Our vdW dielectric film is fabricated via STED (Fig. 1a) at room temperature. We use IMC Sb2O3 powder as the evaporation source. It consists of ultra-small Sb4O6 molecules in the form of bicyclic cages (Sb-O bond length 1.98 Å) [20–22]. All the molecules are bonded together via vdW interaction (Fig. 1b). Due to such a weak intermolecular interaction, Sb2O3 molecules are prone to sublime at elevated temperature in high vacuum before its melting, without the breakage of the Sb-O bonds. Its evaporation temperature is measured to be around (490℃) via thermogravimetric analysis at ambient pressure (Figure S1), much lower than its melting point (656℃), which implies the sublimation process of molecules at elevated temperature (The evaporation temperature in our vacuum chamber is supposed to be much lower according to Clausius-Clapeyron relation). The molecular vapor evaporated from Sb2O3 powder, free of dangling bonds all around, deposit on the substrate to form a vdW substrate (Fig. 1a). Our STED process permits the fabrication of homogeneous wafer-scale film (Fig. 1c and Figure S2) and the film thickness can be precisely controlled (Fig. 1d and Figure S2). Atomic force microscopy (AFM) is used to characterize the morphology of our film. It reveals its great homogeneity and flatness of our deposited film in micro-metric scale, without discernible voids or bumps (Fig. 1e and f). The homogeneity of morphology and elemental distribution are also confirmed by scanning electron microscope (SEM) and elemental maps of energy-dispersive X-ray spectroscopy (EDS) (Figure S3).
To confirm the molecular structure of our deposited film, Raman spectroscopy is employed to probe the vibrational modes of the molecules. We find that all the Raman peaks of Sb2O3 film can be assigned to the intra-molecule Raman mode (see Table S1) and match well with those of Sb2O3 powder used as the deposition source, implying that Sb2O3 evaporates in the form of molecular vapor with the molecular structure preserved. The stability of Sb2O3 molecules can be confirmed from the perspective of theoretical calculations via investigating the vacancy formation energy within the molecule. The formation energies of typical O, Sb, and double O vacancy are all found to be over 5 eV (Table S2), indicative of a robust molecule without the formation of vacancies and dangling bonds. Moreover, we investigate our eventual Sb2O3 film with X-ray diffraction (XRD), high-resolution transmission microscopy (HRTEM) as well selected-area electron diffraction pattern (SAED), which further reveal the polycrystalline structure of our Sb2O3 film(Figure S4). It is worth noting that the benign grain boundaries in our Sb2O3 film generally are free of dangling bonds [23], which in principle would not introduce effective defects into the dielectric.
We now investigate the dielectric properties of our Sb2O3 film. To determine its band gap, a 40 nm Sb2O3 film is deposited on glass substrate for absorption spectroscopy measurement (see Methods). The band gap of Sb2O3 film is determined to be 3.95 eV from its absorption spectrum (Fig. 2a), in good agreement with its density of states (DOS) distribution (Fig. 2b). Such a wide band gap renders our Sb2O3 film an insulator, as ascertained by the conductive test of a two-terminal device (Fig. 2c). Its resistivity of over 109 Ω∙cm at 300 K can be extracted from its I-V curve (Fig. 2c). To test its breakdown electric field, Sb2O3 film is sandwiched between two electrodes (Figure S6). As demonstrated in Fig. 2d, the breakdown voltage of our Sb2O3 film can be estimated to be 180 MV/m, comparable to that of SiO2.
In order to determine the dielectric constant of Sb2O3, we fabricate a series of parallel-plate capacitors, using Sb2O3 of 300 nm as the dielectric sandwiched between degenerately doped Si substrate and metal pads. The capacitance with respect to area S are measured at increasing frequency are respectively measured (Fig. 2e inset). We then extract the static capacitance at low frequency (10 kHz) and estimate the capacitance per unit area to be C = 0.34 nF/mm2. The relative dielectric constant of our Sb2O3 film then can be calculated to be εr = 11.5 via the formula C = εrε0/d where ε0 is dielectric constant of vacuum and d is the thickness of Sb2O3 film. The measured dielectric constant of our Sb2O3 matches well with the reported value [24]. In addition, such a high dielectric constant is comparable to typical high-κ dielectric Al2O3. As the low dielectric constant of hBN (~ 5 [12]) is one of its limitations in the FET application, our Sb2O3 vdW dielectric, in contrast, possesses the highest dielectric constant in all the reported vdW dielectrics (Figure S7).
