This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Gate-tunable trion switch for excitonic device applications

Sarthak Das1, Sangeeth Kallatt2, Nithin Abraham1 and Kausik Majumdar1∗
1Department of Electrical Communication Engineering,
Indian Institute of Science, Bangalore 560012, India
2Center for Quantum Devices, Niels Bohr Institute, University of Copenhagen, Denmark
Corresponding author, email: [email protected]
Abstract

Introduction:

The monolayer semiconducting transition metal dichalcogenides (MoS2, WS2, MoSe2, and WSe2) exhibit strongly bound two-dimensional excitons with a binding energy on the order of few hundreds of meV, making these ultra-thin monolayers an excellent test bed for excitonic manipulation even at room temperature [1, 2, 3, 4, 5]. The neutral excitons (X0X^{0}) show excellent valley polarization and valley coherence properties that can be readily probed through initialization by circularly and linearly polarized photons, respectively, followed by detection through a circular or linear analyzer [6, 7, 8, 9, 10, 11]. However, controlling these excitonic states electrically remains a challenge due to the charge neutral nature of these excitons. In addition, transport of the exciton also remains challenging due to the ultra-fast radiative recombination of exciton[12, 13, 14, 15, 16] resulting from the high oscillator strength[17, 18, 19] - limiting the application of excitonic devices. Recently, this problem has been addressed by creating inter-layer exciton[20, 21, 22, 23, 24] to suppress the fast radiative decay, and exciton transport over several micrometer in the plane of the layered material has been demonstrated [4]. An external gate control has also been achieved by modulating the binding energy of the neutral exciton [4, 5].

In this regard, the charged exciton or trion (XX^{-}) is promising since its intensity can be readily controlled electrically by modulating the doping density using a gate voltage. In addition, the trion, while being optically initiated, can be electrically detected in a spatially nonlocal manner through measuring a charge current [25]. The relatively longer radiative lifetime of trion compared to intra-layer exciton [6, 11] further helps in nonlocal detection. Trions can be valley polarized when initiated through a circularly polarized light [7]. Trion can also carry valley coherence information when resonantly initialized through linearly polarized light [9]. Thus trions are excellent candidates for gate controlled excitonic device applications and also for transferring the valley information into electrical domain. In this work, we demonstrate a fast, gate- and light-tunable vertical trion switch where the trion is coherently initialized through resonant excitation followed by ultra-fast inter-layer transfer and electrical detection - thus generating a gate controlled photocurrent governed by inter-layer trion transport.

Results and Discussions:

Figure 1a shows the schematic diagram of the vertical switch with a monolayer WS2 sandwiched between monolayer graphene (MLG) at the bottom and few-layer graphene (FLG) on the top. The heterostructure is prepared on a heavily doped Si substrate coated with 285285 nm thick thermally grown SiO2. The two electrical contacts are deposited to the top and the bottom graphene. More details of the fabrication process are provided in the Methods section. Figure 1b shows the optical image of the device delineating the different stacked layers by marking with different colors.

The fabricated device is then placed on a thermal stage and the terminals of the device are connected to Keithley 2636B SMUs through micromanipulators for electrical measurements. The temperature (TT) of the device is then increased from room temperature to 423 K in steps of 10 K. At every temperature, the device is illuminated with a linearly polarized laser beam with photon energy of 2.332.33 and 1.95911.9591 eV through an objective with numerical aperture of 0.50.5 at different back gate voltages (VgV_{g}) ranging from 40-40 to +40+40 V. We record the in situ photoluminescence (PL) spectra and photocurrent at each VgV_{g} and TT steps. During measurement, the incident laser power is kept below 8 μ\muW to avoid any unwanted degradation of the device due to laser induced heating.

