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

Auger-type process in ultrathin ReS2

Open Access Open Access

Abstract

The dramatic enhancement of charge carrier interaction makes many-body effects of great prominence in two-dimensional materials. Here we report the defect-assisted Auger scattering combined with band-to-band Auger recombination as playing the dominant recovery mechanism in the charge carriers of atomically thin-layered ReS2. Time resolved transient absorption spectra investigation reveals two different decay processes over the visible and near- infrared range, which is attributed to the shallow and deep defects introduced by the existence of sulfur (S) vacancy. A rate equation system is invoked to rationalize our peculiar pump and temperature dependence of carrier dynamics quantitatively. These findings provide theoretical insights into the significant role played by nonradiative Auger processes and may pave the way for the development of diverse ReS2-based high performance photonic and optoelectronic devices.

© 2020 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

The strong confinement and reduced dielectric screening endow the traditional semiconductors many intriguing properties when they are thinned down to atomic scale [14]. Two-dimensional (2D) materials, pioneered by graphene, have the huge potential to be the building blocks of future nanoelectronic and nanophotonic devices due to their ultrathin size, thickness-dependent bandgap and ultrafast optical response. Among the 2D material family, transition metal dichalcogenides (TMDs) with the formula MX2 (M = Mo, W, Re; X = S, Se, Te) attracted extensive optical research interests due to their exotic properties such as widely tunable bandgap, strong nonlinear optical response and good air stability [58]. Recently, rhenium disulfide (ReS2), as a rising member of TMDs family, has triggered many novel studies in optoelectronic field on account of its peculiar bandgap evolution [9]. Different from other classic 2H TMDs, Peierls distortion of the 1T structure of ReS2 results in the poor interlayer registry and weak interlayer coupling, gives rise to the reduced structure symmetry and strong in-plane anisotropy [7,912]. Hence, when comparing monolayer and bulk, the bandgap renormalization is absent in ReS2 and the indirect bandgap structures are similar [9,13]. Moreover, ReS2 has already shown some unique properties in the application of field-effect transistors, digital devices, photodetectors, phototransistors and nonlinear optical modulators [1421].

However, the key parameter which represents the performance of light-emitting devices in semiconductor, photoluminescence (PL) quantum yield (QY), is in a low range of only ∼0.0001 as reported previously in ReS2 [22]. This value is almost one or two orders less than that in monolayer MoS2 or WS2, suggesting that most of the photoexcited electrons and holes in ReS2 recombine nonradiatively [2224]. According to the previous works, the S vacancies, which accelerate charge carrier recombination nonradiatively and result in the deterioration of the radiative quantum efficiency, are frequently observed in chemical vapor deposition (CVD) fabricated TMDs [2527]. These defects generate mid-gap states within the bandgap and facilitate the energy exchange between charge carriers, opening alternative nonradiative recombination channels such as phonon assisted trapping and Auger type carrier trapping [28]. The earlier flurry of researches in characterizing the exciton absorption in ReS2 by transient absorption (TA) spectroscopy provides us a lot of fundamental information regarding the excitonic nonlinearities [7,29]. Their work mainly focused on carrier dynamics near the exciton line, while the detailed interpretation of the TA spectra in near infrared range closely correlated to nonradiative recombination pathways, is still very scarce and needs to be further explored. Therefore, the prerequisite for utilizing of ReS2-based optoelectronic devices is understanding its optical properties thoroughly. Ultrafast dynamical processes investigation including its recombination mechanisms and the associated timescales is a basic but essential aspect for both fundamental physics of light-matter interaction and further diverse high performance photonic and optoelectronic devices.

In this work, we investigate the carrier dynamics of trilayer ReS2 by pump-probe spectroscopy with broadband ultrafast laser pulses. Different from previously reported ultrafast dynamics studies of ultrathin ReS2 films, the probe wavelength is tuned to be far away from the exciton absorption band (∼810 nm) to avoid the influence of resonant or near-resonant interband excitonic nonlinearity. In our experiment, the relaxation of excited carriers contains a fast and a slow decay component, which are attributed to the shallow and deep defects assisted Auger scattering, respectively. The dependence of transient transmission on pump intensity and temperature rules out most of the recombination mechanism like phonon assisted trapping and bandgap renormalization effect, suggests that band-to-band Auger recombination also plays an important role. With the rate equation simulations, it can be concluded that Auger effect dominates the nonradiative recombination and leads to the deterioration of the radiative quantum efficiency in atomic thin ReS2 film. Our results distinguish the nonradiative recombination pathways and may help to improve the performance of ReS2 based optoelectronic devices via defect engineering.

2. Sample preparation and characterization

Large-scale continuous ReS2 film was synthesized on a sapphire substrate by CVD methods. The layer number and height topography of ReS2 film were determined by atomic force microscopy (AFM). The optical image and height profile shown in Figs. 1(a)–1(b) demonstrate the smooth surface of the ReS2 film with the thickness L of 2.68 nm, corresponding to triple sandwiched S-Re-S layers [9]. Previous studies have pointed out that there are at least 11 Raman modes in the range of 100-400 cm−1 originated from the decoupling of lattice vibrations between adjacent layers, which is completely different from other TMDs with higher crystal symmetries [9]. Here, a confocal microscopy system (LabRAM HR Evolution) excited by 633 nm cw laser was employed to study the Raman vibrational modes of the trilayer ReS2 film. As illustrated in Fig. 1(b), two prominent vibration modes A1g (in plane) and E2g (both the in-plane and out-of-plane) located at 212 cm−1 and 151 cm−1 are observed, respectively. Moreover, the Eg-like mode and second-order Raman peak at 305 cm−1 and 320 cm−1 are also detected, which are in good agreement with the previous reports [7,9,22]. The uniformity of ReS2 film was investigated by Raman mapping in the middle of the film. According to the Raman mapping results for A1g mode (212 cm−1) shown in Fig. 1(e), where the blue areas were factitious scratches and yellow areas represent the ReS2 film, no obvious difference can be observed for each side of the scratch. Figure 1(f) plots the optical image of the Raman mapping areas. These results suggest that our trilayer ReS2 film has a good uniformity.

 figure: Fig. 1.

