• Photonics Research
  • Vol. 7, Issue 7, 762 (2019)
Yafeng Xie1、2, Saifeng Zhang1、6, Yuanxin Li1, Ningning Dong1, Xiaoyan Zhang1, Lei Wang1、2, Weimin Liu3, Ivan M. Kislyakov1, Jean-Michel Nunzi1、4, Hongji Qi1, Long Zhang1, and Jun Wang1、2、5、*
Author Affiliations
  • 1Laboratory of Micro-Nano Optoelectronic Materials and Devices and CAS Key Laboratory of Materials for High-Power Laser, Shanghai Institute of Optics and Fine Mechanics, Chinese Academy of Sciences, Shanghai 201800, China
  • 2Center of Materials Science and Optoelectronics Engineering, University of Chinese Academy of Sciences, Beijing 100049, China
  • 3School of Physical Science and Technology, ShanghaiTech University, Shanghai 201210, China
  • 4Department of Physics, Engineering Physics & Astronomy and Department of Chemistry, Queen’s University, Kingston, Ontario K7L-3N6, Canada
  • 5State Key Laboratory of High Field Laser Physics, Shanghai Institute of Optics and Fine Mechanics, Chinese Academy of Sciences, Shanghai 201800, China
  • 6e-mail: sfzhang@siom.ac.cn
  • show less
    DOI: 10.1364/PRJ.7.000762 Cite this Article Set citation alerts
    Yafeng Xie, Saifeng Zhang, Yuanxin Li, Ningning Dong, Xiaoyan Zhang, Lei Wang, Weimin Liu, Ivan M. Kislyakov, Jean-Michel Nunzi, Hongji Qi, Long Zhang, Jun Wang. Layer-modulated two-photon absorption in MoS2: probing the shift of the excitonic dark state and band-edge[J]. Photonics Research, 2019, 7(7): 762 Copy Citation Text show less

    Abstract

    Questions hovering over the modulation of bandgap size and excitonic effect on nonlinear absorption in two-dimensional transition metal dichalcogenides (TMDCs) have restricted their application in micro/nano optical modulator, optical switching, and beam shaping devices. Here, degenerate two-photon absorption (TPA) in the near-infrared region was studied experimentally in mechanically exfoliated MoS2 from single layer to multilayer. The layer-dependent TPA coefficients were significantly modulated by the detuning of the excitonic dark state (2p). The shift of the quasiparticle bandgap and the decreasing of exciton binding energy with layers were deduced, combined with the non-hydrogen model of excitons in TMDCs and the scaling rule of semiconductors. Our work clearly demonstrates the layer modulation of nonlinear absorption in TMDCs and provides support for layer-dependent nonlinear optical devices, such as optical limiters and optical switches.

    1. INTRODUCTION

    Layered transition metal dichalcogenides (TMDCs) exhibit unique nonlinear optical (NLO) properties, including high-order harmonic generation, ultrafast saturable absorption, optical limiting, multi-photon absorption, and so on [15]. Intrinsically, mono- and few-layered TMDCs possess large excitonic effects. Recently, the existence of excitonic dark states in monolayers (MLs) WS2 and MoS2 has been revealed via two-photon photoluminescence (TPL) spectroscopy, demonstrating their potential application in optoelectronic and quantum devices [69]. However, it should be noted that the TPL technique is applicable only to photoluminescent materials, whereas most TMDCs, especially multilayered ones, are with low-quantum-yield or nonluminous, hampering the investigation of layer modulation (essentially excitonic dark state modulation) on the NLO absorption of TMDCs. In this work, we probed the degenerate two-photon absorption (TPA) of two-dimensional (2D) MoS2 directly utilizing an NLO intensity scan setup (micro-I-scan) for micro/nano materials [10,11]. The TPA coefficients of MoS2 ranging from ML to 31 layers (31L) were obtained, which were deeply modulated by the detuning of the excitonic dark state. Using a non-hydrogen model in combination with the scaling rule of semiconductors [12], the shift of the quasiparticle bandgap and the excitonic dark state with van der Waals layers were deduced [1315]. In addition, the sub-band resonance in the conduction band was also observed in multilayer MoS2.

    2. MATERIALS AND CHARACTERIZATION

    Layered MoS2 nanosheets were mechanically exfoliated from the bulk crystal [16,17], and then transferred to a quartz substrate, which would affect the light absorption extremely weakly [18,19]. As shown in Fig. 1(a), the morphology of the samples can be readily identified using a 50× objective lens. The lateral size of the samples was determined to be 10 to 100 µm, which is suitable for our NLO measurements. Raman spectroscopy and photoluminescence (PL) spectroscopy were performed, shown in Fig. 1. Both the Raman and PL signals were collected using a confocal microscopy setup (Horiba Co., HR Evolution) with a green continuous-wave (CW) laser at 532 nm (1.3  kW/cm2 at focus). The exact thickness of all samples was measured using atomic force microscopy (AFM). Figure 1(b) shows the thickness of the mono- and bilayer MoS2. Additional AFM images are provided in the Appendices (Fig. 8).