Figure 1c shows the transfer characteristics (IdarkI_{dark}-VgV_{g}) of the vertical device under dark condition at 295 K, when a Vd=20V_{d}=20 mV bias is applied across the two terminals. The graphene-like “V” shaped curve suggests that the VgV_{g} dependence primarily arises from the modulation of the chemical potential in the bottom monolayer graphene and the junction acts like a tunneling series resistance. In the inset, the IdarkI_{dark}-VdV_{d} plot is shown at Vg=0V_{g}=0 V confirming the ohmic nature of the junction arising from strong carrier tunneling between the top and the bottom graphene layers through the bandgap of WS2.

The difference in doping induced workfunction between the top FLG and the bottom MLG results in an asymmetry in the device along the vertical direction, which results in a built-in electric field, allowing a detectable photocurrent (Iph=IlightIdarkI_{ph}=I_{light}-I_{dark}) between the two electrodes under Vd=0V_{d}=0 V. This helps us to reduce the average dark current to zero, suppressing the dark noise of the switch dramatically. The sign of the zero-bias vertical IphI_{ph} at different VgV_{g} suggests a net electron flow from the top FLG to the bottom MLG.

Figure 1d depicts the temperature dependent variation of the A1sA_{1s} neutral exciton (X0X^{0}) peak position as obtained using 2.332.33 eV excitation. The individual spectra at each temperature are shown in Supplemental Material S1 [26]. The X0X^{0} and trion (XX^{-}) peak positions, as obtained from a Voigt fitting, are plotted as a function of temperature in Figure 1e. One could readily identify that apart from a red shift of the peak position due to temperature induced reduction in bandgap, the trion peak survives up to the highest temperature (423423 K) used in the experiment - suggesting highly stable trion on the junction. This is further supported by an enhancement in the trion dissociation energy (separation between the X0X^{0} and XX^{-} peaks) at higher temperature in Figure 1e, arising from enhanced doping of monolayer WS2 at higher temperature [25].

The dashed vertical line in Figure 1d indicates the spectral position of the 1.95911.9591 eV excitation. This suggests that using a fixed excitation at 1.95911.9591 eV, with a change in the sample temperature, we can perform a high resolution spectral scan around the X0X^{0} and XX^{-} overlap region. The transient response of the zero-bias IphI_{ph} under 1.95911.9591 eV excitation is shown at different temperatures in Figure 1f. Since Vd=0V_{d}=0, the dark current is ideally zero and practically only limited by the noise of the measurement setup. The vertical jumps in IphI_{ph} when excitation source is toggled between on and off states indicate that the device works as a fast photonic switch. The magnitude of IphI_{ph} exhibits a strong non-monotonic trend with temperature, peaking at T=343T=343 K, which corresponds to the resonant condition between the excitation energy and the XX^{-} peak. This strongly points to the fact that coherently excited trions participate in the detected photocurrent.

The temperature dependent non-monotonic photocurrent generation mechanism is explained in Figure 2. When the temperature is around 300300 K (top panel of Figure 2a), the bandgap of WS2 is higher, and the excitation is well below the X0X^{0} and XX^{-} position, thus the resulting photocurrent is weak. The source of this photocurrent from such non-resonant excitation is the photo-excited electrons from the top FLG being driven to the bottom MLG tunneling through the WS2 barrier by the built-in electric field, as schematically depicted in Figure 2b. Note that if the temperature changes the relative doping between the top and the bottom graphene, the resulting change in the built-in field would in turn cause a monotonic change in the photocurrent with an increase in temperature. The observed non-monotonicity in IphI_{ph} magnitude with temperature thus hints a separate mechanism other than just photoelectron tunneling must contribute to the non-monotonicity.

