Composition-modulated anti-ambipolar behavior enabled by two-dimensional GeSxSe1−x/SnS2 van der Waals heterostructures for high-performance logic inverters

Yanhong Long a, Qunrui Deng a, Shengdi Chen a, Yingbo He a, Yue Wang a, Zhaoqiang Zheng c, Nengjie Huo a, Dongxiang Luo d, Xiao Liu a, Yiming Sun a, Zuxin Chen a, Mengmeng Yang a, Tao Zheng *b and Wei Gao *a
aGuangdong Provincial Key Laboratory of Chip and Integration Technology, School of Electronic Science and Engineering (School of Microelectronics), Faculty of Engineering, South China Normal University, Foshan 528200, P. R. China. E-mail: gaowei317040@m.scnu.edu.cn
bFuture Institute of Technology, Guangdong Provincial Key Laboratory of Nanophotonic Functional Materials and Devices, Guangdong Basic Research Center of Excellence for Structure and Fundamental Interactions of Matter, School of Optoelectronic Science and Engineering, South China Normal University, Guangzhou 510631, P. R. China. E-mail: taozheng@m.scnu.edu.cn
cSchool of Materials and Energy, Guangdong University of Technology, Guangzhou 510006, P. R. China
dHuangpu Hydrogen Innovation Center/Guangzhou Key Laboratory for Clean Energy and Materials, School of Chemistry and Chemical Engineering, Guangzhou University, Guangzhou 510006, PR China

Received 19th July 2025 , Accepted 24th October 2025

First published on 27th October 2025


Abstract

Logic inverters, which lay the foundation for the functionality of large-scale integrated circuits, are achieved using anti-ambipolar transistors (AATs) based on two-dimensional (2D) van der Waals heterojunctions (vdWH). However, the impact of the doping strategy on the figures of merit of logic inverters based on 2D vdWH AATs has not been comprehensively analyzed. Herein, 2D free-standing GeSxSe1−x (0 ≤ x ≤ 0.73) with precisely tunable composition was grown to fabricate GeSxSe1−x/SnS2 vdWH AATs to achieve an optimal logic inverter. By leveraging elemental modulation in GeSxSe1−x, the proposed vdWH was tuned from type-II to type-III band alignment, allowing for a distinctive tunneling process at various bias conditions. The proposed devices with poor S content exhibited a better peak-to-valley ratio of 6.6 × 103 at x = 0.29 and a maximum peak current of 1.4 × 10−7 A at x = 0. Furthermore, the inverter built with the GeSxSe1−x/SnS2 device achieved the highest voltage gain of 8.83 at x = 0.29, while the device with an S-rich AAT delivered a low static power of 12.1 pW, which is attributed to the optimization of band engineering and the low driving voltage under the bottom h-BN/Au structure. This work contributes insights into the expansion of alloy engineering in the construction of high-performance multi-valued logic inverters.



New concepts

Systematic studies show that the energy band alignment of GeSxSe1−x/SnS2 (0≤ x ≤ 0.73) van der Waals heterostructures (vdWHs) can be continuously tuned from type II to type III by varying the composition, which markedly modulates the tunneling window and surface-potential difference. Replacing the conventional SiO2/Si substrate with an h-BN/Au single bottom-gate architecture further boosts the electrical modulation capability, yielding a higher peak-to-valley ratio (PVR) under a lower driving gate voltage (ΔVg). By leveraging band alignment and alloy engineering, composition-modulated anti-ambipolar transport is achieved. The PVR, ΔVg and peak current (Ipeak) could be precisely adjusted from 8.7 × 101 to 6.6 × 103, 15.3 V to 5.8 V and 8.7 × 10−11 A to 1.4 × 10−7 A, respectively. Configurable logic inverters built using these composition-dependent AATs delivered a record voltage gain of 8.83 at x = 0.29 and an ultralow static power consumption of 12.1 pW at x = 0.73, offering a promising route for the rational design of low-power, high-gain multi-valued logic inverters based on 2D vdWHs.

The most significant advantage of logic inverters lies in the use of complementary metal–oxide semiconductors (CMOS) as fundamental device units, which promote efficient digital signal transmission, reduce static power, and facilitate the development of high-performance very-large-scale integrated circuits (VLSICs). Conventional CMOS technology based on complementary p-type and n-type unipolar transistors typically involves extrinsic and complicated doping processes, such as ion implantation and thermal diffusion, to achieve selective doping with shallow donor and acceptor energy levels.1,2 Such doping processes can induce lattice damage within the Si-based semiconductor, necessitating a time-consuming thermal annealing step for recovery. Furthermore, considering the growing demand for VLSICs, the inevitable shrinkage of device dimensions to the nanometer scale leads to a severe short-channel effect,3 which in turn degrades their performance and diminishes the feasibility and controllability of traditional CMOS technology. The advent of atomically thin two-dimensional (2D) materials has enabled the realization of anti-ambipolar transistors (AATs) based on 2D van der Waals heterojunctions (vdWH) either through negative differential resistance (NDR)4–6 or negative differential transconductance (NDT).7–9 For example, AAT devices with NDR (defined as Rdiff = dID/dVD) behavior based on 2D Esaki diodes6,10 and tunneling diodes5,11 were expected to play a pivotal role in multi-valued logic inverters. However, the development of a practical AAT device with NDR behavior has remained elusive, which is primarily ascribed to the operational requirement of cryogenic temperatures. Most of the AAT devices also deliver a relatively low peak-to-valley ratio (PVR). In contrast, AAT devices with NDT (gm = dIds/dVg) behavior typically possess a planar three-terminal structure. They rely on the band alignment at the p–n junction heterointerface and the moderate doping level in the individual channel.1 These devices exhibit Λ-shaped transfer curves at room temperature and enable high PVR larger than 103, which are extremely desirable for practical applications in high-gain logic operations.12 Concurrently, a high peak current (Ipeak) can enhance the signal-to-noise ratio and diminish the interference of external noise on the signal amplitude, which is particularly crucial during low-voltage operation. Therefore, achieving precise control over band alignment and doping level, as well as obtaining a deeper understanding of the composition-mediated modulation of the IdsVds characteristics and anti-ambipolar properties, are highly important for the rational design and practical operation of 2D vdWH-based AATs.

Commonly, the anti-ambipolar properties of AATs based on 2D vdWH can be improved effectively by applying an electrostatic field,9,13–15 strain field,16,17 or a light field.18–20 For example, Miao et al.13 achieved dynamic modulation of conduction polarity and NDT in an MoTe2/BP vdWH device by tuning the type-I and type-II band alignment under an external electric field. Apart from the conventional SiO2/Si bottom gate structure, a HfO2 dielectric layer has been utilized to construct a ternary logic inverter based on a 2D BP/ReS2 vdWH device at lower driving voltages (ΔVg) to realize multifunctional nonvolatile logic-in-memory applications.21 In addition, strain control in an InSe/WSe2 vdWH-based flexible device allowed the fabrication of bi-anti-ambipolar transistors with two tunable threshold voltages (Vth) as a result of the optimization of the conduction pathway, leading to the achievement of multi-value logic inverters.17 Moreover, photo-induced doping in an anti-ambipolar MoTe2/BP vdWH device enabled dynamic adjustment of the peak position of the drain-source current (Ipeak).20 In spite of the advancements and modulation of the anti-ambipolar properties, the above-mentioned strategies ignore the key mechanism in leveraging band alignment for composition modulation to achieve desired IdsVds characteristics and anti-ambipolar properties.

Alloy engineering, an effective strategy for tailoring 2D materials with specific band structures, has demonstrated the potential to modulate band alignment, electronic properties, and transport behavior in 2D vdWHs.22 For instance, through composition modulation of the work functions at the 2D metallic alloy contacts, the contact potential difference in VS2xSe2(1−x)/MoS2 vdWH could be tuned from −71.5 mV to 0 mV to 59.3 mV, that is, from Schottky to ohmic contacts by varying x from 0 to 1.23 Likewise, by introducing controlled amounts of V atoms as dopants, continuous electrical polarity modulation of monolayer WS2 from intrinsic n-type to ambipolar, then to p-type, and ultimately, to a quasi-metallic state was achieved, with obviously regulated current on/off ratios, allowing the construction of homogeneous CMOS inverters based on monolayer WS2 with different doping concentrations.24 Clearly, continuous and precise modulation of band alignment in 2D vdWHs is sought for the construction of multifunctional electronics. Systematically investigating the impact of composition modulation in 2D vdWH-based AATs on the main figures of merit of logic inverters holds great importance25 but remains a daunting challenge.

