With the continuous scaling down of electronics and the urgent need to handle data explosion, the internet of things (IoTs) has risen in use and popularity in the past decade. The IoTs comprise widely distributed devices with multifunctionality that are interconnected, providing real-time communication of the scope of what we want to know and what we want to do.[1–3] The new era of the IoTs not only highlights a productive new avenue for a whole generation of information devices but also provides an opportunity to surpass the limits of Moore's law. A range of novel 2D materials has been actively explored for More Moore and More-than-Moore applications because of their unique structures and optoelectronic properties.[4–6] By vertically or/and laterally assembling the 2D materials, its integration of artificial van der Waals (vdW) heterostructures can be further used to create diverse devices including tunneling transistors,[7,8] atomically p-n junctions,[9–12] light-emitting diodes,[13] reconfigurable field-effect transistors,[14,15] nonvolatile memories,[16,17] and even multifunctional devices.[18–24] These atomically vdW heterostructures-based systems, possessing ultra-low-power, low-loss, and energy-efficient active features, fully satisfy the requirement of the fast-growing IoTs paradigm.
The assembled vdW heterostructures, which do not suffer from impurities or interfacial defects, have high-quality interfaces and are highly suited to use in state-of-the-art electronics with desired multi-functionalities.[25,26] For example, in 2015, a band-to-band tunneling transistor in a MoS2/WSe2 vdW heterostructure was made using a dual-gate device architecture.[18] Under different gate-voltage (Vg) modulations, the vdW system behaved as either an Esaki diode, a backward diode, or a forward rectifying diode, making the device highly versatile. In 2017, a fine band-structure alignment using narrow-bandgap black phosphorus (BP) and large-bandgap MoS2 stacking was achieved, in which a large current rectification ratio of 106 and a high on–off current ratio up to 107 were realized by integrating both the high-performance diode and transistor in a single device.[19] A ternary inverter can be further achieved by matching the partition load and band structure in the vertical heterostructure. In the same year, because the charge-carrier injection can be switched between tunneling and thermal activation under different bias polarities, a multifunctional device, including a diode, transistor, photodetector, and multi-state memory, was reported in an asymmetric MoTe2/MoS2-based vdW heterostructure.[20] A similar asymmetric configuration combined with a floating-gate charge layer was also been adopted in WSe2/hexagonal boron nitride/graphene heterostructures for multi-purpose optoelectronic applications.[22] In addition, in BP/ReSe2 heterostructures, the tunability of diverse current-transport characteristics was demonstrated using the variation of the BP layer thickness.[27] To date, most of the above-related approaches for preparing multifunctional devices have been limited by either complexity in fabrication technology and/or inflexibility in reversible operation, which significantly limit the possibility of multifunctional integrations.
Precisely controlling the conduction behavior, such as conductive polarities or doping degrees, is the key to achieving practical applications for vdW heterostructures.[23,28–31] In general, after choosing specified 2D materials, the vdW devices will exhibit unchanged conduction behaviors, because the conduction behaviors as well as band structure alignments, directly depend on the nature and corresponding thickness of the selected 2D materials. A convertible conducting channel in a single device based on a vdW system, which allows for higher integration and compatibility, is still lacking. Herein, we propose a new and simple strategy for dynamically controlling and optimizing the electronic properties of vdW heterostructures-based electronics, and then realize their multifunctional applications in a “single device.” MoTe2 has been proved to be ultrasensitive to O2.[32–36] Charge carriers of MoTe2 flakes could be easily modulated, which is primarily attributed to the amount of O2 adsorptions on the MoTe2 surface, particularly at its tellurium defects. Such the adsorption process is reversible when MoTe2 flakes are stored in a reversed-O2 condition. Therefore, in this study, through the rapid thermal annealing (RTA) strategy between dry-air (80% N2, and 20% O2, 1 atm) and vacuum conditions, a tunable charge-polarity device based on MoTe2/SnS2 vdW heterostructures were fabricated. Compared with MoTe2, the electrical properties in SnS2 with a suitable thickness exhibit relative stability in different O2 concentrations. The simple RTA strategy in different conditions enables precise tuning of the doping level and effective modulation of the transport behavior in the MoTe2 flakes in a large dynamic range, thus allowing the MoTe2/SnS2 vdW heterostructure to be reversibly switched between unipolar, ambipolar, and anti-ambipolar-dominated transistors. Finally, the flexible multioperation modes of logic-circuit applications, for example, inverter, p-n diode, output polarity controllable (OPC) amplifiers, and ternary-state NMIN and NMAX logics can be precisely engineered. Our proposed strategy for multioperation-mode transistors is highly promising in the prospective of future electronics, as well as the growing IoTs generation.
