Introduction

Atomically thin two-dimensional (2D) semiconductors beyond graphene have emerged as one of the most promising material candidate for next generation electronic devices because of their sizable bandgap1,2,3,4,5,6. Recently, van der Waals heterostructures composed of 2D transition metal dichalcogenides (TMD) semiconductor with a broad range of material diversity have gained tremendous research interest in various device applications, especially in steep-slope tunneling transistors due to their superior properties beyond traditional bulk materials limits1,2. On one hand, van der Waals interactions enable the possibility for diverse heterostructures of highly distinct materials without the constraints of lattice matching, which is different from the covalent bonding in traditional bulk materials1,2,3,4. On the other hand, the dangling-bond-free interfaces of van der Waals heterostructure could mitigate the parasitic trap-assisted tunneling induced by interface states in traditional III-V-based tunneling heterojunction1,2,3,4,5,6,7. Moreover, for 2D-baesd tunneling heterojunctions, the relative band alignment can be modulated through electrostatic gating due to the van der Waals gap between the neighboring layers, theoretically suggesting that the type-II band alignment in the off-state could be modulated to type-III band alignment in the on-state for high drive current8. Consequently, tunneling transistors based on van der Waals heterostructures are expected to realize low leakage current, ultra-steep slope and high on/off current ratio simultaneously, which has been confirmed by lots of theoretical works and shows the great potential in low power electronics9,10,11,12. However, only a few experimental works have been reported regarding the tunneling transistors based on van der Waals heterostructures. Black Phosphorus (BP)/MoS2 tunneling transistors have been fabricated, while its high tunnel barrier (Ebeff) of approximately 0.5 eV at the heterojunction requires relatively large voltage to modulate the band alignment, demonstrating the limited on/off current ratio (~104 at 3 V voltage range)13. Besides, BP material is unstable in the air, which would also cause the degradation of device performance14,15,16. Reported dual-gated tunneling transistors utilizing WSe2/MoS2 also exhibited the unsatisfactory on/off current ratio of 103 due to the large Ebeff8. As an improvement, WSe2/SnSe2 tunneling transistors with Ebeff lowered to 0.4 eV were reported and the on/off current ratio is remarkably enhanced17. However, SnSe2 is very unstable in the ambient environment, and can be easily oxidized18. Besides the Ebeff and stability, the metal-to-2D semiconductor contact resistances would also severely limit the performance of 2D-based transistors19. Therefore, in spite of the optimism created by theoretical works, experimental optimization and demonstration of 2D-2D tunneling transistors with both high on/off current ratio and high material stability are still in urgent need.

In this work, the stable WS2/SnS2 van der Waals heterostructure with theoretically 0.02 eV Ebeff is considered for the first time and selected as the optimal material platform for tunneling transistors. This optimal heterostructure is further experimentally demonstrated, and the WS2 and SnS2 serve as the p-type source layer and the n-type channel and drain layer, respectively. The key challenge of metal-to-2D semiconductor contact is further optimized for integration. Based on the physical insight into the metal/2D interfaces, work-function- and thickness- engineering are conducted for p-type WS2 and n-type SnS2 respectively to reduce the contact resistances. Based on a novel dry transfer technique for vertical heterostructure stack, the bottom-gated WS2/SnS2 tunneling transistor is fabricated and shows the on/off current ratio exceeding 106, which is among the highest in the reported tunneling transistors. Compared with the conventional FET, comparable on-state current and steeper subthreshold slope (SS) are also obtained. The tunnel behaviors are further confirmed by low temperature measurements, and a visible negative differential resistance feature is observed. This work shows the great potential of van der Waals heterostructure for tunneling devices and future-generation energy-efficient electronics.