    (a) Microscopic images of mechanically exfoliated monolayer (1L), bilayer (2L), and trilayer (3L) of MoS2 nanosheets. (b) AFM image of the monolayer and bilayer, where the boundary is marked by yellow dashed-dotted lines. The height profiles are along the direction of the green dashed line. (c)–(f) PL, Raman, and linear absorption spectra of MoS2 nanosheets with different layers, exhibiting the layer-modulated Raman shift, PL intensity, and one-photon absorption.

    Figure 1.(a) Microscopic images of mechanically exfoliated monolayer (1L), bilayer (2L), and trilayer (3L) of MoS2 nanosheets. (b) AFM image of the monolayer and bilayer, where the boundary is marked by yellow dashed-dotted lines. The height profiles are along the direction of the green dashed line. (c)–(f) PL, Raman, and linear absorption spectra of MoS2 nanosheets with different layers, exhibiting the layer-modulated Raman shift, PL intensity, and one-photon absorption.

    As shown in Fig. 1(f), two Raman-active modes, E2g1 and A1g, exhibit significant layer dependence. The Raman shifts of mono-, bi-, and trilayers are 18.99  cm1, 21.98  cm1, and 23.71  cm1, respectively, which are consistent with previous literature results [4,20]. In addition, the two Raman vibration modes provide strong evidence that all the samples used in our work are 2H-MoS2 [4,21]. The PL spectra of MoS2 were obtained under the same incident power. As shown in Fig. 1(c), there are two peaks at 670 nm and 615 nm corresponding to the A and B excitonic (1s) ground states, which come from energy splitting (150  meV) near the valence-band maximum due to spin-orbit coupling [2225]. The linear absorption spectra were achieved using a confocal microscope, which are plotted in Fig. 1(e). It should be noted that there is a slight shift of the two distinct excitonic absorption peaks with layer number, but the strong excitonic effect still exists.

    3. RESULTS AND DISCUSSION

    A. Two-Photon Absorption Properties

    The TPA process in layered MoS2 was investigated using a micro-I scan setup [10], which integrates a microscopic imaging system (Fig. 10). The femtosecond laser pulses (pulse width: 340 fs, repetition rate: 1 kHz) at 1030 nm (1.204 eV) were generated from a mode-locked fiber laser. The laser beam was finally focused down to a spot with a waist radius wp of 5  μm on the surface of the MoS2 (Fig. 7). The intensity of the incident pulses was adjusted continuously with a linearly adjustable attenuator. The process that accounts for one- and two-photon absorption can be described by the propagation equation [2,10,11,2629]: dI(z)dz=αIβ(I)I2(z).Here, z is the propagation distance of the laser beam in the sample. α is the one-photon (linear) absorption coefficient, which cannot contribute to the total absorption significantly owing to the smaller value of the photon energy of 1.2 eV than the optical bandgap [3032]. β(I) is the TPA coefficient. In our experiment, the excitation laser source was a series of Gaussian pulses, i.e., I(r,t)=I0·exp(2r2wp2)·exp(t2τp2),where wp and τp represent the radius of the pump beam waist and the pulse width, respectively.

    Compared with previous research on CdS nanocrystals [33], MoS2 nanosheets have a uniform surface morphology with few defects in such a single crystal. Therefore, the homogeneously broadened saturation model of TPA was taken into account as follows [34,35]: β(I)=β01+(I/Isat)2,where Isat is the saturation intensity of TPA. Based on this model, the numerical fitting curves are shown in Fig. 2(a), which agree well with the experimental results (Fig. 9). The TPA coefficients of the layered MoS2 were extracted, and the detailed parameters are given in Table 1 in the Appendices. As shown in Fig. 2(b), monolayer MoS2 exhibits the largest TPA coefficient, which declines rapidly until approximately four layers. After that, the TPA coefficient gradually increases with layer number. This anomalous tendency reveals the switching of the optical transition from the excitonic type to the interband one, which will be discussed in the following section.

    SampleThickness (nm)TPA Coefficient (β0)(103  cm/GW)Saturation Intensity (Isat)(102GW/cm2)
    ME-1L0.71.88±0.211.98±0.16
    ME-2L1.40.38±0.152.05±0.24
    ME-3L2.10.25±0.132.59±0.13
    ME-4L3.00.19±0.124.08±0.28
    ME-7L5.00.46±0.162.52±0.24
    ME-9L6.30.54±0.172.19±0.21
    ME-12L8.40.62±0.200.97±0.15
    ME-14L9.80.57±0.321.27±0.16
    ME-18L11.91.28±0.261.82±0.17
    ME-25L16.81.81±0.351.18±0.18
    ME-29L20.31.80±0.390.42±0.12
    ME-31L21.71.83±0.280.70±0.15

    Table 1. Complete Parameters of Thickness and NLO Coefficients

    (a) Micro-I-scan results for monolayer, few-layer, and multilayer MoS2, which are fitted using the homogeneously broadened TPA model. (b) TPA coefficients of layered MoS2. The green solid circle stands for the TPA coefficients, which can be separated into two parts: the trailing edge (as indicated by the blue solid line) corresponds to the excitonic resonance part, and the following rising edge corresponds to the interband transition part. The orange dashed line indicates the transition of MoS2 from the two-dimensional prototype to the bulk one.

