Introduction

Two-dimensional (2D) semiconductors hold great promise for fabricating more-than-Moore transistors and exploring the emergent transport properties1,2,3,4,5, but achieving their theoretical performance also requires compatible high-k dielectrics to guarantee efficient gate control6,7,8,9,10. Compared to traditional amorphous dielectrics (Al2O3 and HfO2)11,12,13, single-crystalline gate insulators with dangling-bond-free surfaces are more competitive to fabricate high-performance 2D devices with reduced interfacial scatterings and gate hysteresis14,15,16,17,18. For example, hexagonal boron nitride (h-BN) has been widely used as a van der Waals (vdWs) substrate to improve carrier mobility and investigate exotic properties of 2D materials19,20,21,22,23. However, h-BN is also well known for its shortcomings of low dielectric constant (k = 2~4) and harsh conditions (pressure: >4 G Pa, temperature: >1673 K) for the growth of high-quality single crystals24. Therefore, it is highly desirable to discover new vdWs insulators similar to h-BN, but has a much higher dielectric constant and more facile synthetic conditions, for exploring the emergent transport properties of 2D materials in a high-κ dielectric environment, as well as for fabricating 2D-material-based electronic devices with scaled supply voltage15,16. Even so, very rare have succeeded18,25. Very recently, Peng et al.18 reported the synthesis of Bi2SeO5 bulk crystals grown by chemical vapor transport (CVT), which can be exfoliated into few-nanometer-thick nanosheets and serve as an “h-BN” like high-k dielectric to improve the mobility of 2D materials and enable the observation of quantum Hall effects in Bi2O2Se. Nevertheless, the CVT process for the synthesis of Bi2SeO5 bulk crystals is also very time-consuming (typically ~40 days). Besides, similar to other vdWs dielectrics, transferable Bi2SeO5 nanoflakes with suitable thicknesses and domain sizes were also prepared by a low-efficiency method of mechanical exfoliation for subsequent vdWs device integration.

Compared to mechanical exfoliation, direct growth of free-standing ultrathin 2D insulators with high-k nature by chemical vapor deposition (CVD) is much more efficient, but remains challenging. Typically, CVD-grown atomically thin 2D insulators with a layered crystal structure preferably adopt an in-plane growth mode on substrates26,27,28,29, which will inevitably set obstacles for clean sample transfer and subsequent vdWs integration. However, if an ultrathin 2D insulator can be vertically grown on the substrate, just like the case in Bi2O2Se3,30,31, the transfer problem can be easily overcome due to much reduced interfacial interaction.

Bi2SiO5 is a well-known high-κ dielectrics with an Aurivillius-type layered structure and a large band gap of 3.5~4.4 eV32,33,34,35. According to the previous works regarding the bulk single crystals and polycrystalline powders of Bi2SiO5, Bi2SiO5 shows an anisotropic dielectric constant36,37 and its out-of-plane dielectric constant can be as high as 30~8036,37,38,39, and thus was suggested as a potential candidate for high-temperature dielectrics39. Here, Bi2SiO5 was demonstrated as an excellent gate dielectric for 2D semiconductors. Ultrathin Bi2SiO5 single crystals with thickness down to monolayer were successfully synthesized by a facile CVD method, concurrently owing to the high dielectric constant (>30), large band gap (~3.8 eV), and large breakdown field strength. Remarkably, the preferable CVD growth mode of Bi2SiO5 can be regulated from in-plane to out-of-plane under optimized conditions, showing great feasibility on sample transfer by polymer-free mechanical pressing. Using ultrathin Bi2SiO5 nanoflakes as the vdW dielectrics and screening layers, we can greatly regulate the carrier density and improve the carrier mobility of few-layer MoS2 (almost fifteen times higher than on the SiO2 substrate at 5 K). Besides, the MoS2 field-effect transistors (FETs) using Bi2SiO5 as dielectrics can operate at 0.5 V, exhibiting a large Ion/Ioff (>108), an ignorable hysteresis (~3 mV), low DIBL value (~5 mV/V) and low gate leakage current (~10−13 A).

Results

Structure, CVD growth, and characterization of layered Bi2SiO5

As shown in Fig. 1a, bismuth silicate Bi2SiO5 possesses a monoclinic lattice with Cc space group (quasi-orthogonal, a = 15.12 Å, b = 5.44 Å, c = 5.29 Å, β = 90.07°) and has an Aurivillius-type layered crystal structure with alternatively stacked [Bi2O2]n2n+ and [SiO3]n2n- layers along the a-axis. The first-principle calculations were performed to investigate the band structure of layered Bi2SiO5. As shown in Fig. 1b, Bi2SiO5 exhibits a large direct band gap of ~3.79 eV, whose conduction band minimum (CBM) and valance band maximum (VBM) are both locate at Γ point of the first Brillouin zone and mainly originate from Bi-p and O-p orbits, respectively.