Results and Discussion

Figure 1a shows the schematic structure of bottom-gated vertical tunneling transistors based on the van der Waals heterostructure in this work. The SiO2 and highly n-doped Si are used as the gate dielectric and bottom gate, respectively. The channel layer is designed to be under the source layer, and the electric potential and carrier concentration of the channel layer are modulated by the bottom gate. The band alignment of the van der Waals heterostructure is designed to be type-II, in which the conduction band of the channel layer is above the valence band of the source layer. Figure 1b illustrates the operation mechanism of this n-type tunneling transistor. In the off-state, electrons in the valence band of the source layer cannot tunnel into the conduction band of the channel layer, since there is no tunneling window. As the bottom gate bias increases, the conduction band energy (EC) of the channel layer begins to be lower than the valence band energy (EV) of the source layer, and the tunnel current across the source layer/channel layer heterostructure will increase accordingly, exhibiting n-type characteristics. The band alignment can be tuned from type-II towards type-III due to the van der Waals gap, which would enhance the on/off current ratio of the transistor.

Figure 1
figure 1

(a) Schematic view of the bottom-gated vertical tunneling transistor based on 2D semiconductors. Vertical tunneling occurs across the overlap region between p-type layer and n-type layer. (b) Band alignment in the off-state (left). There is no tunneling window between the valence band of the source layer and the conduction band of the channel layer. Band alignment in the on-state (right). The electric potential and carrier concentration of the channel layer are modulated by the bottom gate, and the tunneling window exists in which electrons can tunnel from the valence band of the source layer to the conduction band of the channel layer. (c) Band diagram of the WS2/SnS2 heterostructure with Ebeff of 0.02 eV. (d) The optical microscope image of the fabricated n-type WS2/SnS2 tunneling transistor. (e) High-resolution STEM image and (f) EDS mapping of the WS2/SnS2 heterostructure. A clean and sharp interface is obtained. (g) Raman characterization of WS2 and SnS2 sheets used for tunneling transistors.

Taking into consideration of both device performance and material stability, the optimal heterostructure for tunneling transistor is selected by the following principles. First, from the perspective of on/off current ratio, the tunnel barrier Ebeff should be considerably reduced so that low voltage is required to modulate the band alignment from type-II to type-III17. Second, tunneling electrons with smaller effective mass is beneficial for the higher tunnel efficiency and the higher tunnel current20. Third, according to our previous work, the lower density-of-state (DOS) of the channel layer would result in the better output characteristics of tunneling transistors with the smaller onset voltage and the better current saturation behavior21. At last, the stability of building materials to form heterostructures in ambient environment should also be taken into account for better device stability and reliability. Based on the above design rules, and according to the band structures of various 2D semiconductors from ab initio calculations22,23, p-WS2/n-SnS2 heterostructure stands out as the superior material platform for tunneling transistor. The Ebeff in this near broken-gap heterostructure can be lowered to 0.02 eV theoretically, as shown in Fig. 1c. Meanwhile, the effective mass of electron in WS2 is relatively small compared to other metal dichalgonedies, and the low DOS of the channel layer SnS2 can improve the output characteristics of tunneling transistors in the meantime24,25. More importantly, WS2 and SnS2 is proved to possess great stability in ambient air26,27.

Figure 1d shows the optical image of a fabricated WS2/SnS2 tunneling transistor in this work. The bottom-gate dielectric of 300 nm-thick SiO2 was firstly grown on the highly n-doped Si substrate. Then, SnS2 and WS2 sheets were vertically stacked to form the van der Waals heterostructure, and after that, source and drain contacts were sequentially defined by electron beam lithography, electron beam evaporation and lift-off process.

For the heterostructure preparation, the SnS2 crystals were then grown by chemical vapor transport method and mechanically exfoliated onto the SiO2/Si substrate. The WS2 sheet was stacked upon SnS2 using a novel dry transfer process to avoid liquid contamination (see the dry transfer process in Supplementary Fig. S1), and the heterostructure was formed via van der Waals interaction. Figure 1e shows the high-resolution scanning TEM (STEM) image of the WS2/SnS2 heterostructure, confirming the clean interface obtained by the dry transfer process. The measured interlayer distance of WS2 is 0.66 nm and that of SnS2 is 0.6 nm, which agree well with values reported in refs 28,29,30. Energy-dispersive X-ray spectroscopy (EDS) composition analysis exhibits the sharp boundary between W and Sn, as shown in Fig. 1f. Raman spectra of these two sheets are shown in Fig. 1g and Raman peaks of both materials are clearly identified. The thicknesses of WS2 and SnS2 in this device are 23 nm and 4 nm, respectively. The characteristic Raman peaks of WS2 and SnS2 can be distinctly observed. WS2 shows the 2LA peak at 349.9 cm−1, E2g peak at 355.8 cm−1, and A1g peak at 420 cm−1. The A1g peak of SnS2 is observed at 313.4 cm−1.

