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

Competition mechanism of exciton decay channels in the stacked multilayer tungsten sulfide

Open Access Open Access

Abstract

The competition mechanism of exciton decay channels in the multilayer TMDs remains poorly understood. Here, the exciton dynamics in the stacked WS2 was studied. The exciton decay processes are divided into the fast and slow decay processes, which are dominated by the exciton-exciton annihilation (EEA) and defect-assisted recombination (DAR), respectively. The lifetime of EEA is on the order of hundreds of femtoseconds (400∼1100 fs). It is decreased initially, followed by an increase with adding layer thickness, which can be attributed to the competition between phonon-assisted effect and defect effect. The lifetime of DAR is on the timescale of hundreds of picoseconds (200∼800 ps), which is determined by the defect density especially in a high injected carrier density.

© 2023 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Two-dimensional (2D) transition metal dichalcogenides (TMDs) exhibit unusual photoelectric properties that can be exploited for novel optoelectronics, such as ultrasensitive phototransistors [1,2], ultrathin photovoltaics [3,4], and flexible photodetection [5,6]. These 2D TMDs assisted by metamaterials also exhibit giant nonlinear optical response [7,8]. For 2D TMDs materials, their energy band structures possess the layer-dependent properties, because there is a transition from an indirect bandgap in the multilayer to a direct gap in the monolayer [9]. Especially, their band structures can be tunable in a variety of ways, such as element doping, defect control, applied strain and electric field etc. [10,11]. What they have in common is that their surfaces are naturally passivated without any dangling bonds, which make it easily to form van der Waals heterostructures or integrate with photonic structures [12]. The difference is that monolayer TMDs absorbs around 10% of vertically incident light at excitonic resonances, while multilayer TMDs can absorb much more of incident light only achieved in the thickness regime [10,13,14]. Therefore, how to balance the relationship between the layer thickness and photoelectric performance is the key for the practical application of 2D materials-based devices. For example, the multilayer TMDs with spiral or pyramid type exhibit stronger second-harmonic intensity than the monolayer TMDs as a novel type of nonlinear optical material [1518].

The carrier lifetimes are critical to all the proposed TMDs-based optoelectronic devices. The carrier lifetimes directly determine the performance of devices, such as luminous intensity, luminous quantum efficiency and photoelectric conversion efficiency, etc. [14,1921]. The excited carriers rapidly decay from the high energy state to the bottom of the conduction band to form excitons, and then decay to the ground state through non-radiative or radiative processes [2224]. So far, most studies have been concentrated on the carrier lifetimes of pure monolayer TMDs [2529]. For the monolayer TMDs, the lifetime of non-radiative processes is usually in the order of hundreds of femtoseconds to a few picoseconds, which are mainly determined by the Auger electron scattering and exciton-exciton annihilation [11,30,31]. While the radiative lifetime of exciton takes a longer time in tens to hundreds of picoseconds depending on the defect density [27,32]. For the mechanically exfoliated multilayer TMDs, basic decay processes of exciton are similar to the monolayer TMDs, the different is that the phonon-assisted effect and the influence of layer thickness [33]. However, the competition mechanism of decay channels in exciton dynamics of the multilayer TMDs remain poorly understood. Developing a better understanding of the exciton decay channels in the multilayer TMDs is especially important, because the longer exciton lifetime is more conducive to carrier separation, thus improving photoelectric conversion efficiency in the TMDs-based optoelectronic device. To our best knowledge, the exciton dynamics of the stacked multilayer tungsten sulfide (WS2) prepared by physical vapor deposition (PVD) has not yet been studied.

In this letter, we reported an experimental observation of exciton dynamics in the stacked multilayer WS2 using a femtosecond pump–probe spectroscopy. The results indicated that exciton decay processes are divided into fast and slow decay processes. The fast decay process is dominated by the exciton-exciton annihilation (EEA) that the lifetime of EEA is decreased initially, followed by an increase with adding layer thickness. The slow decay process is dominated by the defect-assisted recombination (DAR), which is determined by the defect density especially in a high injected carrier density. These results provide a foundation for understanding exciton dynamics in the stacked multilayer WS2 material.

