Introduction

Field-effect transistors (FETs) — three-terminal systems consisting of source, drain and gate electrodes — are integral in many electronic devices, allowing them to achieve energy-efficient high-speed switching. Semiconducting materials, also known as the channel of a FET, span the source and drain electrodes. The channel is electrically isolated from the gate electrode by the gate dielectric. The effective operation of FETs relies on efficient electrostatic coupling between the electric field induced by the gate voltage and the channel, without allowing electrons to flow between them.

Traditional FETs are, most often, based on bulk or 3D semiconductor channels composed of silicon and the III–V semiconductors, GaAs and GaN. These 3D materials have been successfully scaled down to nanoscale dimensions over the past 5 decades following Moore's law, which in 1965 predicted that the density of transistors in a chip will double every 2 years14. The semiconductor industry has established long-term strategies and research goals based on Moore's Law, but this prediction has recently started to falter and substitutes are under consideration5.

Short-channel effects

The dimensions of FETs, and hence the lengths of the channels, continue to decrease in the quest for higher performance. Devices with shorter channel lengths have begun to experience high off-state currents; that is, some charge carriers are able to flow between the source and drain electrodes even on application of a gate voltage that suppreses the flow of current. The presence of an off-state current increases the static power (the product of the drain voltage and off-state current). As a result, the reduction in size of FETs has been limited by the rate at which heat, caused by static power, is dissipated. The current leakage and the resulting challenges associated with heat dissipation caused by an increase in static power are collectively termed short-channel effects6,7. The consequences of short-channel effects are detrimental for device operation and energy efficiency to the extent that the International Technology Roadmap for Semiconductors (ITRS) predicts that transistor densities will double every 3 years rather than every 2 years8. To address these short-channel effects, devices such as multiple gate transistors9,10, FinFETs11,12 and ultrathin body transistors (UTB)1315 are being actively pursued.

There is a need to identify new semiconductor materials that can mitigate short-channel effects. In addition, these materials should be compatible with existing complementary metal oxide semiconductor (CMOS) infrastructure. A starting point for identifying semiconductors for transistor channels is the examination of the FET electrostatics described by Poisson's equation. This equation yields a characteristic channel ‘scaling length’ given by λ= t s t b ( ε s / ε b ) , where ts is the semiconductor thickness, tb is the gate dielectric thickness, and ϵs and ϵb are the semiconductor and gate insulator dielectric constants, respectively6,7. In reality, UTB semiconductors made from decreasing the thickness of 3D (bulk) materials suffer from dangling bonds, leading to scattering of the charge carriers16,17 (Fig. 1a). As a result, the expected enhancement in gate electrostatic properties and substantial degradation of FET performance are observed. More specifically, the mobility, μ, of charge carriers decreases with thickness to the sixth power, μ t6, and the bandgap, Eg, increases by the square of the thickness, ΔEg t2 (Ref. 16). For transistors made from UTB 3D semiconductors, the substantial decrease in performance is observed because of the presence of dangling bonds, undesirable coupling with phonons and the creation of interface states (that occur when the gate dielectric and the source/drain electrodes are deposited on top of the semiconductor). In 3D semiconductor short-channel FETs, current leakage arises from poor electrostatics between some electrons in the channel and the electric field applied by the gate (Fig. 1b). However, in FETs that have a channel made from 2D semiconductors, all electrons are confined in naturally atomically thin channels and, hence, all carriers are uniformly influenced by the gate voltage17 (Fig. 1d). This excellent gate coupling allows the suppression of current leakage if a gate voltage is applied.

Figure 1: Advantages of 2D materials compared with 3D materials for FETs.
figure 1

a | Ultrathin 3D (bulk) semiconductors have dangling bonds that form traps for electrons and reduce the performance of field-effect transistors (FETs). b | Gate electrostatics and mobile charge distribution in 3D semiconductors. V(x) is the gate voltage as a function of the distance x. The potential barrier is the energy level of the gate dielectric and |ψ(x)|2 refers to the probability function of the carriers in the semiconductors. The majority of the mobile charge carriers are located approximately 1.2 nm from the semiconductor gate dielectric interface. c | By contrast, 2D materials have pristine surfaces. d | In 2D materials, charge carriers are confined in the atomically thin semiconductor, resulting in a narrower mobile charge distribution. This confinement of charge carriers allows the carriers to be easily controlled by the gate voltage, leading to excellent gate electrostatics. Figure is adapted with permission from Ref. 17, SPIE Journals and Proceedings.

PowerPoint slide

2D materials

The range of 2D materials for application as the semiconducting component of FETs includes graphene, hexagonal boron nitride (h-BN), transition metal dichalcogenides (TMDs), silicene and phosphorene. Substantial research has been devoted to graphene, the ‘original’ 2D material, which has revealed interesting electronic1821 and photonic22,23 phenomena. However, the absence of an energy bandgap makes graphene less desirable for use in FET switching settings that require high on-state currents but low off-state currents. Hence, the field has expanded into other 2D materials, predominantly semiconducting TMDs because they possess bandgaps in the range of 1–2 eV. The weak, van der Waals interactions between the layers of the corresponding bulk materials enable atomically thin layers of graphite, molybdenum disulfide (MoS2) and black phosphorus to be isolated relatively easily by exfoliation. This approach usually results in a low density of dangling bonds and minimal surface roughness (Fig. 1c). The subsequent transfer of these 2D materials onto thin-layered insulators — for example, in the case of graphene transferred onto h-BN — almost completely eliminates dangling bonds in the insulator24 and mitigates the effect of charged impurities by changing the dielectric environment in 2D semiconductors25. In another example, devices consisting of six-layered MoS2 (semiconductor) encapsulated by few-layered h-BN (gate insulator) show high mobilities, which reach 34,000 cm2 V−1 s−1 at low temperatures26. Recent progress suggests that such atomically thin 2D materials could be one pathway for electronic devices in the future2729.

Characteristic of FETs

To highlight the advantages of using 2D materials as semiconductors in FETs, the important characteristics of the devices must be covered first (Fig. 2). A typical FET is a three-terminal device consisting of a semiconducting channel between the source and drain electrodes. The current between the source and drain electrodes — passing through the semiconducting channel — is modulated by the application of a gate voltage. The transverse electric field created by the gate voltage can either deplete the channel of carriers so that no current flows between the source and drain electrodes (off-state), or enhance the concentration of carriers in the channel, allowing current to flow (on-state). Ideally, the off-state current is as small as possible and the on-state current is >104 compared with the off-state current (that is, an on/off ratio of >104).

