Main

Among the plenitude of advantageous properties, the capability of z-dimensional stacking—in principle of an unlimited number of layers—is believed to be one of the most fascinating perspectives of semiconducting van der Waals (vdW) nanoelectronics. This method of bottom-up three-dimensional (3D) vdW integrability may provide an alternative approach to continue the scaling of transistors in the so-called post-Moore’s-law age, as the silicon technology is approaching its physical limit for further shrinking of the lateral size of transistors29,30. Indeed, for decades, from the very first planar field-effect transistor (FET), to FinFET and to the most advanced gate-all-around FET, the scaling of Si semiconductors has followed an in-plane strategy, as illustrated in Fig. 1a, while achieving 3D integrability remained extremely challenging. Although 3D interconnection of electrodes has been widely implemented in modern silicon integrated circuits, the essential logic gates are yet confined to only the surface of the silicon substrate, which cannot be arranged into multi-layers. Other attempts of face-to-face bonding of two chips require alignments with ultra-high precision, and the gain of room in the z dimension is not that satisfactory31,32. Meanwhile, multilayered 3D flash memory (3D NAND) consists of orthogonally crossed junctions (where floating gate memories are formed) between horizontal and vertical bit and word lines, which, however, do not meet the need for free-design of circuitry33.

Fig. 1: Vertical scaling versus in-plane scaling of semiconducting circuits.
figure 1

a, Schematic illustration of the routes of scaling in Si and vdW technologies. b, Molecular structure of a pristine MoS2 layer. The difference in charge density distributions of slight electron doping with respect to a neutral layer for an isosurface of 10 μe Bohr−3 above the Fermi level is superimposed on its molecular model. c, Schematic diagram of the n-type MoS2 in the conventional pristine state, with the Fermi level (red solid line) positioned in the vicinity of the conduction band minimum (CBM) in MoS2. d, Same plot of differential charge density distributions as b, but in a MoS2–CrOCl heterostructure. In b and d, the charge carrier types of electrons and holes are marked in pink and green, respectively. The atom-symbols used in b and d are illustrated in the bottom part of b. Clear n-type behaviour on electron doping can be seen in b, while in the MoS2–CrOCl heterostructure case in d, most of the doped electrons are transferred to the CrOCl side. e, Schematic band alignment diagram of the MoS2–CrOCl heterostructure under finite vertical negative electric field (corresponding to the negative bottom gate voltage in our experimental configurations), indicating the realization of a p-type semiconducting MoS2 with the Fermi level (red solid line) positioned in the vicinity of the valence band maximum (VBM) in MoS2.

Recently, stacking up vdW semiconductors into 3D vertical circuitry has been a grail pursued with continuous efforts34,35,36. The sizable bandgap and dangling-bond-free surfaces, together with high carrier mobilities and excellent electrostatic control at the ultimate scale (less than 1 nm), make two-dimensional (2D) semiconductors ideal candidates for vertical 3D integration1,2,3,37. The prediction is that advanced monolithic 3D integrated circuits constructed with speedy layer-to-layer signal transmission and efficient heat dissipation will provide much higher integration density38. However, application-wise, 3D integrated circuits of 2D semiconductors have largely been restricted due to the difficulty in obtaining controllable doping of n- and p-type polarities, which is fundamental for complementary logic34. To date, a limited number of examples have been realized in vertical complementary field-effect transistors (CFETs) constructed using 2D semiconductors, such as mixed-dimensional heterostructures4,5 and hetero-2D layers with different carrier polarities6,7,8, among which a maximum of two vertical layers of complementary logic has been demonstrated. Indeed, while n-type 2D semiconductors are advancing rapidly in terms of electrical performance39,40, only a handful of p-doping strategies are known for 2D semiconductors such as WSe2 (refs. 9,10,11,12,13,14,15,16,17) and MoS2 (refs. 11,18,19,20,21,22,23,24,25,26,27,28), using methods including chemical dopants, contact engineering, or oxide coating. Notice that these doping methods may suffer from inhomogeneities or degradation of carrier mobility, and few of them are physically capable of enabling multilayered vertical assemblies of 3D complementary logic.