Fig. 1: Structure, growth, and characterization of ultrathin Bi2SiO5 single crystals.
figure 1

a Crystal structure of Bi2SiO5 (Cc, a = 15.12 Å, b = 5.44 Å, c = 5.29 Å, β = 90.07°) with alternatively stacked [Bi2O2]n2n+ and [SiO3]n2n− layers. b Calculated band structure and density of states (DOS) of Bi2SiO5 with a direct band gap of ~3.79 eV. The first Brillouin zone is inserted in the right panel. c Typical optical micrograph (OM) image of square Bi2SiO5 nanoplates showing an in-plane growth mode on mica substrate. d OM images of Bi2SiO5 nanoplates with thickness-dependent color contrasts on mica. The inset shows the typical atomic force microscope (AFM) image of an ultrathin Bi2SiO5 nanoplate with a thickness of 3.9 nm (5 layers) and an atomically smooth surface. e Scanning electron microscopy (SEM) image of 2D Bi2SiO5 crystals vertically grown on mica substrate. f Thickness-dependent color contrasts for Bi2SiO5 nanoplates transferred onto SiO2/Si substrate by a polymer-free mechanical pressing. The AFM image of a 7.5-nm-thick Bi2SiO5 nanoplate was inserted in f. g Typical AFM image of a terraced Bi2SiO5 nanoplates with a clear step of 0.76 nm. h, i Cross-sectional atomic-resolved high angle annular dark field (HAADF) images (left) and corresponding fast Fourier transform (FFT) diffraction spots (right) of chemical vapor deposition (CVD) grown Bi2SiO5 nanoplates taken along the zone axes of [010] (h) and [011] (i), respectively.

The essential prerequisite for using Bi2SiO5 as single-crystalline dielectrics with strong gate control is to achieve its growth of atomically thin films. However, the CVD growth of ultrathin Bi2SiO5 crystals remains a challenge yet. To our knowledge, SiO2 is chemically inert and has a very high melting point, which prevents it to be used as Si supplier during the CVD growth of Si-based compounds to some extent. Nevertheless, it’s well known that the fluorides will react with SiO2 to form volatile Si-based precursors. As a result, we developed a facile CVD method for the synthesis of Bi2SiO5 ultrathin crystals by using the BiF3 powders as Bi supplier and SiO2 powders or quartz boat as Si supplier (see Supplementary Fig. 1). With this method, ultrathin Bi2SiO5 crystals with various thicknesses can be readily obtained on mica substrates. As shown in Fig. 1c, d and Supplementary Fig. 2, at a relatively high growth temperature (~1023 K), Bi2SiO5 adopts a preferable in-plane growth mode on mica, which is widely observed in the CVD synthesis of layered materials, revealing a transparent square-like shape and atomically thin nature (down to monolayer, Supplementary Fig. 3). Normally, the in-plane growth mode will result in a large attaching area and strong bonding force between the epitaxial layer and the substrate, thereby adding difficulties for sample transfer and subsequent vdWs integration. Remarkably, the preferable growth mode of Bi2SiO5 on mica can be regulated from in-plane to out-of-plane growth modes just by lowering its growth temperature, which is similar to the case of Bi2O2Se30. As shown in Fig. 1e and Supplementary Fig. 2b, c, vertically grown ultrathin Bi2SiO5 crystals gradually emerge on mica at ~973 K, then dominates the CVD growth while further lowering the temperature to ~923 K. Unlike the case in Fig. 1c, vertically grown Bi2SiO5 crystals with thickness ranging from 7.5 to 50 nm can be easily transferred onto various substrates (such as SiO2/Si, Fig. 1e) by a polymer-free mechanical pressing (Supplementary Fig. 4), showing an atomically flat surface even after sample transfer (Fig. 1f). The transferable feature without unfavorable residual polymer contaminations makes free-standing Bi2SiO5 appealing for fabricating vdWs heterojunction device. Occasionally, the CVD-grown Bi2SiO5 nanoflakes revealed a terraced morphology with a step height of ~0.76 nm (Fig. 1g), consistent with the theoretical value for the monolayer step in Bi2SiO5. It’s worth noting that the CVD-grown Bi2SiO5 shows excellent air stability, whose surface morphology and roughness of Bi2SiO5 nanoplates remain almost the same when exposed to air for more than 7 months, which is an important metric for device fabrication as a high-κ gate dielectric (Supplementary Figs. 5, 6). More details about the comparison experiments conducted to understand how the reaction goes in the CVD growth of Bi2SiO5 and the possible reason for its in-plane and out-of-plane growth can be found in the discussion part of supporting information (Supplementary Figs. 7, 8).