In this work, 2D free-standing GeSxSe1−x (0 ≤ x ≤ 0.73) nanoplates, with composition-tunable band structures, were grown by the low-pressure rapid physical vapor deposition (LPRPVD) method reported in our previous work.26 Compared to other 2D materials, these GeSxSe1−x nanosheets offer polymer-free transfer processes and continuously tunable bandgaps, enabling precise control of the electrical properties of AATs. These features overcome the limitations of customizing the electrical parameters of 2D AATs and present a new research direction in the material engineering of ambipolar transistors. Therefore, we further developed a GeSxSe1−x/SnS2 vertical vdWH with a h-BN/Au single bottom gate configuration for AATs. The principle of variation in band alignment from type II to type III and the related tailored anti-ambipolar transport mechanisms were elucidated by modulating the S content and optimizing the dielectric architecture. Particularly, compared with the SiO2/Si bottom dielectric structure, the h-BN/Au structure exhibited a markedly reduced ΔVg and a significantly improved PVR. Thus, as the S composition varied from x = 0 to x = 0.73, the proposed AATs with S-poor compositions (x = 0.29) demonstrated an outstanding PVR and ΔVg of 6.6 × 103 and 5.8 V at Vds = 3 V, significantly outperforming those with S-rich compositions (x = 0.42, and 0.73). Meanwhile, the Ipeak was regulated monotonously from 1.4 × 10−7 A to 8.7 × 10−11 A. Most significantly, the verification of the binary inverters not only highlights the capability of alloy engineering in achieving composition-modulated anti-ambipolar behavior, but also unfolded impressive figures of merit, including a maximum voltage gain of 8.83 at x = 0.29 and a static power of 12.1 pW at x = 0.73. This achievement marks a substantial advancement in the rational design, modulation, and fabrication of high-performance 2D vdWH logic inverters.

Characterization of the GeSxSe1−x and GeSxSe1−x/SnS2 vdWH devices

In the experimental procedure, 2D GeSxSe1−x nanosheets with different S compositions were grown using a LPRPVD technique.26 The nanosheets were characterized by energy-dispersive X-ray spectroscopy (EDS) to identify their atomic stoichiometric ratios, as depicted in Fig. S1 (SI): GeSe (x = 0), GeS0.29Se0.71 (x = 0.29), GeS0.42Se0.58 (x = 0.42), and GeS0.73Se0.27 (x = 0.73), which confirm that the S composition in the alloys could be accurately customized. Fig. 1a shows a schematic of the growth of the nanosheets. Taking x = 0.42 as an example, the 2D GeSxSe1−x nanosheets exhibited free-standing growth on a SiO2/Si substrate, which can be flexibly transferred to other substrates after mechanical pressing using a polydimethylsiloxane (PDMS) film. The optical microscopic (OM) images of the horizontally distributed GeSxSe1−x (0 ≤ x ≤ 0.73) nanosheets on a SiO2/Si substrate are shown in Fig. S2 (SI). The compositional changes in GeSxSe1−x nanosheets were observed from the X-ray diffraction (XRD) patterns, as shown in Fig. 1b. All GeSxSe1−x (0 ≤ x ≤ 0.73) nanosheets exhibited the characteristic peak of (040), which obviously shifted towards large diffraction angles as the S content increased. According to the Bragg equation,27
 
2d[thin space (1/6-em)]sin(θ) = (1)
the interlayer spacing (d) continuously decreases with the substitution of the larger Se atoms by the smaller S atoms, resulting in the shifting of the (040) characteristic peak from 33.7° (x = 0) to 34.3° (x = 0.73).28 The trend of the phonon vibration modes of the 2D GeSxSe1−x nanoplates with varying S composition can be observed from the normalized Raman spectra shown in Fig. 1c. At x = 0, the typical Raman modes A1g (80 cm−1), B3g (149 cm−1), and A3g (185 cm−1) were present. Those three Raman modes underwent a blue-shift as the S composition increased, and the A1g mode disappeared at x = 0.73. Furthermore, another typical Raman mode A2g (264 cm−1) appeared at x = 0.42 and red-shifted to 236 cm−1 as the S composition increased to 0.73. Fig. 1d depicts the evolution of the energy band structure of the GeSxSe1−x and SnS229 monolayers with respect to the vacuum level, as calculated in our previous work26 using the Perdew–Burke–Ernzerhof function. These theoretical results confirm that the bandgap (Eg) of GeSxSe1−x increases with the increment of S composition, while its conduction band minimum (CBM) and valence band maximum (VBM) gradually shift downward. The tunneling window is defined as the energy level difference between the VBM of GeSxSe1−x and the CBM of SnS2, and the calculated results are shown in Fig. 1e. When 0 ≤ x ≤ 0.42, the tunneling window keeps narrowing from 633.9 meV (x = 0) to 195.3 meV (x = 0.42) with increasing S composition and disappears when the S composition increases to 0.73. This demonstrates that the theoretical energy band alignment of GeSxSe1−x/SnS2 vdWH can be switched between type II and type III by adjusting the S composition.

image file: d5nh00508f-f1.tif
Fig. 1 Characterization of the GeSxSe1−x nanoplates and the energy band alignment of the GeSxSe1−x/SnS2 vdWH. (a) Schematic depicting the growth of the GeSxSe1−x nanosheet. (b) XRD patterns and (c) Normalized Raman spectra of the GeSxSe1−x nanosheets with different S compositions. (d) Theoretical band alignment, (e) tunneling window, (f) SPD and (g) height profile data of the GeSxSe1−x/SnS2 vdWH with different S compositions.

Guided by the theoretical results, experiments were carried out to further investigate the electrical properties, including IdsVds and IdsVg curves. First, fresh GeSxSe1−x/SnS2 vdWH devices were fabricated on SiO2/Si substrate with four different S compositions (0, 0.29, 0.42, and 0.73) and characterized. The OM images, atomic force microscopy (AFM) images, Raman spectra, and energy band alignments before and after contact are shown in Fig. S3–S6, respectively (SI). The atomic surfaces of the GeSxSe1−x, SnS2, and overlapped regions were clean and smooth, as observed by OM and AFM. All the prominent Raman peaks of the individual materials appeared in the overlapping area with stronger intensities, indicating that these vdWHs maintained good crystalline quality after dry transfer and the device manufacturing process. The surface potential difference (SPD) across the heterointerface was evaluated by Kelvin probe force microscopy (KPFM) according to the formula shown in Note S1 (SI). As seen in Fig. 1f, the SPD of the GeSxSe1−x/SnS2 vdWHs increased as x increased from 0 to 0.42, with values of 177 mV (x = 0), 249 mV (x = 0.29), and 376 mV (x = 0.42). However, when the S composition increased to 0.73, the SPD of the vdWH decreased to 142 mV, which is related to the transition of its band alignment type. In particular, regardless of the composition of the S dopant, the surface potential of GeSxSe1−x was always higher than that of SnS2, which means that the direction of the built-in electric field is from GeSxSe1−x toward SnS2. Moreover, the GeSe/SnS2 vdWH is used as an example to demonstrate the diffusion and drift process of the charge carriers before and after contact, as shown in Fig. S3d.

Before contact, this vdWH exhibits a type-III energy band alignment, with a Fermi energy level (Ef) difference of 177 meV. After contact, the higher Ef of GeSe causes the electrons of GeSe to diffuse to the SnS2 side and the holes of SnS2 to transfer to the GeSe side until thermal equilibrium. This contributes to the upward band bending of GeSe and the downward band bending of SnS2 at the interface edge. The built-in electric field, with a potential value of 177 mV, points from GeSe to the SnS2 side. The band alignments of the other vdWHs (x = 0.29, 0.42, and 0.73) before and after contact are shown in Fig. S3d, S4d and S5d, respectively (SI). According to the theoretical analyses and KPFM results, the ΔEf between GeSxSe1−x and SnS2 first increases gradually with the increase in S composition, so that the energy band edge bending progresses steeply until the band alignment transforms from type III to type II at x = 0.73. As depicted in Fig. S7, the thicknesses of the h-BN nanosheets used in all the devices were between 55 and 65 nm, which can be easily obtained by mechanical exfoliation and is completely sufficient for a dielectric layer. As shown in Fig. 1g, the thicknesses of GeSxSe1−x and SnS2 were both maintained in the range of 30–50 nm. This ensures that the thickness of these multilayered samples has a negligible influence on the energy band alignment and device performance, thereby precluding the effects of thickness engineering.