2. Results and discussion

2.1 Basic characterization of the stacked WS2 layers

The basic properties of the stacked WS2 layers on a Si/SiO2 substrate prepared by PVD method were first characterized. Figure 1(a) is the optical microscope images of stacked WS2 layers. The samples with triangular shape are typical properties of 2D WS2 layers [26,34,35]. As shown in Supplement 1 Fig. S1, the constructed monolayer WS2 has a side length of approximately 41 µm. Such large-size samples offer a powerful condition for accurate studies of layer-related optics and exciton dynamics features. The corresponding number of layers and measurement positions are marked with symbols (1, 2, 3, 4 and M) in Fig. 1(a). During the subsequent pump power dependent and probe wavelength dependent exciton dynamics examinations, the measurement positions were kept unchanged to minimize the measurement errors caused by position changes. The thickness of the WS2 layers is characterized by atomic force microscopy (AFM) as shown in Fig. 1(b). The height profile of ∼ 7 Å indicates that a monolayer was stacked on the top [23,35,36]. Layer numbers and thickness are indicated in the appropriate positions. The measured microscopic reflectance and photoluminescence (PL) spectra are presented in Fig. 1(c) and (d). The reflectance spectra of stacked WS2 layers are shown in Fig. 1(c). The absorption peaks of excitons appear as valleys in the reflectance spectrum. Due to the limits of the spectral apparatus, the overall linear baseline of the reflection spectrum is altered, but the location of the reflection valley of the exciton is unaffected. All layers display two main valleys of XA and XB excitons. The XA and XB excitons originates from the top splitting of the valence band maximum (VBM) caused by strong spin-orbit coupling and interlayer coupling (only suitable for few-layer WS2) [10,33]. The reflection valley of the XA exciton is approximately 638 nm (∼1.94 eV) in the monolayer and grows with the number of layers which reaches about 648 nm (∼1.91 eV) in the 4-layer WS2. The reflection valley of the XA exciton is, conversely, blue-shifted to 629 nm (∼1.97 eV) in the multilayer WS2. The XB exciton is located at ∼525 nm (∼2.36 eV) for all layers. The PL intensity following excitation at 3.1 eV was normalized around the emission peak of monolayer. The emission peaks of XA and XB excitons were clearly presented in the PL spectra around 642 nm (∼1.93 eV) and 534 nm (∼2.32 eV) which can be observed in all different WS2 layers. The PL intensities of XA and XB excitons decrease as the number of layers increases while the splitting energy (∼400 meV) is almost independent of layer numbers [37]. It needs to be noticed that the PL intensity here is relatively lower compared to the PL from exfoliated WS2 monolayers which indicated the PVD growth stacked WS2 layers have lower quantum efficiency and more defects [10,30]. The PL of bilayer WS2 prepared by the classical method is 2 orders of magnitude less than that of the monolayer [10]. Whereas the PL intensity of bilayer or even multilayer WS2 prepared by PVD method here is of the same order of magnitude as the monolayer due to the 0° stacking [36,38]. Referring to the results of the literature [36] and combining the experimental data of the optical image and PL spectra in this work, we have reason to infer that the stacked multilayer WS2 is a 3R-like phase stacking structure. As the layer number increase, the PL intensity of the 3R-like phase sample weaken more slowly, which mainly attributed to different interlayer coupling strengths [36,39,40]. The interlayer electronic coupling leads to the transition from direct to indirect band gap, the lower the indirect band gap, the stronger the coupling strength [41]. From Fig. 1(d), it is reasonable to infer that the PL intensity of the stacked multilayer WS2 exhibit slower attenuation, which mainly originated from the weak interlayer coupling strengths. The relationship of them and the relevant topics need further study.

 figure: Fig. 1.

Fig. 1. Characterization of the stacked WS2 layers on a Si/SiO2 substrate. (a) Optical microscope images of stacked WS2 layers with scale bar of 10 µm. (b) AFM image of the investigated WS2 layers within the white dotted box in (a). The inset depicts a monolayer AFM picture matching to the black dotted frame in (a). (c) Measured microscopic reflectance spectra of stacked WS2 layers shown in (a). The exciton peaks XA and XB are indicated. (d) Measured microscopic PL spectra of stacked WS2 layers shown in (a). The PL peaks of XA and XB excitons are indicated.

Download Full Size | PDF

2.2 XA exciton dynamics versus layer number

The strategy that is followed to investigate the XA exciton dynamics in stacked WS2 layers consists of exciting the samples with photon energies above their band gaps, then probing the decay dynamics of the XA exciton. The experimental configuration of dichromatic coaxial microscopic pump-probe measurement shown in Fig. S2 is as follows; the excitation photon energy is selected to be 3.1 eV (400 nm) while the probing photon energy is ∼2 eV (∼620 nm). This excitation photon energy can pump electrons with energy levels lower than VBM to energy levels higher than conduction band (CB) bottom level. Noticed that the electron-hole pairs generated by the injection of a 400 nm pump pulse into WS2 layers will rapidly generate excitons due to the large exciton binding energy, which is included in the rising edge of the differential reflection signal [42,43]. Therefore, the photocarrier systems caused by the pump pulse is mainly composed of excitons. From the pump fluence, we can estimate the injected exciton density by considering the absorption on the order of 10% per layer in the visible region of the optical spectrum and assume that every pump photon absorbed excites one exciton [28,44]. The probe energy (∼2 eV) is tuned to around the XA exciton resonance of WS2 layers, so that the dynamic process of XA exciton can revealed. All results reported here were obtained at room temperature ambient conditions. No samples were damaged during the entire study.

The dynamics of XA excitons under stacked WS2 layers are monitored by measuring differential reflection of the time-delayed 620 nm probe as shown in Fig. 2(a). The pump fluences Es here is 0.62 µJ/cm2 which can be derived by the equation ${E_s} = {P_{ex}}/({f \times S} )$, where Pex is the average power of the incident light, which is directly measured by the power meter, S is the area of the pump spot and f is the pulse frequency of the laser. The number of photons n in a single pump pulse can be calculated from equation ${E_s} = nhv$, where h is the Planck’s constant and v is the photon frequency [45]. By considering the absorption on the order of 10% per layer and assume that every pump photon absorbed excites one exciton [28], the corresponding injected carrier density is about 1.24 × 1011 cm−2.The differential reflection is defined as $\Delta R/{R_0}\; = \; ({R\; - \; {R_0}} )/{R_0}$, where R and ${R_0}$ are the reflection of probe signal with and without the presence of the pump injection, respectively. Since differential reflection signal is related to the exciton density as discussed before, its decay reflects exciton dynamics. As shown in Fig. 2(a), the differential reflection signal decays quickly in the first few ps shown in the transparent light blue areas and then slowly over several hundred ps in all layers obviously. Figure 2(b) is the signals extracted from (a) near zero-time delays with 100 fs interval steps. To focus on the rising edge of the signals, with more stacking layers, XA excitons form more slowly. This result means that it takes excited electrons more time in the CB to undergo intraband relaxations before reaching the lowest allowed energy level in the CB to form XA excitons when the layer number increases. The exciton formation is illustrated in Fig. 4(a). Across all layers, the formation time is within 1 ps, and in monolayer, the formation time is around 0.5 ps which is in good agreement with previous literature [29,42,43,46].

 figure: Fig. 2.