As the sample temperature reaches a value of around 343343 K, the linearly polarized photon creates trions in a coherent manner in KK and KK^{\prime} valleys, as explained in Figure 2c [9, 11]. Each of the trion consists of a bright exciton in one valley electrostatically bound with an electron in the lower spin-split conduction band from the opposite valley. Since inter-layer transfer is ultra-fast (\simsub-ps [27, 28, 12, 29]), which is faster than trion radiative decay (\simtens of ps [10, 19, 9, 2, 30, 31, 32]), the whole negatively charged trion can be transferred to the bottom monolayer graphene driven by the built-in electric field, as illustrated in Figure 2d. To account for charge neutrality, the top layer graphene injects an electron to the WS2, completing the circuit. Thus the trion state acts as an intermediate state to provide a favourable path for the flow of the charge current enhancing the photocurrent. Under resonance, this is the dominating photocurrent transport mechanism in the vertical heterojunction. We also note that after resonant excitation, even if the trion recombines radiatively before being transferred to the bottom graphene layer, the released electron in the conduction band of WS2 can still be driven to the bottom MLG through the built-in field to generate the photocurrent, as depicted in Figure 2e. However, noting that the inter-layer transfer process is faster than the radiative lifetime, the latter process of photocurrent is less dominant.

As temperature increases further, the resonance condition breaks and thus the photocurrent also lowers. Around 393393 K , the excitation comes in resonance with X0X^{0} (Figure 2f-g) peak, however the IphI_{ph} is still quite low. The net charge in the neutral exciton being zero, the exciton flow does not contribute to the charge current in spite of being resonantly created and transferred to the bottom graphene layer as evidenced from quenching of photoluminescence in several studies [33, 34, 35]. It is also possible that the resonantly created exciton can form trion with emission of phonon, which could eventually generate a charge current. However, the time it takes to form the trion through phonon emission is much longer than inter-layer exciton transfer, suppressing this process, and hence the photocurrent at exciton resonance is also suppressed compared to trion resonance. With further spectral de-tuning through increase in temperature (Figure 2h), IphI_{ph} is even more suppressed.

Figure 3a shows the transient response of IphI_{ph} with VgV_{g} varying from 40-40 to 4040 V, keeping the temperature fixed at 343343 K, suggesting a monotonic increment of IphI_{ph} with VgV_{g}. Figure 3b shows the strong tunability of IphI_{ph} with TT as well as VgV_{g} through a color plot. With an increase in VgV_{g} at the back gate, the WS2 film can be gated through the bottom MLG film due to its incomplete screening. The gate tunability of IphI_{ph} is maximum when the excitation is around the resonance with the XX^{-} peak, and tunability reduces on both sides. This is a further evidence regarding the strong role of the trion in the photocurrent generation mechanism. With larger positive VgV_{g}, formation of XX^{-} is favoured, which in turn increases the photocurrent. The point is further elaborated in Figure 3c-d, by taking horizontal slices from Figure 3b along Vg=40,0,V_{g}=40,0, and 40-40 V. In Figure 3d, the relative X0X^{0} (solid circles) and XX^{-} (open circles) peak shifts with respect to 1.95911.9591 eV are plotted, with the zero on the vertical axis indicating resonance conditions.

In order to further justify the point that the observed IphI_{ph} results from negatively charged trions, we correlate the photocurrent magnitude with the XX^{-} peak height obtained under 1.95911.9591 eV excitation. Figure 4a shows in a color plot the photoluminescence intensity around the XX^{-} spectral region as a function of VgV_{g}. The vertical axis shows the spectral position with respect to the excitation energy. The exact XX^{-} resonance condition is shown by the open circles (as obtained from 2.332.33 eV excitation), and cannot be reached during PL measurement with 1.95911.9591 eV excitation due to the cut-off of the edge filter (individual spectra are shown in Supplemental Material S2)[26]. Nonetheless, close to the trion resonance, we clearly notice a monotonic increase in the intensity with an increase in VgV_{g}. The individual spectra with 2.332.33 eV excitation at different gate voltages are depicted in Figure 4b, clearly indicating the increasing strength of trion with an increase in the gate voltage. In Figure 4c, we plot the normalized PL intensity as a function of gate voltage along the dashed line (T=343T=343 K) in Figure 4a. The normalization of the PL intensity is performed as PL(Vg)PLminPLmaxPLmin\frac{PL(V_{g})-PL_{min}}{PL_{max}-PL_{min}} where PLmin(max)PL_{min(max)} is the minimum (maximum) PL intensity with varying VgV_{g} at 343343 K. In the right axis of the same figure, we also plot the normalized IphI_{ph} [normalized as Iph(Vg)Iph,minIph,maxIph,min\frac{I_{ph}(V_{g})-I_{ph,min}}{I_{ph,max}-I_{ph,min}} where Iph,min(max)I_{ph,min(max)} is the minimum (maximum) IphI_{ph} with varying VgV_{g}]. The correlation between the two independent measurements is remarkable, unambiguously suggesting the origin of IphI_{ph} from transport of negatively charged trion in the vertical direction.

