Metal−insulator transition in an isoelectronically doped transition metal dichalcogenide and its heterostructures

Yutang Hou , Yu Wang , Lei Yin , Yao Wen , Xinjie Hou , Ruiqing Cheng , Jun He

Front. Phys. ›› 2025, Vol. 20 ›› Issue (4) : 044209

PDF (3335KB)
Front. Phys. ›› 2025, Vol. 20 ›› Issue (4) : 044209 DOI: 10.15302/frontphys.2025.044209
RESEARCH ARTICLE

Metal−insulator transition in an isoelectronically doped transition metal dichalcogenide and its heterostructures

Author information +
History +
PDF (3335KB)

Abstract

Due to the limitations of traditional silicon-based semiconductors at the nanoscale, such as short-channel effects and quantum effects, two-dimensional (2D) transition metal dichalcogenides (TMDs) like MoS2 and MoTe2 are increasingly recognized for their remarkable characteristics. These materials exhibit unique properties, including tunable bandgaps and the ability to mitigate electron scattering. The metal−insulator transition (MIT), a special electrical property found in some 2D materials, holds great potential for various applications. The MIT in TMDs can be induced through external parameters, but challenges like charge inhomogeneity and the detrimental effects of ionic liquid gating complicate device fabrication and measurement. In this work, we report the MIT behavior in an isoelectronic doped transition metal dichalcogenide MoS2(1−x)Se2x. By studying the dependence of conductivity on temperature in MoS2(1−x)Se2x field-effect transistors employing a single back-gate device structure, we observe clear evidence of the metal−insulator transition in the electron carriers. More importantly, we demonstrate that this MIT behavior can be replicated in other 2D material systems that lack such properties by heterostructure engineering. Our research lays the foundation for further enhancing the performance of 2D materials and may lead to broader applications in functional electronic devices.

Graphical abstract

Keywords

2D semiconductors / metal−insulator transition / 2D transistors / van der Waals heterostructures / isoelectronic doping

Cite this article

Download citation ▾
Yutang Hou, Yu Wang, Lei Yin, Yao Wen, Xinjie Hou, Ruiqing Cheng, Jun He. Metal−insulator transition in an isoelectronically doped transition metal dichalcogenide and its heterostructures. Front. Phys., 2025, 20(4): 044209 DOI:10.15302/frontphys.2025.044209

登录浏览全文

4963

注册一个新账户 忘记密码

1 Introduction

Due to the performance limitations encountered by traditional silicon-based semiconductor materials at the nanoscale, such as short-channel effects, increased power consumption, and quantum effects, it has become increasingly difficult to sustain Moore’s Law. Transition metal dichalcogenides (TMDs) like MoS2 and MoTe2 have many exceptional electrical and physical properties because of their atomic-scale thickness and the absence of surface dangling bonds, which has led to significant interest in recent years [111]. Due to their unique layered structure, where the layers are bound by van der Waals interactions, TMDs can also be reduced to stable monolayer structures through methods such as mechanical exfoliation. This process allows their semiconductor properties to be tuned from an indirect bandgap to a direct bandgap [12]. This characteristic makes certain TMDs complementary to graphene in nanoelectronic applications, as the presence of a bandgap enables efficient charge exchange, facilitating logic and optoelectronic operations [1318]. TMDs are poised to become a powerful alternative for future electronic materials, as they mitigate the significant issue of electron surface scattering in channel materials, which becomes more pronounced as device sizes shrink in silicon-based devices.

The metal−insulator transition (MIT) refers to the transformation of a material’s electronic state from a conductive metallic phase to a non-conductive insulating phase, or vice versa. This phenomenon is particularly significant in the study of two-dimensional (2D) materials, as it is much more challenging to achieve MIT in traditional 3D materials due to strong inter-electron interactions and the rigid constraints of their crystal structures. In contrast, in TMDs, the atomic thickness and more flexible electronic structure enable MIT to be induced through external control parameters such as electric fields, temperature, or carrier concentration [1924]. To achieve the MIT, it is usually necessary to tune the chemical potential near the valence or conduction bands in experiments. Monolayer MoS2 has an electronic bandgap of over 2.0 eV when placed on weakly interacting substrates, making it difficult for back-gated devices to induce the MIT transition before breakdown occurs [25, 26]. Previous studies have reported the induction of a shift from an insulating phase to a conductive phase in MoS2 using a dual-gate configuration in a field-effect transistor [27], but these works often require covering the surface of the 2D materials with high-κ dielectric layers to enhance carrier concentration. Furthermore, according to previous reports, the use of ionic liquid gating can also induce the MIT. However, ionic liquid gates often suffer from issues related to charge inhomogeneity, and the freezing of the liquid at low temperatures exacerbates this problem [28]. The ionic liquid gate generates substantial stress on the channel material during temperature fluctuations, which can compromise its structural integrity and damage the device. Furthermore, the potential side reactions between the ionic liquid and the channel material, coupled with the device’s sensitivity to moisture, further complicate both the fabrication and measurement processes of the device. Additionally, the liquid can interfere with technical measurements of the sample surface.