GeSe/SnS2 vdWH FETs on different dielectric gate structures

GeS1−xSex/SnS2 FETs with a h-BN/Au single bottom gate configuration were designed specifically for high-performance logic inverter applications, as depicted in Fig. 2a. Specifically, multilayered SnS2 nanosheets were obtained using the mechanical exfoliation technique together with a PDMS-assist transfer method, while the GeS1−xSex nanoplates were directly transferred through a facile polymer-free pressing transfer method, benefiting from the free-standing growth mode.26 To achieve optimal electrical performance, a h-BN/Au bottom gate structure was designed, in which the h-BN served as the 2D dielectric layer (εr = 3.530). h-BN has an atomically smooth surface without dangling bonds and can form a van der Waals interface with 2D GeSxSe1−x and SnS2, which leads to a low defect density.31 Additionally, h-BN possesses a higher dielectric strength (∼0.8 V nm−1) than SiO2 (∼0.3 V nm−1), allowing for the design of thinner dielectric layers and increased surface capacitance (Cox = εrε0/d), thereby enhancing gate control capability.32 Moreover, it can be more easily fabricated than the HfO2 insulator layer,33 thereby facilitating electrostatic modulation with low static power dissipation for enhanced gain.20Fig. 2b displays the OM image of the typical GeSe/SnS2 vdWH device. Notably, the chosen SnS2 and GeSe nanosheets were successively stacked onto the pre-patterned source (S) and drain (D) electrodes, respectively, to circumvent performance degradation induced by organic solvents and high-energy deposition, which are prone to causing the Fermi pinning effect.34 In addition, the single bottom Cr/Au gate electrode was completely below the whole channel of the device, enabling effective modulation of the carrier concentrations in different doping regions. Then, the electrical transport properties of the GeSe/SnS2 FET were measured using the E1E2 electrodes, while the E1E3 and E2E4 pairs were used to measure the electrical properties of individual GeSe and SnS2, respectively.
image file: d5nh00508f-f2.tif
Fig. 2 Field-effect electrical performance of the GeSe/SnS2 AATs with two different bottom gate structures. (a) 3D Schematic and (b) OM image of the GeSe/SnS2 device with a h-BN/Au bottom gate structure. (c) The IdsVg curves of the individual GeSe and SnS2 FETs with the h-BN/Au bottom gate structure at Vds = 1 V. Comparison of the (d) IdsVg curves, (e) PVR and Ipeak, (f) ΔVg and Vpeak of the vdWHs with two different bottom gate structures at Vds = 1 V.

The transfer curves of the devices in the linear scale and semilogarithmic scale are shown in Fig. 2c and Fig. S8 (SI), respectively. The transfer characteristics at Vds = 1 V indicate the p-type and n-type conducting behaviors of the unipolar GeSe and SnS2 FETs, respectively. The lower on-current observed in the GeSe transistors compared with the SnS2 counterpart is primarily attributed to a difference in intrinsic carrier concentration. Using the linear extrapolation method,35 the current on/off ratio (Ion/Ioff ratio) and the p-type threshold voltage (Vth,p) of GeSe were evaluated to be 102 and 3.9 V, respectively. While this method is widely used in 2D FET research25,36,37 and shares similarity with CMOS, there are no traditional inversion channels in 2D FETs due to the atomic thickness limitation and the electrostatic modulation mechanism. Similarly, for SnS2, the Ion/Ioff ratio and the n-type threshold voltage (Vth,n) were 3 × 106 and −5.6 V, respectively. To highlight the electrical modulation ability of the h-BN/Au single bottom gate structure, the electrical performance of individual GeSe and SnS2 FETs on the SiO2/Si substrate were also measured, as shown in Fig. S9 (SI). For the GeSe (SnS2) FET on the SiO2/Si structure, the values of Vth and Ion/Ioff ratio were 20 V (−58 V) and 5 × 101 (4 × 106), respectively. Clearly, the electrical parameters (Ion/Ioff and Vth) of the FET with the h-BN/Au gate structure were superior to those on the SiO2/Si substrate. The proves that our fabricated h-BN/Au gate structure, with the advantages of fewer scattering centers and high interface quality, can effectively control the carrier concentration under a narrower driving voltage (ΔVg = VoffVon) of 9.4 V without a large gate leakage current.38 The mobility of the SnS2 channel was thereby boosted to 4.21 cm2 V−1 s−1, substantially exceeding the values reported for its counterparts with conventional Si/SiO2 substrates.39 This distinct comparison directly verifies that the h-BN dielectric layer also possesses the capability to enhance carrier mobility. The IdsVg curves of the GeSe/SnS2 AATs with two different gate structures at Vds = 1 V are shown in Fig. 2d (linear scale) and Fig. S10 (semilogarithmic scale), which exhibit distinct Λ-shaped curves (i.e., anti-ambipolar characteristics), demonstrating the switch between p-type and n-type characteristics as the gate voltage (Vg) was swept from 6 V to −6 V. As shown in Fig. 2e and f, the electrical parameters of the AAT device, including the peak-to-valley ratio (PVR), Ipeak, ΔVg, and peak voltage (Vpeak),9 were calculated from the transfer curves presented in Fig. 2d. Obviously, the h-BN/Au gate structure endowed the GeSe/SnS2 AAT with a higher PVR of 250 compared with the SiO2/Si gate structure. Meanwhile, the h-BN/Au single bottom gate structure achieved a twofold higher Ipeak (27.4 nA) than that of the normal SiO2/Si gate structure (13.2 nA). The same tendencies were observed for ΔVg and Vpeak, as seen in Fig. 2f and reported in other similar works,38,40,41 uncovering the superior electrostatic modulation capability of the h-BN/Au gate structure.

The overall advantages of the h-BN/Au gate structure, in terms of the electrical properties of the AAT device, can be ascribed to the smaller Vth and higher carrier mobility (μ) of individual GeSe and SnS2 (Fig. 2c), which enables the carrier concentration to be fully depleted at a small Vg and accumulated rapidly in the ΔVg. It is known that the Von and Voff values of the heterojunction channel in the AAT can be explained by the electrical properties of

the individual p-type and n-type channels. The μ and Vth of the individual GeSe and SnS2 FETs with the h-BN/Au gate structure were calculated, as shown in Note S2 (SI), and thus the μ and Vth of the GeSe/SnS2 AAT were calculated using the following formulas:42

 
image file: d5nh00508f-t1.tif(2)
 
image file: d5nh00508f-t2.tif(3)
where W is the channel width, Ci is the capacitance of the insulating layer between the gate and the semiconductor (i.e., the capacitance Ch-BN of h-BN), and Lp and Ln are the channel lengths of the GeSe and SnS2 FETs, respectively.

The above relevant parameters of the GeSe/SnS2 AAT, as well as the individual GeSe and SnS2 FETs, with the h-BN/Au bottom gate structure are summarized in Table 1. It can be seen that the μn, μp, Vth,n, and Vth,p of the AAT almost match the parameters of the individual GeSe and SnS2 FETs, indicating that the anti-ambipolar behavior in the heterojunction channel is mainly modulated by the Vth values and carrier mobility of the individual channels.

Table 1 The μ and Vth of the GeSe/SnS2 AAT and individual GeSe and SnS2 FETs with the h-BN/Au bottom gate structure at Vds = 1 V
Structure μn (cm2 V−1 s−1) Vth,n (V) μp (cm2 V−1 s−1) Vth,p (V)
AT 4.21 −5.2 5.67 × 10−2 4.2
GeSe N/A N/A 6.78 × 10−2 3.9
SnS2 3.3 −5.6 N/A N/A


