Introduction

Symmetry breaking in low dimensional heterostructures can provide unprecedented possibilities to generate emergent quantum phenomena in condensed matter physics1,2,3,4,5,6,7,8. In general, van der Waals (vdW) dielectric at atomically-sharp semiconductor/dielectric interfaces can break the symmetry of the target materials and form moiré patterns with specific lattice mismatch2,9,10, exhibiting remarkable capabilities to control electronic states and further realize exotic quantum phenomena therein. Examples of these interfacial phenomena such as Chern insulating states11,12, charge density wave states13,14 and topological valley currents15 have been demonstrated in the heterointerfaces based on h-BN dielectric2,9,11,12,13,14,15,16,17, in which both h-BN dielectric and channel materials (graphene or transition metal dichalcogenides, TMDCs) show threefold-rotational symmetry (C3) along the out-of-plane axis at their interface. In contrast, a low-symmetric dielectric material without C3 symmetry (for example, with C2 symmetry) can in principle break the C3 symmetry in monolayer semiconductors by forming anisotropic moiré potentials at the interface5 and result in exotic optical response and anisotropic electronic transport, while retaining the gating capability as a dielectric medium. Therefore, the vdW dielectrics with lower lattice symmetry can generate unique moiré physics and additional device functionalities at the symmetry-mismatched interfaces. However, an experimental confirmation of such a strategy remains elusive.

Herein, we demonstrate a unique anisotropic layered dielectric material SiP2 and reveal its capability to generate giant anisotropy in optical response and electronic transport in isotropic TMDC semiconductors via symmetry engineering. We realize a high-performance SiP2-gating MoS2 transistor with large on/off ratios >105 and low leakage currents (far below the low power limit) and further observe an insulator-to-metal transition in SiP2-gated 1L-MoS2, indicating a great dielectric capability of SiP2 material. Surprisingly, a linearly-polarized photoluminescence and an anisotropic second harmonic generation signals are observed in 1L-MoS2/SiP2 heterostructure, which are in sharp contrast to the isotropic features of pristine 1L-MoS2. Remarkably, we find a large anisotropic conductance in the 1L-MoS2/SiP2 heterostructure and the tunable anisotropy index reaches a considerable value of 1000 with SiP2 gating, which is among the largest values reported so far, including intrinsically anisotropic materials18. Our first-principles calculations reveal that such giant anisotropy in optical response and electronic transport result from the generated anisotropic moiré potential in 1L-MoS2/SiP2 heterostructure that strongly renormalizes the structural and electronic properties of 1L-MoS2 at the heterointerface. Note that the interfacial symmetry engineering by breaking the C3 symmetry of 1L-MoS2 using SiP2 dielectric with C2 symmetry can be regarded as a strategy to tune electronic properties of channel semiconductors and realize moiré physics at the heterointerfaces.

Results

MoS2 transistors gated with SiP2 dielectric

To evaluate the performance of the SiP2 dielectric, we measured the transfer characteristics of MoS2 transistors with dual-gate geometry in which 20-nm-thick SiP2 and 300-nm-thick SiO2 are used as top and bottom gate media (Fig. 1a, b and Supplementary Fig. 1). As shown in Fig. 1c, when sweeping the top gate voltage \({V}_{{{{{{\rm{t}}}}}}{{{{{\rm{g}}}}}}{-{{{{{\rm{SiP}}}}}}}_{2}}\) up to 5 V, the 5-nm-thick MoS2 transistor shows an on/off ratio as high as 105, which is comparable to those values in h-BN-gated MoS2 transistors (Supplementary Table 2) and meets the well-known criterion for practical logic circuit applications19. In contrast, when sweeping the bottom gate voltage \({V}_{{{{{{\rm{bg}}}}}}{-{{{{{\rm{SiO}}}}}}}_{2}}\) to ~5 V, the transistor generates an on/off ratio as small as 10, and requires a large \({V}_{{{{{{\rm{bg}}}}}}{-{{{{{\rm{SiO}}}}}}}_{2}}\) over 80 V to achieve an on/off ratio of 105 (inset of Fig. 1c). This comparison directly demonstrates that, compared to SiO2, the SiP2 gate dielectric with larger dielectric constant and smaller thickness can achieve great capacitive capability. The measured leakage current of SiP2-gated MoS2 transistor is as small as approximately 10–5 A cm–2 at an external electric field strength of 1.5 MV cm–1 (Fig. 1d). Such a low leakage current is comparable to those of transistors gated by high-κ dielectrics19,20,21 such as Al2O3, HfO2 or Bi2SeO5, and better than the criteria of the low-power limit and the standard complementary metal–oxide–semiconductor gate limit19. With increasing \({V}_{{{{{{\rm{t}}}}}}{{{{{\rm{g}}}}}}{-{{{{{\rm{SiP}}}}}}}_{2}}\), the field-effect mobility \(\mu\) at 2 K can reach ~600 cm2 V–1 s–1 when \({V}_{{{{{{\rm{bg}}}}}}{-{{{{{\rm{SiO}}}}}}}_{2}}\) is fixed at 35 V (Fig. 1e). Even for SiP2-gated 1L-MoS2 transistors with the same device geometry (Supplementary Fig. 2), the mobility of 330 cm2 V–1 s–1 at 2 K is better than those reported values in HfO2-gated 1L-MoS2 devices22 (174 cm2 V–1 s–1 at 4 K), indicating that the vdW SiP2 material with a large dielectric constant can effectively reduce the charge scattering and increase the mobility of MoS2 transistors. Such excellent performance with high on/off ratio, low leakage current, and high mobility in MoS2/SiP2 devices suggests that layered SiP2 can be a high-performance dielectric in switching devices.