Fig. 2. (a) Differential reflection signals of stacked WS2 layers measured with pump fluences of 0.62 µJ/cm2 at the center probe spot with 500 fs interval steps. (b) Differential reflection signals extracted from (a) near zero-time delays with 100 fs interval steps. (c) Peak differential reflection signal as a function of WS2 layer number. (d) Dependence of τEEA and τDAR of the dynamics of XA excitons on the WS2 layer number. Error bars are standard deviations of the fitted results. The dashed lines are guides for the eye.

Download Full Size | PDF

The peak $\Delta R/{R_0}$ data versus layer number are extracted from Fig. 2(a) shown in Fig. 2(c). As the pump pulse only changes the probe signal less than 0.7% in all layers, the saturation effect of the pump absorption can be ignored [28,47]. According to the optical configuration, the intensity of the peak $\Delta R/{R_0}$ signal reflects the XA excitons population in the surface atomic layer of the WS2 sample after pump pulse injection [28]. As layers are stacked, the XA excitons population increases; however, with multiple layers, it is attenuated. The reason behind this is due to the linear absorption differences of the layers shown in Fig. 1(c). Since the differential reflection is related to the exciton density, its decay reflects the XA excitonic dynamics. Here, a biexponential decay function was used to fit the decay signal of the XA excitons in Fig. 2(a). The specific fitting process and attention details are shown in Supplement 1 S3. In 2D materials, exciton-exciton annihilation (EEA) can be triggered when the initial injected exciton density exceeds 1010 cm−2 [24,31]. The injected exciton density here reaches up to 1011 cm−2 which can easily trigger the EEA process. Hence, the fast decay lifetime (τEEA) here is considered mainly to be the EEA process as shown in Fig. 4(a). Noted in few and multilayers, which are indirect semiconductors, momentum conservation requires assistance from phonons in the EEA process [33]. As for the slower decay channel, defect-assisted recombination (DAR) is generally considered to be the dominant process since the stacked WS2 layers prepared by PVD method contain many defects which can easily capture excitons into defect state as shown in Fig. 4(b). In Fig. 2(d), the fit values of the fast and slow decay times (τEEA and τDAR) are shown in black and red points, respectively. The dashed lines are guides for the eye. The τEEA decreased from ∼0.7 ps initially, followed by an increase up to ∼1 ps with adding layer thickness. The decrease τEEA can be attributed to the phonon-assistance in few layers which increasing the EEA probability. Whereas, with the further increase of the layer number, the defect density increased which widen the distance between excitons and weaken the exciton diffusion inducing the increase τEEA. Hence, the variation of τEEA can be attributed to the competition between phonon-assisted effect and defect effect. In all layers, the τDAR is on the order of hundred picoseconds, showing a long-lived time for XA excitons under defect-assisted recombination. We also tested the three others stacked WS2 layers to maintain data accuracy and rigorousness as shown in Supplement 1 S4. The results of more systematic exciton lifetime tests and the underlying physical mechanism will be discussed in the following part.

2.3 XA exciton dynamics versus pump fluences

The differential reflection was measured on all layers with different pump fluences to establish a precise correlation between the differential reflection and the exciton density. Figure 3(a) and (b) are the 2D plots of differential reflection signals versus pump fluence for monolayer and multilayer WS2, respectively. The $\Delta R/{R_0}$ scales are normalized from 0 to 1% for both plots. A comparison of Fig. 3(a) and (b) indicates that the τDAR of multilayer WS2 is much longer than that of monolayer. The differential reflected signals under different pump fluences for the remaining samples are shown in Supplement 1 S5. In Fig. 3(c), the peak $\Delta R/{R_0}$ data points under different injected carrier densities are extracted from different layers which are distinguished by different colors. The peak $\Delta R/{R_0}\; $ signals show similar trend with the change of injected carrier densities in all layers. Based on a saturable absorption model, the relationship can be accurately described [28].

$$\frac{{\Delta R}}{{{R_0}}} \propto \frac{N}{{N + {N_S}}}$$

By assuming that every pump photon absorbed excites one exciton, the N and Ns here can be regarded as the XA exciton density and the saturation exciton density, respectively. As shown in Fig. 2(c), the solid lines represent the fitted curves of the saturable absorption model where the XA exciton density saturates more rapidly on multilayer WS2 compared to others. The fitted results Ns in each layer are represented in Fig. 2(d). As layers are stacked, the saturation XA exciton density decreases nonlinearly, from nearly (1.46 ± 0.09) × 1012 cm-2 in monolayer down to (0.23 ± 0.02) × 1012 cm-2 in multilayer, which is consistent with previous literature [48]. The average XA exciton distance is converted from about 8 nm in monolayer increased to 26 nm in multilayer. Considering that the saturation carrier density of monolayer MX2 is in the order of 1 × 1013 cm-2, the results is reasonably consistent with these calculations [28,48]. The impact of surface defect density increase in the uppermost layer is the primary factor causing the saturation exciton density to drop as the number of layers increases [30]. Due to the PVD method, the uppermost atomic layer defect density would grow as layers were stacked, reducing the effective regions where excitons might develop.

 figure: Fig. 3.