The crystalline phase of as-grown samples was confirmed by transmission electron microscopy (TEM), Raman spectroscopy, and X-ray diffraction (XRD). As shown in Fig. 1h, i and Supplementary Fig. 9, we performed high-resolution TEM imaging along three-zone axes [namely (010), (011), and (100)], as well as corresponding fast Fourier transform (FFT) fringes. Based on the atomic-resolved cross-sectional TEM technique, the alternative stacking of [Bi2O2]n2n+ and the [SiO3]n2n− layers with a layer space of 0.76 nm was observed. Besides, the averaged atomic ratio of Bi/Si was measured close to 2: 1 by energy dispersive spectroscopy (EDS, Supplementary Figs. 9, 10), which is consistent with the chemical formula of Bi2SiO5. The active modes at 70, 97, 149, 208, 298, 372, 433 cm−1 of the Raman spectra also matched well with monoclinic Cc phase of Bi2SiO5 (Supplementary Fig. 11)35. Additionally, the sharp XRD peaks, which can be assigned to (h00) crystal planes, further confirmed the high crystalline quality of the CVD-grown Bi2SiO5 nanoplates (Supplementary Fig. 11c).

Dielectric properties of vertically grown Bi2SiO5 nanoplates

The transfer feasibility for vertically grown Bi2SiO5 nanoplates greatly facilitates the evaluation of their intrinsic properties, such as dielectric constant, band gap, and breakdown field strength. As shown in Fig. 2a, metal-insulator-metal (MIM) capacitors were fabricated on quartz substrates to extract the dielectric constant of Bi2SiO5 nanoflakes with varied thicknesses by capacitance-voltage (CV) measurements. Here, the thick graphite and In/Au metals serve as the bottom and top electrodes, respectively. Thanks to high-κ nature, Bi2SiO5 nanoflake with a thickness of 25.6 nm demonstrated a very high capacitance density of 1.12 μF/cm2 at 100 Hz (Fig. 2a), revealing a slight decrease when the measuring frequency is up to 1 MHz. The corresponding capacitance-frequency (Cf) measurements showed the similar results (Fig. 2b). It is worth noting that the absolute capacitance value (>1 × 10−12 F) measured is ~2~3 orders higher than the instrument’s offset and noise level (<1.5 × 10−14 F, Supplementary Fig. 12). Based on the CV and Cf data, we can estimate the effective permittivity (ɛeff) of Bi2SiO5 to be ~32.4 at 100 Hz, which is preferable than commercial amorphous high-κ oxide such as Al2O3 (k = 7~9)13,40,41 and HfO2 (k = 13~17)41,42, and higher than most of reported vdWs single-crystalline dielectrics, such as h-BN (k = ~3.5)43,44, CaF2 (k = 8.4)14, Bi2SeO5 (k = 15.6)18, VOCl (k = 11.7)45 and ZrO2 (k = 8~19)17. Furthermore, we investigated the thickness-dependent ɛeff by fabricating MIM capacitors with different Bi2SiO5 thicknesses, from which we can extract very large ɛeff values of >30 in a wide thickness range (Fig. 2c). The gradual decrease of ɛeff while thinning down can be ascribed to the existence of interfacial “dead layer” in the MIM device, which is similar to other MIM devices15,46,47,48.

Fig. 2: Dielectric constant, band gap, and breakdown field strength of CVD-grown Bi2SiO5 nanoflakes.
figure 2

a Typical Bias-dependent capacitance (C) measurements on CVD-grown Bi2SiO5 nanoflake with a common metal-insulator-metal (MIM) device configuration, where the thick graphite and In/Au metals serve as the bottom and top electrodes (inset), respectively. The dielectric constant (ɛr) of Bi2SiO5 was estimated as ~32.4 at 100 Hz. b Corresponding frequency-dependent capacitance (Cf) characteristics of the MIM device, whose OM image is inset in b. c The thickness-dependent effective permittivity (εeff) of Bi2SiO5 nanoflakes with a measuring frequency of 100 Hz. The dashed blue curve is a visual guide. d Ultraviolet-visible (UV-vis) absorption spectrum of CVD-grown Bi2SiO5 nanoflakes transferred onto quartz substrate with high coverage by mechanical pressing. The inset shows the fitting of its optical band gap (~3.8 eV) by Tauc’s law, where a, h, and ν are the absorption coefficient, Planck constant and frequency, respectively. e Energy band gap versus dielectric constant of representative dielectric materials in literature, showing the coexistence of high dielectric constant and large band gap in Bi2SiO5. The dashed line is a visual guide. f Thickness-dependent current-voltage curves of Bi2SiO5 nanoplates measured by C-AFM, showing a high breakdown field strength of 9.4 MV/cm (8.2 nm).