Finally, we comment on the estimated switching speed of the trion switch. Since trions are coherently generated through optical excitation, the primary step limiting the intrinsic speed of switching is the inter-layer transfer time of the trion. This inter-layer transfer timescale can be roughly estimated through the difference in homogeneous linewidth broadening of the trion emission between WS2 at the junction and from WS2 lying on SiO2. The total homogeneous linewidth of the trion emission can be estimated from the radiative recombination rate (Γr\Gamma_{r}), non-radiative scattering rate (Γnr\Gamma_{nr}) and inter-layer transfer rate (Γtr\Gamma_{tr}) by Γ=Γr+Γnr+Γtr\Gamma=\Gamma_{r}+\Gamma_{nr}+\Gamma_{tr}. The last term is present only for the heterojunction, and absent for a control WS2 sample placed on SiO2. Thus, we have,

τ=2Γtr=ΔΓ\tau=\frac{\hbar}{2\Gamma_{tr}}=\frac{\hbar}{\Delta\Gamma} (1)

where ΔΓ\Delta\Gamma is the difference in the Lorentzian component of the full-width-at-half-maximum of the XX^{-} peaks between the junction and the control sample, after fitting each of them using a Voigt function [18, 33]. We estimate a value of τ65\tau\approx 65 fs suggesting the ultra-fast nature of the trion switch.

In summary, we have demonstrated a gate- and light-controlled trion switch, where trion is optically initiated in a resonant manner and the read out is performed electrically. This can lead to a new paradigm of exciton based optoelectronic switches. The proposed technique exploits the vertical inter-layer transfer of excitonic species, which thus can be extremely fast, compared to relatively slow planar transport of heavy excitons. The efficient controllability through both electrical gating as well as photogating marks an important step towards the realization of trion-based transistor. Since trion can be valley polarized, the valley information can be optically initiated, followed by an electrical sensing, thus transferring the valley information into the electrical domain. The proposed technique can also be used for efficient generation and injection of spin current by resonantly exciting the trion using circularly polarized light - an important step toward realizing spintronic devices.

Methods

Device fabrication: To prepare the heterojunction of MLG/monolayer WS2/FLG we have used dry transfer technique on a highly doped Si substrate covered with 285285 nm thick thermally grown oxide layer. The different layers have been heated subsequently on a hot plate at 7070^{\circ} C for 2 minutes in order to get improved adhesion between layers. Device contacts are fabricated using standard nanofabrication methods. The substrate is spin coated with PMMA 950C3 and baked on a hot plate at 180180^{\circ} C for 2 minutes. This is followed by e-beam lithography with an acceleration voltage of 20 KV, an electron beam current of 220 pA, and an electron beam dose of 200 μ\muCcm-2. Patterns are developed using MIBK:IPA solution in the ratio 1:3. Later samples are washed with IPA and dried in N2 blow. Electrodes are then made by deposition of 10 nm Ni /50 nm Au films using DC magnetron sputtering at 3×1033\times 10^{-3} mBar and subsequent lift-off by dipping the substrate in acetone for 15 minutes, followed by washing in IPA and N2 drying. The oxide at the back side of the wafer is also etched by dilute HF solution.
Photocurrent measurement: Devices are kept on a Linkam thermal stage along with a heater underneath. The laser beam (2.33 eV or 1.9591 eV) is focussed through a 50X objective (NA of 0.5) to the heterostructure with a spot size of approximately 2 μ\mum. The devices are electrically probed using micro manipulators and Keithley 2636B is used as source meter. Temperature of the stage is increased from 295295 K to 423423 K in steps of 1010 K. A gate bias VgV_{g} is applied at the Si substrate ranging from 40-40 to +40+40 Volts in steps of 55 Volts. At each temperature and at each biasing point, photocurrent measurements are carried out and in situ photoluminescence spectra are obtained.