Bulk 2H-phase MoTe2 is an indirect bang gap material, with a band gap in the range of 0.6–1 eV.[37] At the beginning of research on layered MoTe2 transistors, ambipolar behavior in the charge polarity has been revealed.[38] Subsequently, it has been demonstrated that oxygen present in the environment can significantly alter the electrical properties of layered MoTe2 devices. Such related research can be traced back to 2015.[32,33] The presence of tellurium defects and the functionalization of these defect sites with O2 molecules strongly dictate the electronic structure and create deep or shallow states with the optical band gap. Moreover, by varying the vacuum level during the RTA,[34,36,39,40] electrothermal doping process,[41] and scanning visible-light method,[42,43] the transfer characteristics of layered MoTe2 transistors can distinctly shift to show different polarities between unipolar-like and symmetric ambipolar behaviors owing to the physical or chemical adsorption and dissociation of O2 associated with the energy level of these strategies.
Based on the O2-ultrasensitive MoTe2 surface, Figure 1a shows a schematic diagram of reversible charge-polarity control for a multioperation-mode vdW heterostructure, in which two layered materials, MoTe2 and an O2-insensitive n-type flake, were selected to integrate with a vertical stack to form a vdW transistor. First, we considered that in ambient conditions (i.e., relatively high O2 concentration), the O2 vapor naturally adsorbs on the MoTe2 surface, in particular, at the tellurium defects, where charge carriers for MoTe2 would be trapped, and then its polarity becomes hole dominated. An effective p-type doing of MoTe2 was hence developed, as shown in the left-hand panel of Figure 1ai. However, the adsorbed O2 with low adsorption energy on the MoTe2 surface would potentially be removed when the MoTe2-based vdW heterostructure was put in vacuum (i.e., relatively low O2 concentration) condition for a period of time or a given additional energy. Such desorption of O2 could release charge carriers again back to the MoTe2, leading to n-type doping of the vdW heterostructure, as shown in the right-hand panel of Figure 1ai. This process of O2 adsorption/desorption has been proven to be controlled and reversible, freely creating p/n charge polarity in MoTe2. A typical variation of transfer characteristics of a MoTe2 channel is illustrated in Figure 1aii for different degrees of O2 adsorption. Schematic diagrams of the transfer curve shift for the MoTe2-based vdW heterostructure with increasing adsorbed O2 concentration, are shown in Figure 1b for i) unipolar, ii) ambipolar, and iii) anti-ambipolar-dominated charge-polarity transistors. Thus, through O2 adsorption-evacuation cycles, the inflexibility in charge polarity for typical vdW heterostructures can be broken, achieving multioperation-mode transistors.
Figure 1. a) Simplified principle of reversible charge-polarity control for a multioperation-mode transistor. i) Schematic illustration of oxygen adsorption and desorption processes on the surface of a MoTe2-based vdW heterostructure. Owing to the presence of Te vacancies on exfoliated MoTe2 surfaces, the electrical properties of the MoTe2-based vdW heterostructure were highly sensitive to different O2 concentration environments. ii) Corresponding transfer characteristics of a MoTe2 channel with different degrees of oxygen adsorption. b) Schematic diagrams of the transfer-curve shift for a MoTe2-based vdW heterostructure. Under different degrees of O2 adsorption, the vdW heterostructure shows i) unipolar, ii) ambipolar, and iii) anti-ambipolar charge polarity.