    Figure 2.(a) Micro-I-scan results for monolayer, few-layer, and multilayer MoS2, which are fitted using the homogeneously broadened TPA model. (b) TPA coefficients of layered MoS2. The green solid circle stands for the TPA coefficients, which can be separated into two parts: the trailing edge (as indicated by the blue solid line) corresponds to the excitonic resonance part, and the following rising edge corresponds to the interband transition part. The orange dashed line indicates the transition of MoS2 from the two-dimensional prototype to the bulk one.

    Due to the 2D geometry of monolayer MoS2 and poor dielectric screening, the ground-state excitons are well confined with a binding energy of 0.63  eV [25]. This is much larger than that of traditional semiconductors and quantum wells (dozens of meV) [36,37]. As shown in Figs. 1(d) and 1(e), the PL and absorption peaks indicate that the A exciton ground state energy level (1s) of all layered MoS2 is almost the same, with a value of approximately 1.85–1.88 eV. However, the quasi-particle bandgap Eg is much larger than the optical one according to theoretical predictions [38]. For example, Eg is about 2.7 eV for the monolayer [25] and 2.427 eV for the bilayer [38]. Moreover, Qiu et al. have theoretically predicted that there are several excitonic dark states between 1.88 eV and 2.7 eV in monolayer MoS2. According to the transition selection rules, one-photon excitation couples states of opposite parity, as opposed to TPA, which couples states of identical parity [6]. One-photon can excite an electron only to the excitonic ground state (1s) or any other higher-order bright excitonic states. Considering the valley polarization or different valleys in MoS2, we note that there exists the valley-selective circular dichroism arising from its unique symmetry. However, in our experiment, we used a linear polarization excitation source rather than a circular one. Both valleys (K and K) will be populated. In the meanwhile, the valley polarization is obvious only at low temperature. For example, Cao et al. measured the circularly polarized PL of monolayer MoS2 (P50%) at 83 K [39], Zeng et al. gave P32% at 10 K [40], and Mak et al. confirmed that at 14 K [41]. We did the TPA measurements at room temperature (about 300 K), at which the phonon-assisted inter-valley scattering becomes important and the valley polarization effect becomes very weak and can be neglected [42]. Hence, the TPA technique is the most effective method to study 2p excitonic dark state transition in mono- and few-layer MoS2. In multilayer MoS2, the red shift of quasiparticle bandgap continues to increase and so do the high excited excitonic states, which leads to a large detuning between 2p excitonic dark state and two-photon excitation [38,43]. Thus, the TPA coefficient can be modulated by the on–off resonance. In addition, it has been calculated that the transition matrix elements for direct transitions would vary with the number of layers. However, the matrix element at the K point is always the strongest position [44,45].

    B. Theoretical Analysis

    In 2D, the hydrogen model employs an effective mass Hamiltonian H=2r2/2μ+Veh(r), where Veh(r)=e2/εr is locally screened attractive electron–hole interaction. Depending on this model, the excitonic energy level in TMDCs can be calculated using Eq. (4) [46]: Eb(n)=μe422εn2(n12)2=EgEn,where Eb(n) is the binding energy of the n exciton state, Eg is the quasiparticle bandgap, En is the nth-order excitonic energy level, and n is the excitonic principal quantum number. However, the description of 2D excitons is complicated by the fact that the screening cannot be assumed to be local. Therefore, Chernikov et al. considered a nonlocally screened electron–hole interaction described by the potential Veh(r)=πe22r0[H0(rr0)Y0(rr0)], where H0 and Y0 are Struve and Bessel functions, and numerically calculated the radially symmetric excitonic Hamiltonian. From another point of view, the dielectric screening parameter is no longer a constant, which should be an n-dependent value. The analytical expression for the nonlocal 2D screening depends on the excitonic mass and 2D polarizability [15]: εn=12[1+1+32παμ9n(n1)+3].Here, μ=1/(me1+mh1) is the exciton reduced mass, α is the 2D polarizability, the internal polarizability of the 2D dielectric [15,47], and εn is the n-dependent effective screening dielectric parameter.

    Combining Eqs. (4) and (5), we can obtain the n=2 energy levels of the A-excitonic state (2s) of monolayer and bilayer, which are 2.412 eV and 2.175 eV, respectively. The detail of the calculation can be found in the Appendices. It is worth noting that the 2p excitonic dark state energy level is slightly lower than that of 2s, which is another different feature compared to the regular hydrogen model of exciton [4850]. At a given excitonic principal quantum number n, excitonic states with the higher angular momentum have a larger binding energy, which has been observed in the case of the 2H-WS2 monolayer [6,51]. Based on the theoretical estimation, the 2p dark state of A-exciton in a monolayer (slightly lower than 2.412 eV) is almost in resonance with our two-photon energy (2.408 eV) [52]. Thus, we attribute the giant TPA coefficient in the monolayer MoS2 at 1030 nm to this effect. When the number of layers is increased, the detuning between the 2p dark state of A-exciton and the two-photon energy becomes larger, as illustrated in Fig. 3. This leads to the sharp decline of the TPA coefficient. The decline trend can be fitted using a typical TPA resonance model, in the form of β(2ω)=2ωI2W(2ω)=2ωI2W0δ(2ωE2p),where W(2ω) is the two-photon transition rate, W0 is a photon energy-independent constant, and δ(2ωE2p) is an ideal delta function that denotes the resonance with the 2p-excitonic state. However, the excitonic states in monolayer and few-layer MoS2 are not ideally discrete. They are practically broadened with a Lorentzian profile [46,53]. The theoretical fitting of the TPA coefficient is illustrated as the blue solid line in Fig. 2(b), which is consistent with the Lorentzian function. It can be seen that the fitting curve agrees well with the experimental results from monolayer to four layers. Therefore, the sharp decline of the TPA coefficient from monolayer to few-layer can be ascribed to the detuning of 2p excitonic dark state resonance. The contribution of excitonic dark states with larger n is negligible. Such states have much smaller transition matrix elements, and they will be much harder to be populated, which means that the 3p-excitonic state is more difficult to excite than the 2p.