Figure 2: Operating principles of a FET with 2D semiconducting materials forming the channels.
figure 2

a | The cross-section of a typical field-effect transistor (FET) device with source and drain electrodes, a 2D semiconductor channel and a gate electrode that is electrically separated from the channel by an insulator. Vgs is the gate voltage, Vds is the drain bias and VT is the threshold voltage determined by the difference in the work functions of the metal and the semiconductor. b | The energy-band diagram in the vertical direction near the source end of the gate, defining various energy scales and band offsets. b is the metal–dielectric barrier height, EFs is the quasi-Fermi level of the source electrode, qVi is the voltage drop in the insulator, Ec is the energy of the semiconductor conduction band, ΔEc is the conduction band offset and qVgs is the barrier modification by the applied gate voltage. c | A schematic diagram of the switching transfer characteristics of the FET with the drain biased at saturation, showing how the gate voltage controls the electron population in k space in the on and off states. kx,y are the wavevectors that represent the direction of electron wavefunction propagation. df | The energy-band diagram along the channel from the source to the drain electrode, at various drain voltages as indicated in panel g; EFd is the quasi-Fermi level of the drain electrode. g | Output characteristics of the FET. The electron distribution in the k space at the source injection point is shown at different drain voltages. The colours indicate the origin of carriers in the channel: red indicates right-moving carriers injected from the source and blue is for left-moving carriers injected from the drain. At a fixed gate voltage, the net density of the right- and left-going carriers does not change, which is why the area of the two half-circles does not change. The change with the gate voltage is indicated in panel c. At a large drain voltage shown in panel d, it is energetically unfavourable for the drain to inject carriers to the source injection point, which means all carriers at that point are from the source electrode, and the current is saturated beyond this drain voltage (panels d and g).

PowerPoint slide

A FET with a semiconductor channel composed of a single-layer TMD and its operating mechanisms are shown in Fig. 2. The gate barrier is an insulator of thickness tb, the dielectric constant is ϵb and the capacitance is Cb = ϵb/tb. The channel length along the x axis is L and the width along the y axis is W. Mobile electrons in the conduction (or valence) bands of the TMD layer form Bloch waves that must fit in the rectangular channel. This restricts the Bloch waves to kx = nx(2π/L) and ky = ny(2π/W), where nx, ny = …,−1, 0, +1,… are integers, obtained by enforcing periodic boundary conditions, and kx and ky are wave vectors that represent the direction of electron wavefunction propagation in the crystal lattice. The density of mobile carriers in the 2D semiconductor sheet is given by summing all of the occupied allowed electron states over all of the spins and valleys:

(1) n s = ( 1 / W L ) ( k x , k y ) all spin, valley f ( k x , k y )

The central problem of FET physics is determining how two voltages — one between the gate and source electrodes, Vgs, and the other between the drain and source electrodes, Vds — control the occupation of the allowed channel modes, f(kx, ky). The knowledge of this occupation function determines all of the output characteristics of the device. To solve this problem, we make three simplifications without neglecting important features of the FET. First, the gate insulator is ideal and allows no current to leak between the gate and the source and drain electrodes. Second, the contacts are ideal, which means there is no voltage drop at the source and drain contacts. Third, there is no scattering of electrons in the semiconductor channel. On the basis of these assumptions, we develop a detailed quantitative model for the operation of FET-based 2D materials (Supplementary information S1,S2 (box, figure)).

A FET based on 2D semiconductors has several key operating mechanisms30,31. When the source and drain electrodes are grounded (Vds = 0) and a positive voltage is applied on the gate electrode, the gate battery draws electrons from the gate metal and pumps the electrons into the semiconductor channel through the source and drain contacts. The physics of FET operation can be captured by understanding how the electron density of the mobile sheet, ns, in the semiconductor channel depends on the gate voltage. The sheet charge is often assumed to be the gate capacitance, Cb = ϵb/tb, multiplied by the ‘excess’ voltage, qns = Cb(VgsVT), where q is charge and VT is the threshold voltage determined by the difference in the work functions of the metal and the semiconductor. This is an approximation that is only true for a range of voltages when the transistor is deep in the on state — specifically, when VgsVT >> Vth (where the thermal voltage, Vth = kT/q, has a value of 26 mV at room temperature). This assumption fails near VT and for off states of the FET, which may be appreciated by noticing that because ns is the sheet density of filled states, it cannot be negative — however, VgsVT < 0 for Vgs < VT.

The energy levels of the gate metal, insulator and 2D semiconductor along the z direction of the device are plotted in Fig. 2b. The relative position of the energy levels can be represented by B + qVi − ΔEc + (EFsEc) = qVgs, where B is the metal–dielectric barrier height, qVi is the voltage drop in the insulator, ΔEc is the conduction band offset, EFs is the quasi-Fermi level of the source electrode, Ec is the energy of semiconductor conduction band and qVgs is the barrier modification by the applied gate voltage.

The carriers entering the channel through the source move to the right (red in Fig. 2c–f) and are in equilibrium with the energy level of the source electrode. The carriers entering the channel from the drain move to the left (blue in Fig. 2d–f) and are in equilibrium with the drain contact. The carriers cannot change direction in the channels because no scattering of electrons can occur. This results in a modal distribution of electrons in the 2D (kx, ky) space (Fig. 2c–f). By analysing the changes that occur in the energy levels in the presence of applied Vgs and Vds, it is possible to derive the relationship between the carrier density and the gate voltage as,

(2) e n s n b ( e n s n q 1 ) = e V gs V T V th

where ns is the net carrier density, nq is a characteristic 2D quantum concentration defined in the Supplementary information and nb is the characteristic carrier density from the electrostatic capacitance (Supplementary information S1 (box)). This is a transcendental equation, the solution to which gives the mobile carrier density in the semiconductor channel as a function of the gate voltage, ns(Vgs).

Ideal FET output characteristics

We can use the solutions derived in the Supplementary information S1,S2 (box, figure) to calculate the output characteristics of a TMD channel FET in which transport is ballistic — that is, when electrons travel in the channel without being scattered. There are two ways to present the performance of a FET: by the assessment of either transfer characteristics (curve in Fig. 2c; Fig. 3a,b) or output characteristics (curve in Fig. 2g; Fig. 3c). Transfer characteristics are obtained by plotting drain current, Id, as a function of the gate voltage (Fig. 3a,b). The on-state current of FETs can be obtained from transfer characteristics at different source–drain voltages. A subthreshold slope (SS; the gate voltage required to increase the drain current by a factor of ten), which indicates the switching speed of the FET, can be extracted from a transfer curve. In our simulated FET consisting of a 2D channel, the switching properties are notable in that the SS is close to 60 mV dec−1 at 300 K using equation 3 (Ref. 16).