ACKNOWLEDGMENTS

This work was supported in part by a grant under Indian Space Research Organization (ISRO), by the grants under Ramanujan Fellowship, Early Career Award, and Nano Mission from the Department of Science and Technology (DST), and by a grant from MHRD, MeitY and DST Nano Mission through NNetRA.

Competing Interests

The Authors declare no Competing Financial or Non-Financial Interests.

References

  • [1] Mak, K. F. et al. Tightly bound trions in monolayer mos2. Nature materials 12, 207 (2013).
  • [2] Plechinger, G. et al. Trion fine structure and coupled spin–valley dynamics in monolayer tungsten disulfide. Nature communications 7, 12715 (2016).
  • [3] Jadczak, J. et al. Room temperature multi-phonon upconversion photoluminescence in monolayer semiconductor ws2. Nature communications 10, 107 (2019).
  • [4] Unuchek, D. et al. Room-temperature electrical control of exciton flux in a van der waals heterostructure. Nature 560, 340 (2018).
  • [5] Unuchek, D. et al. Valley-polarized exciton currents in a van der waals heterostructure. Nature nanotechnology 1–6 (2019).
  • [6] Singh, A. et al. Coherent electronic coupling in atomically thin mose 2. Physical review letters 112, 216804 (2014).
  • [7] Gao, F. et al. Valley trion dynamics in monolayer mose 2. Physical Review B 94, 245413 (2016).
  • [8] Singh, A. et al. Trion formation dynamics in monolayer transition metal dichalcogenides. Physical Review B 93, 041401 (2016).
  • [9] Hao, K. et al. Trion valley coherence in monolayer semiconductors. 2D Materials 4, 025105 (2017).
  • [10] Wang, G. et al. Valley dynamics probed through charged and neutral exciton emission in monolayer wse 2. Physical Review B 90, 075413 (2014).
  • [11] Hao, K. et al. Coherent and incoherent coupling dynamics between neutral and charged excitons in monolayer mose2. Nano letters 16, 5109–5113 (2016).
  • [12] Ceballos, F., Bellus, M. Z., Chiu, H.-Y. & Zhao, H. Ultrafast charge separation and indirect exciton formation in a mos2–mose2 van der waals heterostructure. ACS nano 8, 12717–12724 (2014).
  • [13] Sun, D. et al. Observation of rapid exciton–exciton annihilation in monolayer molybdenum disulfide. Nano letters 14, 5625–5629 (2014).
  • [14] Hao, K. et al. Direct measurement of exciton valley coherence in monolayer wse 2. Nature Physics 12, 677 (2016).
  • [15] Moody, G., Schaibley, J. & Xu, X. Exciton dynamics in monolayer transition metal dichalcogenides. JOSA B 33, C39–C49 (2016).
  • [16] Robert, C. et al. Exciton radiative lifetime in transition metal dichalcogenide monolayers. Physical Review B 93, 205423 (2016).
  • [17] Palummo, M., Bernardi, M. & Grossman, J. C. Exciton radiative lifetimes in two-dimensional transition metal dichalcogenides. Nano letters 15, 2794–2800 (2015).
  • [18] Gupta, G. & Majumdar, K. Fundamental exciton linewidth broadening in monolayer transition metal dichalcogenides. Physical Review B 99, 085412 (2019).
  • [19] Wang, H. et al. Radiative lifetimes of excitons and trions in monolayers of the metal dichalcogenide mos 2. Physical Review B 93, 045407 (2016).
  • [20] Das, S., Gupta, G. & Majumdar, K. Layer degree of freedom for excitons in transition metal dichalcogenides. Physical Review B 99, 165411 (2019).
  • [21] Miller, B. et al. Long-lived direct and indirect interlayer excitons in van der waals heterostructures. Nano letters 17, 5229–5237 (2017).