In this work, we devise a simple and non-destructive doping method to reconfigure, in a controllable manner, the carrier polarity of 2D semiconductors through vdW interfacial coupling. We found that, unlike the usually manifested n-type nature, few-layered transition metal dichalcogenides (TMDs) (including MoS2, WSe2 and MoSe2) interfaced with few-layered CrOCl systematically turn into p-type and display excellent air stability. Density functional theory (DFT) calculations suggest that this interfacial-coupling-induced polarity inversion is a result of charge transfer from TMDs to CrOCl, followed by subtle ee interactions in the surface state of CrOCl, which should be a universal effect at the interface between TMDs and layered insulators with high work function and large-enough effective mass in its surface band. Taking MoS2 as an example, thanks to the atomically clean interface, the MoS2–CrOCl hybrid exhibits a maximum room-temperature hole mobility reaching approximately 425 cm2 V−1 s−1, with on/off ratio exceeding 106. Further, we construct n- and p-doped logic units by selectively stacking modules of vdW gate, dielectric and semiconducting layers, with and without interfacial coupling layers, defined as vertical inversely polarizable field-effect transistors (VIP-FETs). Our doping strategy can therefore be employed to fabricate self-complemented logic devices vertically, throwing light on the vertical scaling route (Fig. 1a) towards advanced 3D integration of semiconducting circuits.

Modelling of vdW polarity-engineered MoS2

To enable the vertical 3D integration of 2D semiconductors, controlling the n- and p-type polarities is crucial. Therefore, it is essential to identify an effective approach to achieve p-doped 2D semiconductors without compromising the carrier mobility. Recent research shows that CrOCl is one of the candidates for engineering the interfacial coupling in 2D electronic gas systems such as graphene, which gives rise to exotic quantum ground states41,42. The interfacial coupling between TMDs and CrOCl, however, remains unexplored so far. We first consider theoretically a model system of CrOCl coupled to TMD, using MoS2 as an example. By calculating the charge density difference between the slightly electron-doped and neutral MoS2 layer, as shown in Fig. 1b, the pristine state of MoS2 (and most of the TMDs) exhibits n-type behaviour, with the Fermi level close to the conduction band minimum (CBM), illustrated in Fig. 1c. To determine their work functions, we performed DFT calculations of 10-layer MoS2 and 5-layer CrOCl, and the CBM of MoS2 is estimated to be about 0.465 eV above the CBM of CrOCl in the absence of vertical electric fields. Therefore, if MoS2 is doped by additional electrons, we would expect these extra charges to transfer from the CBM of MoS2 to CrOCl. This is more explicitly elucidated by calculating the charge density difference between the electron-doped case and the charge-neutral case at the MoS2–CrOCl interface, as shown in Fig. 1d. Clearly, the doped electron carriers are concentrated at the CrOCl side at the interface, leaving some holes on the MoS2 side.

To further elucidate the experimental observations, we carried out DFT calculations of a three-MoS2-layer + three-CrOCl-layer heterostructure with a supercell of about 200 atoms (more details can be found in Methods). Note that, as experimentally CrOCl still acts as a gate dielectric and hence no free carriers can be found in it, the transfer of electrons (via tunnelling from MoS2 to the Cr 3d orbitals, as suggested by DFT calculations) into the surface states of CrOCl has to be in a localized manner. Indeed, by considering the band structures of MoS2, CrOCl and MoS2–CrOCl heterostructures separately (Supplementary Figs. 1–3 in Supplementary Note 1, and also see Methods), we found that the mechanism here in our system is more than a trivial charge transfer, but rather is further followed by a combination of ee interaction (which drives the charges in the CrOCl surface state into insulator) and self-adjustment of band alignments (Fig. 1e and Supplementary Fig. 4). This is fundamentally different from conventional doping strategies for such TMD semiconductors.