(3) SS = ( 1 + C s C b ) k T q ln10
Figure 3: Calculated FET characteristics and spin–valley locking in the valence band of 2D TMDs.
figure 3

The output characteristics of a transition metal dichalcogenide (TMD) ballistic field-effect transistor (FET) with m c * = 0.5m0, ϵb = 20ϵ0 and tb = 3 nm are calculated (panels ac). a | Transfer characteristics with the drain current, Id, in logarithmic form for two different drain voltages (Vds). b | Transfer characteristics in linear from at different drain voltages. c | Calculated output characteristics of a prototype ballistic 2D semiconductor channel FET at different gate voltages. Note that these calculations do not consider contact resistances. d | The electronic structure of a unit cell — referred to as the first Brillouin zone — of monolayer MoS2. The high symmetry points (K, K′ and Γ) of the unit cell in the momentum space are indicated. The K and K′ points are two inequivalent momentum valleys: the energy maxima and minima. The arrows indicate spins of the valence electrons occupying that valence state. kx,y are the wavevectors that represent the direction of electron wavefunction propagation. e | The valence band splits at these valleys as a consequence of strong spin–orbit coupling. Time reversal symmetry requires spins to be opposite at different valleys. f | Drain current as a function of effective mass for three different values of valley degeneracy (gv). The drain currents were obtained for Vgs and Vds = 0.4 V. The cross-over of current scaling with valleys and effective mass depends on the gate capacitance.

PowerPoint slide

where Cs is the semiconductor capacitance. The current saturates as a function of the drain voltage because it becomes energetically difficult for the drain electrode to inject carriers into the channel (Fig. 2d,g). At a fixed drain voltage, varying the gate voltage changes the number of electron energy levels that are filled at the injection point of the source, because the band edge states are capacitively coupled to the gate electrode. This changes the current exponentially when VgsVT << Vth, and as a power law when VgsVT >> Vth (Refs 30,31). When the transistor is switched on, the drain current in a 2D semiconductor FET is Id Vgs3/2. This is similar to electron transport in a vacuum tube, which is the predecessor of the solid-state switch. This is not a coincidence, because if the transport of electrons is ballistic, the electrons effectively do not see the atoms in their path.

Output characteristics are obtained by plotting Id as a function of Vds for several gate voltages (Fig. 3c). A FET should demonstrate saturation of Id above a certain Vds. This constant current with Vds is important because, in an integrated circuit, several devices are interconnected and the Vds being supplied to each device is variable, but the current from each device will be the same because of the saturation behaviour. In digital circuits, the Id of a FET has to drive the gates of a few other FETs. This is called fan-out and depends on the saturation of the Id of a FET above a certain Vds. If saturation does not occur, Vds will oscillate, and the Id and the output current of the FET will change. This causes failure of fan-out and reduces the gain of the integrated circuit. The output current of a FET should only depend on the input (gate) voltage and not on the voltage of the integrated circuit. This unidirectionality is at the heart of gain and the practical utility of digital operations of FETs.

The variation of the gate voltage provides useful information about output characteristics of the FET. If the gate voltage is modulated from +0.4 to −0.4 V, that is over 0.8 V, an on/off ratio of 108 is possible (Fig. 3a). On-state current densities of 1 mA μm−1 (Fig. 3a,b) are achievable at gate and drain voltages of less than 0.5 V. As a result, FETs comprising TMD semiconductors as the channels can be used for energy-efficient, low-power switching applications. Indeed, the performance of TMD FETs is comparable to silicon and III–V semiconductor FETs with shorter channels1115. Moreover, the short-channel degradation is reduced in 2D channels compared with similar down-scaled FET sizes in silicon and III–V semiconductor devices.

If the swing of the gate voltage is reduced to 0.4 V (from +0.2 to −0.2 V), the achievable on/off ratio reduces to 104, at a current of nearly half of that of the on current. This is an endemic problem with FETs; if the input power is constrained by the supply voltage, the switching speed (proportional to the on current) and the probability of successful operation of the electronic task (proportional to the on/off ratio) decrease. With passive gate dielectrics, the best performance is achieved in FETs in which transport is ballistic, because electrons in the channel do not lose energy as a consequence of scattering. This is because ballistic FETs are capable of achieving higher on/off ratios but maintaining high performance under the constraint of low power. One potential way of achieving this is by using a variant of the FET in which transport is not over the barrier but through it, called a tunnelling FET. 2D semiconductors are believed to be highly attractive for use in tunnelling FETs, especially in the scaling limits.

Influence of effective mass and valleys

Valleys — the local maximum in the valence band or the local minimum in the conduction band — are important features of 2D TMD semiconductors. Valleys could potentially offer additional degrees of freedom beyond the charge and spin of the electrons used in present day electronics. The first Brillouin zone — a primitive unit cell in the reciprocal lattice (or the k space) — of MoS2 is shown in Fig. 3d. Unlike conventional semiconductors, in some 2D TMD semiconductors there is large (approximately hundreds of millielectron volts) splitting of the valence band because of broken crystal symmetries coupled to a strong spin–orbit coupling32 (Fig. 3e). The presence of valleys offers opportunities to explore the relative effects of the effective mass of electrons, valley degeneracy and spin degeneracy on the drive current of a ballistic FET. Naïvely, it may be assumed that a small effective mass (that is, a higher velocity) and a large number of valleys will maximize the current. However, the drain current in a ballistic FET is similar to the drive current of FETs with channels composed of III–V semiconductors33. The dependence of the on-state current on the effective mass is non-linear (Fig. 3f). More specifically, for effective masses larger than 0.2m0, the number of valleys has a small influence on the drive current; however, at smaller effective masses, a higher number of valleys leads to a proportionately higher drive current. The spatial spread of the wavefunction is high for a very small z-directed effective mass, which is undesirable for the scaling of FETs.

Contact resistance in FETs

We now consider the realistic impact of contact resistance, Rc, on FET performance. To reap the benefits of ballistic FETs, the Rc must be reduced far below the state of the art8. The Rc acts as a severe source-choke – that is, it strongly restricts the injection of electrons into the channel. In addition, the effective gate voltage seen at the source injection point is not Vgs but VgsIdRs. This reduction in effective voltage leads to degradation in the performance of the transistor, because the current depends very strongly on the effective gate voltage at the source injection point. To appreciate the importance of the Rc, let us assume that Rc ≈ 1 kΩ μm. To push a current of 1 mA μm−1, the voltage drop across the contacts is 2IdRc ≈ 2V. The Rc must be below 0.1 kΩ μm to reach the ballistic limits in scaled FETs for energy-efficient logic switching. Reduction of Rc will not just benefit the performance of short-channel ballistic FETs, but also long-channel FETs. This is because the maximum drive current of long-channel FETs is severely limited by the source-choking effect in the presence of a large Rc. That is, when the channel becomes very long such that there is a significant chance for scattering in the channel, the drain current is written as Id = qn(x)v(x)W, where n(x) is the sheet density of carriers along the channel at point x and v(x) is the ensemble velocity of the carriers at that point.