Fig. 1. (a) The image of our ReS2 sample. (b) Surface AFM image of 3 L ReS2 film. Inset shows its height profile. (c) Raman spectra of ReS2. (d) Linear transmission and absorption coefficient spectra of ReS2. (e) Raman mapping result of the sample. (f) The optical micrograph of the Raman mapping area.

Download Full Size | PDF

The linear optical properties of 3 L ReS2 were studied through transmission and absorption spectroscopy from 400 to 1100 nm versus an uncoated sapphire substrate as reference. Meanwhile, the linear absorption coefficient α can be deduced by

$$T = \frac{{{T_{sample}}}}{{{T_{sapphire}}}} = {e^{ - \alpha L}},$$
where T is the linear transmission of the 3 L ReS2 film, ${T_{sample}}$ and ${T_{sapphire}}$ are the transmissions of the sample and the uncoated sapphire respectively. As plotted in Fig. 1(d), the transmittance increases with the wavelength while the linear absorption coefficient decreases from visible to near-infrared region. The exciton peak located at 780 nm (1.59 eV) is exhibited both in the transmission and absorption coefficient spectra, while the low energy exciton at 1.53 eV is hardly distinguished due to the temperature broadening.

3. Results and analysis

Transient absorption spectra were established firstly by employing the transient absorption measurements. Laser pulses at 400 nm with 35 fs duration were employed as an excitation source, both the visible (1.85-3.1 eV) and near-infrared (1.38-0.89 eV) region were probed. Since the bandgap of trilayer ReS2 is much smaller than the pump photon energy (3.1 eV), the electrons in the valence band can be excited into the conduction band effectively. As shown in Fig. 2(a), the transient absorption dynamics for visible range shows an initial photobleaching (PB, negative change of differential absorption signal) immediately after the excitation, corresponding to the deep blue areas near zero delay time, followed by a speedy recovery of photoinduced absorption (PIA, positive change of differential absorption signal) and a slow relaxation process. The overall transient behaviors for different wavelength indicate the slow relaxation for the PIA process which lasts for more than 300 ps as depicted in Fig. 2(c). Meanwhile for near-infrared range, the transient absorption spectra demonstrated in Fig. 2(b) and 2(d) contain only an incipient PIA component and the corresponding long recovery process. The possible physical processes that result in the observed dynamics will be discussed later in degenerate and nondegenerate pump-probe part. Tri- and bi-exponential models considering the autocorrelation of the pump and probe pulses were employed to fit the visible and infrared transient absorption decay profiles respectively [30,31].

$$\textrm{g}(t) = \left\{ \begin{array}{c} {D_1}\exp ( - \frac{t}{{{\tau_1}}})erfc(\frac{\sigma }{{\sqrt 2 {\tau_1}}} - \frac{t}{{\sqrt 2 \sigma }}) + {D_2}\exp ( - \frac{t}{{{\tau_2}}})erfc(\frac{\sigma }{{\sqrt 2 {\tau_2}}} - \frac{t}{{\sqrt 2 \sigma }})\\ {D_0}\exp ( - \frac{t}{{{\tau_0}}})erfc(\frac{\sigma }{{\sqrt 2 {\tau_0}}} - \frac{t}{{\sqrt 2 \sigma }}) + {D_1}\exp ( - \frac{t}{{{\tau_1}}})erfc(\frac{\sigma }{{\sqrt 2 {\tau_1}}} - \frac{t}{{\sqrt 2 \sigma }})\\ + {D_2}\exp ( - \frac{t}{{{\tau_2}}})erfc(\frac{\sigma }{{\sqrt 2 {\tau_2}}} - \frac{t}{{\sqrt 2 \sigma }}) \end{array} \right.$$
where g(t) is the pump probe result, D0, D1 and D2 are the amplitudes of each component, “erfc” is the integral error function, σ represents the laser pulse duration (35 fs), t is the delay time, τ0 represents the rise time of the signal from PB to PIA in visible range measurements, τ1 and τ2 are the lifetimes of the excited carriers for prompt and slow components of relaxation process respectively. The optimized fitting parameters at different probe wavelength are summarized in Table 1 and Figs. 2(e) and 2(f). From the fitting results, we can see that the fast recombination time τ1 remains almost unchanged in visible probe range and decreases with the increase of near infrared wavelength. Meanwhile the slow recovery time τ2 shows a monotonic decline trend, demonstrating a faster recombination rate at longer wavelength. The wavelength dependent carrier lifetimes and ultrafast optical response transition from PB to PIA are beneficial to the manipulating of light at selective band [32].

 figure: Fig. 2.

Fig. 2. (a)-(b) Visible and near-infrared transient absorption spectra in a false-color plot under a 35 fs-pulsed laser excitation with pump intensity of 0.8 GW/cm2 at 400 nm. (c)-(d) Decay profiles of transient differential absorption monitored at different wavelength (460-620 and 900-1300 nm). The data were plotted with an offset. (e)-(f) Evaluated lifetimes of the fast and slow component in the recovery of the transient differential absorption spectra, plotted versus probe wavelengths.

Download Full Size | PDF

Tables Icon

Table 1. Fitting results of ultrafast transient absorption spectra measurements.