Rectification behavior and tunneling mechanism in GeSxSe1−x/SnS2 AATs

Fig. S11 (SI) shows the OM images of the other GeSxSe1−x/SnS2 (x = 0.29, 0.42, 0.73) FETs with the h-BN/Au bottom gate structure. The IdsVds curves of individual SnS2 (E2E4 pairs) and GeSxSe1−x (E1E3 pairs) exhibited linear symmetric behavior between the semiconductor and Au, as seen in Fig. S12 and S13 (SI), which indicate that they possessed ohmic contacts, and the IdsVds curve of the heterojunction channel was dominated by the potential barrier region.43 According to the formula σ = L/SR (L is the length of the current path, S is the cross-sectional area, and R is the resistance), the conductivity (σ) of the GeSxSe1−x decreased from 6.25 × 10−1 S m−1 to 3 × 10−4 S m−1 as the S content increased, as depicted in Fig. S14b (SI), which is consistent with previous reports.26 As for GeSxSe1−x/SnS2, the corresponding IdsVds curves at Vg = 0 V are shown in Fig. 3a. It can be observed that the forward Ids at Vds = 3 V significantly decreased from 1.4 × 10−7 A to 4.8 × 10−11 A with an increase in S composition. Fig. S15 (SI) depicts the output curves of the GeSxSe1−x/SnS2 FETs. The rectification ratio (RR) is defined as the ratio of Ids values at Vds = 3 V and −3 V. As shown in Fig. S16, the output curves of the forward and reverse sweeps exhibited the same trend, and the trend of the double-sweep transfer curves was also the same. It is worth mentioning that the hysteresis behavior of the transfer curve may be mainly derived from the metal-semiconductor contact interface and the surface defects in the channel materials. As shown in Fig. S17, by comparing the transfer curves of the GeSe/SnS2 devices with h-BN and SiO2 gate dielectric, it was found that the device supported by the latter has a larger hysteresis window, further demonstrating the superiority of the h-BN/Au gate structure. The plots of the corresponding RR versus Vg are shown in Fig. 3b. In the forward Vg region, the RR at x = 0.29, 0.42, and 0.73 monotonously decreased with the increase of Vg. Nevertheless, the RR at x = 0 showed little variation caused by the high carrier concentration at the GeSe/SnS2 interface, demonstrating the minimal effect of Vg on RR. Under a larger negative Vg, the RR decreased with increasing Vg because of the effect of the strong negative Vg, which can drive more holes from GeSe to accumulate in the heterojunction channel and lead to the domination of the hole drift current. Overall, the Vg of the GeS0.29Se0.71/SnS2 FET had the most significant impact on the RR, which effectively changed from 21.8 to 2.5 × 103, differing by two orders of magnitude. Simmons approximation was used to understand the intrinsic transport mechanism in the forward current region. The equations44,45 used were as follows:
 
image file: d5nh00508f-t3.tif(4)
 
image file: d5nh00508f-t4.tif(5)
where d is the tunneling width, Φ is the potential barrier height, and m* and h denote the effective electron mass and Planck constant, respectively. According to the above equations, the plots of ln (Ids/Vds2) versus 1/Vds are displayed in Fig. 3c. Within the Vds range of 0 V to 3 V, both direct tunneling (DT) and Fowler–Nordheim tunneling (FNT) were present in GeSe/SnS2 and GeS0.29Se0.71/SnS2 FETs, while the devices with x = 0.42 and 0.73 only exhibited the DT process. The GeS0.29Se0.71/SnS2 vdWH FET was chosen as the representative to discuss the transport mechanism. It clearly showed a transition from the logarithmic state of DT at small Vds (red region) to the linear state of FNT with a negative slope at larger Vds (blue region). The transition voltage (Vtrans) was 0.71 V. According to the formula Φ = qVtrans, the tunneling barrier height was concluded to be 0.71 eV.46 Based on the transport mechanism, the band alignment diagrams of the GeS0.29Se0.71/SnS2 vdWH at various Vds values were deduced, as shown in Fig. 3d–f. Under reverse bias, a band-to-band tunneling (BTBT) process was observed. The Ids was ultralow because the minority electrons in GeS0.29Se0.71 flow through the tunneling widow, with the electron depletion region in GeS0.29Se0.71 and the electron accumulation region in SnS2.46 At small forward bias (0 V < Vds < Vtrans), the trapezoidal tunneling potential barrier at the interface dominated charge transfer via DT.44 However, at a larger forward bias (Vds > Vtrans), the potential barrier transformed into a triangular shape, allowing more electrons in SnS2 to easily overcome the potential barrier and transfer to GeS0.29Se0.71. Therefore, the DT and FNT processes determine the electron transfer behavior from the CBM of SnS2 through the potential barrier to the CBM of GeS0.29Se0.71, leading to a continuous increase in the forward Ids. While the GeSe/SnS2 FET exhibited the aforementioned transport behavior, it differed at higher Vds. A narrow sublinear region was obtained in the presence of a space-charge limited region.47,48 The phenomenon could be reproduced in two other GeSe/SnS2 FETs with h-BN/Au and SiO2/Si substrates, as shown in Fig. S18 (SI).

image file: d5nh00508f-f3.tif
Fig. 3 Charge transport properties and intrinsic mechanisms of the GeSxSe1−x/SnS2 (0 ≤ x ≤ 0.73) FETs. (a) IdsVds curves. (b) The RR versus Vg plots. (c) Fowler−Nordheim plots of ln(Ids/Vds2) versus 1/Vds when scanned at Vds > 0 V. Band alignment of the GeS0.29Se0.71/SnS2 FET at (d) Vds < 0 V, (e) 0 V < Vds < Vtrans, and (f) Vtrans < Vds.

Dependence of anti-ambipolar behavior on the S composition in GeSxSe1−x/SnS2 vdWH