An ideal gate dielectric also needs a large band gap to inhibit the current leakage. Here, vertically grown Bi2SiO5 nanoflakes were directly transferred onto the polished quartz substrate by mechanical pressing for the ultraviolet-visible (UV-vis) absorption measurements. As shown in Fig. 2d, the optical band gap of CVD-grown Bi2SiO5 can be extracted as ~3.80 eV by Tauc’s law, which is consistent with the theoretical value (3.79 eV, Fig. 1b). To illustrate the metrics of Bi2SiO5 as a gate dielectric, the relationship between band gap and dielectric constant of representative gate dielectrics in literature was plotted in Fig. 2e, clearly indicating the coexistence of high dielectric constant and large band gap in Bi2SiO5. Moreover, we evaluated the breakdown field strength (Ebd) of the Bi2SiO5 nanoplates by conductive atomic force microscope (C-AFM) measurements on p++ Si. As shown in Fig. 2f, the CVD-grown Bi2SiO5 showed thickness-dependent breakdown field strength ranging from 7.2 (13.5 nm) to 9.4 (8.2 nm) MV cm−1, which is close to that of dielectrics in the silicon industry such as Al2O349 and HfO250 but owns higher dielectric constant. We should emphasize that the C-AFM is a very local and microscopic tool to measure the breakdown voltage of a dielectric insulator, which may be not the same as the global one determined by MIM device. For example, the breakdown experiments based on the MIM devices gave a breakdown field strength of 3~5 MV/cm instead while varying the thickness of Bi2SiO5 from 10.1 to 21.4 nm (Supplementary Fig. 13).

The features of easy-to-transfer, ultra-flat surface, high dielectric constant, large band gap, and breakdown field strength make Bi2SiO5 highly competitive as gate dielectrics and high-κ substrates for dielectric screening. Here, we combine Bi2SiO5 with mechanically exfoliated few-layer MoS2 to construct Bi2SiO5/MoS2 FETs for demonstrating its advantages as back-gate dielectrics (Fig. 3), dielectric screening substrate (Fig. 4) and top-gate dielectrics (Fig. 5).

Fig. 3: MoS2 Hall-bar device with the vertically grown ultrathin Bi2SiO5 nanoflake as the high-κ back-gate dielectrics.
figure 3

a OM image of an as-fabricated MoS2 (~2.7 nm) Hall-bar device on SiOx/Si substrate, in which a thin flake of Bi2SiO5 (~22 nm) and multi-layer graphene were adopted as gate dielectric and back-gate electrode, respectively. b Hall resistance (Rxy) as a function of magnetic field (B) under various gate voltages (Vg) from 0 to 2 V at 300 K. c The extracted Vg-dependent sheet carrier density (n2D), in which a high dielectric constant (εr) of ~36 was derived by linear fitting. The Error bars are based on the standard deviations of slope fitting on Rxy-B curves. d Longitudinal resistance (Rxx) as a function of temperature (T) under different Vg from 0.4 to 2.0 V, showing a gate-induced insulator-to-metal transition. e Two-probe transfer curves of the back-gated MoS2/Bi2SiO5 field-effect transistor (FET) measured by defining the probes of 1 and 4 as source and drain terminals, showing Ion/Ioff > 108, SS ~ 64 mV/decade. The inset shows a small hysteresis of ~15 mV. The extracted two-terminal field-effect mobility of the device is 18.6 cm2 V−1 s−1. f The corresponding output curves measured in a Vg range from 2 to −2 V.

Fig. 4: Dielectric screening and mobility enhancement effects of CVD-grown Bi2SiO5 nanoflakes as the high-κ substrates.
figure 4

a Schematic illustration and OM image of back-gate MoS2 four-probe FETs device. b Linear output curves (IdsVds) of MoS2 FETs on Bi2SiO5 (left) and on SiO2 substrate (right) measured at 300 K. c The 4-probe transfer curves of MoS2 FETs measured at different temperatures (5~300 K) on Bi2SiO5 (left) and SiO2 (right) substrates, respectively. d The extracted temperature-dependent 4-probe FET mobility (μFET,4-probe) of MoS2 on Bi2SiO5 (red) and SiO2 (blue) substrates.