Characterizations of CrOCl-interfaced MoS2

We now construct MoS2–CrOCl vdW heterostructures by stacking few-layer MoS2 onto CrOCl through a standard dry-transfer method43. An optical micrograph of a typical sample is shown in the inset of Fig. 2a, where the left part is MoS2–CrOCl FET and the right part is a control sample of conventional MoS2 FET, with each constituent layer highlighted by red and yellow dashed lines. Here, few-layered hexagonal boron nitride (h-BN), highlighted by a white dashed line in the inset of Fig. 2a, is employed as the gate dielectric for the MoS2–CrOCl and MoS2 FETs. Electrodes of Cr (5 nm) and Au (50 nm) are fabricated via standard lithography and thermal evaporation (fabrication details are available in Methods). Indeed, as shown in Fig. 2a, CrOCl-interfaced MoS2 FET manifests typical p-type semiconducting behaviour (as expected from our simulations) with source–drain current Ids above 1 μA at a source–drain voltage Vds = 0.1 V. The on-state current of p-type MoS2 FET is comparable to that of its counterpart of an n-type MoS2 placed on h-BN, with field-effect curve shown in Fig. 2b. The n-type nature of the latter is often attributed to its electron-donating sulfur vacancies and considerable Fermi level pinning effect at the metal-electrodes–MoS2 interface44,45. Additional characterizations of p-type MoS2–CrOCl FETs can be found in Supplementary Figs. 5 and 6. Output characteristics of a typical MoS2–CrOCl FET with SiO2 serving as gate dielectric are also shown in Supplementary Fig. 7. In general, an on/off ratio exceeding 105 is observed in the MoS2–CrOCl FETs (inset in Fig. 2b), whose transfer curves show negligible hysteresis, indicating the high quality of the MoS2–CrOCl interfaces. It also excludes the trivial scenario of defect-induced effects (control experiment of mild-plasma-treated CrOCl surface can be found in Extended data Fig. 1, where huge hysteresis is seen in the field-effect curves of MoS2 placed on it).

Fig. 2: Electrical performance of MoS2–CrOCl complementary FETs.
figure 2

a,b, Source–drain current Ids as a function of back gate voltage Vbg, measured for a typical MoS2–CrOCl FET (a) (orange line) and a conventional pristine MoS2 FET placed on h-BN (b) (blue line). Trace and retrace (as indicated by the solid black arrows) are recorded in a and b, exhibiting negligible gate hysteresis. Measurements in a and b are carried at Vds = 0.1 V and at room temperature. The inset in a shows an optical micrograph of MoS2 FETs made from the same MoS2 flake, but with different polarities when placed on CrOCl or h-BN. Each constituent layer is highlighted by coloured dashed lines. The inset in b shows the same data as in a and b, but plotted on a log scale. c, False-coloured SEM image of a vertically stacked MoS2 complementary logic inverter. d, Output voltage Vout as a function of input voltage Vin at various supply voltages VDD. e, The gain for each of the curves in d. f, Performance of state-of-the-art p-type MoS2 FETs in the parameter space of on/off ratio and hole mobility at room temperature. Data points of this work (solid red squares) are discussed in more detail in Supplementary Fig. 15. In f, the error bar for the on/off ratio is defined as 1/δIoff, with δIoff being the standard deviation of Ioff in each device, while maximum Ion is fixed. The error bar for the hole mobility is defined as the standard deviation of the γ × dIds/dVg at the vicinity of the maximum, where γ is a coefficient obtained from the sample, written as \(\frac{L}{WC{V}_{{\rm{ds}}}}\). LW, C and Vds are respectively the length, width, the gate capacitance and the source–drain voltage for the measured device. Scale bars, 10 μm (a,c).

Evidence for such a p-type doping characteristic in CrOCl-interfaced MoS2 FETs is further supported by Kelvin probe force microscopy (KPFM) measurements, where an obvious reduction of surface potential is seen in MoS2 in the CrOCl-interfaced region, as shown in Supplementary Fig. 8. From our KPFM experiment, the work function for CrOCl is much larger than that of MoS2, leading to a significantly downward band alignment. This is consistent with the DFT results described in the previous section. It is noticed that, compared with few-layer MoS2, the frequency of the Raman A1g and \({{\rm{E}}}_{2g}^{1}\) modes in the MoS2–CrOCl region are blue-shifted (more discussions can be seen in Supplementary Fig. 9), which might be related to the change of Fermi level46,47, echoing the observed p-type doping effect. Detailed electrical measurements are further conducted on FETs fabricated based on other CrOCl-interfaced TMD 2D semiconductors, including MoSe2 and WSe2 (see Supplementary Figs. 10 and 11). It is seen that, using Cr–Au as electrodes, most of the TMD–CrOCl heterostructures become p-doped, speaking to the universality of our doping strategy. Their electrical performance can be further improved by fine tuning the fabrication procedures, which will be discussed in the next sections.