The transfer curves of the GeSxSe1−x/SnS2 device with four different S compositions (0 ≤ x ≤ 0.73) at Vds = 3 V are shown in Fig. 4a (linear scale) and Fig. S19 (semilogarithmic scale), which demonstrate obvious and composition-dependent Λ-shape transfer characteristics. It is clear that the transfer characteristics of the GeSxSe1−x/SnS2 AAT devices can be divided into three regions based on the charge accumulation state, using which the on-set voltage (Von) and off-set voltage (Voff) of the device can be determined. In region I, the SnS2 layer is depleted, which gives rise to high resistance in the entire channel. The resistance of the entire channel decreases significantly with the rise in Vg, eventually approaching the Vth,n of the individual SnS2 FET, which is denoted as Von. In region II, both the SnS2 and GeSxSe1−x layers are under the sub-threshold regime, and Ids gradually increases with increasing Vg and reaches a peak current. In region III, the GeSxSe1−x layer is gradually depleted, and thus the current decreases more slowly than that in region I, which is mainly attributed to the relatively lower carrier mobility of GeSxSe1−x than that of SnS2. Similarly, the resistance of the entire channel increases gradually with the increase of Vg until it approaches the Vth,p of the individual GeSxSe1−x FET, which is referred to as Voff. Notably, the values of Von (Voff) of the GeSxSe1−x/SnS2 AATs with lower S composition (x = 0 and 0.29, i.e., S-poor) were smaller than those with higher S composition (x = 0.42 and 0.73, i.e., S-rich), resulting in dynamic control of ΔVg. As shown in Fig. S20 (SI), the transconductance (gm = dIds/dVg) drastically varied from a positive to a negative value across the Vpeak, exhibiting the presence of NDT. The Vpeak corresponding to the Ipeak was consistent with that when the gm changes from a positive to negative value.49 Conversely, the values of Vpeak of the GeSxSe1−x/SnS2 AATs with S-poor composition were significantly higher than those with S-rich composition. As shown in Fig. 4b and c, the key electrical performance of the GeSxSe1−x/SnS2 devices with S-poor and S-rich compositions were synthetically compared. The ΔVg and |Vpeak| of the GeS0.29Se0.71/SnS2 AAT with S-poor composition reached the minimum values of 5.8 V and 2.6 V, respectively, which are nearly 3 times lower than those of the GeS0.42Se0.58/SnS2 AAT with S-rich composition. Likely, the GeS0.29Se0.71/SnS2 AAT with S-poor composition also achieved the maximum PVR of 6.6 × 103. Distinctly, The Ipeak decreased monotonously from 1.4 × 10−7 A to 8.7 × 10−11 A with increasing S composition, which may stem from the decrease in the σ and hole mobility (μp) of the individual GeSxSe1−x FETs as the S composition increases (Fig. S21, SI). The transfer curves of GeSxSe1−x FETs under high-vacuum conditions are shown in Fig. S22. It can be observed that the electrical characteristics of the FETs changed with the variation in S composition, which not only rules out the effects of surface adsorption and humidity but also confirms that the changes in electrical characteristics are indeed primarily determined by the ratio of S and Se. Importantly, these results strongly confirm the significant tunability of the electrical properties of GeSxSe1−x/SnS2 AATs through alloy engineering. The optimal comprehensive anti-ambipolar characteristics were obtained at x = 0.29. Table 2 summarizes the key parameters of the GeSxSe1−x/SnS2 (0 ≤ x ≤ 0.73) AATs and provides a comparative evaluation of similar reported devices. It is evident that through sulfur content modulation, the band structure of the GeSxSe1−x/SnS2 AATs can be tuned from type-II to type-III alignments. This behavior dynamically reshapes the carrier transport pathways between the channel and barriers, enabling precise configuration of anti-ambipolar characteristics. Specifically, at a sulfur content of x = 0.29, the device achieved a ΔVg of 5.8 V, a significant reduction compared to conventional devices with the SiO2 gate structure. The low ΔVg will lower the power consumption and can suppress the voltage drift caused by interfacial defects. Concurrently, the Vpeak was optimized to −2.6 V, approaching the zero-bias condition, which reflects balanced electron and hole injection barriers under the Type-III band alignment, thereby enhancing carrier injection efficiency and transport symmetry. Furthermore, the Ipeak was regulated to 1.6 × 10−9 A, striking an optimal balance. This avoids noise sensitivity from insufficient current while preventing excessive static power consumption from large currents. Notably, the device attained a PVR of 103, which is comparable to the values reported in the literature. Leveraging h-BN interfacial passivation to effectively suppress defect scattering, our work achieves synergistic optimization across power consumption, performance, and reliability. This circumvents the multidimensional performance trade-offs of traditional SiO2-based devices and pioneers a new technical pathway for low-voltage, highly reliable anti-ambipolar transistors. To enhance the reliability of the above conclusions, another set of GeSxSe1−x/SnS2 (0 ≤ x ≤ 0.73) AATs was prepared, as shown in Fig. S23–26 (SI). The key parameters were extracted and compared with the first set of devices, as summarized in Table S1 (SI), with each set comprising devices made of four distinct S compositions (x = 0, 0.29, 0.42, and 0.73). In total, eight devices were measured and analyzed in this study. The consistency of the data is evident from the insignificant variation between the two sets, and the same correlation between S doping levels and device performance was observed in both data sets.
image file: d5nh00508f-f4.tif
Fig. 4 Transfer characteristics and transport mechanism of the GeSxSe1−x/SnS2 (0 ≤ x ≤ 0.73) AATs. (a) The Λ-shape transfer curves with three transmission regions at Vds = 3 V. (b) The ΔVg and Vpeakversus S composition plots and (c) PVR and Ipeakversus S composition plots extracted from (a). (d) Equivalent circuit model and the correlating band alignments of the GeS0.29Se0.71/SnS2 vdWH AAT in three distinct Vg regions at Vds = 3 V.
Table 2 Summary of the device performance (ΔVg, Vpeak, Ipeak, and PVR) of the as-fabricated AATs in comparison with previously reported AATs
AATs (p-/n-channels) Dielectric layer ΔVg (V) V peak (V) I peak (A) PVR Band alignment Ref.
GeSe/SnS2 h-BN 9.4 −3.8 1.4 × 10−7 1.5 × 103 Type-III This work
GeS0.29Se0.71/SnS2 h-BN 5.8 −2.6 1.6 × 10−9 6.6 × 103 Type-III This work
GeS0.42Se0.58/SnS2 h-BN 15.3 −7 1.6 × 10−10 5.5 × 102 Type-III This work
GeS0.73Se0.27/SnS2 h-BN 13.8 −5.6 8.7 × 10−11 87 Type-II This work
WSe2/InSe h-BN 20 −14 4 × 10−6 104 Type-II 50
BP/MoTe2 SiO2 45 −5 1 × 10−6 30 Type-I 13
WSe2/ReS2 SiO2 8 −6.7 4 × 10−10 Type-II 51
MoTe2/h-BN SiO2 ≈30 ≈0 3 × 10−7 2 × 102 Type-II 52
WSe2/PdSe2/MoS2 h-BN <1.5 <−0.1 1 × 10−8 1.1 × 105 53
SnSe1.5S0.5/MoTe2 SiO2 ≈46 −38 4.3 × 10−8 6.7 × 103 Type-III 25
SnS2/MoTe2 SiO2 30 −13 8 × 10−10 4.5 Type-III 33
InSe/BP SiO2 2.22 1.5 × 10−10 >10 Type-III 54
WSe2/SnS2 h-BN/SiO2 26 5 3 × 10−8 104 55
MoTe2/CrOCl SiO2 40 −20 3 × 10−8 56


In order to deeply investigate the conductive behavior of the anti-ambipolar characteristics of GeSxSe1−x/SnS2 AAT devices, the equivalent circuit model was utilized.36 The Ids obtained from the transfer characteristics curves can be defined as the shoot-through current in the CMOS inverter, which refers to the overlapping region of Ids in the transfer curves of the individual GeSxSe1−x and SnS2 FETs. Taking the GeS0.29Se0.71/SnS2 AAT device as an example, Fig. 4d demonstrates the equivalent circuit model in three distinct Vg regions at Vds = 3 V. In particular, the p-type GeS0.29Se0.71 FET and n-type SnS2 FET can be regarded as a p-MOS and n-MOS, respectively. These two MOS parts are connected in series to form an equivalent circuit model. The circuit analysis can be sorted into three regions. In the three Vg regions, the direction of current flowing through CMOS coincides with the direction of hole movement described in the circuit diagram, which is depicted by the red arrows. (i) When Vg < Von, the p-MOS is in the conducting state with holes accumulating near the gate, while electrons are turned off. (ii) When Von < Vg < Voff, both p-MOS and n-MOS are in the conductive state, turning the circuit on. Within this region, as Vg monotonically increases, Ids rapidly varies from an upward trend to a downward trend, because of depletion in the n-MOS due to its cutoff state, yielding a Λ-shape transition. (iii) When Vg > Voff, the circuit turns off again due to the complete depletion of p-MOS, even though the n-MOS is in a conductive state. Meanwhile, the band alignment analysis was also employed to further understand the anti-ambipolar transport process. At region i, holes are accumulated in the GeS0.29Se0.71 layer due to its p-type nature, while complete depletion of the n-type SnS2 layer fully shuts off the overall heterojunction, resulting in a low Ids. With increasing Vg (region ii), a moderate decrease in hole concentration occurs in the GeS0.29Se0.71 layer accompanied by a more significant increase in electron concentration in the SnS2 layer, allowing both holes and electrons contribute concurrently to Ids, and thus Ids increases rapidly. Region iii refers to the stage in which the depletion of holes in the GeS0.29Se0.71 layer occurs more rapidly than the accumulation of electrons in the SnS2 layer, which brings about the gradual reduction of Ids until the whole heterojunction is turned off again.

Logic inverters based on GeSxSe1−x/SnS2 vdWH AATs

This study demonstrates the application of GeSxSe1−x/SnS2 (0 ≤ x ≤ 0.73) AATs in binary logic inverters. The schematic in Fig. 5a shows that the inverters consisted of p-type GeSxSe1−x FET and n-type SnS2 FET.57 The supply voltage (Vdd) was applied to the p-FET terminal, while the ground (GND) was applied to the n-FET terminal. The input voltage (Vin) and output voltage (Vout) were connected to the common Vg and the vdWH region, respectively. The corresponding equivalent circuit diagram is shown in Fig. 5b. The transfer characteristics of the GeS0.29Se0.71/SnS2 (x = 0.29) AAT maintained an ‘Λ’ shape across varying Vdd, steepening with an increase in Vdd, as shown in Fig. S27 (SI). The VoutVin diagram of the inverter (x = 0.29) is shown in Fig. 5c, with the inset showing the truth table for the digital logic states. High Vout (i.e., logic “1”) corresponds to low Vin (i.e., logic “0”) and vice versa.58 As Vdd increases, the logic state transitions of the inverter become more rapid, resulting in a higher voltage gain (Gain is calculated using the formula Gain = −dVout/dVin).59 The Gain-Vin curves are illustrated in Fig. 5d, with the maximum gain of 8.83 at Vdd = 3 V. Similarly, the characteristics of the inverters changed with changes in the S composition. The voltage transfer characteristics (VTC) of the inverters at x = 0, 0.42, and 0.73 under Vdd = 3 V are depicted in Fig. S28, and the insets show the corresponding Gain versus Vin plots (SI). Regardless of the S content, all the devices displayed two logic states, and the voltage gain displayed a single peak, confirming their binary inverter functionality. However, with varying S composition, the binary inverters exhibited different transition voltages and transition region widths, which result in variations in voltage gain, as illustrated in Fig. S29 (SI). The binary inverter at x = 0.29 achieved the maximum gain, with decreasing gain observed in devices with increasing S compositions. Notably, the output leads were pre-deposited along with the source and drain electrodes to circumvent performance degradation induced by organic solvents and high-energy deposition steps. The PoutVin curves of the GeSxSe1−x/SnS2 vdWH inverters, as shown in Fig. S30 (SI), were used to extract the maximum static power dissipation (Pstatic).60,61 Consequently, the plot of Pstaticversus the S composition is depicted in Fig. 5e. Specifically, the device with x = 0.73 had the smallest Pstatic, of about 12.1 pW. Moreover, the devices with x = 0.29 and 0.42 showed similar Pstatic, both at around 30 pW. Fig. 5f demonstrates a comparison of the Pstatic of GeSxSe1−x/SnS2 (0 ≤ x ≤ 0.73) binary inverters and previously reported works,62–71 and the Pstatic of inverters in the highlighted blue region exhibit the pW level. In the context of growing demand for power static optimization, the low-power characteristics of GeSxSe1−x/SnS2 AATs, notably their static power consumption at the pW level, are promising for a wide range of applications in the field of high-performance computing and electronic devices and for the development 2D material integration processes, as they are and compatible with existing technologies.
image file: d5nh00508f-f5.tif
Fig. 5 Binary logic inverters based on GeSxSe1−x/SnS2 (0 ≤ x ≤ 0.73) vdWH AATs. (a) The schematic and (b) equivalent circuit diagram of the binary logic inverters. (c) VoutVin curves and (d) gain–Vin curves of the GeS0.29Se0.71/SnS2 (x = 0.29) binary logic inverter with different Vdd. (e) Pstatic of the inverters with different S compositions at Vdd = 3 V. (f) Comparison of Pstatic with other previously reported inverters. The inverters in the highlighted blue area have Pstatic in the pW level.