    Schematic illustration of a TPA-active exciton at the K point in the Brillouin zone. 1s and 2p (∼1.88 eV and ∼2.41 eV in monolayer MoS2) represent ground state and the first excitonic dark state in a 2D non-hydrogen model. (a) In monolayer MoS2, the 2p excitonic dark state is almost in resonance with two-photon energy. (b) In few-layer MoS2, the excitonic resonance detuning increases with layer number. (c) In multilayer MoS2, the red shift of quasiparticle bandgap continues to increase, and the interband TPA occurs.

    Figure 3.Schematic illustration of a TPA-active exciton at the K point in the Brillouin zone. 1s and 2p (1.88  eV and 2.41  eV in monolayer MoS2) represent ground state and the first excitonic dark state in a 2D non-hydrogen model. (a) In monolayer MoS2, the 2p excitonic dark state is almost in resonance with two-photon energy. (b) In few-layer MoS2, the excitonic resonance detuning increases with layer number. (c) In multilayer MoS2, the red shift of quasiparticle bandgap continues to increase, and the interband TPA occurs.

    As the thickness of MoS2 is increased from a single layer to multilayers, the interaction between the van der Waals layers gets stronger, and both the quasiparticle bandgap and the binding energy of the excitons decrease [38]. Consequently, when the quasiparticle bandgap is less than 2.4 eV, TPA can occur from the valence band to conduction band. In this case, the quasiparticle bandgap can be estimated by the scaling rule applicable to most semiconductors, including multilayer TMDCs [5457]: β0=KEpF(2ω/Eg)n02Eg3,where K is a material-independent constant, n0 is the linear refractive index of the material, and Ep (related to the Kane momentum parameter) is nearly material independent for a wide variety of semiconductors. By taking parabolic bands near the K point in the first Brillouin zone into consideration, the function F is given by [54] F(2ω/Eg)=(2ω/Eg1)3/2(2ω/Eg)5.Ep is about 21 eV, the same order of magnitude as previous results [54,55,58]. K is about 106, larger than that of traditional semiconductors [56], implying strong exciton effect in TMDCs. It can be seen in Fig. 2(b) that β0 gradually reaches a constant value when the layer number increases to about 25. It indicates that layered MoS2 with such thickness has a similar TPA performance as the bulk, i.e., they are substantially bulk MoS2.

    Combining Eqs. (7) and (8), the relationship between Eg and layer number was obtained for the multilayer MoS2, as shown in Fig. 4. Then the exciton ground state En=1 and first excited state En=2 can be determined using Eq. (5). The corresponding excitonic dark state E2p in a multilayer can be approximately marked by E2 (i.e. E2s), neglecting the slight difference (smaller than 0.05 eV) [49]. The shift of the conduction band-edge (corresponding to quasiparticle bandgap Eg) and the 2p dark state of A exciton (adjacent to E2s) is clearly demonstrated in Fig. 4. The energy distance between E2 and E1 becomes gradually smaller, meaning that the high excitonic states are compressed and the binding energy weakens when layer number increases. The full set of the parameters, εn=1, εn=2, Eg, and En=2, from monolayer to multilayer are summarized in Table 2. The increasing tendency of the dielectric parameter with layer number is obvious, which implies an increase of the dielectric screening potential. The quasiparticle bandgap Eg and E2s drop to almost a constant when the layer number increases to 25, which is evidence of the transition from 2D to bulk.

    Layersεn=1εn=2Eg(eV)En=1(eV)En=2(eV)Ebn=1Ebn=2
    12.8a1.492.759a1.882.4120.8790.347
    23.551.742.427a1.862.1750.5670.252
    33.961.892.319±0.022N.A.2.105N.A.0.214
    43.901.862.333±0.020N.A.2.113N.A.0.220
    74.161.962.278±0.0211.852.0790.4280.199
    94.231.982.264±0.022N.A.2.069N.A.0.195
    124.312.012.251±0.0251.882.0620.3710.189
    144.261.992.259±0.036N.A.2.066N.A.0.193
    185.002.262.155±0.0301.852.0050.3050.150
    255.622.492.098±0.0421.881.9750.2180.123
    295.602.482.099±0.046N.A.1.976N.A.0.124
    315.642.502.096±0.033N.A.1.974N.A.0.122
    Bulk5.602.482.099a1.881.9760.2190.123