William Shockley introduced the classic model for long-channel FETs by identifying qn(x) = Cb(VgsVTV(x)), where V(x) is the local channel potential, and the ensemble velocity may be written as

(4) v ( x ) = μ F ( x ) 1 + μ F ( x ) / v s a t

where μ is the mobility, vsat is the saturation velocity and F(x) = − ∂V(x)/∂x is the local channel field directed along the channel. In the absence of Rc, the on-state current for a channel length, L, evaluates to

(5) I d = μ C b W L [ ( V gs V T ) V ds V ds 2 2 ] / ( 1 + μ V ds v sat L )

This expression highlights the importance of mobility, saturation velocity and channel length in long-channel FETs. In this scattering-dominated limit, the details of the band structure, such as valleys and spin, are lumped into the parameters of mobility and the saturation velocity. The Rc still has a major role — whenever the voltage drop at the source side is comparable to the applied Vds, the Rc causes a severe source-choke.

Semiconducting 2D materials for FETs

The demonstration of high performance FETs based on <1-nm thick single-layer MoS2 has been an exciting development in the field of 2D electronics34. The first single-layer MoS2 FETs showed on/off ratios of 108, mobility values of 200 cm2 V−1 s−1 and, importantly, SS values of 70–75 mV dec−1 at room temperature34. Although these initial mobility values were subsequently revised downwards35, the SS values were close to the theoretical value of 60 mV dec−1 for standard FET configuration16. In addition, the devices showed respectable on-state currents of 2.5 μA μm−1. These device properties were viewed as promising because sharp turn-on, high drive current and high on/off ratios are important parameters for FETs, even for devices with moderate mobility values. These efficient properties reflect the promise of 2D materials for use in low-standby-power integrated circuits.

Graphene

2D materials can be made by mechanical or chemical exfoliation of bulk layered materials or by chemical vapour deposition of atomically thin layers36,37. Graphene is the most widely studied 2D material because it has an extraordinary mobility of 25,000 cm2 V−1 s−1 at room temperature and lacks dangling bonds. Graphene comprises an sp2-hybridized carbon-atom arrangement in which three valence electrons form the strong in-plane covalent bonds, and the fourth electron remains in the p orbital and forms the out-of-plane π bonds. The electrons in the p orbitals are easily delocalized among the atoms, giving rise to the unique linear dispersion in graphene38. Graphene is often referred to as a zero bandgap semiconductor, because carriers in graphene can be influenced by gate voltage in a FET device39. However, the absence of a bandgap due to the delocalized electrons in graphene means that it is not possible to switch off a FET comprising graphene as the channel material, which results in impractical on/off ratios of typically <10 (Ref. 39). A bandgap of 0.1 eV can be introduced by making narrow (10-nm wide) graphene nanoribbons; however, the fabrication of nanoribbons is challenging. In addition, this approach forms edge states in the 2D structure and thus introduces scattering, resulting in low mobility40. Recently, large arrays of 10-nm graphene nanoribbons have been used to make FETs with a bandgap of 140 meV. These FETs show conventional band transport at room temperature and interband tunnelling at low temperatures41, which suggest that the edges are smooth. For comparison, it should be stated that world-record-breaking n-type and p-type FETs have been demonstrated with InSb, which has a bandgap of 170 meV, mobilities of 30,000–40,000 cm2 V−1 s−1 at room temperature and a carrier concentration of 1012 cm−2 (Refs 42,43). Thus, the optimization of the fabrication of graphene nanoribbons could result in FETs with attractive properties.

In recent years, the FET community has shifted its interest to 2D materials with sizable (>0.3 eV) bandgaps so that low off-state currents and high on-state currents can be realized. More specifically, mono- and few-layered TMDs34,35,4461, phosphorene6273 and silicene74 have been incorporated into FETs.

2D transition metal dichalcogenides

TMDs have the general formula of MX2, where M is a transition metal from group 4, 5 or 6, and X is a chalcogen atom (that is, sulfur, selenium or tellurium). A single layer of these materials consists of three atomic layers in the form of X–M–X (Ref. 37). Adjacent layers of TMDs are weakly held together by van der Waals forces, which allow them to be easily exfoliated. There are approximately 40 possible members in the TMD family, many of which have been synthesized. TMDs possess diverse properties: for example, MoS2, WS2 and MoSe2 are semiconductors; WTe2 and TiSe2 are semimetals; HfS2 is an insulator; and NbS2 and VSe2 are true metals. The bandgap of most semiconducting TMDs changes with layer thickness. Bulk layered materials are indirect bandgap semiconductors, whereas single layered TMDs are direct band semiconductors45,75. In addition, there is an increase in the value of the bandgap for monolayers compared with the bulk; therefore, it is possible to obtain 2D semiconductors with variable band gaps from 1.1 to 2.2 eV (Ref. 37). However, some TMDs, such as ReS2, are direct bandgap semiconductors in the bulk and in single-layer forms76. The energy bands of some important semiconducting TMDs, graphene, h-BN, phosphorene and silicon are shown in Fig. 4a (Ref. 16).

Figure 4: Energy band alignments and device properties of 2D materials.
figure 4

a | Energy levels of various 2D materials compared with that of silicon. The numbers between the valence band (VB) and conduction band (CB) energies indicate the bandgaps of the materials. The energies with respect to the vacuum level (the work function or the electron affinity) are approximate and subject to experimental refinement. b,c | The importance of contact resistance in MoS2 field-effect transistors (FETs). b | Transfer characteristics of a top-gated device measured at a drain voltage, Vds, of 1 V. The red curves represent transfer characteristics of phase-engineered low-resistance contacts and the blue curves are for high-resistance contacts. Higher on-state current and lower subthreshold slopes are apparent in the phase-engineered (optimized) device. The linear fits can be used to extract the values of the subthreshold slopes, which are 95 mV dec−1 for the phase-engineered contacts and 100 mV dec−1 for the high-resistance contacts. A logarithmic scale of the drain current, Id, is shown on the left and a linear scale is on the right. c | Contact resistance as a function of 2D-sheet carrier density for various semiconductors against the quantum limit. The use of metallic 1T phase transition metal dichalcogenides (TMDs) as contacts makes it possible to decrease the contact resistance. A schematic illustration of a device is shown in the inset. Panel a is adapted with permission from Ref. 16, IEEE. Panel b is from Ref. 60, Nature Publishing Group. Panel c is from Ref. 88, Nature Publishing Group.

PowerPoint slide

The electron orbitals of TMDs are different from those in graphene and h-BN. h-BN has an sp2-hybridized atom arrangement and comprises three valence electrons in in-plane sp2-hybridized orbitals, like graphene; however, the two remaining electrons of nitrogen form a highly localized lone pair, thus making h-BN a wide bandgap insulator. In contrast to graphene and h-BN, the electronic properties of TMDs are governed by the d orbitals of the transition metals, and the degree of filling of these orbitals has implications on the electronic structure as well as giving rise to interesting condensed matter phenomena, such as charge density waves, superconductivity, magnetism and hidden states16,7779.