Fig. 5: Hysteresis-free MoS2 FETs and low-power NMOS inverters using Bi2SiO5 as high-κ top-gate dielectrics.
figure 5

a Schematic illustration of the top-gated MoS2 FETs on SiO2/Si substrate with Bi2SiO5 as the top-gate dielectric. b Typical dual-sweep transfer curves of the MoS2/Bi2SiO5 FET measured under different Vds from 0.05 to 1 V, showing an ideal SS value of ~62 mV/decade and ignorable gate hysteresis. The Insert is the OM image of a fabricated MoS2 FET with Bi2SiO5 as the gate dielectric. c Corresponding output characteristics (IdsVds curves) of the device measured by varying the Vg from 0.5 to −0.6 V with a 0.1 V step. d Extracted SS value versus Ids characteristics of the device in b, showing a low SS value (<70 mV/decade) for a wide Ids range. e Dual gated transfer curves of the MoS2/ Bi2SiO5 FET under different back-gate voltages (VBG) from 5 to 0 V. f The extracted threshold voltage Vth from e as a function of VBG. The linear fitting yields a slope of −0.0097 and a high dielectric constant of ~32.3. g The thickness-dependent dielectric constant of Bi2SiO5 extracted by dual-gate measurement on Bi2SiO5/MoS2/SiO2/Si FET. The dotted curve is a visual guide. h Transfer curves of the MoS2 FETs with different thickness Bi2SiO5 as the top-gate dielectrics, showing a trend of smaller Vth for thinner Bi2SiO5 thickness. The Vds is 0.5 V, and no VBG is applied. i Measured output voltage (Vout) and gain as a function of input voltage (Vin) of an NMOS inverter based on two MoS2/ Bi2SiO5 FETs under different supply voltage (Vdd) from 2 V to 0.5 V.

Vertically grown Bi2SiO5 as back-gate dielectric

The small contact area and weak interaction between the vertically grown Bi2SiO5 and the mica substrate make it compatible with the well-developed aligned transfer method (For details, see Supplementary Fig. 14)19,20, allowing us to fabricate complex vdWs heterojunction devices by layer-by-layer stacking. Figure 3a showed the OM image of an as-fabricated MoS2 Hall-bar device using Bi2SiO5 as the back-gate dielectric and multi-layer graphene as a back-gate electrode. Gated Hall measurement is a powerful tool to get a series of key parameters, such as the dielectric constant of the gate dielectric, carrier density, and Hall mobility of the channel semiconductor. Besides, it also enables us to get the transfer and output characteristics by defining two out of six electrodes as source and drain. Figure 3b showed Hall resistance (Rxy) as a function of magnetic field (B) under different gate voltages (Vg) at room temperature, from which we can extract the Vg-dependent sheet carrier densities (n2D). By linear fitting the curve of n2D-Vg (Fig. 3c), the estimated slope equals the Cox/e, where Cox is capacitance density and e is the elementary charge. The Cox for a 22-nm-thick Bi2SiO5 can be as high as 1.45 μF/cm2, suggesting a very high dielectric constant (ɛr ~ 36), which is consistent with the value extracted by CV measurements in a reasonable accuracy. On the other hand, very large capacitance density suggests that we can greatly regulate the carrier density and electrical behavior of the MoS2 semiconductor. As confirmed by the room-temperature Hall measurements (Fig. 3c), a very high carrier density of 1.83 × 1013 cm−2 can be doped into the channel by applying a Vg of 2 V on Bi2SiO5 dielectrics. To this end, the temperature-dependent longitudinal resistance (Rxx-T) behavior of MoS2 can be greatly regulated from insulating to metallic one when sweeping the Vg from 0.4 to 2 V (Fig. 3d), indicating the great potential of Bi2SiO5 dielectric as a powerful tool for physical state regulation.

To examine whether ultrathin Bi2SiO5 can serve as excellent back-gate dielectrics of a FET, we define the electrodes #1, #4 and graphene as source, drain, and gate electrodes, respectively. The dual-sweep transfer curves of MoS2/Bi2SiO5/Gr FET were presented in Fig. 3e, showing a large Ion/Ioff of >108, and a small SS value of ~64 mV/decade. Considering the relatively complex device fabrication process (Supplementary Fig. 14), chemical residuals will inevitably remain at the interfaces between the graphene, Bi2SiO5 and MoS2, which are detrimental to the device performance. Nevertheless, the FET still exhibited small hysteresis of ~15 mV, which is comparable to the value obtained after interface optimization51. Thanks to the Ohmic contact formed by In/Au electrodes (Fig. 3f), two-terminal field-effect mobility of the device can be as high as 18.6 cm2 V−1 s−1 by linear fitting the transfer curve, which is comparable to the Hall mobility measured at room temperature (23.7 cm2 V−1 s−1 at Vg = 2 V, Supplementary Fig. 15). Notably, Bi2SiO5 was also verified as an excellent gate insulator while shrinking the channel length of MoS2 FETs down to 100 nm and even 30 nm (Supplementary Figs. 16, 17).