To realize the concept of reversible charge-polarity control for multioperation-mode transistors based on vdW heterostructures outlined in Figure 1, vertically stacked MoTe2/SnS2 heterostructure transistors were fabricated. A schematic diagram of the structural configuration of a MoTe2/SnS2 heterostructure with the circuit layout is shown in Figure 2a. To enhance the adsorbed/removed O2 effect on the MoTe2 surface, a few-layered MoTe2 flake was intentionally stacked on top of the SnS2 flake. Ti/Au films were used as metallic electrodes to optimize the electrical properties and adhesion. The electrode contact on SnS2 was grounded, while a source–drain voltage (Vd) was applied on the MoTe2. The inset of Figure 2a shows an optical image of the as-fabricated MoTe2/SnS2 vdW heterostructure transistor. Atomic force microscopy (AFM) was performed to measure the thickness of the MoTe2 and SnS2 flakes. The corresponding height profiles (Figure 2b) show that the thicknesses of the MoTe2 and SnS2 flakes were ≈8 and 30 nm, respectively. An AFM image of the vdW heterostructure is shown in Figure S1, Supporting Information. Raman scattering measurement was then carried out on different channel areas of the as-fabricated transistor and the results are shown in Figure 2c. The peak of A1g at 172 cm−1, E12g at 232 cm−1, and B12g at 291 cm−1 correspond to MoTe2 (red line), while the Raman spectrum of the SnS2 shows an A1g peak at 314 cm−1 (blue line). All of these peak positions are consistent with previous reports.[44] In the heterostructure region (black line), the peaks show contributions from both materials, indicating the presence of two distinct materials. Figure 2d further shows transfer characteristics (Id−Vg) of the as-fabricated MoTe2 and SnS2 transistors at Vd = 1 V before any annealing treatment. The transfer in the as-fabricated MoTe2 is dominated by n- and p-type carriers under positive and negative gate biases, respectively, resulting in ambipolar behavior (red curve), while the as-fabricated SnS2 shows n-type transport throughout the entire scan from −60 to 60 V (blue curve). Unlike other anti-ambipolar heterostructures reported in previous work,[45] our as-fabricated MoTe2/SnS2 heterostructure exhibits electron-dominated ambipolar transfer characteristics at different Vd values before annealing, as shown in Figure 2e, primarily because the hole doping of the as-fabricated MoTe2 is far from its balanced level. The corresponding output characteristic of the heterostructure has also been monitored to display no rectification, indicating the formation of good contacts at the semiconductor-metal interfaces (not shown here).
Figure 2. a) Schematic diagram of a van der Waals stacked MoTe2/SnS2 heterostructure transistor on a SiO2/Si substrate. The inset shows an optical image of the as-fabricated transistor, where a few-layer MoTe2 flake was exfoliated on top of an SnS2 flake. The electrode contact on SnS2 was grounded while a source–drain voltage was applied to the MoTe2 contact. The scale bar represents 3 µm. The labels for each electrode are marked in the optical image. b) Height profiles showing the thickness of the MoTe2 (top) and SnS2 (bottom) flakes. c) Raman spectra collected from the different regions corresponding to MoTe2, SnS2, and the heterostructure. d) Transfer characteristic curves of the as-fabricated MoTe2 and SnS2 transistors at Vd = 1 V before annealing treatment. e) Transfer characteristic curves of the as-fabricated MoTe2/SnS2 heterostructure at different Vd values before annealing treatment.
