19118(2020).

    [23] AS Sharbirin, S Akhtar, J Kim. Light-emitting MXene quantum dots. Opto-Electron Adv, 200077(2021).

    [24] H Lin, YW Wang, SS Gao, Y Chen, JL Shi. Theranostic 2D tantalum carbide (MXene). Adv Mater, 1703284(2018).

    [25] A Agresti, A Pazniak, S Pescetelli, Vito A Di, D Rossi et al. Titanium-carbide MXenes for work function and interface engineering in perovskite solar cells. Nat Mater, 1228-1234(2019).

    [26] YC Dong, S Chertopalov, K Maleski, B Anasori, LY Hu et al. Saturable absorption in 2D Ti3C2 MXene thin films for passive photonic diodes. Adv Mater, 1705714(2018).

    [27] YD Wang, YW Wang, KQ Chen, K Qi, TY Xue et al. Niobium carbide MXenes with broad-band nonlinear optical response and ultrafast carrier dynamics. ACS Nano, 10492-10502(2020).

    [28] H Li, SY Chen, DW Boukhvalov, ZY Yu, MG Humphrey et al. Switching the nonlinear optical absorption of titanium carbide MXene by modulation of the surface terminations. ACS Nano, 394-404(2022).

    [29] Y Wang, Y Wang, J He. 2D Transition Metal Carbides (MXenes) for Third Order Nonlinear Optics: Status and Prospects. Laser Photonics Rev, 2200733(2023). https://doi.org/10.1002/lpor.202200733

    [30] Y Liu, NO Weiss, XD Duan, HC Cheng, Y Huang et al. Van der Waals heterostructures and devices. Nat Rev Mater, 16042(2016).

    [31] ZW Li, W Yang, M Huang, X Yang, CG Zhu et al. Light-triggered interfacial charge transfer and enhanced photodetection in CdSe/ZnS quantum dots/MoS2 mixed-dimensional phototransistors. Opto-Electron Adv, 210017(2021).

    [32] JY Chen, T Eul, L Lyu, YL Li, XY Hu et al. Tracing the formation of oxygen vacancies at the conductive LaAlO3/SrTiO3 interface via photoemission. Opto-Electron Sci, 210011(2022).

    [33] G Tagliabue, JS DuChene, M Abdellah, A Habib, DJ Gosztola et al. Ultrafast hot-hole injection modifies hot-electron dynamics in Au/p-GaN heterostructures. Nat Mater, 1312-1318(2020).

    [34] HJ Conley, B Wang, JI Ziegler, RF Jr Haglund, ST Pantelides et al. Bandgap engineering of strained monolayer and bilayer MoS2. Nano Lett, 3626-3630(2013).

    [35] K Ma, H Jiang, Y Hu, C Li. 2D nanospace confined synthesis of pseudocapacitance-dominated MoS2-in-Ti3C2 superstructure for ultrafast and Stable Li/Na-ion batteries. Adv Funct Mater, 1804306(2018).

    [36] ZY Yuan, LL Wang, DD Li, JM Cao, W Han. Carbon-reinforced Nb2CTx MXene/MoS2 nanosheets as a superior rate and high-capacity anode for sodium-ion batteries. ACS Nano, 7439-7450(2021).

    [37] X Wang, H Li, H Li, S Lin, W Ding et al. 2D/2D 1T-MoS2/Ti3C2 MXene heterostructure with excellent supercapacitor performance. Adv Funct Mater, 0190302(2020).

    [38] W Kohn, LJ Sham. Self-consistent equations including exchange and correlation effects. Phys Rev, A1133-A1138(1965).

    [39] V Wang, N Xu, JC Liu, G Tang, WT Geng. VASPKIT: a user-friendly interface facilitating high-throughput computing and analysis using VASP code. Comput Phys Commun, 108033(2021).

    [40] YD Wang, YW Wang, YL Dong, L Zhou, H Wei et al. The nonlinear optical transition bleaching in tellurene. Nanoscale, 15882-15890(2021).

    [41] Y Luo, C Cheng, HJ Chen, K Liu, XL Zhou. Systematic investigations of the electron, phonon and elastic properties of monolayer M2C (M = V, Nb, Ta) by first-principles calculations. J Phys Condens Matter, 405703(2019).