Vertically grown Bi2SiO5 as high-κ screening layer

Dangling bonds on SiO2/Si substrate usually act as the scattering sites of Coulomb impurities (CI) to decrease the mobility of a semiconductor52,53,54. Theoretically speaking, using an h-BN-like high-κ substrate free of dangling bonds can effectively enhance its mobility due to reduced CI scatterings and excellent dielectric screening. Figure 4a shows the scheme and OM image of as-fabricated 4-probe MoS2 FETs on top of Bi2SiO5 and SiO2/Si substrates. To avoid the influence of sample quality variation, one MoS2 sample (4.3 nm) was simultaneously placed on Bi2SiO5 (31.9 nm) and SiO2 substrates for comparison. As demonstrated in Fig. 4b, the Vg-dependent two-probe IV curves (output characteristics) were measured at 300 K, showing a linear behavior over a large voltage window, confirming the Ohmic contact formed by In/Au electrodes. It’s worth noting that the on-state current (Ion) of MoS2 on Bi2SiO5 is about twice as the one on SiO2, suggesting a higher mobility of MoS2 on Bi2SiO5 substrate. To illuminate the influence of contact resistance, 4-probe transfer curves were measured among a temperature range of 5~300 K (Fig. 4c). By linear fitting the transfer curves, we can extract the 4-probe FET mobility (μFET,4-probe) as a function of temperature (Fig. 4d). Among the whole temperature range of 300~5 K, the carrier mobility of MoS2 on Bi2SiO5 is significantly higher than that on SiO2/Si substrate, which can be further confirmed by plotting the Ids as a function of a normalized Vg by threshold voltage (Vg-Vth, Supplementary Fig. 18). Particularly, the mobility of MoS2 on Bi2SiO5 at 5 K is as high as 549.3 cm2 V−1 s−1, which is almost fifteen times higher than the value of MoS2 on SiO2/Si (~37.4 cm2 V−1 s−1). Moreover, the carrier mobility of MoS2 on Bi2SiO5 and SiO2/Si showed totally different temperature dependence. For MoS2/Bi2SiO5, its carrier mobility increased monotonously upon cooling down, suggesting the phonon scattering dominated the whole transport events even at low temperature. In contrast, the carrier mobility of MoS2 on SiO2/Si substrate increased first while cooling down to ~150 K, but gradually decreased upon further cooling. This T-dependent mobility can be well explained by enhanced charge impurities scattering at low temperature in MoS2/SiO2 interface. In a word, using Bi2SiO5 as a substrate greatly improve the performance of the MoS2 FET to get a higher mobility, which can be attributed to the suppressed CI scatterings in the high-κ surroundings and ideal dielectric/semiconductor interface.

Vertically grown Bi2SiO5 as the top-gate dielectric

Top-gate FET is a widely used device configuration in practical applications. In this part, we fabricated Bi2SiO5/MoS2 top-gate FETs to prove the feasibility of Bi2SiO5 as top-gate dielectric layer in hysteresis-free low-power transistors (Supplementary Fig. 19). As shown in Fig. 5a, the Bi2SiO5/MoS2 top-gate FET was placed on the SiO2/Si substrate, which can also serve as the back-gate dielectric and electrode when necessary. Figure 5b showed the typical dual-sweep transfer curves of a 22.9-nm-thick Bi2SiO5-based MoS2 FET measured under different Vds from 0.05 to 1 V. Owing to the excellent gate tunability, the transistor can be effectively turned on and off by applying a Vg within the range of −0.6~0.5 V, showing a large Ion/Ioff of >106 and a low gate leakage current of 10−13 A (approaching the detection limit). Furthermore, the ideal SS value (~62 mV/decade), ignorable gate hysteresis (~3 mV), and low drain-induced barrier lowering (DIBL, ~5 mV/V) suggest the perfect interface was formed at MoS2/Bi2SiO5 interface (Supplementary Fig. 20). The MoS2 transistor showed a linear Ids-Vds curve at low Vds region, then gradually saturated at high Vds region (Fig. 5c). Particularly, the SS remains low (<70 mV/decade) for different Ids of several orders of magnitude for both forward and reverse top-gate sweeping (Fig. 5d). More importantly, the top-gate Bi2SiO5/MoS2 short-channel FET, whose channel length was defined by the gap distance between two graphene electrodes (~180 nm), still showed a small DIBL value of ~22 mV/V and SS value of ~79 mV/decade (Supplementary Fig. 21).