  • [22] Nagler, P. et al. Interlayer exciton dynamics in a dichalcogenide monolayer heterostructure. 2D Materials 4, 025112 (2017).
  • [23] Yu, H., Wang, Y., Tong, Q., Xu, X. & Yao, W. Anomalous light cones and valley optical selection rules of interlayer excitons in twisted heterobilayers. Physical review letters 115, 187002 (2015).
  • [24] Okada, M. et al. Direct and indirect interlayer excitons in a van der waals heterostructure of hbn/ws2/mos2/hbn. ACS nano 12, 2498–2505 (2018).
  • [25] Kallatt, S., Das, S., Chatterjee, S. & Majumdar, K. Interlayer charge transport controlled by exciton–trion coherent coupling. npj 2D Materials and Applications 3, 15 (2019).
  • [26] See supplemental material at [url will be inserted by publisher] for (1) temperature dependent PL spectra with 2.33 eV laser excitation, (2) back gate voltage dependent PL spectra with 1.9591 eV laser excitation at 343 K. .
  • [27] Wang, H. et al. The role of collective motion in the ultrafast charge transfer in van der waals heterostructures. Nature communications 7, 11504 (2016).
  • [28] Zheng, Q. et al. Phonon-coupled ultrafast interlayer charge oscillation at van der waals heterostructure interfaces. Physical Review B 97, 205417 (2018).
  • [29] Hong, X. et al. Ultrafast charge transfer in atomically thin mos 2/ws 2 heterostructures. Nature nanotechnology 9, 682 (2014).
  • [30] Godde, T. et al. Exciton and trion dynamics in atomically thin mose 2 and wse 2: Effect of localization. Physical Review B 94, 165301 (2016).
  • [31] Mouri, S., Miyauchi, Y. & Matsuda, K. Tunable photoluminescence of monolayer mos2 via chemical doping. Nano letters 13, 5944–5948 (2013).
  • [32] Lundt, N. et al. Valley polarized relaxation and upconversion luminescence from tamm-plasmon trion–polaritons with a mose2 monolayer. 2D Materials 4, 025096 (2017).
  • [33] Hill, H. M. et al. Exciton broadening in ws 2/graphene heterostructures. Physical Review B 96, 205401 (2017).
  • [34] Froehlicher, G., Lorchat, E. & Berciaud, S. Charge versus energy transfer in atomically thin graphene-transition metal dichalcogenide van der waals heterostructures. Physical Review X 8, 011007 (2018).
  • [35] Xie, L., Ling, X., Fang, Y., Zhang, J. & Liu, Z. Graphene as a substrate to suppress fluorescence in resonance raman spectroscopy. Journal of the American Chemical Society 131, 9890–9891 (2009).
Refer to caption
Figure 1: Photocurrent in FLG/monolayer WS2/MLG vertical heterojunction. (a) Schematic of the experimental device structure where 1L-WS2 is sandwiched between monolayer graphene (MLG) in the bottom and few-layer graphene (FLG) on the top. (b) The optical micrograph of the heterostructure where the MLG, 1L-WS2 and the FLG are marked with blue, yellow and black outline, respectively. (c) The transfer characteristics of the device under dark condition at different temperatures with Vd=20V_{d}=20 mV. Inset: The dark current-voltage characteristics at 293 K with Vg=0V_{g}=0 V. (d) The color plot of temperature dependent PL spectra with 2.332.33 eV laser excitation. The bandgap decreases monotonically with an increase in temperature. The vertical red dashed line corresponds to 1.95911.9591 eV excitation. (e) Variation of exciton (X0X^{0}) and trion (XX^{-}) peak position as a function of temperature showing enhanced separation (and hence higher trion stability) between the two at higher temperatures. (f) The transient photocurrent at zero external bias with 1.95911.9591 eV laser excitation at different temperatures starting from 303303 K to 423423 K in steps of 2020 K.