    [42] H Lin, SS Gao, C Dai, Y Chen, JL Shi. A two-dimensional biodegradable niobium carbide (MXene) for photothermal tumor eradication in NIR-I and NIR-II biowindows. J Am Chem Soc, 16235-16247(2017).

    [43] XM Geng, YL Zhang, Y Han, JX Li, L Yang et al. Two-dimensional water-coupled metallic MoS2 with nanochannels for ultrafast supercapacitors. Nano Lett, 1825-1832(2017).

    [44] CJ Shen, LB Wang, AG Zhou, B Wang, XL Wang et al. Synthesis and electrochemical properties of two-dimensional RGO/Ti3C2Tx nanocomposites. Nanomaterials, 80(2018).

    [45] T Schultz, NC Frey, K Hantanasirisakul, S Park, SJ May et al. Surface termination dependent work function and electronic properties of Ti3C2Tx MXene. Chem Mater, 6590-6597(2019).

    [46] H Zong, L Hu, SJ Gong, K Yu, ZQ Zhu. Flower-petal-like Nb2C MXene combined with MoS2 as bifunctional catalysts towards enhanced lithium-sulfur batteries and hydrogen evolution. Electrochim Acta, 139781(2022).

    [47] J Halim, KM Cook, M Naguib, P Eklund, Y Gogotsi et al. X-ray photoelectron spectroscopy of select multi-layered transition metal carbides (MXenes). Appl Surf Sci, 406-417(2016).

    [48] A Darlinski, J Halbritter. Angle-resolved XPS studies of oxides at NbN, NbC, and Nb surfaces. Surf Interface Anal, 223-237(1987).

    [49] DD Sarma, CNR Rao. XPES studies of oxides of second- and third-row transition metals including rare earths. J Electron Spectrosc Relat Phenom, 25-45(1980).

    [50] HC Woo, IS Nam, JS Lee, JS Chung, KH Lee et al. Room-temperature oxidation of K2CO3MoS2 catalysts and its effects on alcohol synthesis from CO and H2. J Catal, 525-535(1992).

    [51] JK El-Demellawi, S Lopatin, J Yin, OF Mohammed, HN Alshareef. Tunable multipolar surface plasmons in 2D Ti3C2Tx MXene flakes. ACS Nano, 8485-8493(2018).

    [52] MG Helander, MT Greiner, ZB Wang, ZH Lu. Pitfalls in measuring work function using photoelectron spectroscopy. Appl Surf Sci, 2602-2605(2010).

    [53] MG Helander, MT Greiner, ZB Wang, WM Tang, ZH Lu. Work function of fluorine doped tin oxide. J Vac Sci Technol A, 011019(2011).

    [54] A Kahn. Fermi level, work function and vacuum level. Mater Horiz, 7-10(2016).

    [55] YY Liu, H Xiao, III WA Goddard. Schottky-barrier-free contacts with two-dimensional semiconductors by surface-engineered MXenes. J Am Chem Soc, 15853-15856(2016).

    [56] KP Wang, J Wang, JT Fan, M Lotya, A O’Neill et al. Ultrafast saturable absorption of two-dimensional MoS2 nanosheets. ACS Nano, 9260-9267(2013).

    [57] XT Jiang, SX Liu, WY Liang, SJ Luo, ZL He et al. Broadband nonlinear photonics in few-layer MXene Ti3C2Tx (T = F, O, or OH). Laser Photon Rev, 1700229(2018).

    [58] QL Bao, H Zhang, Y Wang, ZH Ni, YL Yan et al. Atomic-layer graphene as a saturable absorber for ultrafast pulsed lasers. Adv Funct Mater, 3077-3083(2009).

    [59] M Breusing, S Kuehn, T Winzer, E Malić, F Milde et al. Ultrafast nonequilibrium carrier dynamics in a single graphene layer. Phys Rev B, 153410(2011).

    Yiduo Wang, Yingwei Wang, Yulan Dong, Li Zhou, Jianlong Kang, Ning Wang, Yejun Li, Xiaoming Yuan, Zhengwei Zhang, Han Huang, Mengqiu Long, Si Xiao, Jun He. 2D Nb2CTx MXene/MoS2 heterostructure construction for nonlinear optical absorption modulation[J]. Opto-Electronic Advances, 2023, 6(10): 220162
    Download Citation