    Table 2. Dielectric Parameters (εn=1, εn=2), Quasiparticle Bandgap Eg, Energy Level, and Binding Energy for n=1 and n=2 Excitonic States from Monolayer to Multilayer

    Shift of the quasiparticle bandgap Eg and the first excited excitonic state (En=2) versus number of layers. As a result, the exciton binding energy (Ebn=1 and Ebn=2) decreases with layer number (bottom inset). The 1s-excitonic state locates at ∼1.88 eV [obtained from the linear absorption spectra in Fig. 1(e)] and is independent of the layer number. Here, n is the principal quantum number in the non-hydrogen model. Top inset: increase of the dielectric parameter (εn=1, εn=2) with layers. The dashed lines are intended as a visual guide.

    Figure 4.Shift of the quasiparticle bandgap Eg and the first excited excitonic state (En=2) versus number of layers. As a result, the exciton binding energy (Ebn=1 and Ebn=2) decreases with layer number (bottom inset). The 1s-excitonic state locates at 1.88  eV [obtained from the linear absorption spectra in Fig. 1(e)] and is independent of the layer number. Here, n is the principal quantum number in the non-hydrogen model. Top inset: increase of the dielectric parameter (εn=1, εn=2) with layers. The dashed lines are intended as a visual guide.

    To further confirm the excitonic energy compression and the shift of the conduction band-edge, we investigated the excitation photon energy-dependent TPA process in 25LMoS2, which can be treated as a bulk semiconductor. A wavelength-tunable femtosecond system (35 fs, 1 kHz, TOPAS, Coherent Co.) pumped by a Ti:sapphire laser at 800 nm was used for these measurements. As shown in Fig. 5, the TPA coefficient approximately increases with the excitation photon energy, in good agreement with the scaling rule in Eq. (7). We noticed that there is a distinct peak at about 1.21 eV. Since the quasiparticle bandgap Eg of 25L MoS2 is below 2.1 eV as given in Table 2, TPA (at photon energy 2.42  eV) of the electrons in the valence band can transit deep into the conduction band; this peak was ascribed to the sub-band resonance [59].

    Green circles represent the TPA coefficients of ∼25LMoS2 obtained at different excitation photon energies. The peak at ∼1.21 eV is ascribed to the sub-band resonance in the conduction band. (Inset: schematic of two-photon resonant transition from valence band to a sub-band in the conduction band.) The dashed-dotted line is intended as a visual guide.

    Figure 5.Green circles represent the TPA coefficients of 25LMoS2 obtained at different excitation photon energies. The peak at 1.21  eV is ascribed to the sub-band resonance in the conduction band. (Inset: schematic of two-photon resonant transition from valence band to a sub-band in the conduction band.) The dashed-dotted line is intended as a visual guide.

    4. CONCLUSION

    In summary, we prepared monolayer, few-layer, and multilayer MoS2 nanosheets through mechanical exfoliation. Using the micro-I-scan technique, we studied the TPA properties in 2D layered MoS2 and found that the TPA coefficient decreased rapidly from monolayer to few-layer, and increased gradually from few-layer to multilayer up to 25 layers. Combining the non-hydrogen model of excitons in TMDCs with the scaling rule in semiconductors, we deduced the red shift of the 2p excitonic dark state and the band-edge of conduction band with increasing layer number. The quasiparticle bandgap and excitonic dark levels, as well as the binding energy of the exciton dark state dropped dramatically until the layer number increased to 25L, which implies this is the 2D-to-bulk transition point. The sharp decline of the TPA coefficient from monolayer to 2L–4L was ascribed to the on–off resonance of 2p excitonic dark states. In addition, the sub-band resonance in the conduction band was demonstrated. Our results provide valuable insights into the mechanism of the evolution of band structure and excitonic levels in layered MoS2, which can be applied to other TMDCs.

    Acknowledgment

    Acknowledgment. The authors acknowledge the utilization of equipment in the School of Physical Science and Technology in ShanghaiTech University. Y. F. Xie is grateful for the detailed discussion with Prof. Jean-Michel Nunzi.

    Appendix A

    In our TPA measurement, the femtosecond pulses centered at 1030 nm are generated from a mode-locked fiber laser. The pulse width is about 10 nm, i.e., the spectral width is 10 meV, which is measured using the PG2000 (Ideaoptics) fiber optic spectrometer (see Fig. 6). It is much narrower than that of a Ti:sapphire femtosecond laser. Therefore, resonance excitation should occur between the conduction band maximum (CBM) and the excited excitonic state.

    Spectrum of the femtosecond laser pulse in the nonlinear optical measurements.

    Figure 6.Spectrum of the femtosecond laser pulse in the nonlinear optical measurements.

    Identification of the size of laser beam spot in our nonlinear optical measurements; the radius is ∼5 μm.

    Figure 7.Identification of the size of laser beam spot in our nonlinear optical measurements; the radius is 5  μm.

    AFM images of few-layer and multilayer MoS2.

    Figure 8.AFM images of few-layer and multilayer MoS2.