Here, we report a method to induce the metal-insulator transition in an isoelectronic doped transition metal dichalcogenide MoS2(1−x)Se2x, in which the sulfur (S) atoms were substituted with isoelectronic selenium (Se) atoms. By studying the dependence of conductivity on temperature in MoS2(1−x)Se2x field-effect transistors with a single back-gate device structure, we observe clear evidence of the metal−insulator transition in the electron carriers. More importantly, we demonstrate that this MIT behavior can be replicated in other 2D material systems that lack such properties by heterostructure engineering. This also reduces the Schottky barrier of MoTe2 while increasing its carrier concentration and carrier mobility, opening a new path for further enhancing the performance of 2D semiconductor materials.

2 Experimental methods

2.1 Material characterizations

Few-layer MoS2(1−x)Se2x flakes were prepared by mechanically peeling off the bulk crystal (2D semiconductor). The morphology of the few-layer MoS2(1−x)Se2x flakes were analyzed using optical microscopy (OM, Olympus BX51M) and atomic force microscopy from Dimension Icon. Atomic-resolution high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) images were acquired using JEOL ARM200F microscope, with an acceleration voltage of around 80 kV. Raman spectroscopy was conducted with a LabRAM HR Evolution instrument.

2.2 FET fabrication and electrical measurements

First, few-layer MoS2(1−x)Se2x flakes were exfoliated onto 300 nm SiO2/Si substrates using the conventional typical scotch method. Then, We exposed and developed the source and drain regions using standard electron beam lithography (EBL), followed by depositing 5/50 nm Cr/Au through metal thermal evaporation. The electrical performance of the fabricated FET was measured using a probe station with a vacuum pump, in conjunction with a Keithley 4200 semiconductor analyzer. As for the heterostructure devices, MoTe2 flakes were first exfoliated onto the 300 nm SiO2/Si substrates. MoS2(1−x)Se2x flake was mechanically peeled onto the PPC-coated substrates, followed by its precise transfer onto the underlying MoTe2. The sample was heated at 115 °C for 12 minutes, followed by dissolving the PPC layer with acetone. Same methods as above were adopted for the device fabrication and electrical measurements.

3 Results and discussion

A Ball-and-stick model representation of the MoS2(1−x)Se2x, where Se atoms randomly substitute the positions of S atoms, is shown in Fig.1(a). Raman spectroscopy was performed on the sample to confirm this isoelectronic substitution. As shown in Fig.1(b), the Raman peaks at 271 cm−1, 402 cm−1, and 376 cm−1 correspond to the MoSe2-like A1g mode, MoS2-like A1g mode, and MoS2-like E2g mode, respectively, which are consistent with the previous reports [29]. Fig.1(c) gives the corresponding PL spectrum, and the bandgap emission energy of MoS2(1−x)Se2x is approximately 1.7 eV (i.e., 714 nm), which is smaller than that of MoS2 (~1.86 eV) [30]. Fig.1(d) displays the HAADF-STEM image of MoS2(1−x)Se2x, where the yellow arrows indicate the positions of Se atoms. The clearly defined hexagonal atomic arrangement shows distinct contrast, further revealing the substitution of S atoms by Se atoms. Fig.1(e) shows the selected area electron diffraction (SAED) pattern with hexagonal symmetry, revealing the high crystallinity and robust atomic configuration of MoS2(1−x)Se2x [31]. Additionally, the atomic composition of MoS2(1−x)Se2x is displayed in the energy-dispersive X-ray spectroscopy (EDS) in Fig.1(f), where the elemental percentages of Mo, S, and Se are 22.68%, 28.43%, and 48.88%, respectively. More EDS measurement information is provided in Fig. S1.