MoS2 and other TMD FETs. Despite the large number of semiconducting TMDs, research on these materials for use in electronic devices has almost exclusively focused on MoS2, although devices with WS2, WSe2 and other TMDs have also been demonstrated47,50,5459. MoS2 has attracted this attention because it is readily available (in its natural form), and high quality 2D crystals can be obtained relatively easily. MoS2 is also mechanically and chemically robust. In 2007, mechanically exfoliated crystals of MoS2 (8–40 nm in thickness) were used as a channel material in very thin TMD FETs80. In a back-gated FET configuration and at a gate voltage of −50 V, n-type transport was observed with mobilities of up to 50 cm2 V−1 s−1 and an on/off ratio of >105. In 2010, top-gated monolayer MoS2 transistors with HfO2 as the gate dielectric showed mobilities in the range of tens of cm2 V−1 s−1 (Refs 34,35), high on/off ratios, high on currents at a source–drain voltage of 0.5 V and a low SS of 74 mV dec−1. These initial results demonstrated that useful FET properties could be achieved from very thin semiconductor channels and provided encouragement for the community to pursue 2D materials for electronics.

Quantum transport simulations using the non-equilibrium Green's function have been performed to determine the scaling limits of single-layer MoS2 transistors with high on/off ratios and good short-channel behaviour, suggesting that MoS2 transistors could be suitable for low-power applications46. To demonstrate that MoS2 transistors are immune to short-channel effects, FETs with channel lengths of up to 100 nm have been fabricated81. In addition, the performance limit of these MoS2 transistors is a consequence of the high Rc between the contacts and the MoS2 channel, and as such, contacts that allow full transmission of electrons from the electrodes to the channel are required for the realization of high performance short-channel devices. In a further study, the performance metrics of 5-nm-channel FETs were shown to be comparable to the ITRS 2026 low operating power technology requirements82. MoS2 transistors have been fabricated with a range of metals as the contact electrodes, and it has been shown that Fermi-level pinning at the conduction band of MoS2 strongly influences the metal/MoS2 interface49. The best device performance was achieved for scandium contacts owing to high carrier injection and a low Rc of 0.65 kΩ μm (Ref. 49). Another interesting approach for ensuring a low Rc, and hence high performance devices, is to degenerately dope the contacts of MoS2 with elements, such as potassium52. However, doping with materials or chemicals that readily react with the environment or that evolve just above room temperature introduces instability and can cause the properties of the device to deteriorate over time. In addition to single-layer MoS2 FETs, interesting characteristics — including mobilities of 100 cm2 V−1 s−1 — can be obtained from multilayer MoS2 FETs83. Multilayer MoS2 has other attractive features, such as high current modulation, low SS and the ease of growing multilayer MoS2 over a large area.

Challenges for MoS2 FETs. MoS2 FETs exhibit promising results; however, fundamental and practical challenges remain. For example, although the effective masses of the conduction and valence band edges have been calculated to be approximately symmetric (that is, the electron effective mass is 0.57 and the hole effective mass is 0.66), MoS2 FETs usually exhibit n-type characteristics, suggesting that the channel material has been unintentionally doped34. The exact origin of the n-type behaviour is unclear, but the presence of impurities, such as rhenium and gold, sulfur vacancies and Fermi-level pinning near the conduction band, are possible causes84,85. It is also important to note that 2D MoS2 is non-stoichiometric with a variable Mo/S ratio from 1:1.8 to 1:2.3 (Ref. 85) and that the energy of the Fermi level can vary over the surface. The structural defects, such as sulfur vacancies86, along with impurities can limit the performance of 2D MoS2 FETs. Synthesis of high purity and stoichiometric MoS2 should allow the realization of ambipolar FETs. Furthermore, development of controlled doping techniques is required for tuning the polarity of the devices. These limitations can be partially overcome by sandwiching MoS2 between layers of h-BN, which reduces the effect of charged impurities on transport properties and allows mobilities that are close to what is theoretically possible (30,000 cm2 V−1 s−1) to be obtained, although at low temperatures26. Other TMDs — for example, WSe2 — exhibit primarily p-type characteristics47, and WS2, which tends to be more stoichiometric than MoS2, exhibits ambipolar behaviour87.

Improving contact resistance. The resistance of contact electrodes is a major limitation in short-channel FETs comprising silicon and III–IV semiconductors. It has also limited the performance of FETs incorporating TMD or other 2D materials as the channel material88,89. Resistance at the source and drain contacts is around 10 kΩ μm in MoS2-based FETs49,60 — over 100 times higher than that of silicon-based electronics8 (<20 Ω μm). Contacts in silicon or III–IV electronics are degenerately doped or regrown to achieve exceptionally low resistance and sharp interfaces15. These methodologies pose challenges in 2D materials because of their atomically thin nature. Future FETs will require single-layer channels and multilayer heavily doped contacts. In multilayered MoS2 FETs, scandium electrodes49 and doping by physisorbed molecules52 have been successfully implemented to reduce the Rc. In another study, graphene has been shown to be an effective contact electrode material for ultrathin MoS2 FETs26,90,91. Phase engineering has also been used to decrease the Rc in 2D MoS2 devices down to 200 Ω μm (Ref. 60). More specifically, TMDs can exist in either the semiconducting 2H phase or the metallic 1T phase, and interfacing these two phases in a controllable manner results in atomically sharp metal/semiconductor junctions92. The decrease in Rc using phase engineering allows current capacities to exceed 100 μA μm−1 in TMD FETs60 (Fig. 4c). Recently, phase engineering in MoTe2 by laser irradiation has been shown to fabricate Ohmic contacts93. Phase engineering, therefore, offers a route for building lateral heterostructures with low Rc, but additional work is necessary to achieve values comparable to state-of-the-art materials and to approach the quantum limit88,89 (Fig. 4d). In addition, methods that are compatible with electronic processing and have sufficiently good spatial resolution for device patterning are required.

Calculations have revealed that the effective masses of electrons and holes pose fundamental challenges for achieving high switching speeds in MoS2 and other TMD-based FETs46. However, the large effective masses of TMDs are beneficial for switching devices, such as tunnelling FETs. Nevertheless, the excellent performance of FETs outlined above offers opportunities for realizing large area, transparent and flexible electronics90,91,9498. Recent progress on the wafer-scale synthesis of high quality MoS2 suggests that large-area monolayer MoS2-based FETs99 could find use in electronics for which exceptionally high processing speeds are not required.