The excited carrier dynamics was studied by employing both the degenerate and nondegenerate pump-probe experiments to explore the ultrafast optical responses and carrier relaxation processes at different energy states. Laser pulses at 520 nm with 380 fs duration were employed as the pump, while the probe was applied at both 520 nm (degenerate) and 1040 nm (nondegenerate). The precision of the time-resolved pump-probe setup has been confirmed by our previous reports of MoS2, PtS and PtSe2 films [5,33,34]. By using the knife edge method, the beam radius of 520 nm pump pulse was determined to be ∼130 µm, while the spot sizes of probe beam at 520 nm and 1040 nm were ∼50 µm and ∼56 µm respectively. Assuming that one pump photon at 520 nm only excites one pair of photocarriers, the maximal carrier density N can be calculated by

$$N = \frac{{{F_{peak}}}}{{h\nu }}(1 - {e^{ - \alpha L}}), $$
where Fpeak is the fluence of pump excitation, L is the film thickness [35]. As the pump fluence was tuned in 285–1140 µJ/cm2, knowing the linear absorption coefficient α equaled to 1.45 × 106 cm−1 and the thickness L of the trilayer ReS2 was 2.68 nm, the injected carrier density N was estimated to be in 2.38–9.58 × 1014 cm−2 range, which means that high-order nonlinear optical effects like bandgap renormalization and Auger effect need to be considered [25,36,37]. Both the degenerate and nondegenerate transient differential transmission ΔT/T results are depicted in Fig. 3. Here, ΔT/T = (T-T0)/T0, where T0 is the linear transmission. Since the visible and near-infrared transient absorption results are nearly identical to the nonlinear ultrafast responses at degenerate and nondegenerate probe wavelength, we can attribute this observed phenomenon to the same physical process. For probing at 520 nm, as illustrated in Fig. 3(a), the whole process shows four different temporal regions: ($\textrm{I}$) Firstly, an initial PB immediately after the excitation, which reaches its maximum value within 1 ps. ($\textrm{II}$) Then, a rapid recovery and increase of PIA that occurs within 1.5 ps. Finally, ($\textrm{III}$) fast and ($\textrm{IV}$) slow decay components that last more than several hundred picoseconds. The initial PB signal can be attributed to the filling of states and Pauli blocking of electrons. On account of the ground state bleaching, the linear dependence between the maximum peak values of ΔT/T and the densities of photocarriers N is consistent with the rule of saturable absorption as shown in Fig. 3(c), which also suggests that higher order nonlinear excitonic interaction is negligible for the pump fluences employed in our experiment. The following PIA signal may be originated from many possible mechanisms such as bandgap renormalization, intra-band carrier thermalization or hot carrier intra-band relaxation, and free carrier absorption (FCA). Generally, the bandgap renormalization can be defined as the shrink of electronic gaps resulted from the strong interaction between excited state carriers, which creates a new transition pathway for the probe photons and increases the transient absorption signal. For bandgap renormalization process, the anticipated dependence for ΔT/T should be N1/3 at weak carriers interaction or N1/2 when the carrier-carrier interactions become more violent [36,38,39]. However, as illustrated in Fig. 3(c), the linear dependence of ΔT/T versus carrier density N (taken at maximum values of PIA) indicates that the contribution from bandgap renormalization is negligible. After the initial photoexcitation, a large quantity of excited carriers will be thermalized and then cooled instantly, and this process generally occurs within 500 fs (close to the resolution of our experiment), which is much faster than the delay time in region II. Moreover, the intra-band relaxation process of hot carriers will only lead to the initial growing of PIA but not contribute to the negative transient transmission [37]. Therefore, in addition to the hot carriers relaxation and thermalization process, FCA of the intra-band transition may well be the mechanism which results in the observed PIA signals. As reported previously, FCA coefficient scaled linearly with N [28], which is in consistent with our observation shown in Fig. 3(c). Note that FCA is generally sensitive to temperature, but such a dependence is not reliable when $\hbar \omega \gg k{\cal T}$ occurs, where k is the Boltzmann constant and ${\cal T}$ is the electron temperature [40]. Hence it is reasonable that we did not observe any significant temperature dependence in the transient results shown in Figs. 4(b) and 4(d).

 figure: Fig. 3.

Fig. 3. The zoomed-in transient differential transmission dynamics at 520 nm femtosecond pulse excitation, probed at (a) 520 nm and (d) 1040 nm. The entire transient dynamics probed at (b) 520 nm and (e) 1040 nm with different pump intensity plotted versus delay time between pump and probe pulse. The solid lines represent the fittings using two exponential decay functions. (c) The maximum and minimum values of ΔT/T versus the pump intensity at zero-time delay probed at 520 nm. (f) The maximum values of |ΔT/T| versus the pump intensity at zero-time delay probed at 1040 nm. The linear dependence rules out the existence of bandgap renormalization effect in our dynamics.

Download Full Size | PDF

 figure: Fig. 4.

Fig. 4. Normalized transient transmission ΔT/T of 3 L ReS2 film with different pump intensities probed at (a) 1040 nm and (b) 520 nm. Temperature-dependent transient transmission dynamics of 3 L ReS2 film probed at (c)1040 nm and (d) 520 nm.

Download Full Size | PDF

Conversely, the 1040 nm transient dynamics depicted in Fig. 3(d) is appreciably different. The transient transmission signal dips immediately upon the pump pulse excitation and reaches its maximum negative value within 1 ps. This process can be attributed to the enhanced intra-band absorption of probe photons. The excited carriers are thermalized and relaxed efficiently to the conduct band minimum (CBM) after the initial pump pulse excitation, then give rise to the transition from CBM to an energetically higher state. The PIA signal possesses fast and slow recovery with timescales around 2 ps and 400 ps.

Figures 3(b) and 3(e) present the entire ReS2 transient dynamics for 520 nm and 1040 nm probe with different pump intensities. And the same exponential model discussed in the transient absorption part was employed to fit the pump-probe results, where the laser pulse duration is 380 fs, τ1 and τ2 are the lifetimes of the excited carriers for relaxation region II and III respectively. For probing at 520 nm, the results of decay time are 0.7 ± 0.2 ps (${\tau _1}$) and 150 ± 50 ps (${\tau _2}$), while for probing at 1040 nm, they change to 0.9 ± 0.1 ps (${\tau _1}$) and 200 ± 50 ps (${\tau _2}$).