FETs were fabricated to study the electrical properties of MoS2(1−x)Se2x flakes, as shown schematically in Fig.2(a). The optical and AFM images of the device are given in Fig. S2. The transport characteristics of MoS2(1−x)Se2x device were first measured at 300 K under different VDS, as shown in Fig.2(b), exhibiting outstanding n-type characteristics and excellent gate tunability with On/Off current ratio ranging from 108 to 109. The output characteristics were also measured [Fig.2(c)], where the typical linear characteristic curve reveals good ohmic contact between the electrodes and MoS2(1−x)Se2x [32].

Fig.3(a) shows the transfer characteristics of MoS2(1−x)Se2x back-gate field-effect transistor at different temperatures. IDS rises as the temperature increases at low VGS, indicating the insulating behavior of MoS2(1−x)Se2x. When VGS reaches approximately 45 V, MoS2(1−x)Se2x transitions into a metallic state, and IDS increases as the temperature decreases. This MIT transition is manifested by the crossing of IDSVGS curves obtained at different temperatures, indicating the existence of two distinct states [33]. In addition, the threshold voltage (VTH) of the MoS2(1−x)Se2x FET also increases as the temperature decreases, and the carrier concentration increases as indicated by

n2 D= CO X (VG SVT H )e .

This further reveals the transition of MoS2(1−x)Se2x to the metallic state [Fig.3(b)]. Figure S3 shows the IDSVDS curves for the MoS2(1−x)Se2x FET at different temperatures, with the typical linear output characteristic revealing good ohmic contact between MoS2(1−x)Se2x and the electrodes. Additionally, at VGS = 80 V, the current decreases with increasing temperature, which also reveals the MIT in MoS2(1−x)Se2x.

To quantitatively obtain the Schottky barrier, we performed temperature-dependent electrical measurements between 200 K and 320 K. According to the thermionic emission theory, IDS can be expressed as [34, 35]

I2 D=A2DT1.5e( Φ B kBT)[1 e (VD SkBT)],

where A*2D = q(8πkB3m*)0.5/h2 is the Richardson constant (where m* is the effective electron mass), q represents the elementary charge, ΦB is the effective contact barrier height, kB denotes the Boltzmann constant, and T is the temperature. To determine the Schottky barrier height (SBH), the values of ln(IDS/T1.5) at various VGS under VDS = 1 V were plotted on an Arrhenius plot [(Fig.3(c)]. The barrier height was obtained by extracting the slope of the Arrhenius curve [Fig.3(d)]. ΦB varies linearly with VGS when VGS is less than approximately −31 V. With further increase in VGS, the plot line satisfies the flat-band voltage condition at the deviation from the linear relationship, corresponding to an extracted Schottky barrier height (SBH) of 118 meV.

Fig.3(e) shows the relationship between the conductivity σ (in units of e2/h) and temperature under different VGS in the range of 10 V to 80 V. At lower VGS, the σ increases as the temperature decreases, corresponding to the insulating behavior. As the VGS increases, the σ gradually exhibits a phenomenon where it first increases and then decreases with the decreasing temperature. When VGS reaches 80 V, the conductivity increases with the decreasing temperature, corresponding to metallic behavior. This trend is also consistent with Fig.2(a). Furthermore, as VGS increases, the sample conductivity approaches the quantum conductivity e2/h, which is the minimum value for metallic conductivity. This indicates that with the increase of VGS, not only does the sample’s conductivity approach metallic behavior, but its temperature-dependent conductivity behavior also gradually shifts from insulating to metallic behavior [36, 37].

The conductivity is defined as G = IDSL/(VDSW), where L and W represent the channel length and width, respectively. We measured the change in the sample conductivity G with temperature under a gate voltage range of −20 V to 20 V, and found that within the temperature range of 80 K to 320 K, the sample conductivity G follows the thermally activated transport model:

G=G0eEa kBT,

where Ea is the energy barrier that electrons need to overcome in order to transition from the Fermi level to the conduction band, and G0 is the relevant parameter extracted from the fitting curve [38]. By plotting the logarithm of G as the vertical axis and 1/T as the horizontal axis, a fit of G was performed. The fitting line shows a good agreement with the experimental data of G [Fig.3(f)], indicating that the experimental data follow the thermally activated transport model well, which suggests that the charge transport in the sample is thermally activated [39]. Studies on the mobility also show that the mobility follows a μTγ dependence above 200 K, consistent with the electron−phonon scattering transport model (Fig. S4) [40]. The thermal activation energy Ea at various VGS values was obtained by analyzing the slope of the fitted line according to the thermal activation model equation. As shown in Fig. S5, Ea decreases as the VGS increases, making it easier for carriers to be excited, which in turn increases the conductivity of the sample.