It has been demonstrated that the binding energy arising from the interaction between O2 (and/or H2O) and the layered MoTe2 surface is only in the range of 30–60 meV,[32] which is comparable to room-temperature energy and strongly implies that adsorbed molecules could be desorbed by suitable thermal energy. Hence, we deliberately adopted the RTA process to accelerate the O2 reaction under dry-air and vacuum conditions, then developed a versatile O2 doping approach that enables highly effective and nonvolatile band modulation of the MoTe2-based vdW heterostructures. Doping was done by applying the RTA process on the vdW transistor at a fixed temperature for 1 min under dry-air conditions. The vdW transistor was then put into the probe station and a transfer curve was recorded to characterize its transport behavior. The transfer characteristics for different annealing temperatures from 150 to 200 °C are shown in Figures 3a and 3b for the MoTe2 and SnS2 transistors, respectively. The solid and dashed lines are semilog and linear plots, respectively. It is noticed that no meaningful hysteresis loop was observed in the transfer curves, either activation condition exists or not. The magnitude of the change was qualitatively proportional to the annealing temperature. As shown in Figure 3a, the MoTe2 transistor transforms from having electron-dominated ambipolar behavior before annealing to a highly doped unipolar p-type transistor as the annealing temperature increases from 150 to 200 °C. In contrast, the SnS2 transistors underwent the same RTA treatment and showed very little response despite a gradual reduction in on-state current (Id) from 5.2 to 4.2 µA at Vg = 60 V (see Figure 3b). Such a relatively weak response to the RTA process mainly results from a lower surface-to-volume ratio of the thicker SnS2 channels because the molecule adsorption/desorption is a surface effect. The thickness evidence is shown in Figure S2, Supporting Information. It was found that the Id values at a fixed Vg condition for a thin SnS2 transistor significantly dropped by two orders of magnitude after the RTA process, as shown in Figure S2a,b, Supporting Information, which provides direct evidence to support our expectation. Therefore, a thicker SnS2 flake was selected as an O2-insensitive n-type material to give reversible charge-polarity vdW heterostructures. Here, to maintain a suitable current on–off modulation, the range of the SnS2 thickness was typically chosen from 20 to 50 nm for vdW p-n stackings.
Figure 3. Transfer characteristic curves for the a) MoTe2 and b) SnS2 transistors on semilog (solid line) and linear (dashed line) scales. c,d) Transfer characteristic curves of the MoTe2/SnS2 vdW heterostructure on semilog and linear scales, respectively. The curves in different colors correspond to different annealing temperatures from 150 to 200 °C under dry-air conditions during the RTA process. The height of maximum Id values is marked as Ipeak in (d). e) Threshold voltage (Vth) of the MoTe2 transistor (red) and the voltage position of the Ipeak of the MoTe2/SnS2 heterostructure (black) as a function of annealing temperature in dry-air conditions. The inset shows the Vth of the SnS2 transistor as a function of the annealing temperature. f) Transfer characteristic curves of the MoTe2, SnS2, and MoTe2/SnS2 vdW heterostructure after 200 °C RTA treatment. All the conditions were measured at Vd = 1 V.
Figures 3c and 3d show the transfer characteristics of the MoTe2/SnS2 vdW heterostructure with different annealing temperatures from 150 to 200 °C under a dry-air condition in semilog and linear scales, respectively. The vdW transistor evolved from having electron-dominated ambipolar behavior before annealing (see Figure 2e) to highly balanced ambipolarity (gray line in Figure 3c), and then to anti-ambipolar charge transport (red line in Figure 3c) as the annealing temperature increased from 150 to 200 °C. In the linear plot of the transfer curves, the height of maximum Id values is marked as Ipeak (see Figure 3d). As the annealing temperature increases, the Ipeak gradually increases, bringing a more obvious anti-ambipolar charge transport, as well as p-type doping in the MoTe2. In addition, the positions of the corresponding Ipeak (i.e., Vpeak) and threshold voltage (Vth) for the MoTe2 and SnS2 transistors are displayed as a function of annealing temperature in Figure 3e. The Vpeak and the Vth of MoTe2 increased linearly with the annealing temperature from 150 to 200 °C, whereas the SnS2 transistor maintained a Vth of around −45 V (inset), regardless of thermal treatment, suggesting again that the charge-polarity change of the vdW transistor is attributed to the doped MoTe2. Through the RTA doping treatment, we were able to reconfigure the MoTe2 transistor to the desired properties for a suitable device application. The outcome in Figure 3f verifies the clear formation of the anti-ambipolar vdW heterostructure at the junction of two opposite polarity transistors under 200 °C RTA treatment, in which the off-state voltages around −60 and 36 V are displayed. It is emphasized that the maximum of applied Vd to the vdW heterostructure is not higher than 2 V, which makes no meaningful impact on Joule heating to alter the charge polarity. In short, using the simple RTA strategy, we can precisely control the transition from the electron-dominated ambipolar to anti-ambipolar states and accurately modulate both the Ipeak height, as well as its position, which is critical for building more functional blocks using the vdW heterostructures.