MoS2-based vertical CFETs

In the following, we demonstrate the feasibility of MoS2-based self-complemented circuits, where the combination of n- and p-type transistors enables logic function with high noise immunity and low static power consumption. To start with, a conventional planar inverter, namely, a ‘NOT’ logic gate that outputs a voltage representing the opposite logic level to its input12,48, is constructed by using laterally adjacent n- and p-type transistors of MoS2, as indicated by the optical image in Supplementary Fig. 12a. Electrical performance of the inverter is shown in Supplementary Fig. 12b, where standard inverter action is observed for switching between logic ‘1’ (close to the supply voltage VDD) and logic ‘0’ (close to 0 V).

The key achievement of this study is the obtaining of vertically stackable and polarity-invertible high mobility 2D FETs, which is of significance in increasing the integration density by scaling in the z dimension. For instance, vertically assembled 2D semiconducting inverters that consist of n- and p-type transistors vertically assembled with each other, can improve the integration level by 42–50% compared with those conventional logic circuits with planar architecture using bulk Si semiconductors4,49. As illustrated by the drawing in Extended data Fig. 2, using the described controllable p-doping strategy, a 3D-integrated 2D inverter with both n- and p-type transistors vertically stacked using modules of vdW gate, dielectric and MoS2, with and without CrOCl coupling layers, can be realized, which we previously defined as VIP-FETs and is also a typical vertical nanoarchitecture of CFET.

False-colour scanning electronic microscope (SEM) image of such a typical 3D-integrated 2D inverter with 6 vdW layers is shown in Fig. 2c, with its circuit diagram being similar to that depicted for conventional lateral inverter in Extended data Fig. 2d. Here, a gate electrode (vertically shared by the upper p-channel and lower n-channel) serves as the input voltage (Vin) terminal, and the n-FET is grounded while a supply voltage VDD is applied to the p-FET. The transfer curves, showing output voltage Vout versus Vin, of the vertically stacked 2D CFET inverter with various VDD, are plotted in Fig. 2d. Clear signal inversion is observed: the ON-states in p-FET and n-FET compete with each other, yielding a Vout switched from VDD (logic ‘1’) to ground (logic ‘0’) at different ranges of the input voltage Vin. The sharp transition between the logic states can be characterized by the voltage gain (defined as dVout/dVin), which is a crucial metrics presenting the sensitivity of Vout to the change in Vin. As shown in Fig. 2e, the resulted voltage gain is approximately 23 for VDD = 7 V and approximately 2 for VDD = 1 V.

We evaluate the noise margins of the MoS2-based vertical CFETs by using the expressions NMH = VOH − VIH and NML = VIL − VOL, where NMH and NML represent the high- and low-state noise margins, respectively, and VIL and VIH are the input voltages at which the slope of the voltage transfer curve is −1, whereas VOH and VOL are the corresponding output voltages (Supplementary Fig. 13). The calculated total noise margin of the CFET is around 83% at VDD = 3 V. The CFET devices can operate at low power consumption, as shown in Extended data Fig. 3, where the peak power consumption is 518 pW for VDD = 1 V. In addition, the dynamic inverting performance was also investigated by applying an a.c. voltage on the input terminal. Supplementary Fig. 14 displays the input voltage sequence and the corresponding output signals, clearly showing the output states are opposite to the input signals. It is worth mentioning that, as depicted in Extended data Fig. 4, the electrical property of the device shows minimal degradation within 12 months at room temperature in air, highlighting the long-term stability of our vdW interfacial doping strategy. Among those MoS2–CrOCl FETs tested, the best performance exhibits a high on/off ratio reaching 106 and hole mobility of approximately 425 cm2 V−1 s−1 at room temperature (see Supplementary Fig. 15), indicating that our doping method gives better device performance compared with previous reports11,18,19,20,21,22,23,24,25,26,27,28, as shown in Fig. 2f.