In addition to observing the MIT transition in MoS2(1−x)Se2x flakes within a single back-gate device structure, we also discovered that this MIT behavior can be replicated in MoTe2, a material that does not inherently exhibit such properties, through heterostructure engineering. Fig.4(a) shows the optical image and schematic diagram of the bare MoTe2 device and the MoTe2 device overlapped with MoS2(1−x)Se2x flake (abbreviated as MoTe2−MoS2(1−x)Se2x device). Fig.4(b) and (d) show the relationship between conductivity and back-gate voltage for both the MoTe2 device and MoTe2−MoS2(1−x)Se2x device, measured at temperatures ranging from 80 K to 320 K with a fixed bias voltage VDS = 1 V. The transfer curves of MoTe2−MoS2(1−x)Se2x device cross at different temperatures, a typical hallmark of MIT, while the bare MoTe2 device does not exhibit this behavior. This indicates that the presence of MoS2(1−x)Se2x flake could duplicate its MIT behavior to other 2D material systems that have no such properties themselves by heterostructure engineering. As a control, we also measured the electrical properties of the independent MoTe2 device to rule out the lateral effect of MoS2(1−x)Se2x on MoTe2 (Fig. S6). Furthermore, we measured the electrical properties of MoS2(1−x)Se2x in the MoTe2−MoS2(1−x)Se2x device, and the results indicated that MoS2(1−x)Se2x still retains this metal−insulator transition behavior (Fig. S7).

Fig.4(c) and (e) are the Arrhenius plots for the MoTe2 and MoTe2−MoS2(1−x)Se2x device, respectively. The corresponding Schottky barriers were extracted in Fig.4(f) and (g). Compared to the bare MoTe2, the MoTe2 device overlapped with MoS2(1−x)Se2x flake shows a significantly reduced Schottky barrier. Fig.4(h) shows the transfer curves for the MoTe2 device and MoTe2−MoS2(1−x)Se2x device at room temperature (300 K) with a fixed bias voltage VDS = 1 V. The linear portion of the curve was fitted to obtain the threshold voltage VTH for both devices. It was found that the additional MoS2(1−x)Se2x results in a lower threshold voltage. The electron concentration can be calculated using the parallel plate capacitor model:

n2 D= CO XΔVq,

where n2D is the electron concentration, COX = ε 0 εr/d, ε 0 is the vacuum permittivity, ε r represents the relative dielectric constant of the silicon dioxide layer on the back gate, d denotes the thickness of the SiO2, while ΔV = VGSVTH. By measuring the threshold voltage VTH, the electron concentration can be obtained. The MoTe2−MoS2(1−x)Se2x device exhibits a lower threshold voltage VTH, signifying a higher electron concentration. Moreover, the MoTe2−MoS2(1−x)Se2x device also demonstrates a considerably lower gate voltage at the minimum current, which further corroborates its higher carrier concentration (Fig. S8). In addition, we fabricated more MoTe2−MoS2(1−x)Se2x devices to verify the universality of the metal−insulator transition behavior (Fig. S9).

To explain this phenomenon, we evaluated the conduction band position of the MoTe2−MoS2(1−x)Se2x device. At a gate voltage of VGS = 45 V and room temperature T = 300 K, we calculated the 2D electron concentration n2D = 3.17 × 1012 cm−2. The thickness of the 2D electron gas layer is less than 2 nm, so the 3D electron concentration at the heterojunction interface is n3D>1.58 × 1019 cm−3.

The electronic density of states at the lower edge of the conduction band for MoS2(1−x)Se2x is given by the equation:

NC= 2( 2π m e kBTh2) 32=9× 1018c m 3,

where the effective electron mass for MoS2(1−x)Se2x m e = 0.521m0 [41], the Planck constant h is 6.626 × 10−34 Js.

The relationship between the electron concentration n3D and the density of states NC is expressed as

n3 D=NCeEC EF k BT.