Phosphorene FETs

In 1914, black phosphorus was synthesized by compressing red phosphorus under high pressure and at high temperatures100. Black phosphorus is a layered material with covalent bonds in the layers and weak, van der Waals bonds between the layers62,101 (Fig. 5). The layers have a puckered honeycomb structure that, if stretched laterally, would result in a graphene-like hexagonal structure. Each layer consists of phosphorus atoms with three valence p-orbital electrons that form three covalent bonds to each other. Because all valence electrons are involved in covalent bonding, black phosphorus is a semiconducting material with a bandgap of 0.3 eV (Ref. 102). It typically behaves as a p-type semiconductor, but the origin of this behaviour is unclear, although the polarity of the FETs can be tuned by selecting contact metals with appropriate work functions. The weak bonding between each layer allows the material to be exfoliated down to a few layers with ease6273. The bandgap varies as a function of the number of layers and reaches a value of 2 eV for single-layer black phosphorus103. Single- and few-layered black phosphorus is referred to as phosphorene. The bulk electronic properties of black phosphorus are intriguing. In particular, electron and hole mobilities of 1,000 cm2 V−1 s−1 have been observed62,66. At low temperatures, mobilities as high as 15,000 cm2 V−1 s−1 for electrons and up to 50,000 cm2 V−1 s−1 for holes have been predicted104. This has motivated research in electronic properties of 2D phosphorene for applications in photodetectors105 and solar cells65, as well as in FETs6268.

Figure 5: 2D phosphorene for FETs.
figure 5

a | The orthorhombic crystal structure of black phosphorus (top) from which few-layered phosphorene is obtained. Each phosphorus atom is bonded to three other phosphorus atoms to form a six-membered ring. The individual puckered sheets are linked by weak van der Waals bonding with a spacing of 5.3 Å. A schematic illustration of the back-gated field-effect transistor (FET) is depicted below. The x direction in the FET device is correlated with the x direction of the phosphorene lattice. b | Drain–source current, Ids, as a function of back gate bias, Vbg, transfer curves for a 5-nm thick phosphorene channel of length L = 1 μm, showing p-type behaviour. c | Ids versus drain-source voltage (Vds) output characteristics showing current saturation. d | An optical micrograph of a 30-nm black phosphorus flake (left); black phosphorus flake with 12 electrodes spaced 30° apart (middle); and direct-current conductivity and relative infrared extinction (right) measured along the six directions shown in the image on the left. The circles represent direct current conductance and the squares represent polarization-resolved extinction at 2,700 cm−1. The colours of the dots and squares correspond to the colours in the image on the left. e | Angle-resolved Hall mobility, μx,y, measurements for phosphorene films show that thicker films have higher mobilities. Also, mobility along the x direction is 1.8 times higher than in the y direction. The hole carrier concentration is constant at 6.7 × 1012 cm−2. f | The voltage transfer curve of an inverter device consisting of n-type metal-oxide semiconductor (NMOS) MoS2 FET and few-layer phosphorene p-type metal-oxide semiconductor (PMOS) FET (left). A schematic diagram of the inverter device is shown in the middle, and a circuit representing the inverter device is depicted on the right. Vdd, power supply voltage; Vin, input voltage; Vout, output voltage. Panels ae are from Ref. 66, Nature Publishing Group. Panel f (left) is courtesy of P. Ye, Purdue University, USA. Panel f (middle and right) is published with permission from Ref. 62, American Chemical Society.

PowerPoint slide

Single-layer phosphorene FET devices have not been widely reported because there has not been a pressing need to isolate phosphorene monolayers for this application. This is because although the value of the bandgap varies with the number of layers, phosphorene remains a direct bandgap semiconductor. The thickness of few-layered phosphorene in electronic devices is around 10 nm (Ref. 106). Isolation and eventual growth of stable single- or few-layered phosphorene is essential for the development of high performance, scalable FETs. Typical transfer and output characteristics of back-gated phosphorene FETs are shown in Fig. 5b,c (Ref. 66). Most reported FETs incorporating phosphorene have been in the back-gated configuration; these devices yield mobilities ranging from 100 to 1,000 cm2 V−1 s−1 with on/off ratios of 104 and a SS of 1 V dec−1 (Refs 6268). Top-gated devices improve the SS values to 100 mV dec−1 (Ref. 107). The drive current capacity of phosphorene is 100 μA μm−1 at a drain bias of 0.5 V, which is comparable to that achieved in TMDs. Saturation in the output characteristics, similar to the output curves of MoS2 FETs, is also observed in phosphorene FETs. A Hall mobility, μ, reaching 6,000 cm2 V−1 s−1 has been achieved in few-layer phosphorene, which has enabled the observation of the quantum Hall effect at temperatures of <10 K (Ref. 108).

An interesting feature of phosphorene is that it is electronically anisotropic; that is, transport in the armchair direction (x direction in Fig. 5a) is much more efficient than in the zigzag direction (y direction in Fig. 5a). Polarization-resolved infrared spectroscopy and angle-resolved direct current conductance measurements have been performed to elucidate the degree of anisotropy in phosphorene (Refs 62,66). Using the directional dependence of low-field conductivity for an anisotropic material given by σθ = σxcos2(θφ) + σysin2(θφ) to calculate the conductance, an excellent fit with the experimentally measured data that is observed (Fig. 5d). In the equation, σx and σy are the conductivities in the x and y directions in the crystal (Fig. 4a), θ is the angle with respect to 0° along which the conductance is measured, σθ is the conductivity in the θ direction and, finally, φ is the angle between the x direction and the 0° reference. It has been reported that φ = 8°, which is in agreement with results using polarized infrared spectroscopy66. The high and low conductivity directions also represent the high and low mobility directions, and the ratio of σx/σy can be extracted to be 1.5, which is slightly smaller than the experimentally measured Hall mobility values (Fig. 5e). Hall mobility measurements in the x and y directions yield a μx/μy ratio of 1.8, because the calculations do not consider the spreading of current in the material66.

A CMOS logic circuit has been demonstrated that consists of top-gated p-type phosphorene and few-layered n-type MoS2 FETs69 (Fig. 5f). Voltage transfer characteristics of these FETs show a clear transition to zero within the input voltage range from −5 to +5 V. These results also highlight the integration of two heterogeneous 2D materials for FETs.

A principal challenge associated with phosphorene is its environmental stability109114. The surface of phosphorene is hydrophilic as a consequence of a permanent out-of-plane dipole moment and, as a result, it readily oxidizes111. More specifically, it has been recently elucidated that a combination of oxygen, light and moisture leads to rapid degradation of the material110. Some progress has been made on capping phosphorene using h-BN112, atomic layer deposited Al2O3 (Refs 113,114) and a double-layer comprising Al2O3 and a hydrophobic fluoropolymer110. The latter double-layer approach appears to offer indefinite stability by ensuring conformal sidewall coverage to prevent diffusion from edges and moisture resistance from an effective hydrophobic surface.