In order to reduce the contact resistances for better device performance, contact-engineering is conducted for p-type WS2 and n-type SnS2, separately. To date, contacts of metal-to-2D semiconductors in the majority of reported metal dichalcogenide transistors are Schottky contacts instead of Ohmic contacts due to the difficulty of heavy doping in thin 2D semiconductors, which is also the case in WS2 and SnS228,31,32,33. In principle, low work functions (Wm) of contact metals lead to the small Schottky barrier (SB) height for electrons, and high work function metals are beneficial for low hole barriers. In order to investigate the SB height at metal-to-2D semiconductor contacts, bottom-gated SnS2 and WS2 SB-FETs were fabricated and characterized firstly as shown in Fig. 2a. For SnS2 SB-FETs, Ti (Wm = 4.33 eV) and Sc (Wm = 3.5 eV) were adopted for realizing n-type contacts due to their low Wm34. Both Ti- and Sc-contacted SnS2 SB-FETs exhibit n-type transistor behaviors (Fig. 2b). Yet compared with Ti, Sc would be easily oxidized in the air, resulting in severe degradation of transfer characteristics for Sc-contacted SB-FETs over time (Fig. 2c and Supplementary Fig. S2). Therefore, Ti with the better stability is chosen as the electrode for contacts to SnS2 in the n-type tunneling transistors in this work. Figure 2d shows the measured output characteristics of Ti-contacted SB-FETs, and the linear dependence of current on drain voltage further confirms the low contact resistance at Ti/SnS2 interfaces33. The extracted Schottky barrier height is as low as 0.181 eV (Fig. 2e), and the detailed extraction method can be seen in Supplementary Fig. S3. Addtionally, no significant hysteresis is observed in the transfer characteristics of SnS2 FET with Ti contacts (Fig. 2f).

Figure 2
figure 2

Optimization of n-type contact for SnS2. (a) Schematic view of bottom-gated SnS2 or WS2 SB-FETs. (b) Measured transfer characteristics of n-type SnS2 SB-FETs using Ti and Sc as contacts (VBG is the voltage applied to the highly n-doped Si substrate). Inset is the optical microscope image of SnS2 SB-FET. (c) Transfer characteristics of SnS2 SB-FET with Ti contacts after a period of time, (d) Measured output characteristics of n-type SnS2 SB-FETs using Ti contacts. Linear dependence of ID on VDS indicates low contact resistance. (e) Extracted SB height for the Ti-to-SnS2 contact. (f) Measured hysteresis characteristics of the SnS2 SB-FET with Ti contacts, showing no significant hysteresis.

Apart from n-type contacts to SnS2, the resistance of p-type contacts to WS2 was also investigated. High work function metals, Pd (Wm = 5.12 eV) and Pt (Wm = 5.65 eV), were chosen as the metal electrodes to realize low-resistance p-type contacts34. Figure 3a shows the measured transfer characteristics of WS2 SB-FETs. Although there is a significant difference (0.53 eV) of work functions between Pd and Pt, hole current in the Pd-contacted SB-FET is comparable to that in Pt-contacted device, indicating the strong Fermi-level pinning at the metal/WS2 interface. In order to suppress Fermi-level pinning at the metal-to-2D material interface, inserting a thin layer of substoichiometric molybdenum trioxide (MoOx) between metal and 2D materials has been verified as an effective way to facilitate hole injection, which can be attributed to the high work function of MoOx and its excellent interface properties with 2D materials31. However, the work function of MoOx is very sensitive to ambient gas exposure, and thus the device performance will degrade in the air. Therefore, a novel approach is proposed in this work to reduce the SB height for holes at metal-to-WS2 contacts. As we know, the bandgaps of 2D semiconductors are thickness-sensitive due to the influence of quantum confinement effect23. According to results from ab initio calculations, as the thickness of WS2 layers increases, the valence band moves upward while the position of conduction band remains nearly unchanged23. As a consequence, the SB height for holes can be decreased with thicker WS2. In this work, SB-FETs with different numbers of WS2 layers were fabricated and Fig. 3b shows the corresponding hole currents. As the number of WS2 layers increases from 3 to 12, hole currents are increased by three decades. When the thickness of WS2 layers is increased to 23 nm, WS2 p-type SB-FET exhibits high ION exceeding 0.2 μA μm−1 (Fig. 3c), suggesting low resistance of fabricated p-type contacts. Figure 3d shows the atomic force microscope (AFM) image of the 23 nm-thick WS2 SB-FET, and the height difference between the WS2 sheet and the substrate.