Fig. 3. 2D plot of differential reflection signals versus pump fluence for monolayer (a) and multilayer (b) WS2. The $\Delta R/{R_0}$ scales are normalized from 0 to 1% for both diagrams. (c) An overview of the peak $\Delta R/{R_0}$ signals as a function of the injected carrier density in all layers. The solid lines represent the fitting curves. (d) Variation of saturation carrier concentration with the number of layers. Error bars are standard deviations of the fitted results.

Download Full Size | PDF

The decay curves of differential reflection signals under varied injected carrier densities were collected and fitted with biexponential functions using the 2D diagrams of WS2 layers in Fig. 3 and Supplement 1 S4. EEA and DAR processes illustrated in Fig. 4(a), respectively, are dominant in the fitted lifetimes of fast and slow decay processes, as was previously noted. The individual τEEA and τDAR as a function of the injected carrier density are represented in Fig. 4(b) and (d). As observed in Fig. 4(b), τEEA generally keeps at 0.6 ps in the case of few layers stacking with just 1 to 4 layers, and is therefore basically independent of the injected carrier density. For the multilayer WS2, τEEA declines nonlinearly from 1.6 ps to around 0.6 ps as the injected carrier density rises, which finally is equivalent to 1-4 layers. In multilayer WS2, the additional defect denistiy widen the space between excitons and slow down exciton diffusion [30]. At low carrier injection densities, these effects lead to a lower probability of exciton-exciton annihilation and are attenuated at high injection carrier densities. Figure 4(d) shows the variation of τDAR with the injected carrier density under all layers. Compared with monolayer, the superposition of layers leads to higher defect density and more relaxation of indirect excitons are caused by defect-assisted recombination, thus prolonging the lifetime of τDAR [24,32,46]. As the injected carrier density increases for WS2 with few layers, τDAR drops from 300–600 ps to below 200 ps. This is due to the fact that more non-radiative recombination pathways get involved in exciton recombination as the density of injected carriers climbs [24,33]. However, the τDAR of the multilayer exhibits a phenomenon of nonlinear increasing from ∼200 ps to ∼ 800 ps. On the multilayer, the probe spot covers several stages. The actual result is the superposition of the exciton lifetime at the edges of different layers in the multilayer. More defects in the multilayer and the isolation of exciton diffusion between layers can be the root cause of the rising long τDAR [32]. As a result, the rate of exciton recombination is falling, and the lifetime of excitons is further increasing. The entire process can be very complex and competitive.

 figure: Fig. 4.

Fig. 4. (a) Schematic of XA exciton dynamics in defective WS2 layers. (b) and (d) are the τEEA and τDAR as a function of the injected carrier density in all layers, respectively. The biexponential fitted values with standard deviations are shown as solid color points with error bars. (c) The lifetime weights of τEEA and τDAR, marked as A1 and A2, as a function of the injected carrier density in each layer independently. The solid lines serve as eye-guides while the intersection positions are marked with green points.

Download Full Size | PDF

A1 and A2 in Fig. 4(c) are the amplitudes of the biexponential components which represents the lifetime weight of EEA and DAR process, respectively. The variations of Ai with increasing injected carrier density can reveal information on the predominant decay channel for excitons because the amplitudes Ai are closely correlated with the excited population [32]. The lifetime weight A1 of the EEA process for few layers decreases as the injected carrier density increases, whereas A2 has the opposite tendency. This competition phenomenon has been reported in the literature before [32], the trend can be explained by as follows. When in low exciton density, the exciton decay channel is dominated by the phonon-assisted EEA process for the defects can be nearly negligible. As the density of excitons rises, there are many defect states available inducing the DAR pathway is preferred while the phonon-assisted effect is attenuated leading the weight of EEA process falls. For few layers, the intersection position of A1 and A2 advances with the increase of the layer number, that is, the increase of defect density further increases the weight of DAR, lending further credence to the aforementioned explanation [32]. For multilayer, the weights of A1 and A2 remain basically unchanged with the injected carrier density for the extremely higher defect density and lower saturation exciton density.

2.4 XA exciton dynamics versus probe wavelength

In addition to studying the variation of XA exciton dynamics with layer number and pump fluence, this work continues to further explore the XA exciton dynamics as a function of probe wavelength. Here, a low carrier injection density of 0.62 µJ/cm2 with injected photoenergy of 3.1 eV at all probe wavelengths were maintained. The probe wavelengths, which covered absorption peaks of the XA exciton in each layer, varied from 610 nm (2.03 eV) to 650 nm (1.91 eV). From (a) and (b) in Fig. 5, it is clear that the change in peak $\Delta R/{R_0}$ intensity with probe wavelength corresponds to the reflectance spectrum. Please see Supplement 1 S6 for the 2D diagrams of all layers. This indicates that pump injected excitons lessen the strength of the exciton transition in all stacked layers at the responding absorption resonance peaks [28,32,46]. The individual τEEA and τDAR as a function of the probe wavelength are shown in Fig. 5(c) and (d). Figure 5(c) shows that in the case of few layers stacking with just 1 to 4 layers in the overall probe photoenergy, the whole τEEA decreased initially, followed by an increase with adding layer thickness. The reason has been revealed in the former part. The overall lifetime of the DAR process in Fig. 5(d), which is measured to be hundreds of ps and has minor dependence on the probe wavelength, does not appear to be regularly distributed. This result shows that the EEA and DAR processes of stacked WS2 layers prepared by PVD method can be obviously detected when the detecting energy near the XA exciton energy band. The lifetime of two processes is consequently unchanged with probe wavelength at low carrier injection densities, implying that minor variations in the observation energy levels have no obvious impact on the characterization of these processes. Additionally, the detection spot is restricted to the matching layer surface and the moving region is very limited, the diffusion related to the exciton has not been examined in this work.

 figure: Fig. 5.