When n3D > NC, it indicates that EF > EC. At the MoTe2−MoS2(1−x)Se2x heterojunction interface, due to band bending, the minimum of the conduction band of MoS2(1−x)Se2x is positioned below the Fermi level, enabling electrons to populate the conduction band and consequently increasing the electron concentration at the interface. The 2D electron gas at the interface caused by band bending and electron filling should be the key to the property change brought about by MoS2(1−x)Se2x.

In the context of the MIT in MoS2(1−x)Se2x, an increase in gate voltage or temperature results in a rise in electron concentration, prompting the system to shift progressively from the insulating state to the metallic state. The heterojunction of MoTe2−MoS2(1−x)Se2x also gives rise to a higher electron concentration, the increase in electron concentration is always closely related to the MIT phenomenon, indicating that the enhancement of electron concentration has a pivotal impact on the occurrence of MIT, which might be the underlying cause of the MIT phenomenon in the MoTe2−MoS2(1−x)Se2x device [42].

To characterize the Coulomb interactions among electrons in the material, the ratio [43] of the electron potential energy to kinetic energy is defined as follows:

rs= ECEF= n V a B πn2D= nV me e24πε 2π n2 D,

where nV represents the number of degenerate valleys, a B=( 4πε 2)/( m e e2) is the effective Bohr radius, m e is the effective electron mass, and ε is the system’s dielectric constant. When ε lies within 1 to 3 times that of MoS2, the value of rs varies from 8.55 to 2.85, signifying strong Coulomb interactions among the electrons. Typically, electrons in metals are in delocalized states and the interactions among them are mutually screened. In contrast, electrons in insulators tend to be in localized states. Nevertheless, for the MoTe2−MoS2(1−x)Se2x device, the electrons can possess rather potent interactions and concomitantly experience a transition from an insulating state to a metallic state, which implies that the electron interactions in this system are fundamentally distinct from those in common metals. Such strong interactions can prompt the system to shift from a strongly localized insulating state to a weakly localized metallic state [44].

Fig.4(i) shows the mobility μ of the MoTe2 device and the MoTe2−MoS2(1−x)Se2x device at different temperatures under VDS = 1 V. The mobility μ is calculated using the formula:

μ=L( dID S dVG S)WC O X VD S.

The mobility can be determined from the output curve. The incorporation of MoS2(1−x)Se2x remarkably enhances the mobility. As the temperature rises, the mobility of the MoTe2−MoS2(1−x)Se2x device first increases and then decreases. At elevated temperatures, the device exhibits metallic behavior, while the MoTe2 device solely demonstrates insulating properties. The change in the mobility of the MoTe2−MoS2(1−x)Se2x device at higher temperatures conforms to the electron−phonon scattering model, indicating that electron−phonon scattering becomes the dominant scattering mechanism at higher temperatures. Under the mode of electron−phonon scattering, the variation of the mobility with temperature follows the relationship μ~Tγ, where the exponent γ depends on the dominant phonon scattering mechanism. Therefore, we used the general formula of the temperature dependence of the mobility μ~Tγ to fit the part of the curve at higher temperatures. Through fitting the curve, we found that γ ≈ 1.01. This stark contrast in temperature-dependent behavior between the MoTe2−MoS2(1−x)Se2x device and MoTe2 devices correlates with the occurrence or absence of the MIT phenomenon, thereby confirming that MoS2(1−x)Se2x imparts MIT characteristics to MoTe2. In addition, this also reflects that MoS2(1−x)Se2x has brought about performance enhancements in various aspects to the MoTe2 device, providing insights for other research directions beyond the study of the MIT phenomenon, such as device performance.

4 Conclusion

To summarize, we propose a strategy for inducing the MIT in isoelectronic doped transition metal dichalcogenide MoS2(1−x)Se2x by substituting sulfur atoms with isoelectronic selenium atoms. The dependence of MIT on gate voltage and temperature in MoS2(1−x)Se2x was also investigated. In the study, the observed trend of increasing electron concentration coincided with the occurrence of the MIT, indicating that electron interactions play a crucial role in this transition. More importantly, we demonstrate that this MIT behavior can be replicated in other 2D material systems, which do not inherently exhibit such properties, through heterostructure engineering. The difference in electron concentration between MoTe2 and MoS2(1−x)Se2x devices suggests that the inclusion of MoS2(1−x)Se2x results in a higher electron concentration. Strong electron interactions within the system account for the observed MIT. Furthermore, the integration of MoS2(1−x)Se2x reduces the Schottky barrier, enhancing the device’s mobility and carrier concentration, thereby improving its overall performance. This work contributes to a deeper understanding of the MIT phenomenon in 2D transition metal dichalcogenide systems and provides valuable strategies for improving the efficiency of 2D material-based transistors.