To visualize the polarity change process during the RTA treatment, the conventional equation μn,p = (L/W)CggmVd was used, where μe,h is the field-effect mobility for electrons and holes, L/W is the channel length-to-width ratio, Cg = 1.15 × 10−8 F cm−2 is the gate-oxide capacitance per unit area for a 300-nm SiO2 dielectric and is transconductance. The MoTe2 mobility for holes was found to increase by one order of magnitude from 0.4 to 4.3 cm2 V−1 s−1 at Vd = 1 V at the highest annealing temperature (i.e., 200 °C), while that for electrons decreased from 16.9 to 0.8 cm2 V−1 s−1 at Vd = 1 V, as depicted in Figure 4a. Further taking the relation of n, p = (Id · (L/W))/(qVdμn,p) into consideration, where q is the elementary charge, the corresponding carrier concentration (p) for holes increased by 1.4 times from 3.5 × 1012 to 4.3 × 1012 cm−2, while that (n) for electrons decreased by 1.2 times from 6 × 1012 down to 4.7 × 1012 cm−2, which provides consistent evidence for the adsorption of O2 molecules on the MoTe2 surface resulting in the down-shifted Fermi level (EF) in MoTe2 (see Figure 4b). In contrast, both the mobility and carrier concentration for SnS2 remained constant to be 1.1 cm2 V−1 s−1 and 6.6 × 1013 cm−2, respectively, as a function of annealing temperature (Figure 4c), again confirming the O2-insensitive properties of SnS2 at suitable flake thickness. In this work, we monitored more than ten devices treated using the same RTA strategy. There were unavoidable variations, such as doping mobility and carrier concentration for individual material or Ipeak (Vpeak) position and on-state current for anti-ambipolar transfer characteristics due to differences in channel quality, geometry, and minute process-to-process deviations, but all devices behaved the same qualitatively in terms of reversibility between the ambipolar and anti-ambipolar states and tunability of the peak position under the RTA treatment (see Figure S3, Supporting Information). In addition, the RTA treatment doping was bidirectional and reversible when the MoTe2/SnS2 vdW heterostructure was exposed cyclically between the dry-air and vacuum conditions to allow O2 adsorption/desorption on the MoTe2 surface, as shown in Figure 4d. The corresponding reversibility for the MoTe2 and SnS2 transistors is shown in Figure S4, Supporting Information. Notice that the transfer curve of the vdW heterostructure can be back to the as-fabricated case, like the result presented in Figure 1e, if the RTA temperature applied was higher than 200 °C and the device could withstand an extended treatment time. However, this introduced a greater possibility of damage to the 2D materials, which is not within the scope of this work and is not discussed here. Since the RTA improvement of contact resistance at the metal-semiconductor interface[46] or the formation of oxidation[47,48] is irreversible, the above description in electrical properties strongly supports again that the doping/de-doping reversibility is due to the adsorption/desorption processes of O2 molecules onto the surface of MoTe2. Other supplementary are provided in Figures S5 and S6, Supporting Information. Let us consider the energy-band diagrams to explain the detail of the cycle (i.e., dry-air and vacuum) of the RTA strategy for the MoTe2/SnS2 vdW heterostructure, as illustrated in Figure 4e. The adsorbed/desorbed O2 molecules act as carrier acceptors/donors to donate additional holes/electrons to the MoTe2 flakes, which results in its downward/upward EF shift, as well as the polarity change of the vdW heterostructure. Additionally, notice that MoS2, the most used 2D material, is an intrinsically n-type semiconducting channel because of the existence of sulfur vacancies and is relatively O2-stable. It is widely believed that the charge polarity of MoS2 cannot be easily changed, compared with ambipolar-type MoTe2. Therefore, the MoS2 potential is not considered in this work.