Fig. 5. 2D plots of differential reflection signals versus probe wavelength for monolayer (a) and multilayer (b) WS2. The $\Delta R/{R_0}$ scales are normalized from 0 to 1.43% for both diagrams. (c) and (d) are the τEEA and τDAR as a function of the probe wavelength in all layers, individually. The biexponential fitted values with standard deviations are shown as solid color points with error bars. The dashed lines serve as eye-guides.

Download Full Size | PDF

3. Conclusion

In summary, a thorough investigation of the XA exciton dynamics in stacked WS2 layers grown by PVD method was studied. Firstly, as the layer stacks, the defect density increases, the saturation XA exciton density drops nonlinearly, and the exciton formation time rises. All WS2 layers present the fast and slow decay processes, which are dominated by the EEA and DAR processes, respectively. Secondly, the lifetime of EEA decreased from ∼0.7 ps initially, followed by an increase up to ∼1 ps with adding layer thickness, which can be attributed to the competition between phonon-assisted effect and defect effect. Thirdly, the lifetime of DAR is on the long-lived timescale of hundreds of picoseconds (200∼800 ps), which is mainly determined by the defect density, extremely in a high injected carrier density. Lastly, the layer number and injected carrier density affect the weights of the EEA and DAR processes. This is primarily because the competition of decay channel between phonon-assisted effect and defect density. The study of the competition mechanism of exciton decay process for the stacked WS2 materials, which is useful for its optical characteristics and 2D TMDs-based optoelectrical devices. The choice of relaxation channels and lifetime of exciton directly affect its luminescence characteristics, including the PL, second or third harmonic emission etc. Moreover, the photoelectric conversion efficiency is a key indicator for the applications of 2D TMDs-based optoelectrical devices. Extending the carrier lifetimes is more beneficial to the collection of charge. This work revealed that the carrier lifetimes can be efficiently modulated by controlling the exciton decay channels, which lays foundation for the analysis of the carrier relaxation process at the heterojunction interface, and is also helpful to guide the design of 2D TMDs-based optoelectronic devices.

4. Experimental section

Preparation of the stacked 2D-WS2 layers: An alumina boat loaded with WS2 powder was first placed at the heating center of a 1-in. quartz tube, and a piece of Si wafer (with 300 nm of SiO2) was placed 13 cm downstream of the center of the furnace. Argon gas flow was introduced into the system at a flow rate of 60−100 sccm for 60 min to exhaust the oxygen inside the tube before heating. The furnace was then rapidly heated to 1150 °C and maintained at this temperature for about 30 min, and WS2 nanosheets with stacked multilayer structures were deposited on the SiO2 surfaces.

Microscopic Reflectance and PL spectroscopy measurements: A customized microscopic optical spectroscopy system was used to measure the reflectance and PL spectra which were recorded by the spectrometer (Princeton Instruments SP2500) with a × 100 objective lens (Olympus MPlanFL, NA = 0.9). The samples were located on an automated XYZ stage with 0.5 µm step size with commercial Kohler lighting system. The reflectance spectra were represented by the relative reflectivity, which was calculated by dividing the reflected intensity of the sample to that of the substrate. The same procedure was performed for the PL spectra with excited photoenergy of 3.1 eV. Suitable filters were applied on the optical path to remove unnecessary signals.

Dichromatic coaxial microscopic pump-probe spectroscopy: The entire pump-probe measuring apparatus was constructed by ourselves, and Fig. S2 shows the schematic diagram that goes with it. The 800-nm fundamental frequency light generated in Chameleon Ti:Sapphire Laser with a repetition rate of 80 MHz and a pulse duration of 100 fs was passed into the Chameleon Compact OPO-vis device. High quality 400 nm pump light and tunable probe light were emitted from the two ports of the instrument, individually. The pump light chopped at 1000 Hz went through a linear polarizer with the polarization direction along the x axis then passed through a delay line controlled by a stepper linear stage (Newport Motion Controller ESP301). The two laser beams were collinear through a dichroic mirror and normally incident on the sample at room temperature by a × 40 objective lens (0.65 NA). The spot diameter of laser on the sample surface is about 3 µm. The sample can be precisely positioned via a three-axis high-precision translation stage with a self-made Kohler illumination device. Reflected probe light from the sample was detected by a high-sensitivity photomultiplier (Thorlabs PMM02), while reflected pump light was filtered by a long pass filter of 450 nm. The SNR was further improved through the chopper with lock-in amplifier SR830.The acquisition of transient differential reflection signals were controlled by Matlab program.

Funding

National Natural Science Foundation of China (11804408, 11874276, 5227130389); Fundamental Research Program of Shanxi Province (20210302124027); Zhejiang International Studies University scientific research fund (090530232021).

Acknowledgments

This work was supported by the National Natural Science Foundation of China (NSFC) (11804408, 11874276, 5227130389), the Fundamental Research Program of Shanxi Province (20210302124027) and the Zhejiang International Studies University scientific research fund (090530232021).

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Supplemental document

See Supplement 1 for supporting content.