3D-integrated 2D logic over 10 vdW layers

In the following, by vertically integrating these VIP-FETs described in the previous text, we benchmark the bottom-up scaling of complementary logic circuitry with a true 3D architecture up to 14 vdW-stacked layers. We take a four-transistor SRAM (4T-SRAM) and a four-transistor NAND (4T-NAND) logic as examples. The schematic images of these devices can be arranged in a 3D vertical manner, as shown in Fig. 3a,b. It consists in building blocks of two n-type and two p-type FETs. When realized by the fabrication of VIP-FETs as devised in this work, a NAND logic can be further illustrated in the schematics in Fig. 3c, where the MoS2 layer (blue), with and without interfacial coupling layer of CrOCl (pink), are the p- and n-type semiconducting channels, with few-layered h-BN (not shown) serving as gate dielectric and few-layered graphene (black) as gate electrode. Inputs A and B, supply voltage VDD and the ground electrode of the NAND logic are labelled correspondingly in Fig. 3b,c. The detailed fabrication process of the final devices can be found in Extended data Fig. 5.

Fig. 3: 3D-integrated logic gates with more than ten vdW layers.
figure 3

a,b, Illustrative schematic images of logic gates arranged in a 3D manner, leading to a SRAM (a) and a NAND gate (b). c, Artist rendering of the 3D-integrated NAND gate based on MoS2 VIP-FETs, which consists of building blocks of four transistors, with two n-type and two p-type FETs. d, Bright-field scanning transmission electron microscopy image of the cross-section of a typical 3D-integrated NAND gate, with the stacking order for the 14 constituent vdW flakes indicated at each layer. The two graphite gate layers are rendered with false colour in yellow, for better visibility. e, High resolution high-angle annular dark-field STEM image of a typical MoS2–CrOCl interface encapsulated between two h-BN layers. f, Zoomed-in image of the blue-boxed area in e. g, EELS mappings of the boxed area in d, which highlights the distribution of S, Cr, B and C in each layer. The EELS maps are elongated along the horizontal direction for better visualization. Scale bars, 100 nm (d), 5 nm (e), 1 nm (f).

Figure 3d displays the bright-field scanning transmission electron microscopy image of the cross-section of a typical vertically integrated NAND gate based on MoS2 VIP-FETs. It is clearly seen that all the functional layers illustrated in Fig. 3c are realized by vdW layers, with each FET spaced by an h-BN dielectric spacer. A total of 14 vdW layers are included in the NAND logic device, as labelled in Fig. 3d. Zoomed-in STEM images in Fig. 3e,f further illustrate atomically sharp interfaces between the MoS2, CrOCl and h-BN layers. Notice that the distance between MoS2 and CrOCl is about 0.2 nm, which is taken as a reference for our DFT calculations. The corresponding electron energy loss spectroscopy (EELS) mappings (Fig. 3g) also confirm a clean vdW heterostructure with high-quality interfaces.

We achieved the milestone of vertical integration of more than ten layers of 2D complementary FETs and demonstrated such 3D-integration of 2D semiconductors at the device level. Figure 4a shows the logic truth table of such a typical 14-vdW-layer 3D NAND gate operation. Figure 4b exhibits a dynamic performance for the device, where two different input voltages, Vin-A and Vin-B, are fed with a rectangle wave in a time sequence but phase shifted, and the supply voltage VDD and ground are fixed at +3 V and −3 V, respectively, during the measurements. The device outputs a logic state ‘0’ only if both input states are ‘1’, firmly demonstrating the functionality of a NAND gate (more details on the performance of the 3D NAND gate are provided in Supplementary Fig. 16).

Fig. 4: Towards future 3D integration of 2D semiconducting complementary logic.
figure 4