    Complete micro-I-scan fitting results of monolayer, few-layer, and multilayer MoS2, according to the homogeneously broadened TPA model.

    Figure 9.Complete micro-I-scan fitting results of monolayer, few-layer, and multilayer MoS2, according to the homogeneously broadened TPA model.

    Schematic diagram of the setup of micro-intensity scan.

    Figure 10.Schematic diagram of the setup of micro-intensity scan.

    Appendix B

    In monolayer, the first-order effective screening is given by Ref. [38] in the text: Eb(1)=Eg En=1=2.759 1.88=0.879 eV,ε1=2.8,εn=12[1+1+32παμ9n(n 1)+3],Eb(n)=μe42 2εn2(n 12)2.Combined with Eqs. (B1) and (B2), we can get that ε2=1.49 and Eb(2)=0.347 eV.

    Meanwhile, we can obtain the value of αμ of 0.60, and μe42 2 of 1.72 eV, which are assumed as constants in few-layer and multilayer MoS2. Hence, we can calculate the ε1, ε2, and Eb(2) of few-layer and multilayer MoS2, as shown in Table 2.

    References

    [1] X. B. Yin, Z. L. Ye, D. A. Chenet, Y. Ye, K. O’Brien, J. C. Hone, X. Zhang. Edge nonlinear optics on a MoS2 atomic monolayer. Science, 344, 488-490(2014).

    [2] K. P. Wang, J. Wang, J. T. Fan. Ultrafast saturable absorption of two-dimensional MoS2 nanosheets. ACS Nano, 7, 9260-9267(2013).

    [3] H. C. Kim, H. Kim, J. U. Lee, H. B. Lee, D. H. Choi, J. H. Lee, W. H. Lee, S. H. Jhang, B. H. Park, H. Cheong, S. W. Lee, H. J. Chung. Engineering optical and electronic properties of WS2 by varying the number of layers. ACS Nano, 9, 6854-6860(2015).

    [4] X. L. Li, W. P. Han, J. B. Wu, X. F. Qiao, J. Zhang, P. H. Tan. Layer-number dependent optical properties of 2D materials and their application for thickness determination. Adv. Funct. Mater., 27, 1604468(2017).

    [5] F. Zhou, W. Ji. Giant three-photon absorption in monolayer MoS2 and its application in near-infrared photodetection. Laser Photon. Rev., 11, 1700021(2017).

    [6] Z. Ye, T. Cao, K. O’Brien, H. Zhu, X. Yin, Y. Wang, S. G. Louie, X. Zhang. Probing excitonic dark states in single-layer tungsten disulphide. Nature, 513, 214-218(2014).

    [7] K. He, N. Kumar, L. Zhao, Z. Wang, K. F. Mak, H. Zhao, J. Shan. Tightly bound excitons in monolayer WSe2. Phys. Rev. Lett., 113, 026803(2014).

    [8] J. Xiao, Z. Ye, Y. Wang, H. Zhu, Y. Wang, X. Zhang. Nonlinear optical selection rule based on valley-exciton locking in monolayer WS2. Light: Sci. Appl., 4, e366(2015).

    [9] H. M. Hill, A. F. Rigosi, C. Roquelet, A. Chernikov, T. C. Berkelbach, D. R. Reichman, M. S. Hybertsen, L. E. Brus, T. F. Heinz. Observation of excitonic Rydberg states in monolayer MoS2 and WS2 by photoluminescence excitation spectroscopy. Nano Lett., 15, 2992-2997(2015).

    [10] M. Rumi, J. W. Perry. Two-photon absorption: an overview of measurements and principles. Adv. Opt. Photon., 2, 451(2010).

    [11] D. A. Fishman, C. M. Cirloganu, S. Webster, L. A. Padilha, M. Monroe, D. J. Hagan, E. W. Van Stryland. Sensitive mid-infrared detection in wide-bandgap semiconductors using extreme non-degenerate two-photon absorption. Nat. Photonics, 5, 561-565(2011).

    [12] M. Reichert, A. L. Smirl, G. Salamo, D. J. Hagan, E. W. Van Stryland. Observation of nondegenerate two-photon gain in GaAs. Phys. Rev. Lett., 117, 073602(2016).

    [13] B. Taheri, H. Liu, B. Jassemnejad, D. Appling, R. C. Powell, J. J. Song. Intensity scan and two photon absorption and nonlinear refraction of C60 in toluene. Appl. Phys. Lett., 68, 1317-1319(1996).

    [14] Y. Li, N. Dong, S. Zhang, X. Zhang, Y. Feng, K. Wang, L. Zhang, J. Wang. Giant two-photon absorption in monolayer MoS2. Laser Photon. Rev., 9, 427-434(2015).

    [15] T. Olsen, S. Latini, F. Rasmussen, K. S. Thygesen. Simple screened hydrogen model of excitons in two-dimensional materials. Phys. Rev. Lett., 116, 056401(2016).

    [16] K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang, S. V. Dubonos, I. V. Grigorieva, A. A. Firsov. Electric field effect in atomically thin carbon films. Science, 306, 666-669(2004).