Phosphorene also suffers from a high Rc, thus limiting the overall performance of the FETs. Using Ti–Au contacts, Schottky barrier heights of around 0.2 eV have been measured. New graphene side contacts have been achieved but the overall Rc remains in the kΩ-μm range112. However, unlike MoS2, the Fermi level is not pinned and both p- and n-type devices can be achieved. This also suggests that phosphorene has good structural integrity and a low defect concentration, as long as the material is not exposed to the environment.

Silicene FETs

Silicene is an allotrope of silicon with a buckled, six-membered ring 2D structure that is analogous to graphene115 (Fig. 6a). It is, like phosphorene, highly unstable under ambient conditions because of mixed sp2sp3 hybridized bonding116. The use of silicene in electronic devices is particularly exciting because, if high performance devices can be realized, the existing CMOS infrastructure that is based almost entirely on the processing of silicon could be easily applied.

Figure 6: 2D silicene FETs.
figure 6

a | A schematic illustration of silicene showing the buckled six-membered ring structure. b | A fabrication route to a silicene back-gated field-effect transistor (FET) showing epitaxial growth of silicene on an Ag(111) crystalline film, in situ capping with Al2O3, delamination and encapsulation of the film, and the use of the native Ag film to produce the contact electrodes. c | Scanning tunnelling microscope image of Si overlayers with 4 × 4 superstructures. d | A transfer curve plotted of response (R = Vd/Id, where Vd is the drain voltage and Id is the drain current) as a function of overdrive voltage (VgVdirac, where Vg is the gate voltage and Vdirac is the voltage at which Rmax occurs). The mobility values for electrons and holes are indicated. The device indicates the range of electron and hole mobility values that can be obtained from silicene FETs. The devices fabricated74 indicate that the bandgap in silicene is 0.2 eV. Panels ad are from Ref. 74, Nature Publishing Group.

PowerPoint slide

Silicene deposited under ultrahigh vacuum conditions on bulk single-crystal Ag(111) or thin-film substrates has been reported (Fig. 6c). An effective encapsulation and transfer process (Fig. 6b) has allowed the fabrication of silicene-based FETs, which have good mobilities at room temperature74 (Fig. 6d). More fundamentally, the density of states in silicene is predicted to be tuned between gapped (semiconducting) and gapless (metallic) to gapped states by applying a vertical electric field. The gapless version of silicene is predicted to host topologically non-trivial states, which is why its physics, in some sense, is richer than that of graphene or sp3-hybridized silicon. The reported transfer characteristics of silicene FETs are similar to those obtained for graphene with an on/off ratio of 10, and hole and electron mobilities of 129 cm2 V−1 s−1 and 58 cm2 V−1 s−1, respectively, using the ambipolar diffusive transport model74. The intrinsic carrier concentration in silicene (5 × 109 cm−2) is believed to be an order of magnitude less than that in graphene, although the Fermi velocity is comparable117,118. These factors suggest that silicene has a small but finite bandgap of 210 meV (Ref. 74). The room-temperature oscillatory conductivity behaviour reported for silicene (Fig. 6d) could potentially point towards rich transport physics, which is sure to be explored in the future.

Tunnelling FETs based on 2D Materials

In the discussion of ballistic and long-channel FETs, constraints on the voltage swing were shown to severely restrict the on/off ratio and the current drive. The combined metric for such Boolean (on/off-based) logic operations is termed the energy–delay product, which must be minimized. The energy–delay product to complete the same circuit-level operation for several electronic and magnetic logic devices are shown in Fig. 7a (Ref. 119). Some of these devices (for example, CMOS) are common, some have been experimentally demonstrated and others have only been proposed. The delay in a charge-based device is of the form τ = CV/I, where I is the current drive and Q = CV is the net charge to be switched. The most desirable situation (or preferred corner shown in Fig. 7a) is for FETs to perform very fast computation (lowest delay) with the least amount of energy. We see that a high-performance CMOS is fast but takes more energy to perform logic operations; these devices form the core of powerful microprocessors in electronics today. By contrast, a low-voltage CMOS consumes less energy (by an order of magnitude), but is slower and has enabled the recent explosion of computational devices for portable electronics, such as smart phones and tablets. These slower devices lack the number-crunching capability of desktop and notebook computers but can be operated using battery power for at least 24 hours. A major challenge in semiconductor device design today is to build a device that gets closer to the preferred corner (Fig. 7) of a low-energy and low-delay product.

Figure 7: Future FET technologies.
figure 7

a | Intel's benchmarking of future field-effect transistor (FET) technologies comparing complementary metal-oxide semiconductor (CMOS) high-performance and CMOS low-voltage in the energy–delay product metric for digital logic circuits. 2D semiconductor-based tunnelling FETs, such as the graphene nanoribbon tunnelling FET and the thin-tunnelling FET, show significant promise. b | A schematic illustration of the thin-tunnelling FET projected performance of the device. The device structure consists of vertically stacked p- and n-type transition metal dichalcogenides (TMDs) connected to drain and source electrodes, respectively. Vtg is the gate voltage in a tunnelling FET and Vds is the drain voltage. c | The transfer properties reveal exceptionally low subthreshold slopes and high on/off ratios. d | The output characteristics exhibit excellent ambipolar modulation with tunnel gate voltage as well as current saturation. ASL, all spin logic; CMOS HP, high-performance silicon CMOS; CMOS LV, low-voltage silicon CMOS; CSL, charge spin logic; EOT, equivalent oxide thickness; ExFET, excitonic FET; FEFET, ferroelectric FET; GaNFET, gallium nitride tunelling FET; gnrTFET, graphene nanoribbon TFET; GpnJ, graphene p–n junction; HetJTFET, heterojunction III–IV TFET; HomJTFET, Homojunction III–IV TFET; MITFET, metal–insulator transistor FET; NCFET, negative capacitance FET; NML, nanomagnetic logic; PiezoFET, piezoelectric FET; SMG, spin majority gate; SpinFET, spin FET (Sughara–Tanaka); STO, spin torque oscillator; STT/DW, spin torque domain wall; SWD, spin wave device; ThinTFET, 2D heterojunction interlayer TFET; vdW FET, van der Waals solids (or 2D Materials) FET. Panel a is adapted with permission from Ref. 119, IEEE. Panels bd are adapted with permission from Ref. 120, IEEE.

PowerPoint slide

In a traditional metal oxide semiconductor FET, thermal electrons in a semiconductor band go over a gate-controlled barrier and, as a result, there is a tail of high-energy electrons that do not see the barrier and trickle from the source to the drain electrode30,31. This gives rise to a leakage current in the off state of the device. The electrostatics that drive the flow of charges in a FET give a SS close to 60 mV dec−1 (as calculated in Fig. 3a and described in Fig. 8a). To reach the desired energy–delay product, the device switch needs to be made with a steeper SS (Fig. 8b). This can be achieved by changing from a passive gate insulator to an active one, for example, by using a ferroelectric layer that could boost the gate potential and possibly enable a steeper SS120,121.