Figure 4. a) Mobilities and b) carrier concentrations for holes and electrons in the MoTe2 transistors as a function of annealing temperature in dry-air conditions. c) Mobilities and carrier concentrations for electrons in the SnS2 transistors as a function of annealing temperature in dry-air conditions. d) Reversibility of the RTA process, including cyclic exposure of the MoTe2/SnS2 vdW heterostructure to dry air and vacuum. Transfer properties of the vdW heterostructure before the RTA process (first step, solid black line). Transfer properties of the vdW heterostructure processed by the RTA process in dry air (second step, solid red line). Transfer features of the vdW heterostructure after the RTA process in a vacuum (third step, dashed red line) and dry air (fourth step, dashed black line). e) Corresponding band diagrams of the MoTe2/SnS2 vdW heterostructure during the cycle (i.e., dry air and vacuum) of the RTA process.
Integration of an O2-ultrasensitive MoTe2 material and an O2-insensitive SnS2 material to form a vdW heterostructure allows us to modulate and control the device under different degrees of oxygen adsorption. As proof of concept to multifunctionality in a single device, several operation modes of logic-circuit applications were designed and demonstrated according to unipolar-, ambipolar-, or anti-ambipolar-mode MoTe2/SnS2 vdW heterostructure transistors, as illustrated in Figure 5. First, an inverter function, which is one of the typical binary-states circuits and can be normally realized in as-fabricated MoTe2/SnS2 vdW transistors, was achieved using a unipolar vdW transistor connected to a load resistor. Figure 5a shows the transfer characteristics and the voltage gains (−dVout/dVin) of the inverter at different Vdd values, where Vout, Vin, and Vdd denote the output, input, and supply voltage, respectively. With a low Vin (i.e., logic 0), a high Vout (i.e., logic 1) is achieved, and vice versa. For the ambipolar-mode feature, an OPC amplifier based on a single active device is presented. A sinusoidal signal Vin was applied to the gate electrode, while the output signal was detected using an oscilloscope. When the ambipolar MoTe2/SnS2 vdW transistor exhibited n-type behavior, for larger/smaller Vin, more conductive/insulating states, as well as smaller/larger Vout, can be achieved, which is called the out-of-phase or common-drain mode (see Figure 5bi). In contrast, when the vdW transistor worked as a p-type channel, Vout increased synchronously with the increase of Vin, which is called the in-phase or common-source mode for the OPC amplifier, as shown in Figure 5bii. The demonstration of both modes in the vdW transistors suggests that more complicated circuits could be developed such as phase-shift key. The schemes of an inverter and OPC amplifier circuits equipped with an off-chip resistor are shown in Figure S7, Supporting Information. The MoTe2/SnS2 vdW heterostructure transistor treated with the highest annealing treatment was further used to demonstrate the potential of the anti-ambipolar mode feature. It is easy to understand as the vdW heterostructure is formed by a balanced p- and n-transistor in series, in which the total current reaches a maximum when both transistors are turned on, and then the anti-ambipolar behavior occurs. Therefore, a simple application using the anti-ambipolar mode transistor is a rectifying p-n diode, as shown in Figure 5ci. The inset displays the feature of half-wave characteristics by applying sinusoidal Vdd at 5 Hz. Finally, we demonstrated a ternary-states inverter, which is a basic block in multivalued logic applications. Figure 5cii shows the equivalent circuit of the vdW heterostructure-based ternary inverter, which uses the MoTe2 transistor as a load resistor and the vdW heterostructure as a driver. The inset shows the corresponding optical microscopy (OM) image. Figure 5ciii plots the Vout variation and the voltage gains of the ternary inverter as a function of Vin at different Vdd values. When Vin was low, the vdW heterostructure region was fully turned off, but the MoTe2 transistor provided a small resistance path between Vdd and Vout, leading to the output of logic “1.” When a large Vin was applied, the load resistor (i.e., MoTe2 transistor) was turned off, but the heterostructure reached a high conductance, and thus, the device outputs a logic “0.” At a moderate Vin, the resistances of the load resistor and the driver were almost equal, which resulted in the output values of the logic state “1/2.” It is also observed that the width of the “1/2” state plateau can be controlled by applying Vdd values. The “1/2” logic state and its gain monotonically increase with Vdd, which is attributed to how obvious is the anti-ambipolar “Λ” shape in the transfer characteristics.[19] To demonstrate the applicability of the ternary-state devices for complex logic operations, NMIN and NMAX gates are further simulated. The logic circuit diagrams and the corresponding truth tables of the NMIN and NMAX are illustrated in Figure 5d, which are similar to AND and OR operations in the binary system, respectively. From Figure 5e, the transient responses of NMIN and NMAX are shown under variation of applied input voltages. The versatility described definitely demonstrates the great potential of our proposed RTA strategy for future device applications based on 2D vdW heterostructures.