Conclusions

In summary, we propose a series of composition-tunable GeSxSe1−x/SnS2 (0 ≤ x ≤ 0.73) vdWH AATs with a h-BN/Au single bottom gate configuration for high-performance logic inverter applications. The S-poor AATs (x = 0, 0.29) exhibited rectification behaviors with DT and FNT transport mechanisms, while the S-rich AATs (x = 0.42 and 0.73) exhibited rectification behaviors with only DT transport mechanism, because adjusting the doping level of GeSxSe1−x through alloy engineering modulated the width of the tunneling window of the vdWH. Furthermore, all the anti-ambipolar characteristics, including ΔVg, Vpeak, Ipeak, and PVR, could be dynamically modulated by adjusting S composition. Significantly, these devices exhibited a remarkable PVR of 6.6 × 103 (x = 0.29) and the largest Ipeak of 1.4 × 10−7 A (x = 0), which is attributed to the altered energy band alignment of vdWH induced by composition modulation. Therefore, the inverters fabricated using our device exhibited the highest voltage gain of 8.83 (x = 0.29) and the lowest static power consumption of 12.1 pW (x = 0.73). This work envisions composition-dependent tuning of anti-ambipolar transistors as a reliable strategy for realizing configurable multi-valued logic inverters, which offer a great opportunity for harnessing the full potential of the emerging parallel computing scheme.

Experimental methods

Material preparation

The 2D free-standing GeSxSe1−x nanoplates were synthesized using a two-inch quartz tube slide rail furnace (OTF-1200X-S50-SL, Hefei Kejing Materials Technology Co., Ltd). GeSe (purity 99.999%, Alfa Metal Materials Co., Ltd) and GeS (purity 99.999%, Alfa Metal Materials Co., Ltd) powders served as the growth precursors. The S composition of the GeSxSe1−x nanoplates was varied by adjusting the molar mass ratio of the powders. Specifically, the GeSe and GeS mixtures were placed in a quartz boat at the center of the furnace. The furnace was sealed, evacuated, and filled with Ar to expel oxygen, and set to 570 °C. Once the target temperature was reached, the heating source was quickly positioned in the center of the furnace for 4–5 minutes and then swiftly removed to the cooling zone after the growth process was complete. Moreover, the SnS2 and h-BN nanoplates were obtained by mechanical exfoliation of the bulk crystal materials bought from Taizhou SUNANO New Energy Co., Ltd.

Device fabrication

All electrodes were patterned and deposited using the ARP-5350 positive photoresist (purchased from Taizhou SUNANO New Energy Co., Ltd), AR 300-26 developing solution (ALLRESIST GmbH, purchased from Taizhou SUNANO New Energy Co., Ltd), an Ultraviolet Maskless Lithography machine (TuoTuo Technology, UV Litho-ACA), and electron beam evaporation. First, an Au metal gate (with a thickness of 50 nm and an area of 40 µm × 40 µm) with an extended contact lead was fabricated on a 300 nm SiO2/Si substrate. After that, the h-BN nanoplate was used to cover the entire metal gate area by the dry transfer method. To ensure a tight stack, the bottom gate structure was annealed in a 250 °C argon environment for 20 minutes to remove the organic and glue residues from the interface. Then, source-drain electrodes (10 nm Cr/40 nm Au) were prepared on the h-BN/Au bottom gate structure. Lastly, a PDMS stamp (17 mil, Gel park, bought from Shanghai Onway Technology Co., Ltd) was used to timely transfer GeSxSe1−x and SnS2 nanoplates sequentially onto the designated electrode areas via a 3D transfer platform (Shanghai Onway Technology Co., Ltd), forming a vertically-stacked GeSxSe1−x/SnS2 vdWH device. The entire process of device preparation was carried out in a clean room.

Characterization and electrical measurements

The OM images were captured using an optical microscope (SOPTOP BH200M, Ningbo Sunny Instruments Co., Ltd). The thicknesses and surface potential were obtained using a scanning probe microscope (Dimension FastScan from Bruker Co., Ltd) equipped with an atomic force microscope and a Kelvin probe force microscope. The Raman spectra were measured by confocal microscopy (Nost Technology Co., Ltd) using a 532 nm laser at room temperature and in an ambient atmosphere. All electrical properties of the devices were measured using a four-probe stage equipped with Keithley 2636B and Keithley 2611B and a three-probe stage equipped with a semiconductor parameter analyzer (Fs-Pro, Primarius Co., Ltd bought from Guangdong Avit Technology Co., Ltd).

Author contributions

W. Gao and T. Zheng: supervision of the entire project. Y. Long: methodology and writing-original draft. Q Deng, S Chen, Y. He, Y. Wang, Z. Zheng, N. Huo, D. Luo, X. Liu, Y. Sun, Z. Chen, and M. Yang: data curation. W. Gao, T. Zheng, and Q Deng: writing-review and editing.

Conflicts of interest

There are no conflicts to declare.

Data availability

The data supporting this article have been included as part of the Supplementary Information.

Supplementary information (SI): note S1: calculation of surface potential difference (SPD). Note S2: calculation of carrier mobility (μ) and threshold voltage (Vth). Fig. S1–30 and Table S1. See DOI: https://doi.org/10.1039/d5nh00508f.

Acknowledgements

We acknowledge the financial support from the Guangzhou Science and Technology program (No. 2025A04J1892), the National Natural Science Foundation of China (No. 62004071), the Science and Technology Program of Guangdong (No. 2024A0505040026), and the Basic and Applied Basic Research Foundation of Guangdong Province (No. 2024A1515110204).