Refer to caption
Figure 2: Resonant trion mediated IphI_{ph} generation with 1.95911.9591 eV excitation. (a) Temperature dependent PL spectra of the heterojunction with 2.332.33 eV excitation. The XX^{-} and X0X^{0} energy states, fitted with two voigt curves (cyan and violet, respectively) are shown separately. The vertical red dashed line shows the spectral position of 1.95911.9591 eV excitation. The XX^{-} state comes in resonance with 1.95911.9591 eV excitation at 343343 K while for the X0X^{0}, it is around 393393 K (not shown in the figure). (b-h) Schematic of the photocurrent generation mechanism with 1.95911.9591 eV excitation at different temperatures. The off-resonance condition is shown in (b) and (h). (b) shows the situation at 303303 K with the excitation is in the sub-optical-bandgap range of 1L WS2 (at 303303 K, the optical bandgap is 2.0142.014 eV), while (h) shows the situation at 423423 K where the excitation is above the optical bandgap. The excitation is in resonance with the XX^{-} at 343343 K (c-e). (c) shows the coherent formation of bright inter-valley trion with linear polarization. The resonantly formed XX^{-} can either be transferred directly to the bottom MLG [shown in (d)] or it can recombine radiatively, releasing an electron, which in turn get transferred to the bottom MLG [shown in (e)]. Both of the processes generate photocurrent, however, the later process is of weaker efficiency. (f-g) show the X0X^{0} resonance condition at 393393 K, which does not contribute to the photocurrent generation due to charge neutral nature of X0X^{0}.
Refer to caption
Figure 3: Gate voltage modulation of generated photocurrent. (a) Modulation of transient photocurrent at 343343 K (trion resonance condition) with different gate voltages (VgV_{g}) ranging from 40-40 to +40+40 V, showing a monotonic increase in the photocurrent (IphI_{ph}). (b) The color plot of IphI_{ph} for the simultaneous modulation of temperature (TT) and VgV_{g}, which reveals stronger VgV_{g} modulation of IphI_{ph} under XX^{-} resonance. (c) Horizontal line cuts [shown in dashed lines in (b)] at three different VgV_{g} values (40-40, 0 and +40+40 V). (d) Relative separation of X0X^{0} (in solid symbols) and XX^{-} (in open symbols) from 1.95911.9591 eV excitation at different temperatures The resonance conditions with X0X^{0} and XX^{-} are indicated by the vertical dashed lines.
Refer to caption
Figure 4: Correlation of VgV_{g} modulation with XX^{-} intensity under resonance. (a) The color plot of the PL intensity obtained with 1.95911.9591 eV excitation as a function of VgV_{g} (in horizontal axis) and the relative emission energy positions of XX^{-} with respect to the excitation energy (in vertical axis). The green open symbols denote the position of the XX^{-} peak obtained from 2.332.33 eV excitation. (b) Normalized PL spectra with 2.332.33 eV excitation at 343343 K showing the XX^{-} and X0X^{0} energy states separately at five different VgV_{g} conditions. The red arrows indicate the position of the 1.95911.9591 eV excitation, with which photocurrent is measured. (c) Normalized IphI_{ph} (in brown symbols, left axis) and trion intensity along the horizontal dashed line in (a) (in orange symbols, right axis) for different VgV_{g} with 1.95911.9591 eV excitation at 343343 K, showing strong correlation between the two independent measurements.