Figure 5. a) Transfer characteristics and voltage gains for an inverter operation, integrated by a unipolar MoTe2/SnS2 vdW heterostructure connected to a load resistor (1 GΩ). b) Oscillating signals of i) the common-drain and ii) the common-source mode measured at the output of an ambipolar MoTe2/SnS2 vdW heterostructure connected to a load resistor. ci) Output characteristics of diode operation at Vg = −60 V in an anti-ambipolar MoTe2/SnS2 vdW heterostructure. The inset shows the features of a half-wave rectifier. ii) Equivalent circuit for the ternary inverter and the corresponding OM image based on an anti-ambipolar MoTe2/SnS2 vdW heterostructure connected to a MoTe2 transistor. iii) Transfer characteristics and voltage gains for the ternary inverter operation. d) Equivalent circuits and truth tables of NMIN and NMAX logic gates. e) Output voltages of NMIN and NMAX logic gates under applied input voltages (VA and VB).
In summary, we used a simple and effective RTA doping/de-doping strategy under dry-air and vacuum conditions, to realize a tunable and reversible charge-polarity device based on MoTe2/SnS2 vdW heterostructures from unipolar, ambipolar to anti-ambipolar behaviors. The core of the doping/de-doping reversibility is responsible for the adsorption/desorption processes of O2 molecules onto the surface of MoTe2. Using the MoTe2/SnS2 vdW heterostructures, flexible multioperation functions, such as inverter, OPC amplifier, p-n diode, and ternary-state logic (NMIN and NMAX gates) were achieved, which provides a new doping strategy for layered electronics before integrated-circuit packaging. Our work definitely demonstrates the potential for use in next-generation electronics and will transcend the limits of Moore's law.
Experimental Section Device FabricationMoTe2/SnS2 vdW heterostructure transistors were fabricated by the dry-transfer method. First, thin flakes of SnS2 were mechanically exfoliated from bulk SnS2 crystals onto a 300 nm thick SiO2 dielectric/Si substrate, while MoTe2 flakes were exfoliated from the bulk form onto a PDMS film (Gel-Pak, PF-40-X4). The PDMS film attaching the exfoliated MoTe2 flakes was positioned on the arm of a micromanipulator and precisely transferred onto the top of the selected SnS2 flake. Subsequently, several pairs of contact electrodes were defined via conventional electron-beam lithography. Ti/Au (20/60 nm) metal layers were then deposited at a base pressure of 2 × 10−6 torr, followed by a lift-off approach in acetone to complete the device fabrication.
CharacterizationThe morphology and thickness of the MoTe2/SnS2 vdW heterostructure were investigated using OM and AFM. Raman spectra were collected on a LabRam HR-800 Raman spectrometer (Jobin Yvon), with a laser wavelength of 532 nm. Electrical measurements were performed under vacuum (<10−5 torr) in a cryogenic probe station (Lakeshore TTPX) with an Agilent B1500A semiconductor parameter analyzer.
AcknowledgementsC.-F.C. and S.-H.Y. contributed equally to this work. This work was supported by the Taiwan Ministry of Science and Technology (Grant Nos. MOST 109-2112-M-005-013-MY3 and 110-2119-M-007-003-MBK). C.F.C. and S.-T.L. acknowledge additional support from the Young Scholar Fellowship Program by the Ministry of Science and Technology (Grant No. MOST 110-2636-M-009-009) and the Center for Emergent Functional Matter Science of National Yang Ming Chiao Tung University from The Featured Areas Research Center Program within the framework of the Higher Education Sprout Project by the Ministry of Education (MOE) in Taiwan. W.-W.W. acknowledges the support from the “Center for Semiconductor Technology Research of National Yang Ming Chiao Tung University” from The Featured Areas Research Center Program within the framework of the Higher Education Sprout Project by the MOE in Taiwan. This work was also supported in part by the Taiwan Ministry of Science and Technology (Grant No. MOST 110-2634-F-009-027), the Natural Science Foundation of Shanghai (Grant No. 19ZR1473400), the NSAF Foundation of China (Grant No. U1830130), and the Young Scientist Project of MOE Innovation Platform, China.