References

[1]

L. Yin, R. Cheng, J. Ding, J. Jiang, Y. Hou, X. Feng, Y. Wen, and J. He, Two-dimensional semiconductors and transistors for future integrated circuits, ACS Nano 18(11), 7739 (2024)

[2]

Q. H. Wang, K. Kalantar-Zadeh, A. Kis, J. N. Coleman, and M. S. Strano, Electronics and optoelectronics of two-dimensional transition metal dichalcogenides, Nat. Nanotechnol. 7(11), 699 (2012)

[3]

M. Chhowalla, D. Jena, and H. Zhang, Two-dimensional semiconductors for transistors, Nat. Rev. Mater. 1(11), 16052 (2016)

[4]

M. Y. Li, S. K. Su, H. S. P. Wong, and L. J. Li, How 2D semiconductors could extend Moore’s law, Nature 567(7747), 169 (2019)

[5]

S. Manzeli,D. Ovchinnikov,D. Pasquier,O. V. Yazyev,A. Kis, 2D transition metal dichalcogenides, Nat. Rev. Mater. 2(8), 17033 (2017)

[6]

K. S. Novoselov,A. Mishchenko,A. Carvalho,A. H. Castro Neto, 2D materials and van der Waals heterostructures, Science 353(6298), aac9439 (2016)

[7]

T. Song, X. Cai, M. W. Y. Tu, X. Zhang, B. Huang, N. P. Wilson, K. L. Seyler, L. Zhu, T. Taniguchi, K. Watanabe, M. A. McGuire, D. H. Cobden, D. Xiao, W. Yao, and X. Xu, Giant tunneling magnetoresistance in spin-filter van der Waals heterostructures, Science 360(6394), 1214 (2018)

[8]

D. Akinwande, C. Huyghebaert, C. H. Wang, M. I. Serna, S. Goossens, L. J. Li, H. S. P. Wong, and F. H. L. Koppens, Graphene and two-dimensional materials for silicon technology, Nature 573(7775), 507 (2019)

[9]

R. Cheng, L. Yin, R. Hu, H. Liu, Y. Wen, C. Liu, and J. He, Modulation of negative differential resistance in black phosphorus transistors, Adv. Mater. 33(25), 2008329 (2021)

[10]

L. Yin, P. He, R. Cheng, F. Wang, F. Wang, Z. Wang, Y. Wen, and J. He, Robust trap effect in transition metal dichalcogenides for advanced multifunctional devices, Nat. Commun. 10(1), 4133 (2019)

[11]

R. Yang, J. Fan, and M. Sun, Transition metal dichalcogenides (TMDCs) heterostructures: Optoelectric properties, Front. Phys. (Beijing) 17(4), 43202 (2022)

[12]

C. H. Lui, K. F. Mak, J. Shan, and T. F. Heinz, Ultrafast photoluminescence from graphene, Phys. Rev. Lett. 105(12), 127404 (2010)

[13]

L. Yin, R. Cheng, X. Wan, J. Ding, J. Jia, Y. Wen, X. Liu, Y. Guo, and J. He, High-κ monocrystalline dielectrics for low-power two-dimensional electronics, Nat. Mater. 24(2), 197 (2025)

[14]

R. Cheng, F. Wang, L. Yin, Z. Wang, Y. Wen, T. A. Shifa, and J. He, High-performance, multifunctional devices based on asymmetric van der Waals heterostructures, Nat. Electron. 1(6), 356 (2018)

[15]

H. Liu, A. T. Neal, and P. D. D. Ye, Channel Length Scaling of MoS2 MOSFETs, ACS Nano 6(10), 8563 (2012)

[16]

Z. Y. Yin, H. Li, H. Li, L. Jiang, Y. M. Shi, Y. H. Sun, G. Lu, Q. Zhang, X. D. Chen, and H. Zhang, Single-layer MoS2 phototransistors, ACS Nano 6(1), 74 (2012)

[17]