Figure 8: Operating mechanism of tunnelling FETs.
figure 8

a | An energy band diagram showing the Boltzmann tail of electrons in a semiconductor band that leads to the 60 mV dec−1 subthreshold slope (SS) and drain current, Id, leakage in a normal field-effect transistor (FET). The barrier for injection from the source can be lowered by the application of a gate voltage, Vgs. b | The transfer characteristics (drain current as a function of gate voltage of a FET) of a normal FET and a tunnelling FET. The tunnelling FET can switch at a SS of <60 mV dec−1 that can reduce the voltage swing (Vswing) — the voltage difference between the off and on states — but maintain the same on/off ratio. c | If the source is replaced by a filled band, the Boltzmann tail is cut off by the energy gap (top). Interband Zener tunnelling and the density-of-states overlap of the filter electrons of the valence and conduction bands in a narrow energy window, enabling a sharper turn-off, are depicted in the middle and bottom. d | The on current can be significantly boosted by choosing two 2D crystals with type-II or type-III heterostructure band alignment, and by interlayer tunnelling of carriers between them. The tunnelling FET can potentially allow for energy-efficient switching, because it reduces the voltage swing necessary to achieve high on current by virtue of the steep SS. Ec, energy of the conduction band; Ev, energy of the valence band; Ioff, off-state current; Ion, on-state current; VT, threshold voltage.

PowerPoint slide

Another method to achieve a steep SS is to remove the high-energy tail that leaks from the source to the drain electrode in a normal FET. This ‘energy filtering’ can be achieved by introducing an energy gap at the appropriate energies in the density of states of the source of the FET. For example, inserting superlattices in 1D materials to break up the continuous energy band into a set of discrete minibands with gaps can lead to steep subthreshold FETs122. An easier way is if the source of the electrons is from the top of the valence band, then the valence band edge and the gap above it are a natural energy filter (Fig. 8c). However, for electrons to flow, they must overcome the barrier of the energy gap. Electrons have a wave-like nature that allows them to use evanescent states to tunnel through the gap at a low-energy cost. Transistors based on interband tunnelling, termed tunnelling FETs, are expected to enable low-power transistor switches that beat the SS limit of 60 mV dec−1 (Ref. 123) (Fig. 8b). Getting below the traditional SS value has not been difficult; semiconductor tunnelling FETs that show steeper than 60 mV dec−1 switching have been demonstrated using silicon, germanium and carbon nanotubes124. The current drives in many such tunnelling FETs have been extremely low, leading to long delays, CV/I, making them less attractive than, for example, the low-voltage CMOS in Fig. 7a. The interband tunnelling current decreases as an exponential function of the bandgap. Narrow-band-gap semiconductors allow larger tunnelling currents, but because they are scaled down to nanoscale dimensions, the bandgaps in thinned 3D materials increase owing to strong quantum confinement effects16.

In this field of tunnelling FETs, crystalline 2D semiconductors are expected to provide a unique solution. Although the bandgap of a single-layer 2D crystal semiconductor is well defined, by choosing a suitable heterojunction of two layers — one n-type layer and one p-type layer (Fig. 7b) — it is possible to maximize the interlayer tunnelling current but maintain a very low SS. The tunnelling current can be boosted by the proper choice of heterojunction band offsets between the two 2D crystal layers. For example, in III–V heterostructures, the tunnelling current has been significantly increased by the choice of type-II (staggered) or type-III (broken) band alignments125 (Fig. 8d). Similar heterostructure design opportunities exist for 2D crystal interlayer tunnelling FETs. This type of tunnelling transistor could be the thinnest possible manifestation of a tunnelling FET, which can also be scaled to the smallest dimensions, far below what may be feasible with traditional 3D semiconductors. The race is currently on to realize this proposed device, called the thin-tunnelling FET. A first step in this direction is the demonstration of interlayer tunnelling in layered 2D crystal gapped semiconductor heterostructures126128. A tunnelling FET with a 2D MoS2 channel contacted with highly doped germanium source electrodes has been reported with a SS of 32 mV dec−1 and a supply voltage of <0.1 V, although at low current values129. The goal of achieving such tunnelling FETs is also forcing controlled growth, doping and heterostructure design, and will open several new opportunities in the future of 2D crystal electronics.

Conclusions

The inherent advantages of 2D materials — for example, the absence of dangling bonds and excellent gate electrostatics — make them promising candidates for high performance electronics. However, the implementation of 2D semiconductors in electronic devices remains in its infancy, and some basic techniques of modern electronics are yet to be developed.

Research on 2D materials for electronics has largely focused on achieving high carrier mobilities, and this parameter is often used as a figure of merit for evaluating potential materials for FETs. However, parameters such as low SS across a wide range of gate biases, resulting in large on/off ratios at low input voltages, are more relevant for addressing the challenges of the field of electronics. In addition, contact resistance is a major factor limiting performance of 2D semiconductor FETs, especially at short-channel lengths, and must be lowered by at least an order of magnitude to <100 Ω μm. This requires the contact regions to be degenerately doped by either the direct growth of metallic multilayer 2D materials or implantation. In 2D TMDs, using metallic 2D phases is promising for the fabrication of low resistance side contacts. More generally, controllable and stable doping of 2D semiconductors is a major challenge that must be overcome in the realization of both n-type and p-type semiconductor FETs. Enhancing our knowledge on doping and uniform growth of 2D semiconductors and heterostructures should enable the next generation of low-power, high-speed devices, such as thin-tunnelling FETs, to be fabricated.

In addition to device-related challenges, uniform growth and environmental stability of 2D TMDs in their semiconducting and metallic phases are topics for future research. Although tremendous progress has been made in the uniform growth of graphene by chemical vapour deposition, large-area growth of two-element compounds (for example, MoS2) has proven more difficult. Although polycrystalline MoS2 and WS2 have been grown successfully using metal–organic chemical vapour deposition, the stoichiometry, chemical purity and yield must be optimized. A fact about TMDs that is frequently overlooked is that they are rarely stoichiometric or chemically pure as a consequence of impurities in the precursors. The ability to precisely control the phases and the number of layers of TMDs also remain challenges to be addressed. Epitaxial growth methods are currently being used to fabricate 2D materials for electronic applications. These methods could be particularly important for the development of 2D semiconductors for high performance FETs as well as thin-tunnel FETs that require high-quality heterostructures. In addition, large-area growth of high-quality 2D semiconductors by chemical vapour deposition or solution-phase exfoliation could fulfil the requirements for flexible electronics.