References

1. N. Huo and G. Konstantatos, “Ultrasensitive all-2D MoS2 phototransistors enabled by an out-of-plane MoS2 PN homojunction,” Nat. Commun. 8(1), 572 (2017). [CrossRef]  

2. K. F. Mak and J. Shan, “Photonics and optoelectronics of 2D semiconductor transition metal dichalcogenides,” Nat. Photonics 10(4), 216–226 (2016). [CrossRef]  

3. M. Bernardi, M. Palummo, and J. C. Grossman, “Extraordinary sunlight absorption and one nanometer thick photovoltaics using two-dimensional monolayer materials,” Nano Lett. 13(8), 3664–3670 (2013). [CrossRef]  

4. F. Bonaccorso, L. Colombo, G. Yu, M. Stoller, V. Tozzini, A. C. Ferrari, R. S. Ruoff, and V. Pellegrini, “2D materials. Graphene, related two-dimensional crystals, and hybrid systems for energy conversion and storage,” Science 347(6217), 1246501 (2015). [CrossRef]  

5. 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]  

6. A. Allain, J. Kang, K. Banerjee, and A. Kis, “Electrical contacts to two-dimensional semiconductors,” Nat. Mater. 14(12), 1195–1205 (2015). [CrossRef]  

7. L. Lin, J. Xue, H. Xu, Q. Zhao, W. Zhang, Y. Zheng, L. Wu, and Z.-K. Zhou, “Integrating lattice and gap plasmonic modes to construct dual-mode metasurfaces for enhancing light–matter interaction,” Sci. China Mater. 64(12), 3007–3016 (2021). [CrossRef]  

8. Z. K. Zhou, H. F. Xu, Y. Yu, L. Lin, and X. H. Wang, “Giant Nonlinear Response of Monolayer MoS2 Induced by Optimal Field-Enhancement Gain Mode on the Surface of Hyperbolic Metamaterials,” Laser Photonics Rev. 15, 2100281 (2021). [CrossRef]  

9. W. Zhao, Z. Ghorannevis, L. Chu, M. Toh, C. Kloc, P.-H. Tan, and G. Eda, “Evolution of Electronic Structure in Atomically Thin Sheets of WS2 and WSe2,” ACS Nano 7(1), 791–797 (2013). [CrossRef]  

10. C. Cong, J. Shang, Y. Wang, and T. Yu, “Optical Properties of 2D Semiconductor WS2,” Adv. Opt. Mater. 6(1), 1700767 (2018). [CrossRef]  

11. R. Perea-Causin, D. Erkensten, J. M. Fitzgerald, J. J. P. Thompson, R. Rosati, S. Brem, and E. Malic, “Exciton optics, dynamics, and transport in atomically thin semiconductors,” APL Mater. 10(10), 100701 (2022). [CrossRef]  

12. K. S. Novoselov, D. Jiang, F. Schedin, T. J. Booth, V. V. Khotkevich, S. V. Morozov, and A. K. Geim, “Two-dimensional atomic crystals,” Proc. Natl. Acad. Sci. 102(30), 10451–10453 (2005). [CrossRef]  

13. Q. Li, J. Lu, P. Gupta, and M. Qiu, “Engineering Optical Absorption in Graphene and Other 2D Materials: Advances and Applications,” Adv. Opt. Mater. 7(20), 1900595 (2019). [CrossRef]  

14. A. Maniyar and S. Choudhary, “Visible region absorption in TMDs/phosphorene heterostructures for use in solar energy conversion applications,” RSC Adv. 10(53), 31730–31739 (2020). [CrossRef]  

15. X. Fan, Y. Jiang, X. Zhuang, H. Liu, T. Xu, W. Zheng, P. Fan, H. Li, X. Wu, X. Zhu, Q. Zhang, H. Zhou, W. Hu, X. Wang, L. Sun, X. Duan, and A. Pan, “Broken Symmetry Induced Strong Nonlinear Optical Effects in Spiral WS2 Nanosheets,” ACS Nano 11(5), 4892–4898 (2017). [CrossRef]  

16. X. Lin, Y. Liu, K. Wang, C. Wei, W. Zhang, Y. Yan, Y. J. Li, J. Yao, and Y. S. Zhao, “Two-Dimensional Pyramid-like WS2 Layered Structures for Highly Efficient Edge Second-Harmonic Generation,” ACS Nano 12(1), 689–696 (2018). [CrossRef]  

17. X.-L. Li, W.-P. Han, J.-B. Wu, X.-F. Qiao, J. Zhang, and P.-H. Tan, “Layer-Number Dependent Optical Properties of 2D Materials and Their Application for Thickness Determination,” Adv. Funct. Mater. 27(19), 1604468 (2017). [CrossRef]  

18. J. Peng, D. Yang, C. Ren, Y. Jiang, X. Zhu, F. Jing, H. Qiu, H. Liu, and Z. Hu, “Electronic Properties and Carrier Dynamics at the Alloy Interfaces of WS2xSe2-2x Spiral Nanosheets,” Adv. Mater. 34(11), 2107738 (2022). [CrossRef]  

19. D. Jariwala, V. K. Sangwan, L. J. Lauhon, T. J. Marks, and M. C. Hersam, “Emerging device applications for semiconducting two-dimensional transition metal dichalcogenides,” ACS Nano 8(2), 1102–1120 (2014). [CrossRef]  

20. B. W. Baugher, H. O. Churchill, Y. Yang, and P. Jarillo-Herrero, “Optoelectronic devices based on electrically tunable p-n diodes in a monolayer dichalcogenide,” Nat. Nanotechnol. 9(4), 262–267 (2014). [CrossRef]  