a, Logic operation of VinVout characteristics of a typical 3D-NAND device with 14 vertically integrated vdW layers as shown in Fig. 3c. A VDD of 3 V, GND of −3 V and Vin = ±3 V for inputs A and B were applied during measurements. b, Dynamic NAND performance of the same sample as in a tested at a pulse period of 400 ms. c,d, show the input waveforms (c) and output level (d) of a typical 3D-SRAM device with 14 vertically integrated vdW layers. e, Field-effect curves of p-type MoSe2 (red curves, with the statistics of 12 devices) and p-type MoS2 (green curves, with the statistics of 8 devices) induced by coupling of the CrOCl substrates. Vds = 0.1 V is used in the measurements. f, Output performance of a typical p-type MoSe2–CrOCl transistor, with a maximum Ids reaching 0.3 mA at Vds = 2 V. The channel length and width of the tested device are 3 μm and 5 μm, respectively. g, Logarithmic plot of transfer curves of typical p-type MoSe2–CrOCl transistors, measured using the transfer length method (TLM). A Vds of 0.1 V was used in the measurement. hj, False-colour SEM images of typical 4T-SRAM devices based on the planar-FET (h), CFET (i) and VIP-FET (j) architectures, with distinctive colour-coding for their VDD, GND, Vin and Vout electrodes in green, purple, yellow and blue, respectively. km, art illustrations of the devices pictured in h (k), i (l) and j (m). Devices in ij are plasma-etch patterned into square areas for visual clarity. n, An outlook of future 3D integration of 2D VIP-FETs, based on the technology described in this work. Scale bars, 10 µm (hj).

Outlook of the VIP-FETs

Before going further in the demonstration of 3D logic, we discuss in the Methods further details (Supplementary Figs. 17–21 in Supplementary Note 2) of the interfacial-coupling-induced p-doping effect and the improvement of their electrical performance. The vertically free arrangement of CFETs demonstrated in our work can in principle be expanded into any 3D-integrated circuitries. For example, by re-wiring the vertically piled up four transistors, functionalities of SRAM with 14 vdW layers can also be realized, as shown in Fig. 4c,d (more details can be found in Extended data Figs. 6 and 7). To show the universality, statistics of field-effect curves of p-type MoS2 and MoSe2 are shown in Fig. 4e. Characterizations of output performance of a typical device and the effects of channel length are illustrated in Fig. 4f,g (see Methods).

The interfacial-coupling-induced p-doping and the resulting VIP-FETs are key techniques invented in this study and are conceptually suitable for the vertical scaling of future 2D semiconducting complementary metal–oxide–semiconductor circuits. To visualize the envisioned picture, we compare side-by-side the 4T-SRAM devices based on the planar-FET, CFET and VIP-FET architectures, shown in the SEM images in Fig. 4h–j, with their technical illustrations in Fig. 4k–m, respectively. Clearly, their footprint sizes are sequentially decreasing from 4 unit areas to 1 unit area, with a z-dimensional accumulation from 3 vdW to 14 vdW layers. We can thus expect future 3D integration of 2D VIP-FETs as shown in Fig. 4n. And this makes the most distinct feature of our technique, though there are very interesting alternative ways of p-doping TMDs9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28. Nevertheless, we emphasize that there are yet further technical challenges that need to be addressed in the longer-term perspective for the 3D-integration of 2D semiconductors in future nanoelectronics at a large scale application. (We have become aware that recent efforts on wafer-scale vdW vertical CFETs have been reported during our submission2,3). These challenges include heat dissipation, large-area growth of p-type film (MoS2–CrOCl heterostructure for example) with high uniformity suitable for industrial production, as well as the possible interlayer parasitic capacitance. We here briefly discuss the possibility of chemical-vapour-deposition growth of large scale CrOCl thin films in Supplementary Note 3 (Supplementary Figs. 22–26), and we also performed technology computer-aided design simulations of the parasitic capacitance due to the interlayer overlapping of electrodes, shown in Supplementary Note 4 (Supplementary Figs. 27–30).

Rather than pursue increased in-plane scaling of Si-based semiconducting devices, we took the route of vdW vertical scaling and have demonstrated its feasibility. We report a facile and stable p-type doping strategy for 2D semiconductors in order to have access to complementary FETs that are compatible with vertical integration. Our studies show that by stacking TMDs (MoS2, WSe2 and MoSe2) onto a vdW insulator CrOCl, the dominant carrier type can be effectively modulated from electrons to holes. First-principles calculations further reveal that such behaviour may originate from the strong vdW interfacial coupling. It indicates a cooperative effect of gate-tunable band alignment, charge transfer and ee interactions, which could essentially be different from conventional p-doping strategies for semiconducting TMDs. It is worthwhile noting that a similar mechanism already leads to a number of exotic quantum electronic states reported previously in graphene–CrOCl systems41,42. FETs fabricated based on this approach exhibit excellent electrical properties with on/off ratios reaching 106, and the extracted room-temperature hole mobility reaches 425 cm2 V−1 s−1 in MoS2 with outstanding long-term air stability. Furthermore, based on our doping method, advanced 3D logic circuits, such as vertically constructed inverters with 6 vdW layers, NANDs with 14 vdW layers and SRAMs with 14 vdW layers, are implemented, confirming that our vdW interfacial-coupling-induced p-type doping may be a potent strategy for the design of future vertical scaling to realize ultra-high 3D-integration of advanced logic circuits.