References

  1. Y. Meng, W. Wang, W. Wang, B. Li, Y. Zhang and J. Ho, Adv. Mater., 2024, 36, 2306290 CrossRef CAS.
  2. H. Wang, L. Bao, R. Guzman, K. Wu, A. Wang, L. Liu, L. Wu, J. Chen, Q. Huan, W. Zhou, S. T. Pantelides and H. J. Gao, Adv. Mater., 2023, 35, 2301067 CrossRef CAS.
  3. W. Cao, H. Bu, M. Vinet, M. Cao, S. Takagi, S. Hwang, T. Ghani and K. Banerjee, Nature, 2023, 620, 501–515 CrossRef CAS.
  4. S. Seo, J. I. Cho, K. S. Jung, M. Andreev, J. H. Lee, H. Ahn, S. Jung, T. Lee, B. Kim, S. Lee, J. Kang, K. B. Lee, H. J. Lee, K. S. Kim, G. Y. Yeom, K. Heo and J. H. Park, Adv. Mater., 2022, 34, e2202799 CrossRef.
  5. Y. Luo, J. Chen, A. Abbas, W. Li, Y. Sun, Y. Sun, J. Yi, X. Lin, G. Qiu, R. Wen, Y. Chai, Q. Liang and C. Zhou, Adv. Funct. Mater., 2024, 34, 2407253 CrossRef CAS.
  6. C. Y. Liu, K. Y. Tien, P. Y. Chiu, Y. J. Wu, Y. Chuang, H. S. Kao and J. Y. Li, Adv. Mater., 2022, 34, e2203888 CrossRef PubMed.
  7. S. Seo, J. Koo, J.-W. Choi, K. Heo, M. Andreev, J.-J. Lee, J.-H. Lee, J.-I. Cho, H. Kim, G. Yoo, D.-H. Kang, J. Shim and J.-H. Park, npj 2D Mater. Appl., 2021, 5, 32 CrossRef.
  8. J. H. Kim, B. H. Moon and G. H. Han, Appl. Phys. Lett., 2024, 124, 123104 CrossRef CAS.
  9. R. J. Mathew, K. H. Cheng, C. R. P. Inbaraj, R. Sankar, X. P. A. Gao and Y. T. Chen, Adv. Electron. Mater., 2023, 9, 2300095 CrossRef CAS.
  10. S. Fan, Q. A. Vu, S. Lee, T. L. Phan, G. Han, Y. M. Kim, W. J. Yu and Y. H. Lee, ACS Nano, 2019, 13, 8193–8201 CrossRef CAS PubMed.
  11. L. Britnell, R. V. Gorbachev, A. K. Geim, L. A. Ponomarenko, A. Mishchenko, M. T. Greenaway, T. M. Fromhold, K. S. Novoselov and L. Eaves, Nat. Commun., 2013, 4, 1794 CrossRef CAS.
  12. Y. Wakayama and R. Hayakawa, Adv. Funct. Mater., 2019, 30, 1903724 CrossRef.
  13. M. Zubair, H. Wang, Q. Zhao, M. Kang, M. Xia, M. Luo, Y. Dong, S. Duan, F. Dai, W. Wei, Y. Li, J. Wang, T. Li, Y. Fang, Y. Liu, R. Xie, X. Fu, L. Dong and J. Miao, Small, 2023, 19, 2300010 CrossRef CAS PubMed.
  14. D. Tan, X. Wang, W. Zhang, H. E. Lim, K. Shinokita, Y. Miyauchi, M. Maruyama, S. Okada and K. Matsuda, Small, 2018, 14, 1704559 CrossRef PubMed.
  15. W. Sang, D. Xiang, Y. Cao, F. Tan, Z. Han, W. Songlu, P. Zhou and T. Liu, Adv. Funct. Mater., 2023, 34, 2307675 CrossRef.
  16. K. Liang, J. Wang, X. Wei, Y. Zhang, J. Fan, L. Ni, Y. Yang, J. Liu, Y. Tian, X. Wang, C. Yuan and L. Duan, J. Phys.: Condens. Matter, 2023, 35, 315501 CrossRef CAS PubMed.
  17. C. R. Paul Inbaraj, R. J. Mathew, R. K. Ulaganathan, R. Sankar, M. Kataria, H. Y. Lin, Y. T. Chen, M. Hofmann, C. H. Lee and Y. F. Chen, ACS Nano, 2021, 15, 8686–8693 CrossRef CAS PubMed.
  18. J. Shim, S. H. Jo, M. Kim, Y. J. Song, J. Kim and J. H. Park, ACS Nano, 2017, 11, 6319–6327 CrossRef CAS PubMed.
  19. H. Han, B. Zhang, Z. Zhang, Y. Wang, C. Liu, A. K. Singh, A. Song, Y. Li, J. Jin and J. Zhang, Nano Lett., 2024, 24, 8602–8608 CrossRef CAS PubMed.
  20. E. Wu, Y. Xie, Q. Liu, X. Hu, J. Liu, D. Zhang and C. Zhou, ACS Nano, 2019, 13, 5430–5438 CrossRef CAS.
  21. X. Xiong, J. Kang, Q. Hu, C. Gu, T. Gao, X. Li and Y. Wu, Adv. Funct. Mater., 2020, 30, 1909645 CrossRef CAS.
  22. J. Yao and G. Yang, Adv. Sci., 2022, 9, 2103036 CrossRef.
  23. X. Li, H. Long, J. Zhong, F. Ding, W. Li, Z. Zhang, R. Song, W. Huang, J. Liang, J. Liu, R. Wu, B. Li, B. Zhao, X. Yang, Z. Zhang, Y. Liu, Z. Wei, J. Li and X. Duan, Nat. Electron., 2023, 6, 842–851 CrossRef CAS.
  24. B. Gao, W. Wang, Y. Meng, C. Du, Y. Long, Y. Zhang, H. Shao, Z. Lai, W. Wang, P. Xie, S. Yip, X. Zhong and J. C. Ho, Small, 2024, 20, 2402217 CrossRef CAS PubMed.
  25. X. Luo, Y. Liu, T. Zheng, L. Huang, Z. Zheng, J. Huang, Z. Lan, L. Zhao, J. Ma, N. Huo, Y. Yan, Y. Berencén, W. Gao and J. Li, ACS Appl. Mater. Interfaces, 2024, 16, 42491–42501 CrossRef CAS PubMed.
  26. T. Zheng, Y. Pan, M. Yang, Z. Li, Z. Zheng, L. Li, Y. Sun, Y. He, Q. Wang, T. Cao, N. Huo, Z. Chen, W. Gao, H. Xu and J. Li, Adv. Mater., 2024, 36, 2313721 CrossRef CAS.
  27. C. G. Pope, J. Chem. Educ., 1997, 74, 129 CrossRef CAS.
  28. D. Hu, C. Ye, X. Wang, X. Zhao, L. Kang, J. Liu, R. Duan, X. Cao, Y. He, J. Hu, S. Li, Q. Zeng, Y. Deng, P. F. Yin, A. Ariando, Y. Huang, H. Zhang, X. R. Wang and Z. Liu, Nano Lett., 2021, 21, 5338–5344 CrossRef CAS PubMed.
  29. C. Xia, J. Du, M. Li, X. Li, X. Zhao, T. Wang and J. Li, Phys. Rev. Appl., 2018, 10, 054064 CrossRef CAS.
  30. G.-H. Lee, X. Cui, Y. D. Kim, G. Arefe, X. Zhang, C.-H. Lee, K. W. Fan Ye, T. Taniguchi, P. Kim and J. Hone, ACS Nano, 2015, 9, 7019–7026 CrossRef CAS.
  31. C. R. Dean, A. F. Young, I. Meric, C. Lee, L. Wang, S. Sorgenfrei, K. Watanabe, T. Taniguchi, P. Kim, K. L. Shepard and J. Hone, Nat. Nanotechnol., 2010, 5, 722–726 CrossRef CAS PubMed.
  32. M. K. Joo, B. H. Moon, H. Ji, G. H. Han, H. Kim, G. Lee, S. C. Lim, D. Suh and Y. H. Lee, Nano Lett., 2016, 16, 6383–6389 CrossRef CAS PubMed.
  33. R. Yao, Z. Zheng, M. Xiong, X. Zhang, X. Li, H. Ning, Z. Fang, W. Xie, X. Lu and J. Peng, Appl. Phys. Lett., 2018, 112, 103503 CrossRef.
  34. L. Liu, L. Kong, Q. Li, C. He, L. Ren, Q. Tao, X. Yang, J. Lin, B. Zhao, Z. Li, Y. Chen, W. Li, W. Song, Z. Lu, G. Li, S. Li, X. Duan, A. Pan, L. Liao and Y. Liu, Nat. Electron., 2021, 4, 342–347 CrossRef CAS.
  35. A. Ortiz-Conde, F. J. G. a Sánchez, J. J. Liou, A. Cerdeira, M. Estrada and Y. Yue, Microelectron. Reliab., 2002, 42, 583–596 CrossRef.
  36. S. Xiao, T. Zheng, W. Chen, J. Zhang, M. Yang, Y. Sun, Z. Zheng, D. Hao, N. Huo, Z. Chen and W. Gao, Adv. Funct. Mater., 2024, 34, 2403509 CrossRef CAS.