To better understand the mechanism hiding behind the observed transient dynamics, we investigated the dependence of transient transmission signals on pump intensity and temperature as depicted in Fig. 4. The normalized transient transmissions of ReS2 at different pump intensities for 1040 nm and 520 nm probe are illustrated in Figs. 4(a) and 4(b), respectively. No significant pump intensity dependence for the fast relaxation process is observed in the entire range of the pump intensity values employed in our experiments. While for the long recombination process, the recovery becomes marginally faster at higher pump intensity. Therefore, the hot phonon effect can be excluded since it will result in increasing relaxation time at high pump fluence [41,42]. Besides, the different dependences for fast and slow relaxations suggest that our decay dynamics cannot be explained merely by one single recombination mechanism. Figures 4(c) and 4(d) plot the transient results probed at temperature of 77 K and 300 K of 1040 nm and 520 nm with an incident pump intensity of 0.75 and 1.63 GW/cm2 respectively. There is no observable temperature dependence in the signal amplitude and timescale for two different temperatures. The lack of temperature dependence implies that phonon assisted relaxation processes such as phonon assisted trapping and phonon cascade process are not so significant in our experiment [25,43,44]. Based on these pump intensity and temperature dependent dynamic results, we propose a compound mechanism of defect assisted Auger scattering and band-to-band Auger recombination to rationalize all features of our experiment data.

Recent experimental advances of note which are similar to our results include (1) Wang et al. investigated the below-gap behaviors of monolayer MoS2 and attributed the pump fluence independent dynamics to fast and slow traps assisted Auger scattering [45] and (2) Paul et al. invoked that the Auger recombination was the main mechanism which resulted in the decreases of exciton lifetime at high pump fluence for monolayer WS2 [46]. However, we find that our data cannot be explained by either trap assisted or pure band-to-band Auger recombination alone, since the measured timescales in Figs. 4(a) and 4(b) exhibit different pump fluence dependence for the fast and slow recombination regions. Hence, we propose a compound model which contains both the occupied traps assisted Auger scattering and the band-to-band Auger recombination for describing the capture process of excited carriers. For our sample, the atomic ratio of Re/S is about 1/1.86 that obtained from the X-ray photoelectron spectroscopy (XPS) measurement, which reveals the S vacancies are the dominant defects type. As reported previously, the Re atoms around the S vacancy would introduce two different trap states, shallow and deep traps within the bandgap [47]. These two traps will dominate the decay time of the excited carriers in turn, where the recovery time assisted by shallow traps can be shortened by several orders of magnitude compared with deep traps. Meanwhile, the band-to-band Auger recombination requires two or more charge carriers involved and always exhibits a faster recombination rate at higher pump intensity. Therefore, we consider that both the trap assisted Auger scattering and band-to-band Auger recombination are influential in our dynamics. Similar to the previous research of monolayer MoS2, the shallow and deep defect states are all assumed to be fully occupied before the pump pulse excitation [45]. Thus, the rate equations can be written as follows,

$$ \left\{ \begin{array}{l} \frac{{dn}}{{dt}} ={-} {D_{ns}}{n_s}{n^2}(1 - {F_s}) - {D_{nd}}{n_d}{n^2}(1 - {F_d}) - A{n^2}p + g(t)\\ \frac{{dp}}{{dt}} ={-} {D_{ps}}{n_s}np{F_s} - {D_{pd}}{n_d}np{F_d} - Bn{p^2} + g(t)\\ {n_s}\frac{{d{F_s}}}{{dt}} = {D_{ns}}{n_s}{n^2}(1 - {F_s}) - {D_{ps}}{n_s}np{F_s}\\ {n_d}\frac{{d{F_d}}}{{dt}} = {D_{nd}}{n_d}{n^2}(1 - {F_d}) - {D_{pd}}{n_d}np{F_d} \end{array}\right.$$

Here Dns (Dnd) is the shallow (deep) defect assisted coefficient for electron capture, Dps (Dpd) is the shallow (deep) defect assisted coefficient for hole capture, ns (nd) is the density of shallow (deep) defect level, Fs (Fd) is the electron occupation of the shallow (deep) defect level. A (B) is the rate coefficient for band-to-band Auger recombination of electron-hole pair and energy transfer to another electron (hole). g(t) is the pump induced generation rate of the free carriers.

According to previous works, the temporal transient transmission signal ΔT/T can be written as

$$\frac{{\triangle T}}{T} ={-} \frac{{{\eta _0}}}{2}[n(t) + p(t)]e\mu ,$$
where we assume the same mobility for electron and hole [48,49]. Here η0 is the impedance of free space and equals to 377 Ω, the carrier mobility µ = 12 cm2/V·s [15], n(t) and p(t) are the temporal areal density of electron and hole respectively. It is worthy to note that the Eq. (5) can only apply to the intraband absorption process, which is the case of our nondegenerate pump-probe results. Since the excitonic nonlinearities and band filling effects have a complicated dependence on the probe energy, the quantitative study of pump probe data is a difficult task when the probe energy is either near an exciton resonance line or near a band-edge [45]. We chose the fitting parameters in the simulation based on the same methods reported previously [45]. All the simulation parameters shown in Table 2 are suitable for different pump excitation intensity. Figure 5(b) demonstrates the simulation results for normalized transient dynamics at 520 nm pump and 1040 nm probe with 0.75 GW/cm2 and 3 GW/cm2, in which we find the model exhibits a good agreement with experiment results in our pump intensity range. Therefore, we are able to deduce that the ultrafast carrier dynamics consists of several main decay processes corresponding to different timescales demonstrated in Fig. 5(a). After the initial optical pulse excitation, plentiful electron-hole pairs were created at pump photon energy. These excited carriers will thermalize and cool instantly and the timescale of this process is usually less than 500 fs. Subsequently, the recovery process can be divided into a fast and a slow component respectively. Most of the excited carriers will be captured by the shallow traps within 1 ps, corresponding to the fast decay time τ1. At the same time, only a small portion of excited carriers is captured by deep traps or transfers energy to another electron or hole via band-to-band Auger recombination process. Right after the filling of all shallow traps, the deep traps assisted Auger scattering and band-to-band Auger recombination are taken into account, and this decay process is responsible for the slow decay time τ2.

 figure: Fig. 5.