Conflict of InterestThe authors declare no conflict of interest.
Data Availability StatementResearch data are not shared.
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
© 2022. This work is published under http://creativecommons.org/licenses/by/4.0/ (the “License”). Notwithstanding the ProQuest Terms and Conditions, you may use this content in accordance with the terms of the License.
Abstract
Van der Waals (vdW) heterostructures—in which layered materials are purposely selected to assemble with each other—allow unusual properties and different phenomena to be combined and multifunctional electronics to be created, opening a new chapter for the spread of internet‐of‐things applications. Here, an O2‐ultrasensitive MoTe2 material and an O2‐insensitive SnS2 material are integrated to form a vdW heterostructure, allowing the realization of charge‐polarity control for multioperation‐mode transistors through a simple and effective rapid thermal annealing strategy under dry‐air and vacuum conditions. The charge‐polarity control (i.e., doping and de‐doping processes), which arises owing to the interaction between O2 adsorption/desorption and tellurium defects at the MoTe2 surface, means that the MoTe2/SnS2 heterostructure transistors can reversibly change between unipolar, ambipolar, and anti‐ambipolar transfer characteristics. Based on the dynamic control of the charge‐polarity properties, an inverter, output polarity controllable amplifier, p‐n diode, and ternary‐state logics (NMIN and NMAX gates) are demonstrated, which inspire the development of reversibly multifunctional devices and indicates the potential of 2D materials.
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
Details

1 Department of Electrophysics and Center for Emergent Functional Matter Science (CEFMS), National Yang Ming Chiao Tung University, Hsinchu, Taiwan; Department of Physics, National Chung Hsing University, Taichung, Taiwan
2 Department of Physics, National Chung Hsing University, Taichung, Taiwan; International Collaborative Laboratory of 2D Materials for Optoelectronics Science and Technology (Ministry of Education), Engineering Technology Research Center for 2D Material Information Functional Devices and Systems (Guangdong Province), Institute of Microscale Optoelectronics, Shenzhen University, Shenzhen, China
3 Department of Physics, National Chung Hsing University, Taichung, Taiwan
4 Department of Physics, National Chung Hsing University, Taichung, Taiwan; Department of Materials Science and Engineering, National Yang Ming Chiao Tung University, Hsinchu, Taiwan
5 Department of Physics, National Chung Hsing University, Taichung, Taiwan; Institute of Electronics Engineering, National Tsing Hua University, Hsinchu, Taiwan
6 Institute of Electronics Engineering, National Tsing Hua University, Hsinchu, Taiwan
7 Department of Chemistry, Graduate School of Science and Engineering, Saitama University, Saitama, Japan
8 International Collaborative Laboratory of 2D Materials for Optoelectronics Science and Technology (Ministry of Education), Engineering Technology Research Center for 2D Material Information Functional Devices and Systems (Guangdong Province), Institute of Microscale Optoelectronics, Shenzhen University, Shenzhen, China
9 Department of Materials Science and Engineering, National Yang Ming Chiao Tung University, Hsinchu, Taiwan; Center for the Intelligent Semiconductor Nano‐system Technology Research, National Yang Ming Chiao Tung University, Hsinchu, Taiwan
10 Shanghai Frontiers Science Research Base of Intelligent Optoelectronics and Perception, Zhangjiang Fudan International Innovation Center, Institute of Optoelectronics, Department of Materials Science, Fudan University, Shanghai, China
11 Department of Electrophysics and Center for Emergent Functional Matter Science (CEFMS), National Yang Ming Chiao Tung University, Hsinchu, Taiwan
12 Department of Physics, National Chung Hsing University, Taichung, Taiwan; Department of Materials Science and Engineering, Institute of Nanoscience, i‐Center for Advanced Science and Technology (i‐CAST), National Chung Hsing University, Taichung, Taiwan