21. L. Gao, “Flexible Device Applications of 2D Semiconductors,” Small 13(35), 1603994 (2017). [CrossRef]  

22. H. Liu, C. Wang, D. Liu, and J. Luo, “Neutral and defect-induced exciton annihilation in defective monolayer WS2,” Nanoscale 11(16), 7913–7920 (2019). [CrossRef]  

23. Yu Yang, W.-J. Zhang, W.-Y. Zhao, X. Lin, Z.-M. Jin, W.-M. Liu, and G.-H. Ma, “Dynamics of A-exciton and spin relaxation in WS2 and WSe2 monolayer,” Acta Phys. Sin. 68(1), 017201 (2019). [CrossRef]  

24. H. Liu, C. Wang, Z. Zuo, D. Liu, and J. Luo, “Direct Visualization of Exciton Transport in Defective Few-Layer WS2 by Ultrafast Microscopy,” Adv. Mater. 32(2), 1906540 (2020). [CrossRef]  

25. Y. Fu, D. He, J. He, A. Bian, L. Zhang, S. Liu, Y. Wang, and H. Zhao, “Effect of Dielectric Environment on Excitonic Dynamics in Monolayer WS2,” Adv. Mater. Interfaces 6(23), 1901307 (2019). [CrossRef]  

26. S. Boandoh, F.-H. Hsiao, B. Shin, M. Mack, L. P. Grote, R. Wang, Y. Han, J. Lou, N. Koch, N. H. Nickel, C.-W. Luo, and Y.-T. Tsai, “Ultrafast Pump–Probe Microscopy on 2D Transition Metal Dichalcogenides,” Photonics Res. 3(8), 2200046 (2022). [CrossRef]  

27. H. Wang, C. Zhang, and F. Rana, “Ultrafast dynamics of defect-assisted electron-hole recombination in monolayer MoS2,” Nano Lett. 15(1), 339–345 (2015). [CrossRef]  

28. N. Kumar, Q. Cui, F. Ceballos, D. He, Y. Wang, and H. Zhao, “Exciton-exciton annihilation in MoSe2 monolayers,” Phys. Rev. B 89(12), 125427 (2014). [CrossRef]  

29. F. Ceballos, Q. Cui, M. Z. Bellus, and H. Zhao, “Exciton formation in monolayer transition metal dichalcogenides,” Nanoscale 8(22), 11681–11688 (2016). [CrossRef]  

30. L. Yuan, T. Wang, T. Zhu, M. Zhou, and L. Huang, “Exciton Dynamics, Transport, and Annihilation in Atomically Thin Two-Dimensional Semiconductors,” J. Phys. Chem. Lett. 8(14), 3371–3379 (2017). [CrossRef]  

31. D. Sun, Y. Rao, G. A. Reider, G. Chen, Y. You, L. Brezin, A. R. Harutyunyan, and T. F. Heinz, “Observation of rapid exciton-exciton annihilation in monolayer molybdenum disulfide,” Nano Lett. 14(10), 5625–5629 (2014). [CrossRef]  

32. Z. Ezgi Eroglu, D. Contreras, P. Bahrami, N. Azam, M. Mahjouri-Samani, and A. Boulesbaa, “Filling Exciton Trap-States in Two-Dimensional Tungsten Disulfide (WS2) and Diselenide (WSe2) Monolayers,” Nanomaterials 11(3), 770 (2021). [CrossRef]  

33. L. Yuan and L. Huang, “Exciton dynamics and annihilation in WS2 2D semiconductors,” Nanoscale 7(16), 7402–7408 (2015). [CrossRef]  

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]  

35. H. Liu, J. Lu, K. Ho, Z. Hu, Z. Dang, A. Carvalho, H. R. Tan, E. S. Tok, and C. H. Sow, “Fluorescence Concentric Triangles: A Case of Chemical Heterogeneity in WS2 Atomic Monolayer,” Nano Lett. 16(9), 5559–5567 (2016). [CrossRef]  

36. G. Shao, X.-X. Xue, X. Liu, D. Zhang, Y. Jin, Y. Wu, B. You, Y.-C. Lin, S. Li, K. Suenaga, X. Wang, A. Pan, H. Li, J. Hong, Y. Feng, and S. Liu, “Twist Angle-Dependent Optical Responses in Controllably Grown WS2 Vertical Homojunctions,” Chem. Mater. 32(22), 9721–9729 (2020). [CrossRef]  

37. H. Zeng, G. B. Liu, J. Dai, Y. Yan, B. Zhu, R. He, L. Xie, S. Xu, X. Chen, W. Yao, and X. Cui, “Optical signature of symmetry variations and spin-valley coupling in atomically thin tungsten dichalcogenides,” Sci. Rep. 3(1), 1608 (2013). [CrossRef]  

38. J. Hong, C. Jin, J. Yuan, and Z. Zhang, “Atomic Defects in Two-Dimensional Materials: From Single-Atom Spectroscopy to Functionalities in Opto-/Electronics, Nanomagnetism, and Catalysis,” Adv. Mater. 29(14), 1606434 (2017). [CrossRef]  

39. S. M. Shinde, K. P. Dhakal, X. Chen, W. S. Yun, J. Lee, H. Kim, and J.-H. Ahn, “Stacking-controllable interlayer coupling and symmetric configuration of multilayered MoS2,” NPG Asia Mater. 10(2), e468 (2018). [CrossRef]  