    [17] K. F. Mak, C. Lee, J. Hone, J. Shan, T. F. Heinz. Atomically thin MoS2: a new direct-gap semiconductor. Phys. Rev. Lett., 105, 136805(2010).

    [18] M. Drüppel, T. Deilmann, P. Krüger, M. Rohlfing. Diversity of trion states and substrate effects in the optical properties of an MoS2 monolayer. Nat. Commun., 8, 2117(2017).

    [19] T. Deilmann, K. S. Thygesen. Important role of screening the electron-hole exchange interaction for the optical properties of molecules near metal surfaces. Phys. Rev. B, 99, 045133(2019).

    [20] H. Li, Q. Zhang, C. C. R. Yap, B. K. Tay, T. H. T. Edwin, A. Olivier, D. Baillargeat. From bulk to monolayer MoS2: evolution of Raman scattering. Adv. Funct. Mater., 22, 1385-1390(2012).

    [21] C. G. Lee, H. G. Yan, L. E. Brus, T. F. Heinz. Anomalous lattice vibrations of single and few-layer MoS2. ACS Nano, 4, 2695-2700(2010).

    [22] R. Coehoorn, C. Haas, J. Dijkstra, C. J. F. Flipse, R. A. de Groot, A. Wold. Electronic structure of MoSe2, MoS2, and WSe2. I. Band-structure calculations and photoelectron spectroscopy. Phys. Rev. B, 35, 6195-6202(1987).

    [23] R. Coehoorn, C. Haas, R. A. de Groot. Electronic structure of MoSe2, MoS2, and WSe2. II. The nature of the optical band gaps. Phys. Rev. B, 35, 6203-6206(1987).

    [24] A. Splendiani, L. Sun, Y. Zhang, T. Li, J. Kim, C. Y. Chim, G. Galli, F. Wang. Emerging photoluminescence in monolayer MoS2. Nano Lett., 10, 1271-1275(2010).

    [25] D. Y. Qiu, F. H. da Jornada, S. G. Louie. Optical spectrum of MoS2: many-body effects and diversity of exciton states. Phys. Rev. Lett., 111, 216805(2013).

    [26] M. Sheik-Bahae, A. A. Said, T. H. Wei, D. J. Hagan, E. W. Van Stryland. Sensitive measurement of optical nonlinearities using a single beam. IEEE J. Quantum Electron., 26, 760-769(1990).

    [27] F. Zhou, W. Ji. Two-photon absorption and subband photodetection in monolayer MoS2. Opt. Lett., 42, 3113-3116(2017).

    [28] F. Zhou, J. H. Kua, S. Lu, W. Ji. Two-photon absorption arises from two-dimensional excitons. Opt. Express, 26, 16093-16101(2018).

    [29] G. S. He, Q. Zheng, A. Baev, P. N. Prasad. Saturation of multiphoton absorption upon strong and ultrafast infrared laser excitation. J. Appl. Phys., 101, 083108(2007).

    [30] A. R. Beal, H. P. Huges. Kramers-Kronig analysis of the reflectivity spectra of 2H-MoS2, 2H-MoSe2, and 2H-MoTe2. J. Phys. C, 12, 881-890(1979).

    [31] C. Yim, M. O’Brien, N. McEvoy, S. Winters, I. Mirza, J. G. Lunney, G. S. Duesberg. Investigation of the optical properties of MoS2 thin films using spectroscopic ellipsometry. Appl. Phys. Lett., 104, 103114(2014).

    [32] S. F. Zhang, N. N. Dong, N. McEvoy, M. O’Brien, S. Winters, N. C. Berner, C. Yim, Y. X. Li, X. Y. Zhang, Z. H. Chen, L. Zhang, G. S. Duesberg, J. Wang. Direct observation of degenerate two-photon absorption and its saturation in WS2 and MoS2 monolayer and few-layer films. ACS Nano, 9, 7142-7150(2015).

    [33] J. He, J. Mi, H. P. Li, W. Ji. Observation of interband two-photon absorption saturation in CdS nanocrystals. J. Phys. Chem. B, 109, 19184-19187(2005).

    [34] R. L. Sutherland. Handbook of Nonlinear Optics(2003).

    [35] N. Dong, Y. Li, S. Zhang, N. McEvoy, R. Gatensby, G. S. Duesberg, J. Wang. Saturation of two-photon absorption in layered transition metal dichalcogenides: experiment and theory. ACS Photon., 5, 1558-1565(2018).

    [36] D. A. B. Miller, A. M. Fox. Excitons in resonant coupling of quantum wells. Phys. Rev. B, 42, 1841-1844(1989).

    [37] A. M. Fox, D. A. B. Miller. Excitonic effects in coupled quantum wells. Phys. Rev. B, 44, 6231-6242(1991).

    [38] T. Cheiwchanchamnangij, W. R. L. Lambrecht. Quasiparticle band structure calculation of monolayer, bilayer, and bulk MoS2. Phys. Rev. B, 85, 205302(2012).

    [39] T. Cao, G. Wang, W. Han, H. Ye, C. Zhu, J. Shi, Q. Niu, P. Tan, E. Wang, B. Liu, J. Feng. Valley-selective circular dichroism of monolayer molybdenum disulphide. Nat. Commun., 3, 887(2012).