G. Eda and S. A. Maier, Two-dimensional crystals: Managing light for optoelectronics, ACS Nano 7(7), 5660 (2013)

[18]

Z. Wang, Y. Yang, B. Hua, and Q. Ji, Synthetic two-dimensional electronics for transistor scaling, Front. Phys. (Beijing) 18(6), 63601 (2023)

[19]

J. T. Ye, Y. J. Zhang, R. Akashi, M. S. Bahramy, R. Arita, and Y. Iwasa, Superconducting dome in a gate-tuned band insulator, Science 338(6111), 1193 (2012)

[20]

B. H. Moon, J. J. Bae, G. H. Han, H. Kim, H. Choi, and Y. H. Lee, Anomalous conductance near percolative metal−insulator transition in monolayer MoS2 at low voltage regime, ACS Nano 13(6), 6631 (2019)

[21]

X. Chen, Z. Wu, S. Xu, L. Wang, R. Huang, Y. Han, W. Ye, W. Xiong, T. Han, G. Long, Y. Wang, Y. He, Y. Cai, P. Sheng, and N. Wang, Probing the electron states and metal−insulator transition mechanisms in molybdenum disulphide vertical heterostructures, Nat. Commun. 6(1), 6088 (2015)

[22]

J. H. Lee, H. Z. Gul, H. Kim, B. H. Moon, S. Adhikari, J. H. Kim, H. Choi, Y. H. Lee, and S. C. Lim, Photocurrent switching of monolayer MoS2 using a metal−insulator transition, Nano Lett. 17(2), 673 (2017)

[23]

C. L. Wu, H. Yuan, Y. Li, Y. Gong, H. Y. Hwang, and Y. Cui, Gate-induced metal−insulator transition in MoS2 by solid superionic conductor LaF3, Nano Lett. 18(4), 2387 (2018)

[24]

Y. Li, M. Yang, Y. Lu, D. Cao, X. Chen, and H. Shu, Reversible doping polarity and ultrahigh carrier density in two-dimensional van der Waals ferroelectric heterostructures, Front. Phys. (Beijing) 18(3), 33307 (2023)

[25]

P. V. Nguyen, N. C. Teutsch, N. P. Wilson, J. Kahn, X. Xia, A. J. Graham, V. Kandyba, A. Giampietri, A. Barinov, G. C. Constantinescu, N. Yeung, N. D. M. Hine, X. Xu, D. H. Cobden, and N. R. Wilson, Visualizing electrostatic gating effects in two-dimensional heterostructures, Nature 572(7768), 220 (2019)

[26]

Z. Qiu, M. Trushin, H. Fang, I. Verzhbitskiy, S. Gao, E. Laksono, M. Yang, P. Lyu, J. Li, J. Su, M. Telychko, K. Watanabe, T. Taniguchi, J. Wu, A. H. C. Neto, L. Yang, G. Eda, S. Adam, and J. Lu, Giant gate-tunable bandgap renormalization and excitonic effects in a 2D semiconductor, Sci. Adv. 5(7), eaaw2347 (2019)

[27]

B. Radisavljevic and A. Kis, Mobility engineering and a metal−insulator transition in monolayer MoS2, Nat. Mater. 12(9), 815 (2013)

[28]

Y. Ren, H. Yuan, X. Wu, Z. Chen, Y. Iwasa, Y. Cui, H. Y. Hwang, and K. Lai, Direct imaging of nanoscale conductance evolution in ion-gel-gated oxide transistors, Nano Lett. 15(7), 4730 (2015)

[29]

Q. Feng, Y. Zhu, J. Hong, M. Zhang, W. Duan, N. Mao, J. Wu, H. Xu, F. Dong, F. Lin, C. Jin, C. Wang, J. Zhang, and L. Xie, Growth of large-area 2D MoS2(l−x)Se2x semiconductor alloys, Adv. Mater. 26(17), 2648 (2014)

[30]

K. F. Mak, C. Lee, J. Hone, J. Shan, and T. F. Heinz, Atomically thin MoS2: A new direct-gap semiconductor, Phys. Rev. Lett. 105(13), 136805 (2010)

[31]

M. Li,J. Yao,X. Wu,S. Zhang,B. Xing, X. Niu,X. Yan,Y. Yu,Y. Liu,Y. Wang, P-type doping in large-area monolayer MoS2 by chemical vapor deposition, ACS Appl. Mater. Interfaces 12(5), 6276 (2020)