In analogy to typical vdW dielectric h-BN, our Sb2O3 film of molecular crystal is free of dangling bonds, naturally holding low-density charge scattering centers and charge trap states. Our vdW film can potentially support high-performance 2D electronic devices of higher mobility and smaller hysteresis as well. To demonstrate its advantages in this regard, we fabricate 2D FETs respectively on our Sb2O3 vdW substrate and standard SiO2 substrate, then systematically investigate their temperature-dependent device characteristics.
The well-studied 2D semiconductor MoS2 is chosen as the representative channel materials and all the FET devices are fabricated using the same processes (see Methods). For a clear comparison, our Sb2O3 substrate is composed of 40-nm film deposited on the standard SiO2/Si substrate to fully screen the charged center and trap states on SiO2 substrate (Fig. 3a). 2D MoS2 flakes are prepared via mechanical exfoliation and transferred onto the substrates. Their thickness are confirmed via optical measurement (Figure S8) and AFM (Figure S9a and b) before the device fabrication. Degenerately doped Si serves as FET back gate in our measurement (Fig. 3a). To minimize the effect of contact resistance in our devices, we use In/Au metal to form low-resistance contact with MoS2 [25]. Moreover, we also degas all the devices in high vacuum (10− 6 torr) for 3 hours to reduce air adsorption on our device surface before the tests are carried out [26]. The FET based on monolayer MoS2 supported on Sb2O3 substrate is demonstrated in Fig. 3b.
For monolayer MoS2, the band offsets at valance band maximum (VBM) and conduction band minimum (CBM) can be estimated from its band alignment to Sb2O3 (the vacuum level at 0 eV while the well-known VBM and CBM of MoS2 are extracted from reference [27]) (Fig. 3c). These band offsets over 1 eV effectively confine the charges within the MoS2 channel during the device measurement. The Ohmic contact of electrodes to 2D MoS2 and great gate control of our FET can be verified from the linear output curves (Ids-Vds) both at 40 K (Fig. 3d) and 300 K (Figure S9c). From the typical double-sweep transfer characteristics curves (Ids-Vgs) of monolayer MoS2/Sb2O3 FET measured at 40 K( Fig. 3e), we observe a negligibly small hysteresis window (sweep rate in all our measurements ~ 1 V/S) and estimate the electron mobility µFE at the linear range to be over 80 cm2/Vs (see Methods). In contrast, monolayer MoS2 FET on SiO2 exhibits a considerable hysteresis and much lower mobility (~ 20 cm2/Vs) despite the same device fabrication processes and measurement conditions (Figure S13). For a clear comparison of dielectric effect (Sb2O3 and SiO2) on FET mobility, we plot together their electron mobility at various gate voltage (Vgs) (Fig. 3f). The maximum electron mobility appears at around Vgs = 50V, corresponding to the charge carrier density n = 4.0 × 1012/cm2 (see Methods), in agreement with the reported MoS2 FETs [25]. The FET mobility monotonically decreases at higher temperature for our devices (Fig. 3g), presumably due to rising phonon scattering. In comparison to the reported measurements at room temperature, our monolayer MoS2/Sb2O3 FET exhibits a mobility µFE of over 70 cm2/Vs, even higher than the reported value of MoS2 supported on hBN [12, 16] (Fig. 3g).
Similar contrast experiments are also carried out using few-layer MoS2 as channel materials. The thickness-dependent mobility of supported different substrates are measured at various temperature (Fig. 3g). The comparative advantages of our vdW substrate for thicker MoS2 can also be clearly identified though the mobility improvement becomes less apparent, by a factor of 2 (from 40 cm2/Vs on SiO2 to 90 cm2/Vs on Sb2O3 for trilayer MoS2 at 40 K ). This is generally attributed to the rise of screening effect for thicker MoS2 to the charged disorders on the underlying dielectric. Interestingly, the MoS2 thickness more sensitively affect the mobility for the SiO2 supported devices (mobility changes by a factor of 2 for monolayer and trilayer MoS2) while MoS2 of various thickness on Sb2O3 demonstrate similar mobility (see Fig. 3g). This also implies the low-density of disorders on our Sb2O3 substrate, without apparent scattering to the carrier transport of all the MoS2 channel.