40. J. van Baren, G. Ye, J.-A. Yan, Z. Ye, P. Rezaie, P. Yu, Z. Liu, R. He, and C. H. Lui, “Stacking-dependent interlayer phonons in 3R and 2 H MoS2,” 2D Mater. 6(2), 025022 (2019). [CrossRef]  

41. K. Liu, L. Zhang, T. Cao, C. Jin, D. Qiu, Q. Zhou, A. Zettl, P. Yang, S. G. Louie, and F. Wang, “Evolution of interlayer coupling in twisted molybdenum disulfide bilayers,” Nat. Commun. 5(1), 4966 (2014). [CrossRef]  

42. C. Trovatello, F. Katsch, N. J. Borys, M. Selig, K. Yao, R. Borrego-Varillas, F. Scotognella, I. Kriegel, A. Yan, A. Zettl, P. J. Schuck, A. Knorr, G. Cerullo, and S. D. Conte, “The ultrafast onset of exciton formation in 2D semiconductors,” Nat. Commun. 11(1), 5277 (2020). [CrossRef]  

43. P. Steinleitner, P. Merkl, P. Nagler, J. Mornhinweg, C. Schuller, T. Korn, A. Chernikov, and R. Huber, “Direct Observation of Ultrafast Exciton Formation in a Monolayer of WSe2,” Nano Lett. 17(3), 1455–1460 (2017). [CrossRef]  

44. X. Marie and B. Urbaszek, “2D materials: Ultrafast exciton dynamics,” Nat. Mater. 14(9), 860–861 (2015). [CrossRef]  

45. Y. Yu, Z. Ji, S. Zu, B. Du, Y. Kang, Z. Li, Z. Zhou, K. Shi, and Z. Fang, “Ultrafast Plasmonic Hot Electron Transfer in Au Nanoantenna/MoS2 Heterostructures,” Adv. Funct. Mater. 26(35), 6394–6401 (2016). [CrossRef]  

46. Z. E. Eroglu, O. Comegys, L. S. Quintanar, N. Azam, S. Elafandi, M. Mahjouri-Samani, and A. Boulesbaa, “Ultrafast dynamics of exciton formation and decay in two-dimensional tungsten disulfide (2D-WS2) monolayers,” Phys. Chem. Chem. Phys. 22(30), 17385–17393 (2020). [CrossRef]  

47. G. Vecchi, V. Giannini, and J. Gómez Rivas, “Shaping the Fluorescent Emission by Lattice Resonances in Plasmonic Crystals of Nanoantennas,” Phys. Rev. Lett. 102(14), 146807 (2009). [CrossRef]  

48. J. He, D. He, Y. Wang, Q. Cui, F. Ceballos, and H. Zhao, “Spatiotemporal dynamics of excitons in monolayer and bulk WS2,” Nanoscale 7(21), 9526–9531 (2015). [CrossRef]  

Supplementary Material (1)

NameDescription
Supplement 1       Supplemental document

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

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 (5)

Fig. 1.
Fig. 1. Characterization of the stacked WS2 layers on a Si/SiO2 substrate. (a) Optical microscope images of stacked WS2 layers with scale bar of 10 µm. (b) AFM image of the investigated WS2 layers within the white dotted box in (a). The inset depicts a monolayer AFM picture matching to the black dotted frame in (a). (c) Measured microscopic reflectance spectra of stacked WS2 layers shown in (a). The exciton peaks XA and XB are indicated. (d) Measured microscopic PL spectra of stacked WS2 layers shown in (a). The PL peaks of XA and XB excitons are indicated.
Fig. 2.
Fig. 2. (a) Differential reflection signals of stacked WS2 layers measured with pump fluences of 0.62 µJ/cm2 at the center probe spot with 500 fs interval steps. (b) Differential reflection signals extracted from (a) near zero-time delays with 100 fs interval steps. (c) Peak differential reflection signal as a function of WS2 layer number. (d) Dependence of τEEA and τDAR of the dynamics of XA excitons on the WS2 layer number. Error bars are standard deviations of the fitted results. The dashed lines are guides for the eye.
Fig. 3.
Fig. 3. 2D plot of differential reflection signals versus pump fluence for monolayer (a) and multilayer (b) WS2. The $\Delta R/{R_0}$ scales are normalized from 0 to 1% for both diagrams. (c) An overview of the peak $\Delta R/{R_0}$ signals as a function of the injected carrier density in all layers. The solid lines represent the fitting curves. (d) Variation of saturation carrier concentration with the number of layers. Error bars are standard deviations of the fitted results.
Fig. 4.
Fig. 4. (a) Schematic of XA exciton dynamics in defective WS2 layers. (b) and (d) are the τEEA and τDAR as a function of the injected carrier density in all layers, respectively. The biexponential fitted values with standard deviations are shown as solid color points with error bars. (c) The lifetime weights of τEEA and τDAR, marked as A1 and A2, as a function of the injected carrier density in each layer independently. The solid lines serve as eye-guides while the intersection positions are marked with green points.
Fig. 5.
Fig. 5. 2D plots of differential reflection signals versus probe wavelength for monolayer (a) and multilayer (b) WS2. The $\Delta R/{R_0}$ scales are normalized from 0 to 1.43% for both diagrams. (c) and (d) are the τEEA and τDAR as a function of the probe wavelength in all layers, individually. The biexponential fitted values with standard deviations are shown as solid color points with error bars. The dashed lines serve as eye-guides.

Equations (1)

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

Δ R R 0 N N + N S
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.