[32]

J. Gao, Y. D. Kim, L. Liang, J. C. Idrobo, P. Chow, J. Tan, B. Li, L. Li, B. G. Sumpter, T. M. Lu, V. Meunier, J. Hone, and N. Koratkar, Transition-metal substitution doping in synthetic atomically thin semiconductors, Adv. Mater. 28(44), 9735 (2016)

[33]

I. Castillo, T. Sohier, M. Paillet, D. Cakiroglu, C. Consejo, C. Wen, F. W. Klein, M. Q. Zhao, A. Ouerghi, S. Contreras, A. T. C. Johnson, M. J. Verstraete, B. Jouault, and S. Nanot, Metal−insulator crossover in monolayer MoS2, Nanotechnology 34(33), 335202 (2023)

[34]

X. Cui, G. H. Lee, Y. D. Kim, G. Arefe, P. Y. Huang, C. H. Lee, D. A. Chenet, X. Zhang, L. Wang, F. Ye, F. Pizzocchero, B. S. Jessen, K. Watanabe, T. Taniguchi, D. A. Muller, T. Low, P. Kim, and J. Hone, Multi-terminal transport measurements of MoS2 using a van der Waals heterostructure device platform, Nat. Nanotechnol. 10(6), 534 (2015)

[35]

L. Yu, Y. H. Lee, X. Ling, E. J. G. Santos, Y. C. Shin, Y. Lin, M. Dubey, E. Kaxiras, J. Kong, H. Wang, and T. Palacios, Graphene/MoS2 hybrid technology for large-scale two-dimensional electronics, Nano Lett. 14(6), 3055 (2014)

[36]

S. Das Sarma, M. P. Lilly, E. H. Hwang, L. N. Pfeiffer, K. W. West, and J. L. Reno, Two-dimensional metal−insulator transition as a percolation transition in a high-mobility electron system, Phys. Rev. Lett. 94(13), 136401 (2005)

[37]

H. J. Chuang, X. Tan, N. J. Ghimire, M. M. Perera, B. Chamlagain, M. M. C. Cheng, J. Yan, D. Mandrus, D. Tomanek, and Z. Zhou, High mobility WSe2 p- and n-type field-effect transistors contacted by highly doped graphene for low-resistance contacts, Nano Lett. 14(6), 3594 (2014)

[38]

B. Radisavljevic, A. Radenovic, J. Brivio, V. Giacometti, and A. Kis, Single-layer MoS2 transistors, Nat. Nanotechnol. 6(3), 147 (2011)

[39]

S. Ghatak, A. N. Pal, and A. Ghosh, Nature of electronic states in atomically thin MoS2 field-effect transistors, ACS Nano 5(10), 7707 (2011)

[40]

K. Kaasbjerg, K. S. Thygesen, and K. W. Jacobsen, Phonon-limited mobility in n-type single-layer MoS2 from first principles, Phys. Rev. B 85(11), 115317 (2012)

[41]

M. Farkous, M. El-Yadri, H. Erguig, L. M. Perez, D. Laroze, C. V. Nguyen, N. T. T. Binh, N. N. Hieu, H. V. Phuc, M. Sadoqi, G. Long, and E. Feddi, Anisotropy of effective masses induced by strain in Janus MoSSe and WSSe monolayers, Physica E 134, 114826 (2021)

[42]

A. Punnoose and A. M. Finkel’stein, Metal−insulator transition in disordered two-dimensional electron systems, Science 310(5746), 289 (2005)

[43]

L. H. Ho, A. P. Micolich, A. R. Hamilton, and O. P. Sushkov, Ground-plane screening of Coulomb interactions in two-dimensional systems: How effectively can one two-dimensional system screen interactions in another, Phys. Rev. B 80(15), 155412 (2009)

[44]

E. Abrahams, P. W. Anderson, D. C. Licciardello, and T. V. Ramakrishnan, Scaling theory of localization: Absence of quantum diffusion in two dimensions, Phys. Rev. Lett. 42(10), 673 (1979)

RIGHTS & PERMISSIONS

Higher Education Press

AI Summary AI Mindmap
PDF (3335KB)

Supplementary files

fop-25076-of-chengruiqing_suppl_1

1360

Accesses

0

Citation

Detail

Sections
Recommended

AI思维导图

/