Fig. 5. (a) Schematic of charge carriers recombination in 3 L ReS2 film. Different recombination processes include carriers captured by ① shallow traps and ② deep traps through Auger scattering, and ③ band-to-band Auger recombination. (b) The simulation and experiment results of transient dynamics are plotted versus the probe delay at two extreme values of pump intensities used in our experiments, 0.75 and 3 GW/cm2. The inset shows the details for the first 7 ps. The simulated (c) electron and hole densities and (d) electron occupations of shallow and deep defects are plotted versus time delay with 3 GW/cm2 pump excitation.

Download Full Size | PDF

Tables Icon

Table 2. Fitting results used in the simulation of transient data.

The temporal evolution of electron density n(t), hole density p(t), the electron occupations of shallow traps Fs(t) and deep traps Fd(t) are plotted in Figs. 5(c) and 5(d) up to 500 ps at 3 GW/cm2 photoexcitation. The shallow traps start to capture carriers immediately after the photoexcitation and are almost fully occupied within 2 ps. Thereafter, the band-to-band Auger recombination along with deep trap capture process dominates the decay of the remaining charge carriers. Thus, we find that the deep traps are still not completely occupied at 500 ps delay time. The simulation results well reproduce the pump intensity dependence of the experimental data, in which the fast recovery rate is independent of pump intensity, while the slow decay transient becomes faster at higher pump intensity, further supporting the existence of defect assisted Auger scattering and band-to-band Auger recombination in trilayer ReS2 film. These results are of profound significance for the understanding of the nonradiative recombination mechanism in ReS2 and the defect engineering of ReS2 based optoelectronic devices.

4. Conclusions

In summary, the visible to near-infrared ultrafast dynamics of 3 L ReS2 have been investigated by transient absorption spectra together with degenerate and nondegenerate pump probe techniques. We have invoked the defect assisted Auger scattering and band-to-band Auger recombination to analyze the relaxation dynamics. The rate equation simulation matches well with the experiment results and reveals the leading role of Auger recombination played in the decay process. This inference can also be well confirmed by the temperature and pump intensity dependence of ultrafast dynamics results. Our work provides a detailed understanding of recombination mechanism in visible and near-infrared range of ReS2, and will be helpful in evaluating and enhancing the performance of ReS2 based optoelectronic devices.

5. Methods

5.1 Materials preparation and characterization

The sapphire substrate (1 cm × 1 cm) was sonicated in acetone for 30 minutes firstly, then it was cleaned by deionized water for 10 minutes and dried in vacuum drying oven for 1 hour. The substrate was placed on a ceramic boat containing 10 mg ReO3 powder (Sigma-Aldrich, 99.5%) and loaded at the center of furnace. The furnace temperature was ramped to 450 °C at 20 °C min-1 with 500 sccm Ar gas. And then the H2S gas with a flow rate of 4 sccm was injected for 15 minutes. After the growth of ReS2 film, the furnace was cooled down naturally to the room temperature. We found that the ReS2 film grew at a rate of one layer every 5 minutes under these experiment conditions. Therefore, we can get a trilayer ReS2 sample with a growth time of 15 minutes.

The AFM height profile was acquired using FM-Nanoview 6800 and operated in tapping mode. VG Scientific ESCAlab Mkll system using Al Kα X-rays was employed to obtain the XPS spectra with an analyzer pass energy of 20 eV. Raman characterization of ReS2 films were carried out using a confocal microscopy system (LabRAM HR Evolution) excited by 633 nm cw laser. The optical transmission spectra of ReS2 film were measured using a PerkinElmer Lambda 950 instrument.

5.2 Ultrafast pump probe measurements

The transient absorption spectra were investigated using femtosecond transient absorption spectrometer (HELIOS, Ultrafast System LLC, USA). The pump source was a 35 fs-pulsed laser at 400 nm with 1 kHz repetition rate. The probe light was a spectrum with a wavelength range from 430 to 640 nm and 900 to 1400 nm. The degenerate and non-degenerate pump probe measurements were performed with a 380 fs pump pulse at 520 nm and probe at 520 nm as well as the double frequency at 1040 nm with a repetition rate of 1 kHz. The intensity of probe pulses was always one order of magnitude less than pump pulses to avoid the self-induced nonlinearities of the probe pulses. The polarization of probe pulses was rotated to be perpendicular to the pump pulses, which could minimize the coherent artifacts [50]. The temperature dependent differential-transmission signal ΔT/T0 was measured in a vacuum cryostat.

Funding

National Natural Science Foundation of China (11874370, 61675217, 61875213, 61975221); Strategic Priority Research Program of Chinese Academy of Sciences (XDB16030700); CAS Center for Excellence in Ultra-intense Laser Science, the Key Research Program of Frontier Science of CAS (QYZDB-SSW-JSC041); Program of Shanghai Academic Research Leader (17XD1403900); Natural Science Foundation of Shanghai (18ZR1444700); Shanghai Rising-Star Program (19QA1410000); President’s International Fellowship Initiative of CAS (2017VTB0006, 2018VTB0007); Ministry of Science and Technology of the People's Republic of China (G20190161002); Program of Shanghai International Science and Technology Cooperation (19520710200).

Disclosures

The authors declare no conflicts of interest relate to this article.