Figure 3
figure 3

Optimization of p-type contact for WS2. (a) Measured transfer characteristics of p-type WS2 SB-FETs using Pd and Pt contacts (VBG,T = VBG − Vonset(Pt), Vonset(Pt): the onset voltage at which hole current begins to increase in the Pt-contacted SB-FET). Similar currents are observed in these two devices. (b) Dependence of on-state current density on WS2 thickness in WS2 SB-FETs. The on-state hole current increases with the number of WS2 layers. (c) Measured transfer characteristics of SB-FET with 23 nm-thick WS2. Large hole current in the on-state indicates the low resistance p-type contact. (d) AFM height profile of the 32-layer WS2 sheet. Inset is the corresponding AFM image of fabricated devices.

Based on the optimized n-type and p-type contacts, the tunneling transistors based on WS2/SnS2 van der Waals heterostructures are experimentally demonstrated with Pt as the source contact and Ti as the drain contact, and electrical characterization at different temperatures and under different bias conditions are performed. The thicknesses of the WS2 flakes are designed according to the above contact optimization, and are measured to be 23 nm by AFM. In contrast, the contact resistance of n-type metal-to-SnS2 is thicknesses insensitive due to the nearly unchanged position of conduction band when layer number increases, and the SnS2 used in tunneling transistor is measured to be 4 nm. Figure 4a shows the measured typical transfer characteristics of the n-type WS2/SnS2 tunneling transistor at room temperature. The results are obtained by applying the bias on SnS2 contact, with WS2 contact grounded. With the optimized design of metal-to-2D semiconductor contacts, the n-type WS2/SnS2 tunneling transistor exhibits high ION of 3.7 μA. The on/off current ratio of the fabricated device is over 106 and the on-state current density is 186 nA μm−2. The high on/off current ratio, and high on-state current which is comparable with the value obtained in the SnS2 SB-FET, further confirm the optimized band alignment in WS2/SnS2 tunneling transistors. Compared with conventional SB-FET in Fig. 2b, the subthreshold slope of the fabricated WS2/SnS2 device is much steeper and also increases with gate voltage which is a typical feature of tunnel transistors. Since the transistor is fabricated based on bottom-gated structure with 300 nm-thick SiO2, the value of SS is relatively large, and can be further optimized by reducing the gate oxide thickness or incorporating with high-κ dielectrics. In order to further verify the BTBT mechanism of this n-type transistor, the dependence of transfer characteristics on temperature is studied. As shown in Fig. 4b, ION shows the positive dependence on temperature and SS changes little with temperature, exhibiting the typical features of BTBT operation mechanism35,36,37,38. The weak dependence of SS on temperature also indicates the suppression of trap-assisted tunneling, which benefits from the clean interface obtained from the dry transfer process.

Figure 4
figure 4

Electrical characteristics of the WS2/SnS2 tunneling transistor. (a) Measured transfer characteristics of the bottom-gated WS2/SnS2 tunneling transistor. The on-state current (ION) is 3.7 μA with an area of 20 μm2 and on/off current ratio is over 106. (b) Measured temperature characteristics of the bottom-gated WS2/SnS2 tunneling transistor.