One step further, nearly hysteresis-free transfer curves were preserved in a dual-gate FET configuration, in which a back-gate voltage (VBG) was employed to modulate the threshold voltage (Vth) of the device. As a result, the top-gate transfer curves gradually shifted as the VBG varied from 5 to 0 V (Fig. 5e and Supplementary Fig. 22). By linear fitting the Vth as a function of VBG, we can extract a slope of −0.0097, which equals to the ratio of the bottom-gate to top-gate capacitance, namely C(SiO2)/C(Bi2SiO5), when the parallel-plate capacitor model is assumed for both top and bottom gates16,55. For a 285 nm SiO2 dielectric (ɛr = 3.9), its capacitance can be calculated as 0.0121 μF/cm2. In this case, the capacitance and dielectric constant for a 22.9-nm-thick Bi2SiO5 were derived as 1.25 μF/cm2 and 32.3, both of which matched well with the value obtained by CV (Fig. 2a) and gated Hall measurements (Fig. 3c). Next, the interface trap density Dit was extracted based on the following equation:15

$${SS}={{{{{\rm{ln}}}}}}\left(10\right)\frac{{k}_{B}T}{q}\left(1+\frac{q{D}_{{it}}}{{C}_{{ox}}}\right)$$
(1)

where SS is the subthreshold swing, kB is Boltzmann constant, T is absolute temperature, q is the elementary charge, Cox is the gate capacitance obtained from MOS capacitance measurements. As a result, a low Dit value of 2.88 × 1011 cm−2/eV was extracted, verifying the high quality of the vdWs interface.

Moreover, the dielectric constant of Bi2SiO5, extracted by dual-gate transfer measurements, also showed similar thickness dependence with the CV results (Fig. 5g). Figure 5h plotted the dual-sweep transfer curves of MoS2 FETs with different Bi2SiO5 thicknesses (10.1~67.5 nm). Apparently, smaller Vg is needed to switch the transistor on and off when a thinner Bi2SiO5 dielectric is used. We should emphasize that the EOT value for a 10.1 nm Bi2SiO5 is as small as 1.3 nm, but its gate leakage current is still on the order of 10−13 A, signifying substantial room space for further scaling of EOT and great potential applications in low-power devices. The Ion of MoS2 FETs in Fig. 5h seems to decrease with decreasing the thickness of Bi2SiO5, which may originate from the contact issues existing in the MoS2 FETs with the thin Bi2SiO5 as gate insulators. However, we should emphasize that, the Ion of ultrathin Bi2SiO5-gated MoS2 FETs can be greatly improved by optimizing the device fabrication process. As shown in Supplementary Figs. S2325, the Ion of another 10-nm-thick Bi2SiO5-gated MoS2 can be similar to the value of the thick Bi2SiO5-gated FETs (0.11 μA/μm, Fig. 5b). It is worth noting that the Ion is greatly limited by the remaining ungated channel in the top-gate device configuration. Additionally, as confirmed by the dual-sweep transfer curves, a nearly ideal SS value and a low normalized gate hysteresis can indeed be obtained in the Bi2SiO5 gated MoS2 FET.

The low operating voltage is essential to fabricate low-power logic circuits. As demonstrated in Fig. 5i, we used two n-type transistors as the load and driver terminal to construct 2D inverter with a high voltage gain of 22.0. The inverter can demonstrate the logic state 0 and 1 within ±0.5 V, showing a low dynamic power consumption of <0.7 nW (Supplementary Fig. 26).

In summary, our work achieved the direct CVD growth of ultrathin free-standing high-k single-crystalline dielectrics, which is much more efficient than traditional mechanical exfoliation. The vertically grown Bi2SiO5 2D crystals present the metrics for a gate dielectric, as evidenced by the coexistence of high dielectric constant, large band gap, high breakdown field strength, as well as the characteristic of easy transfer by facile polymer-free mechanical pressing. These features make Bi2SiO5 attractive as an inert vdWs substrate superior to h-BN for exploring exotic transport properties under reduced interfacial scatterings and stronger gate control, as well as for fabricating hysteresis-free 2D transistors with scaled supply voltage.

Methods

CVD growth of ultrathin Bi2SiO5 nanoplates on mica substrate

2D Bi2SiO5 crystals were synthesized inside a homemade CVD system equipped with a single heating zone tube furnace and 30 mm diameter quartz tube. Typically, the BiF3 powders (purity 99.999%, Macklin) were placed in an empty quartz boat or on top of the SiO2 powders located in the heating center, and the freshly cleaved fluorophlogopite mica substrates were placed above the quartz boat. The heating temperature of the source was 600–750 °C and the growth time was 20 min. The 50 s.c.c.m Ar and 5 s.c.c.m mixed Ar/O2 (1‰) gas were introduced into the CVD system as carrier gas. The system pressure was kept constant as 760 Torr during the whole growth process.