Methods

Sample fabrication

The vdW few-layers of MoS2 and CrOCl were obtained by mechanically exfoliating high-quality bulk crystals. The vertical assembly of vdW-layered compounds was fabricated using the dry-transfer method. Electron-beam lithography was done using a Zeiss Sigma 300 SEM with a Raith Elphy Quantum graphic writer. Input gates, as well as contacting electrodes, were fabricated with an e-beam or a thermal evaporator, with typical thicknesses of Cr and Au of approximately 5 nm and 50 nm, respectively.

STEM characterizations

STEM characterizations were carried out on a monochromated Nion U-HERMES 100 microscope equipped with an alpha-type monochromator and a fifth-order aberration corrector and operated at 60 kV. Cross sections of as-prepared devices were made using a focused ion beam tool, Thermo Scientific Helios G4 CX DualBeam system cut at 30 kV and milled with the voltage gradually decreasing from 30 kV to 5 kV to minimize sample damage.

Morphology tests

A Bruker Dimension Icon AFM was used for thickness and morphology tests, as well as KPFM characterizations. Optical images were collected by a Nikon LV-ND100 microscope.

Electrical measurements

The high precision of current measurements of the devices were provided by measurement using a Cascade M150 probe station at room temperature, with an Angilent B1500A Semiconductor Device Parameter Analyzer. For the Dynamic NAND performance measurement in Fig. 4, a pulse train is adopted by using the waveform generator fast measurement unit.

Density functional theory calculations

The first-principles calculations based on DFT were carried out with Vienna Ab initio Simulation Package with a projector augmented wave method50,51. The plane-wave energy cutoff was set to be 600 eV, the generalized gradient approximation by Perdew, Burke and Ernzerhof (PBE) was taken as the exchange-correlation potential52. As Cr is a transition metal element with localized 3d orbitals,the so-called fully localized limit of the spin-polarized DFT + U functional was adopted, as suggested by Liechtenstein and co-workers53. The on-site Hubbard U = 3.0 eV parameter was used in the calculations, and the magnetic configurations of CrOCl is an antiferromagnetic order as shown in Supplementary Fig. 1f. This leads to a gap of 2.12 eV for CrOCl. For the work-function calculations, ten layers of MoS2, five layers of CrOCl and a three + three-layer heterostructure were used, and the crystal structure was fully relaxed until the residual forces on the atoms were less than 0.01 eV Å−1. To avoid any artificial interactions, 15 Å vacuum layers are added. In the calculation of heterostructure, a 7 × 1 supercell for the CrOCl and a 5\(\sqrt{3}\) × 1 supercell for MoS2 were adopted. To construct such commensurate supercells, the lattice constant of CrOCl has been slightly compressed by 0.3% in one direction (changed from 3.2 Å to 3.188 Å) and has been slightly expanded by 1.5% in the other direction (changed from 3.88 Å to 3.94 Å). MoS2 remains unstrained. A 12 × 12 × 1 Γ-centred k-grid mesh for layered structures and a 6 × 1 × 1 mesh for heterostructure were used. The DFT + D2 type of vdW correction was adopted to properly describe the interlayer interactions54,55.