In order to further validate the BTBT mechanism in this WS2/SnS2 heterostructure, the output characteristics in forward bias region are investigated. Figure 5a shows the output characteristics at 100 K, which are measured by applying the voltage on WS2 contact, with SnS2 contact grounded. The distinct negative differential resistance (NDR) is observed and confirms that a heavily-doped p-n junction is formed at the interface due to charge transfer17. The band alignments of the WS2/SnS2 herterostructure under different bias conditions are illustrated in Fig. 5b–e. In the equilibrium state, the Fermi level in p-WS2 and n-SnS2 is aligned, as shown in Fig. 5b. As the source-drain voltage VSD increases, the energy band of WS2 is pulled down and a finite tunneling window is created for electrons in the conduction band of SnS2 to tunnel into the empty states in the valence band of WS2. The tunnel current reaches its peak when the Fermi level of WS2 aligns with the conduction band minimum of SnS2, as shown in Fig. 5c. With VSD further increasing, the tunneling window is gradually switched off and the reduction of tunnel current leads to NDR (Fig. 5d). After that, thermal injection current begins to dominate the total current due to the reduced thermal barrier, and increases with VSD (Fig. 5e). It should be noted that the peak voltage can be modulated by the bottom-gate voltage VBG in this device. The conduction band of SnS2 varies with VBG, and the peak voltage needed to align the Fermi level of WS2 with the conduction band of SnS2 changes consequently.

Figure 5
figure 5

Negative differential resistance (NDR) feature of the WS2/SnS2 tunneling transistor. (a) Measured output characteristics of the bottom-gated WS2/SnS2 tunneling transistor in the forward bias region. Clear NDR is observed for positive gate voltages. (b) Band alignment at VSD = 0 V. Fermi level in WS2 and SnS2 is aligned. (c) Band alignment at VSD > 0 V. The band of p-type WS2 is pulled down and a finite tunneling window is created for electrons in the conduction band of SnS2 to tunnel into the empty states in the valence band of WS2. The tunnel current reaches its peak when the Fermi level of WS2 aligns with the conduction band minimum of SnS2. With further increasing VSD (d), the tunnel window is gradually switched off and the reduction of tunnel current leads to NDR. When VSD continues to increase (e), the thermal injection current dominates and increases with the reduced energy barrier.

Conclusions

In conclusion, the optimal WS2/SnS2 van der Waals heterostructure for tunneling transistors is presented and elaborately engineered, taking into consideration both electric properties and material stability. Besides, the key challenge of metal-to-2D semiconductor contact is optimized to achieve low-resistance n-type and p-type contacts for SnS2 and WS2, respectively. Ti contact with low work function and superior stability can induce small Schottky barrier height for electrons at metal-to-SnS2 contacts. Low-resistance p-type contacts are obtained at the metal/WS2 interface through the thickness optimization of WS2. With the optimized metal-to-2D semiconductor contacts and a proposed new dry transfer technique for vertical heterostructure stack, the fabricated n-type WS2/SnS2 tunneling transistor exhibits superior performance with the high on/off current ratio over 106, as well as comparable on-state current and steeper subthreshold slope compared with conventional FET, showing the great potential of van der Waals heterostructure for future energy-efficient devices.

Methods

Device Fabrication

The highly n-type Si substrate with 300 nm thermal silicon oxide is prepared as the bottom gate structure. The starting materials used for the fabrication of n-type WS2/SnS2 tunneling transistors were high-quality bulk crystals of WS2 and SnS2. The process flow of the bottom-gated WS2/SnS2 tunneling transistor has been described in the Supplementary Fig. S1. In details, 10 nm Ti/20 nm Au was deposited to form the Ti-to-SnS2 contact, and the Pt-to-WS2 contact was generated with 20 nm Pt/40 nm Au.

Physical Characterization

AFM and Raman spectra were used to characterize the thicknesses of WS2 and SnS2. Raman spectra were excited by 514 nm laser with the spot diameter about 1 μm. The laser power was kept less than 0.1 mW to avoid sample heating and oxidation in the air. Structural characterization by scanning TEM (STEM) was performed in JEM-ARM200F with the acceleration voltage of 100 keV. The STEM sample was prepared by focused ion beam (FIB) using the gallium beam.

Electrical Measurements

N-type WS2/SnS2 tunneling transistors were electrically characterized in the vacuum chamber using the Agilent B1500A semiconductor parameter analyzer.