References

1. S. Sim, D. Lee, M. Noh, S. Cha, C. H. Soh, J. H. Sung, M.-H. Jo, and H. Choi, “Selectively tunable optical Stark effect of anisotropic excitons in atomically thin ReS2,” Nat. Commun. 7(1), 13569 (2016). [CrossRef]  

2. X. Xu, W. Yao, D. Xiao, and T. F. Heinz, “Spin and pseudospins in layered transition metal dichalcogenides,” Nat. Phys. 10(5), 343–350 (2014). [CrossRef]  

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

4. A. Castellanos-Gomez, “Why all the fuss about 2D semiconductors?” Nat. Photonics 10(4), 202–204 (2016). [CrossRef]  

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

6. Y. Zhao, J. Qiao, Z. Yu, P. Yu, K. Xu, S. P. Lau, W. Zhou, Z. Liu, X. Wang, W. Ji, and Y. Chai, “High-electron-mobility and air-stable 2D layered PtSe2 FETs,” Adv. Mater. 29(5), 1604230 (2017). [CrossRef]  

7. X. Wang, K. Shinokita, H. E. Lim, N. B. Mohamed, Y. Miyauchi, N. T. Cuong, S. Okada, and K. Matsuda, “Direct and indirect exciton dynamics in few-layered ReS2 revealed by photoluminescence and pump-probe spectroscopy,” Adv. Funct. Mater. 29(6), 1806169 (2019). [CrossRef]  

8. F. Ceballos and H. Zhao, “Ultrafast laser spectroscopy of two-dimensional materials beyond graphene,” Adv. Funct. Mater. 27(19), 1604509 (2017). [CrossRef]  

9. S. Tongay, H. Sahin, C. Ko, A. Luce, W. Fan, K. Liu, J. Zhou, Y.-S. Huang, C.-H. Ho, J. Yan, D. F. Ogletree, S. Aloni, J. Ji, S. Li, J. Li, F. M. Peeters, and J. Wu, “Monolayer behaviour in bulk ReS2 due to electronic and vibrational decoupling,” Nat. Commun. 5(1), 3252 (2014). [CrossRef]  

10. H.-X. Zhong, S. Gao, J.-J. Shi, and L. Yang, “Quasiparticle band gaps, excitonic effects, and anisotropic optical properties of the monolayer distorted 1 T diamond-chain structures ReS2 and ReSe2,” Phys. Rev. B 92(11), 115438 (2015). [CrossRef]  

11. S. Sim, D. Lee, A. V. Trifonov, T. Kim, S. Cha, J. H. Sung, S. Cho, W. Shim, M.-H. Jo, and H. Choi, “Ultrafast quantum beats of anisotropic excitons in atomically thin ReS2,” Nat. Commun. 9(1), 351 (2018). [CrossRef]  

12. M. Z. Bellus, M. Li, S. D. Lane, F. Ceballos, Q. Cui, X. C. Zeng, and H. Zhao, “Type-I van der Waals heterostructure formed by MoS2 and ReS2 monolayers,” Nanoscale Horiz. 2(1), 31–36 (2017). [CrossRef]  

13. X. Meng, Y. Zhou, K. Chen, R. H. Roberts, W. Wu, J.-F. Lin, R. T. Chen, X. Xu, and Y. Wang, “Anisotropic saturable and excited-state absorption in bulk ReS2,” Adv. Opt. Mater. 6(14), 1800137 (2018). [CrossRef]  

14. F. Liu, X. Zhao, X.-Q. Yan, J. Xie, W. Hui, X. Xin, Z.-B. Liu, and J.-G. Tian, “Ultrafast nonlinear absorption and carrier relaxation in ReS2 and ReSe2 films,” J. Appl. Phys. 125(17), 173105 (2019). [CrossRef]  

15. E. Liu, M. Long, J. Zeng, W. Luo, Y. Wang, Y. Pan, W. Zhou, B. Wang, W. Hu, Z. Ni, Y. You, X. Zhang, S. Qin, Y. Shi, K. Watanabe, T. Taniguchi, H. Yuan, H. Y. Hwang, Y. Cui, F. Miao, and D. Xing, “High responsivity phototransistors based on few-layer ReS2 for weak signal detection,” Adv. Funct. Mater. 26(12), 1938–1944 (2016). [CrossRef]  

16. E. Zhang, Y. Jin, X. Yuan, W. Wang, C. Zhang, L. Tang, S. Liu, P. Zhou, W. Hu, and F. Xiu, “ReS2-based field-effect transistors and photodetectors,” Adv. Funct. Mater. 25(26), 4076–4082 (2015). [CrossRef]  

17. C. M. Corbett, C. McClellan, A. Rai, S. S. Sonde, E. Tutuc, and S. K. Banerjee, “Field effect transistors with current saturation and voltage gain in ultrathin ReS2,” ACS Nano 9(1), 363–370 (2015). [CrossRef]  

18. L. Du, G. Jiang, L. Miao, B. Huang, J. Yi, C. Zhao, and S. Wen, “Few-layer rhenium diselenide: an ambient-stable nonlinear optical modulator,” Opt. Mater. Express 8(4), 926–935 (2018). [CrossRef]  

19. J. S. Ross, P. Klement, A. M. Jones, N. J. Ghimire, J. Yan, D. G. Mandrus, T. Taniguchi, K. Watanabe, K. Kitamura, W. Yao, D. H. Cobden, and X. Xu, “Electrically tunable excitonic light-emitting diodes based on monolayer WSe2 p-n junctions,” Nat. Nanotechnol. 9(4), 268–272 (2014). [CrossRef]  

20. F. Xia, T. Mueller, Y.-m. Lin, A. Valdes-Garcia, and P. Avouris, “Ultrafast graphene photodetector,” Nat. Nanotechnol. 4(12), 839–843 (2009). [CrossRef]  

21. E. Liu, Y. Fu, Y. Wang, Y. Feng, H. Liu, X. Wan, W. Zhou, B. Wang, L. Shao, C.-H. Ho, Y.-S. Huang, Z. Cao, L. Wang, A. Li, J. Zeng, F. Song, X. Wang, Y. Shi, H. Yuan, H. Y. Hwang, Y. Cui, F. Miao, and D. Xing, “Integrated digital inverters based on two-dimensional anisotropic ReS2 field-effect transistors,” Nat. Commun. 6(1), 6991 (2015). [CrossRef]  

22. N. B. Mohamed, K. Shinokita, X. Wang, H. E. Lim, D. Tan, Y. Miyauchi, and K. Matsuda, “Photoluminescence quantum yields for atomically thin-layered ReS2: Identification of indirect-bandgap semiconductors,” Appl. Phys. Lett. 113(12), 121112 (2018). [CrossRef]  