  37. H. Lu, Y. Wang, X. Han and J. Liu, ACS Nano, 2024, 18, 23403–23411 CrossRef CAS PubMed.
  38. L. Wu, W. Gao, Y. Sun, M. Yang, Z. Zheng, W. Fan, K. Shu, Z. Dan, N. Zhang, N. Huo and J. Li, Adv. Mater. Interfaces, 2022, 9, 2102099 CrossRef CAS.
  39. S. De Stefano, O. Durante, A. Sessa, A. Politano, G. D’Olimpio, T. Dadiani, E. Faella, A. Dinescu, C. Parvulescu, C. Hetherington, C. N. Kuo, C. S. Lue, M. Aldrigo, M. Passacantando and A. D. Bartolomeo, ACS Appl. Mater. Interfaces, 2025, 17, 50901–50915 CrossRef CAS.
  40. M. Andreev, J. W. Choi, J. Koo, H. Kim, S. Jung, K. H. Kim and J. H. Park, Nanoscale Horiz., 2020, 5, 1378–1385 RSC.
  41. S. Ghosh, A. Varghese, H. Jawa, Y. Yin, N. V. Medhekar and S. Lodha, ACS Nano, 2022, 16, 4578–4587 CrossRef CAS.
  42. R. Hayakawa, K. Honma, S. Nakaharai, K. Kanai and Y. Wakayama, Adv. Mater., 2022, 34, e2109491 CrossRef.
  43. Y. Yan, S. Li, J. Du, H. Yang, X. Wang, X. Song, L. Li, X. Li, C. Xia, Y. Liu, J. Li and Z. Wei, Adv. Sci., 2021, 8, 1903252 CrossRef CAS.
  44. S. Chen, J. Ma, N. Bu, T. Zheng, J. Chen, J. Huang, X. Luo, Z. Zheng, N. Huo, J. Li and W. Gao, ACS Appl. Mater. Interfaces, 2024, 16, 33740–33751 CrossRef CAS PubMed.
  45. X. Cong, Y. Zheng, F. Huang, Q. You, J. Tang, F. Fang, K. Jiang, C. Han and Y. Shi, Nano Res., 2022, 15, 8442–8450 CrossRef CAS.
  46. X. Zhou, X. Hu, S. Zhou, H. Song, Q. Zhang, L. Pi, L. Li, H. Li, J. Lu and T. Zhai, Adv. Mater., 2018, 30, 1703286 CrossRef PubMed.
  47. C. Park, J. Lee, M. J. Kim, N. T. Duong, M. S. Jeong and S. C. Lim, Appl. Surf. Sci., 2021, 558, 149870 CrossRef CAS.
  48. Y. Pan, T. Zheng, F. Gao, L. Qi, W. Gao, J. Zhang, L. Li, K. An, H. Gu and H. Chen, Small, 2024, 20, 2311606 CrossRef CAS PubMed.
  49. D. Tan, X. Wang, W. Zhang, H. E. Lim, K. Shinokita, Y. Miyauchi, M. Maruyama, S. Okada and K. Matsuda, Small, 2018, 14, 1704559 CrossRef.
  50. C. R. Paul Inbaraj, R. J. Mathew, R. K. Ulaganathan, R. Sankar, M. Kataria, H. Y. Lin, Y.-T. Chen, M. Hofmann, C.-H. Lee and Y.-F. Chen, ACS Nano, 2021, 15, 8686–8693 CrossRef CAS PubMed.
  51. Y. Shingaya, T. Iwasaki, R. Hayakawa, S. Nakaharai, K. Watanabe, T. Taniguchi, J. Aimi and Y. Wakayama, ACS Appl. Mater. Interfaces, 2024, 16, 33796–33805 CrossRef CAS PubMed.
  52. Z. Zhang, S. Huo, Q. Tian, F. Meng, Z. Yang, Y. Ma, Y. Wang, Y. Xie, X. Hu, W. Gao, E. Wu and C. Pan, Adv. Funct. Mater., 2025, 35, 2424728 CrossRef CAS.
  53. Y. Pang, Y. Zhou, S. Qiu, L. Tong, N. Zhao and J. B. Xu, Nat. Commun., 2025, 16, 3188 CrossRef CAS PubMed.
  54. H. Xu, R. Xie, J. Miao, Z. Zhang, H. Ge, X. Shi, M. Luo, J. Wang, T. Li, X. Fu, J. C. Ho, P. Zhou, F. Wang and W. Hu, Light: Sci. Appl., 2025, 14, 72 CrossRef CAS PubMed.
  55. Y. Lv, C.-Y. Wu, Y. Zhao, G. Wu, M. Abid, J. Cho, M. Choi, C. O. Coileain, K.-M. Hung and C.-R. Chang, ACS Appl. Electron. Mater., 2022, 4, 5487–5497 CrossRef CAS.
  56. S. Li, J. Zhang, Y. Li, K. Zhang, L. Zhu, W. Gao, J. Li and N. Huo, Appl. Phys. Lett., 2023, 122, 083503 CrossRef CAS.
  57. S. H. A. Jaffery, M. Riaz, Z. Abbas, G. Dastgeer, S. Aftab, S. Hussain, M. Ali and J. Jung, EcoMat, 2022, 5, e12307 CrossRef.
  58. C. F. Chen, S. H. Yang, C. Y. Lin, M. P. Lee, M. Y. Tsai, F. S. Yang, Y. M. Chang, M. Li, K. C. Lee, K. Ueno, Y. Shi, C. H. Lien, W. W. Wu, P. W. Chiu, W. Li, S. T. Lo and Y. F. Lin, Adv. Sci., 2022, 9, 2106016 CrossRef CAS.
  59. J. Bi, X. Zou, Y. Lv, G. Li, X. Liu, Y. Liu, T. Yu, Z. Yang and L. Liao, Adv. Electron. Mater., 2020, 6, 2000291 CrossRef CAS.
  60. C. Yang, C. Jiang, W. Niu, D. Hao, H. Huang, H. Fu, J. Miao, X. Liu, X. Zou, F. Shan and Z. Yang, Appl. Phys. Lett., 2024, 124, 073504 CrossRef CAS.
  61. C. Jiang, H. W. Choi, X. Cheng, H. Ma, D. Hasko and A. Nathan, Science, 2019, 363, 719–723 CrossRef CAS.
  62. G. Gao, B. Wan, X. Liu, Q. Sun, X. Yang, L. Wang, C. Pan and Z. L. Wang, Adv. Mater., 2018, 30, 1705088 CrossRef PubMed.
  63. A. Pezeshki, S. H. Hosseini Shokouh, P. J. Jeon, I. Shackery, J. S. Kim, I. K. Oh, S. C. Jun, H. Kim and S. Im, ACS Nano, 2016, 10, 1118–1125 CrossRef CAS.
  64. J. Y. Lim, A. Pezeshki, S. Oh, J. S. Kim, Y. T. Lee, S. Yu, D. K. Hwang, G. H. Lee, H. J. Choi and S. Im, Adv. Mater., 2017, 29, 1701798 CrossRef.
  65. H. Zhang, C. Li, J. Wang, W. Hu, D. W. Zhang and P. Zhou, Adv. Funct. Mater., 2018, 28, 1805171 CrossRef.
  66. A. J. Yang, K. Han, K. Huang, C. Ye, W. Wen, R. Zhu, R. Zhu, J. Xu, T. Yu, P. Gao, Q. Xiong and X. Renshaw Wang, Nat. Electron., 2022, 5, 233–240 CrossRef CAS.
  67. L. Kong, X. Zhang, Q. Tao, M. Zhang, W. Dang, Z. Li, L. Feng, L. Liao, X. Duan and Y. Liu, Nat. Commun., 2020, 11, 1866 CrossRef CAS PubMed.
  68. C. Pan, C.-Y. Wang, S.-J. Liang, Y. Wang, T. Cao, P. Wang, C. Wang, S. Wang, B. Cheng, A. Gao, E. Liu, K. Watanabe, T. Taniguchi and F. Miao, Nat. Electron., 2020, 3, 383–390 CrossRef CAS.
  69. N. Li, Q. Wang, C. Shen, Z. Wei, H. Yu, J. Zhao, X. Lu, G. Wang, C. He, L. Xie, J. Zhu, L. Du, R. Yang, D. Shi and G. Zhang, Nat. Electron., 2020, 3, 711–717 CrossRef CAS.
  70. X. Wei, X. Zhang, H. Yu, L. Gao, W. Tang, M. Hong, Z. Chen, Z. Kang, Z. Zhang and Y. Zhang, Nat. Electron., 2024, 7, 138–146 CrossRef.
  71. P. J. Jeon, J. S. Kim, J. Y. Lim, Y. Cho, A. Pezeshki, H. S. Lee, S. Yu, S. W. Min and S. Im, ACS Appl. Mater. Interfaces, 2015, 7, 22333–22340 CrossRef CAS PubMed.

Footnote

Y. L. and Q. D. contributed equally to this work.

This journal is © The Royal Society of Chemistry 2026
Click here to see how this site uses Cookies. View our privacy policy here.