We now focus on the hysteresis of our MoS2 FET to investigate the charge trapping states of our vdW Sb2O3 and SiO2 substrates. The hysteresis of transfer characteristic curves features the instability of a FET at work, usually caused by the trapping states located in channel semiconductors, dielectric and at their interface [28]. Using vdW dielectric has been proved effective to minimize the hysteresis of 2D semiconductor FETs [12, 29]. In Fig. 3e, we already demonstrated the ultra-small hysteresis in the double-sweep transfer curves. Such a small hysteresis demonstrate a clear contrast to that obtained from a typical SiO2-supported monolayer MoS2 FET (Fig. 4a). The variation of the onset voltage ΔVon, which is usually used to quantify the FET hysteresis, reduces over an order of magnitude from 5.1 V for MoS2/SiO2 to 0.24 V for MoS2/Sb2O3 FET at 40 K. We also investigated the temperature-dependent transfer characteristics of our FET (Fig. 4b and d). We firstly note that the onset voltage (Von) position apparently shifts toward lower voltage with the increasing temperature, presumably resulted from the Fermi level downward shift due to rising thermal excitation in MoS2, in agreement with the reported works [30, 31]. As to the amount of hysteresis ΔVon, the temperature variation from 40 K to 300 K leads to a slight increase of hysteresis for MoS2/Sb2O3 FET. Such a small dependence implies to a low density of effective trap states within our Sb2O3 dielectric as the hysteresis is generally induced by the charge carriers trapped into the trap states during the FET on/off switching (Fig. 4). In contrast, the hysteresis of monolayer MoS2/SiO2 FET sensitively depends on the temperature (Figure S13). The double-sweep transfer curves exhibit a large hysteresis window at 300 K and its ΔVon reaches 12 V. This observation accordingly points to a high density of trap states on SiO2 substrate, which can even be thermally activated at higher temperature [31].
To confirm the source of trap states, we also investigate the thickness-dependent hysteresis of MoS2 FETs. For all the MoS2/Sb2O3 FETs at various temperature, the transfer characteristic curves exhibit small hysteresis window and ΔVon negligibly depends on MoS2 thickness (Fig. 4d). Such an observation implies that the trap states are not caused by the bulk defects in MoS2 considering that the density of such trap states are in principle thickness-dependent. For MoS2/SiO2 FETs, as one can anticipate, the hysteresis demonstrates no obvious dependence on MoS2 thickness but increases considerably with temperature due to the thermal activation of trap states at higher temperature. As our Sb2O3 substrates are exposed in air for long time (a few days), air adsorption onto our Sb2O3 film may introduce some the trap states and leads to the hysteresis [25]. Our experimental results, however, ruled out this possibility. It turns out that the typical gas can hardly adsorb onto Sb2O3 molecules, as revealed by our theoretical calculations owing to its inert surface (see Table S3).
In order to quantitate the trap states density on Sb2O3 and SiO2 substrates, we investigate the variation of threshold voltage ΔVth in double-sweep transfer characteristic curves, which correlates to the charge of trap states density ΔQ according to ΔVth = ΔQ × C [28], where C and ΔQ respectively stand for the gate capacitance and trapped charges. The values of ΔVth for all our MoS2/Sb2O3 devices can hardly be precisely extracted by linearly extrapolating the transfer curves due to the almost negligible hysteresis window (see Fig. 3e, Figure S9 d). We estimate ΔVth for monolayer MoS2/Sb2O3 device to be lower than 0.1 V without observable dependence on temperature, thus corresponding to a trap states density of 6.9 × 109/cm2. In contrast, the trap states density on SiO2 demonstrates a trap charge states of 4V at 40 K. It furthermore increases to 9V at 300 K under thermal activation, corresponding to a trap states density 4.3 × 1011/cm2, which matches well with the reported value extracted from MoTe2 FET [28]. We thus confirm a remarkable reduction of trap states by nearly two orders of magnitude for our Sb2O3 substrate with respect to SiO2.
Our approach substantially relies on the particular structure of Sb2O3 and its excellent insulating properties. Our results may merely open up the opportunities for the scalable fabrication of vdW dielectrics via compatible processes. Evidently, such an approach of compatible fabrication is not limited to the Sb2O3, but applicable to other IMCs. In this regard, it would be of great interest to explore other IMCs (for instance with large bandgap and higher dielectric constants). As to our Sb2O3 film, the fabrication process can still be optimized via the modulation of substrate temperature and deposition rate in more advanced deposition systems. For instance, the deposition at low temperature may lead to formation of amorphous Sb2O3 film, which may furthermore modulate the film morphology (such as film roughness) as well as dielectric properties.
As a representative example, the monolayer MoS2/Sb2O3 FETs demonstrated a mobility enhancement of 4 times using a FET measurement. However, the contact issue at the InAu-MoS2 interface may still exists for two-terminal devices, possibly leading to some underestimation of the vdW dielectric effect on the FET mobility enhancement. Its full potentials may be realized via the measurement of Hall devices. In principle, such a vdW dielectric may improve the device performance based on other 2D materials [2] and can be potentially used in other device architectures [32–34].