Characterization of CVD-grown ultrathin Bi2SiO5 single-crystalline dielectric

The morphologies of as-synthesized in-plane and free-standing 2D Bi2SiO5 nanoplates were characterized by optical microscopy (Olympus BX53), scanning electron microscopy (SEM, JSM-7800F) and atomic force microscope (AFM, Bruker dimension icon). With a polymer-free method of mechanical pressing, vertically grown Bi2SiO5 nanoflakes were transferred onto Cu grid, glass, and optical quartz substrates to perform the characterizations of transmission electron microscopy (TEM, JEM 2800), X-ray diffraction (XRD, Rigaku Smart Lab 30 KW) and absorption spectrum (SHIMADZU, UV-2600), respectively. The Raman spectroscopy was measured on WITec alpha300R with a laser of 532 nm. The cross-sectional TEM samples of in-plane and vertically grown Bi2SiO5 nanoflakes were both prepared by using a focused ion/electron dual beam system (FEI, Helios 5 CX). All cross-sectional scanning transmission electron microscopy (STEM) imaging was obtained on an aberration-corrected TEM operating at 300 kV (FEI Titan cubed Themis G2 300). The breakdown field strength of Bi2SiO5 was measured with the C-AFM module of Bruker Dimension Icon.

First-principles calculations

The structural relaxation of Bi2SiO5 is performed within the framework of density function theory (DFT) using the projector augmented wave pseudopotential and the Perdew–Burke–Ernzerhof exchange-correlation functional as implemented in the VASP. The energy cutoff for the plane-wave expansion is set to 500 eV, and a Monkhorst–Pack k-mesh of 3 × 9 × 9 is used in the Brillouin zone. The energy convergence threshold is 10−6 eV and the force 10−3 eV Å−1 in the structural optimization. In order to overcome the underestimation of energy gap from the generalized gradient approximation (GGA), we use the method of modified Becke–Johnson potential (mBJ) to calculate the electronic structure.

Device fabrication

To illuminate the possible capacitance coupling, the Bi2SiO5-based metal-insulator-metal (MIM) capacitors were fabricated on quartz substrates rather than SiO2/Si substrates. First, the thick graphite was exfoliated onto the quartz substrate as bottom electrodes for its ultrasmooth surface. Next, the vertically grown Bi2SiO5 nanoflakes were directly picked up from the mica substrate by a polypropylene carbonate/polydimethylsiloxane (PPC/PDMS) stamp, followed by aligned transfer onto the specific graphite bottom electrode by a high-precision transfer platform. Subsequently, the standard electron-beam lithography (EBL) process and thermal evaporation were used to pattern the top electrodes ((In/Au, 5/40 nm).

For fabricating the back-gated MoS2 Hall-bar device using graphite as bottom electrodes, the detailed process was listed as follows. First, SiO2/Si substrates (285 nm SiO2) were pretreated with O2 plasma (Diener Pico plasma cleaner) for 5 min at a power of 50 W. Next, few-layer MoS2 and graphite were exfoliated onto different SiO2/Si substrates. With the help of the PPC/PDMS stamp and high-precision transfer platform, the free-standing Bi2SiO5 and MoS2 nanoflakes were sequentially stacked on top of the graphite. The six-terminal electrode legs for the Hall-bar and the bottom metal electrodes were simultaneously written by one-step EBL and following thermal metal deposition (In/Au, 5/40 nm). The device fabrication process of 4-terminal MoS2 FET includes the following parts: (1) transfer the Bi2SiO5 onto SiO2/Si substrate by mechanical pressing; (2) place part of the few-layer MoS2 on top of the Bi2SiO5 nanosheet; (3) EBL and metal deposition (In/Au, 5/40 nm).

For the top-gated MoS2 device, few-layer MoS2 nanosheets were exfoliated onto SiO2/Si substrate, followed by stacking Bi2SiO5 nanoflakes in the middle of the MoS2 nanosheets as the gate dielectrics. The source, drain, and top-gate electrodes of MoS2 FETs were patterned together with one-step EBL process and thermal metal evaporation (In/Au, 5/40 nm).

Electrical transport measurements

2-probe electrical properties of the back-gate and top-gate MoS2 FETs, including the Figs. 3e, f, 4b, 5b, c, e, h, were carried out by a semiconductor analyzer (FS-Pro) in a shielded vacuum chamber (<0.1 Torr) at room temperature, whose noise level is ~1 × 10−13 A within the voltage range of ±2 V. The 4-probe transfer curves and gated 4-probe measurements (such as gated Rxx-T and Hall data), including Figs. 3b, d, 4c, were carried out in a Physical Properties Measurement Systems (PPMS-9T, Quantum Design) equipped with a homemade electrical measurement system, which is composed of 2 Keithley 2400, and 2 Keithley 2182 A nanovoltmeter and has a noise level of ~10−10 A within the voltage range of ± 2 V. The CV and Cf measurements were carried out on a FS336 LCR Meter.