23. C. Wuerth, M. Grabolle, J. Pauli, M. Spieles, and U. Resch-Genger, “Relative and absolute determination of fluorescence quantum yields of transparent samples,” Nat. Protoc. 8(8), 1535–1550 (2013). [CrossRef]  

24. C. Wuerth, D. Geissler, and U. Resch-Genger, “Quantification of anisotropy-related uncertainties in relative photoluminescence quantum yield measurements of nanomaterials - semiconductor quantum dots and rods,” Z. Phys. Chem. 229(1-2), 153–165 (2015). [CrossRef]  

25. L. Li, M.-F. Lin, X. Zhang, A. Britz, A. Krishnamoorthy, R. Ma, R. K. Kalia, A. Nakano, P. Vashishta, P. Ajayan, M. C. Hoffmann, D. M. Fritz, U. Bergmann, and O. V. Prezhdo, “Phonon-suppressed auger scattering of charge carriers in defective two-dimensional transition metal dichalcogenides,” Nano Lett. 19(9), 6078–6086 (2019). [CrossRef]  

26. L. Li, R. Long, T. Bertolini, and O. V. Prezhdo, “Sulfur adatom and vacancy accelerate charge recombination in MoS2 but by different mechanisms: time-domain ab Initio analysis,” Nano Lett. 17(12), 7962–7967 (2017). [CrossRef]  

27. W. Zhou, X. Zou, S. Najmaei, Z. Liu, Y. Shi, J. Kong, J. Lou, P. M. Ajayan, B. I. Yakobson, and J.-C. Idrobo, “Intrinsic structural defects in monolayer molybdenum disulfide,” Nano Lett. 13(6), 2615–2622 (2013). [CrossRef]  

28. J. Linnros, “Carrier lifetime measurements using free carrier absorption transients. I. Principle and injection dependence,” J. Appl. Phys. 84(1), 275–283 (1998). [CrossRef]  

29. Q. Cui, J. He, M. Z. Bellus, M. Mirzokarimov, T. Hofmann, H.-Y. Chiu, M. Antonik, D. He, Y. Wang, and H. Zhao, “Transient absorption measurements on anisotropic monolayer ReS2,” Small 11(41), 5565–5571 (2015). [CrossRef]  

30. G. Wang, K. Wang, B. M. Szydlowska, A. A. Baker-Murray, J. J. Wang, Y. Feng, X. Zhang, J. Wang, and W. J. Blau, “Ultrafast nonlinear optical properties of a graphene saturable mirror in the 2 µm wavelength region,” Laser Photonics Rev. 11(5), 1700166 (2017). [CrossRef]  

31. K. Wang, B. M. Szydlowska, G. Wang, X. Zhang, J. J. Wang, J. J. Magan, L. Zhang, J. N. Coleman, J. Wang, and W. J. Blau, “Ultrafast nonlinear excitation dynamics of black phosphorus nanosheets from visible to mid-infrared,” ACS Nano 10(7), 6923–6932 (2016). [CrossRef]  

32. F. Zhang, X. Jiang, Z. He, W. Liang, S. Xu, and H. Zhang, “Third-order nonlinear optical responses and carrier dynamics in antimonene,” Opt. Mater. 95, 109209 (2019). [CrossRef]  

33. J. Huang, N. Dong, N. McEvoy, L. Wang, C. O. Coileain, H. Wang, C. P. Cullen, C. Chen, S. Zhang, L. Zhang, and J. Wang, “Surface-state assisted carrier recombination and optical nonlinearities in bulk to 2D nonlayered PtS,” ACS Nano 13(11), 13390–13402 (2019). [CrossRef]  

34. L. Wang, S. F. Zhang, N. McEvoy, Y. Y. Sun, J. W. Huang, Y. F. Xie, N. N. Dong, X. Y. Zhang, I. M. Kislyakov, J. M. Nunzi, L. Zhang, and J. Wang, “Nonlinear optical signatures of the transition from semiconductor to semimetal in PtSe2,” Laser Photonics Rev. 13(8), 1900052 (2019). [CrossRef]  

35. Q. Cui, Y. Li, J. Chang, H. Zhao, and C. Xu, “Temporally resolving synchronous degenerate and nondegenerate two-photon absorption in 2D semiconducting monolayers,” Laser Photonics Rev. 13(2), 1800225 (2019). [CrossRef]  

36. Z. Guo, Y. Wan, M. Yang, J. Snaider, K. Zhu, and L. Huang, “Long-range hot-carrier transport in hybrid perovskites visualized by ultrafast microscopy,” Science 356(6333), 59–62 (2017). [CrossRef]  

37. X. Miao, G. Zhang, F. Wang, H. Yan, and M. Ji, “Layer-dependent ultrafast carrier and coherent phonon dynamics in black phosphorus,” Nano Lett. 18(5), 3053–3059 (2018). [CrossRef]  

38. J. G. Lu, S. Fujita, T. Kawaharamura, H. Nishinaka, Y. Kamada, T. Ohshima, Z. Z. Ye, Y. J. Zeng, Y. Z. Zhang, L. P. Zhu, H. P. He, and B. H. Zhao, “Carrier concentration dependence of band gap shift in n-type ZnO : Al films,” J. Appl. Phys. 101(8), 083705 (2007). [CrossRef]  

39. S. C. Jain, J. M. McGregor, and D. J. Roulston, “Band-gap narrowing in novel III-V semiconductors,” J. Appl. Phys. 68(7), 3747–3749 (1990). [CrossRef]  

40. W. P. Dumke, “Quantum Theory of Free Carrier Absorption,” Phys. Rev. 124(6), 1813–1817 (1961). [CrossRef]  

41. B. Gao, G. Hartland, T. Fang, M. Kelly, D. Jena, H. Xing, and L. Huang, “Studies of intrinsic hot phonon dynamics in suspended graphene by transient absorption microscopy,” Nano Lett. 11(8), 3184–3189 (2011). [CrossRef]  