    [40] H. Zeng, J. Dai, W. Yao, D. Xiao, X. Cui. Valley polarization in MoS2 monolayers by optical pumping. Nat. Nanotechnol., 7, 490-493(2012).

    [41] K. F. Mak, K. He, J. Shan, T. F. Heinz. Control of valley polarization in monolayer MoS2 by optical helicity. Nat. Nanotechnol., 7, 494-498(2012).

    [42] G. Kioseoglou, A. T. Hanbicki, M. Currie, A. L. Friedman, D. Gunlycke, B. T. Jonker. Valley polarization and intervalley scattering in monolayer MoS2. Appl. Phys. Lett., 101, 221907(2012).

    [43] J. H. Choi, P. Cui, H. P. Lan, Z. Y. Zhang. Linear scaling of the exciton binding energy versus the band gap of two-dimensional materials. Phys. Rev. Lett., 115, 066403(2015).

    [44] H.-P. Komsa, A. V. Krasheninnikov. Effects of confinement and environment on the electronic structure and exciton binding energy of MoS2 from first principles. Phys. Rev. B, 86, 241201(2012).

    [45] T. Deilmann, K. S. Thygesen. Interlayer excitons with large optical amplitudes in layered van der Waals materials. Nano Lett., 18, 2984-2989(2018).

    [46] A. Chernikov, T. C. Berkelbach, H. M. Hill, A. Rigosi, Y. Li, O. B. Aslan, D. R. Reichman, M. S. Hybertsen, T. F. Heinz. Exciton binding energy and nonhydrogenic Rydberg series in monolayer WS2. Phys. Rev. Lett., 113, 076802(2014).

    [47] P. Cudazzo, C. Attaccalite, I. V. Tokatly, A. Rubio. Strong charge-transfer excitonic effects and the Bose-Einstein exciton condensate in graphane. Phys. Rev. Lett., 104, 226804(2010).

    [48] B. Zhu, X. Chen, X. Cui. Exciton binding energy of monolayer WS2. Sci. Rep., 5, 9218(2015).

    [49] T. C. Berkelbach, M. S. Hybertsen, D. R. Reichman. Bright and dark singlet excitons via linear and two-photon spectroscopy in monolayer transition-metal dichalcogenides. Phys. Rev. B, 92, 085413(2015).

    [50] M. M. Glazov, L. E. Golub, G. Wang, X. Marie, T. Amand, B. Urbaszek. Intrinsic exciton-state mixing and nonlinear optical properties in transition metal dichalcogenide monolayers. Phys. Rev. B, 95, 035311(2017).

    [51] D. Y. Qiu, F. H. da Jornada, S. G. Louie. Screening and many-body effects in two-dimensional crystals: monolayer MoS2. Phys. Rev. B, 93, 235435(2016).

    [52] G. Wang, C. Robert, M. M. Glazov, F. Cadiz, E. Courtade, T. Amand, D. Lagarde, T. Taniguchi, K. Watanabe, B. Urbaszek, X. Marie. In-plane propagation of light in transition metal dichalcogenide monolayers: optical selection rules. Phys. Rev. Lett., 119, 047401(2017).

    [53] G. D. Mahan. Theory of two-photon spectroscopy in solids. Phys. Rev., 170, 825-838(1968).

    [54] B. S. Wherrett. Scaling rules for multiphoton interband absorption in semiconductors. J. Opt. Soc. Am. B, 1, 67-72(1984).

    [55] E. W. V. Stryland, M. A. Woodall, H. Vanherzeele, M. J. Soileau. Energy band-gap dependence of two-photon absorption-stryland. Opt. Lett., 10, 490-492(1985).

    [56] M. Sheik-Bahae, D. J. Hagan, E. W. Van Stryland. Dispersion and band-gap scaling of the electronic Kerr effect in solids associated with two-photon absorption. Phys. Rev. Lett., 65, 96-99(1990).

    [57] N. Dong, Y. Li, S. Zhang, N. McEvoy, X. Zhang, Y. Cui, L. Zhang, G. S. Duesberg, J. Wang. Dispersion of nonlinear refractive index in layered WS2 and WSe2 semiconductor films induced by two-photon absorption. Opt. Lett., 41, 3936-3939(2016).

    [58] A. Shimizu. Two-photon absorption in quantum-well structures near half the direct band gap. Phys. Rev. B, 40, 1403-1406(1989).

    [59] G. B. Serapiglia, E. Paspalakis, C. Sirtori, K. L. Vodopyanov, C. C. Phillips. Laser-induced quantum coherence in a semiconductor quantum well. Phys. Rev. Lett., 84, 1019-1022(2000).

    Yafeng Xie, Saifeng Zhang, Yuanxin Li, Ningning Dong, Xiaoyan Zhang, Lei Wang, Weimin Liu, Ivan M. Kislyakov, Jean-Michel Nunzi, Hongji Qi, Long Zhang, Jun Wang. Layer-modulated two-photon absorption in MoS2: probing the shift of the excitonic dark state and band-edge[J]. Photonics Research, 2019, 7(7): 762
    Download Citation