We note that although Heyd–Scuseria–Ernzerhof hybrid functional56 calculation gives a gap of approximately 3 eV for CrOCl, which is more consistent with optical measurement57, the calculation yields a highly over-estimated gap of 1.55 eV for MoS2 (compared with experimental gap 1.29 eV (ref. 58)). Therefore, we adopt a PBE functional for MoS2 and the DFT + U method for CrOCl. When U = 3.0 eV, the CBM of CrOCl is lower than that of MoS2 (calculated using DFT + U and PBE) by 0.465 eV based on work-function calculations, while the CBM of CrOCl is 0.37 eV lower than that of MoS2 based on hybrid-functional-based work-function calculations. Therefore, the CBM energy positions obtained from our calculations are quite consistent with those obtained from hybrid functional calculations. The underestimated gap of CrOCl is attributed to the higher VBM energy position, which is always far (approximately 1–2 eV) below the chemical potential anyway, thus is irrelevant with regard to the mechanism discussed in this work.

Based on the above, we then calculated the band structures of MoS2–CrOCl heterostructure without electric field (Supplementary Fig. 3a), as well as those of MoS2 slabs and CrOCl slabs under electric fields (Supplementary Fig. 3b). The detailed evolution behaviour of the band edges of CrOCl and MoS2 as a function of vertical electric field in these scenarios can be seen in Supplementary Table 1. When a negative electric field of approximately 0.15 V nm−1 is applied (corresponding to the situation of negative bottom gate in our setup), the CBM of CrOCl is slightly lowered in energy, by 0.16 eV compared to the case without electric field, while the VBM of MoS2 is dramatically increased in energy, by approximately 0.25 eV, such that it is only about 0.074 eV above the CBM of CrOCl. Meanwhile, the electron carriers that are transferred to the surface CBM of CrOCl are expected to be frozen to form an electronic crystal state driven by ee interactions, owing to the large effective mass and small carrier density. And the chemical potential of the heterostructure would be lowered and getting closer to the VBM of MoS2 as illustrated in Fig. 1e, which thus can be easily tuned to be p-type upon further increasing the negative gate voltage. Therefore, it is very likely the subtle interplay among the gate-tunable band alignment, charge transfer and ee interactions that results in a chemical potential resident in the vicinity of MoS2 VBM and eventually leads to the effective gate tuning from n-type to p-type carriers. Such a mechanism is believed to be universal and applicable to a wide range of 2D semiconducting materials (M. L. and J .Liu, manuscript in preparation).

Electrical performance of the p-FETs

We note that, as shown in Supplementary Fig. 17, in the scenario of TMD–CrOCl, the thinnest working p-type doped MoSe2 layer was found to be around 3.2 nm (about four layers). This might be due to the fact that thinner TMDs have larger band gaps with lower VBMs, which do not fulfil the required Fermi level down-shift as calculated by our DFT results, as also illustrated in Supplementary Fig. 4. We did find that other replacements, such as few-layered Cr2Ge2Te6, can effectively dope monolayered MoS2 into a p-type FET, as shown in Supplementary Fig. 19. In this case, even the on-state threshold voltage Vth can be tuned to different values as compared to the value in TMD–CrOCl devices. Further, TMD–CrOCl FETs with different gate dielectric materials have also been fabricated, showing remarkable p-type field-effect behaviour, as shown in Supplementary Fig. 20. Our investigations also reveal that using a delicate annealing process with Ti–Au electrodes in an Ar–H2 (30:4 sccm) mixture, the output currents of both p-type MoS2–CrOCl and MoSe2–CrOCl can be significantly improved (Supplementary Fig. 21) compared to the typical Cr–Au contacted device shown in Fig. 2a.

Characterization of contact resistance

It is important to have an estimation of temperature dependence of the hole mobility of these TMD–CrOCl systems. Taking CrOCl-coupled few-layer MoS2 and MoSe2, for example, we found a significant suppression of the hole mobility upon cooling below 200 K (Extended data Fig. 8). This suggests that, although the IV curves are rather linear (Fig. 4f) in the Vds range of ±0.1 V at room temperature, there is still a tiny contact barrier, which is detrimental in low-temperature performance. TLM measurements show that this barrier is detrimental when the channel is approaching sub-200 nm, as illustrated in Fig. 4g. Nevertheless, it is the best performance among TMD-based p-type transistors, to our knowledge. The contact resistance was determined to be 8.8 KΩ μm using the TLM, as shown in Extended data Fig. 9. Notice that to obtain ultrascaled sub-50-nm channel lengths, a higher precision lithography tool may be needed. In the scenario of very small samples, manual alignment will also pose limitations. Larger-sized films would be preferable for potential practical applications.