42. H. Shi, R. Yan, S. Bertolazzi, J. Brivio, B. Gao, A. Kis, D. Jena, H. G. Xing, and L. Huang, “Exciton dynamics in suspended monolayer and few-Layer MoS2 2D crystals,” ACS Nano 7(2), 1072–1080 (2013). [CrossRef]  

43. C. H. Henry and D. V. Lang, “Nonradiative capture and recombination by multiphonon emission in GaAs and GaP,” Phys. Rev. B 15(2), 989–1016 (1977). [CrossRef]  

44. M. Lax, “Cascade capture of electrons in solids,” Phys. Rev. 119(5), 1502–1523 (1960). [CrossRef]  

45. 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]  

46. P. D. Cunningham, K. M. McCreary, and B. T. Jonker, “Auger recombination in chemical vapor deposition-grown monolayer WS2,” J. Phys. Chem. Lett. 7(24), 5242–5246 (2016). [CrossRef]  

47. J. Jiang, C. Ling, T. Xu, W. Wang, X. Niu, A. Zafar, Z. Yan, X. Wang, Y. You, L. Sun, J. Lu, J. Wang, and Z. Ni, “Defect engineering for modulating the trap states in 2D photoconductors,” Adv. Mater. 30(40), 1804332 (2018). [CrossRef]  

48. S. Kar, Y. Su, R. R. Nair, and A. K. Sood, “Probing photoexcited carriers in a few-layer MoS2 laminate by time-resolved optical pump-terahertz probe spectroscopy,” ACS Nano 9(12), 12004–12010 (2015). [CrossRef]  

49. J. H. Strait, P. Nene, and F. Rana, “High intrinsic mobility and ultrafast carrier dynamics in multilayer metal-dichalcogenide MoS2,” Phys. Rev. B 90(24), 245402 (2014). [CrossRef]  

50. P. D. Olszak, C. M. Cirloganu, S. Webster, L. A. Padilha, S. Guha, L. P. Gonzalez, S. Krishnamurthy, D. J. Hagan, and E. W. Van Stryland, “Spectral and temperature dependence of two-photon and free-carrier absorption in InSb,” Phys. Rev. B 82(23), 235207 (2010). [CrossRef]  

Cited By

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

Alert me when this article is cited.


Figures (5)

Fig. 1.
Fig. 1. (a) The image of our ReS2 sample. (b) Surface AFM image of 3 L ReS2 film. Inset shows its height profile. (c) Raman spectra of ReS2. (d) Linear transmission and absorption coefficient spectra of ReS2. (e) Raman mapping result of the sample. (f) The optical micrograph of the Raman mapping area.
Fig. 2.
Fig. 2. (a)-(b) Visible and near-infrared transient absorption spectra in a false-color plot under a 35 fs-pulsed laser excitation with pump intensity of 0.8 GW/cm2 at 400 nm. (c)-(d) Decay profiles of transient differential absorption monitored at different wavelength (460-620 and 900-1300 nm). The data were plotted with an offset. (e)-(f) Evaluated lifetimes of the fast and slow component in the recovery of the transient differential absorption spectra, plotted versus probe wavelengths.
Fig. 3.
Fig. 3. The zoomed-in transient differential transmission dynamics at 520 nm femtosecond pulse excitation, probed at (a) 520 nm and (d) 1040 nm. The entire transient dynamics probed at (b) 520 nm and (e) 1040 nm with different pump intensity plotted versus delay time between pump and probe pulse. The solid lines represent the fittings using two exponential decay functions. (c) The maximum and minimum values of ΔT/T versus the pump intensity at zero-time delay probed at 520 nm. (f) The maximum values of |ΔT/T| versus the pump intensity at zero-time delay probed at 1040 nm. The linear dependence rules out the existence of bandgap renormalization effect in our dynamics.
Fig. 4.
Fig. 4. Normalized transient transmission ΔT/T of 3 L ReS2 film with different pump intensities probed at (a) 1040 nm and (b) 520 nm. Temperature-dependent transient transmission dynamics of 3 L ReS2 film probed at (c)1040 nm and (d) 520 nm.
Fig. 5.
Fig. 5. (a) Schematic of charge carriers recombination in 3 L ReS2 film. Different recombination processes include carriers captured by ① shallow traps and ② deep traps through Auger scattering, and ③ band-to-band Auger recombination. (b) The simulation and experiment results of transient dynamics are plotted versus the probe delay at two extreme values of pump intensities used in our experiments, 0.75 and 3 GW/cm2. The inset shows the details for the first 7 ps. The simulated (c) electron and hole densities and (d) electron occupations of shallow and deep defects are plotted versus time delay with 3 GW/cm2 pump excitation.

Tables (2)

Tables Icon

Table 1. Fitting results of ultrafast transient absorption spectra measurements.

Tables Icon

Table 2. Fitting results used in the simulation of transient data.

Equations (5)

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

T = T s a m p l e T s a p p h i r e = e α L ,
g ( t ) = { D 1 exp ( t τ 1 ) e r f c ( σ 2 τ 1 t 2 σ ) + D 2 exp ( t τ 2 ) e r f c ( σ 2 τ 2 t 2 σ ) D 0 exp ( t τ 0 ) e r f c ( σ 2 τ 0 t 2 σ ) + D 1 exp ( t τ 1 ) e r f c ( σ 2 τ 1 t 2 σ ) + D 2 exp ( t τ 2 ) e r f c ( σ 2 τ 2 t 2 σ )
N = F p e a k h ν ( 1 e α L ) ,
{ d n d t = D n s n s n 2 ( 1 F s ) D n d n d n 2 ( 1 F d ) A n 2 p + g ( t ) d p d t = D p s n s n p F s D p d n d n p F d B n p 2 + g ( t ) n s d F s d t = D n s n s n 2 ( 1 F s ) D p s n s n p F s n d d F d d t = D n d n d n 2 ( 1 F d ) D p d n d n p F d
T T = η 0 2 [ n ( t ) + p ( t ) ] e μ ,
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.