Fig. 1: High-performance MoS2 transistors gated with a SiP2 dielectric with non-symmorphic C2 rotational symmetry.
figure 1

a Schematic structure of the cross-section of a MoS2/SiP2 heterostructure. Orange, gray, light blue, red, and dark blue spheres represent Si, PA, PB, S, and Mo atoms, respectively. The PB atoms form the unique quasi-1D PB–PB chains along the y direction in SiP2 crystal lattice. b Schematic illustration of MoS2/SiP2 dual-gated device. The top gate medium is 20-nm-thick SiP2, and the bottom gate medium is 300-nm-thick SiO2. D and S represent the drain and source electrodes. \({V}_{{{{{{\rm{tg}}}}}}{-{{{{{\rm{SiP}}}}}}}_{2}}\) and \({V}_{{{{{{\rm{bg}}}}}}{-{{{{{\rm{SiO}}}}}}}_{2}}\) are the top gate voltage and back gate voltage, respectively. And \({V}_{{{{{{\rm{ds}}}}}}}\) is the source-drain voltage applied on the MoS2 channel material. c Transfer curves of the 5-nm-MoS2-based transistor at 2 K when sweeping top (red curve) and bottom (blue curve) gate voltages via the SiP2 and SiO2 dielectric media. The inset is a transfer curve with scanning \({V}_{{{{{{\rm{bg}}}}}}{-{{{{{\rm{SiO}}}}}}}_{2}}\). d Leakage current \({I}_{{{{{{\rm{gs}}}}}}}\) as a function of \({V}_{{{{{{\rm{t}}}}}}{{{{{\rm{g}}}}}}{-{{{{{\rm{SiP}}}}}}}_{2}}\) on a 5-nm MoS2 device at 2 K. \({I}_{{{{{{\rm{gs}}}}}}}\) and \({V}_{{{{{{\rm{t}}}}}}{{{{{\rm{g}}}}}}{-{{{{{\rm{SiP}}}}}}}_{2}}\) are rescaled to the leakage current density (\({J}_{{{{{{\rm{Leak}}}}}}}\)) and the electric field strength for a better comparison (right and top axes). Horizontal red lines mark the limits of leakage current density for various types of integrated circuits. e The field-effect mobility \(\mu\) as a function of \({V}_{{{{{{\rm{t}}}}}}{{{{{\rm{g}}}}}}{-{{{{{\rm{SiP}}}}}}}_{2}}\) and \({V}_{{{{{{\rm{bg}}}}}}{-{{{{{\rm{SiO}}}}}}}_{2}}\) for the 5-nm-MoS2/SiP2 device. The white dashed lines highlight that \(\mu\) equals 5, 100, and 500 cm2 V–1 s–1. f Sheet resistances (\({R}_{{{{{{\rm{s}}}}}}}\)) versus temperature of a 1L-MoS2 device under different \({V}_{{{{{{\rm{tg}}}}}}{-{{{{{\rm{SiP}}}}}}}_{2}}\) values. g Comparison of the anisotropy ratio of the effective mass for dielectrics with different relative dielectric constants (\({\varepsilon }_{{{{{{\rm{r}}}}}}}\)). The anisotropy ratio is defined as the ratio of the electron effective mass (me) along the y and x directions of the corresponding lattice.

To demonstrate the great gate tunability of SiP2 dielectric, we measured the temperature-dependent sheet resistance (\({R}_{{{{{{\rm{s}}}}}}}\)\(T\)) and observed a gating-induced insulator–metal transition in the 1L-MoS2/SiP2 transistor. As shown in Fig. 1f, the \({R}_{{{{{{\rm{s}}}}}}}\)\(T\) curves show typical insulating behavior with negative temperature coefficients \({{{{{\rm{d}}}}}}{R}_{{{{{{\rm{s}}}}}}}/{{{{{\rm{d}}}}}}T\) and follow a thermal activation dependence23 when \({V}_{{{{{{\rm{tg}}}}}}{-{{{{{\rm{SiP}}}}}}}_{2}}\) < 3.8 V and \({V}_{{{{{{\rm{bg}}}}}}{-{{{{{\rm{SiO}}}}}}}_{2}}\) = 0 V (Supplementary Fig. 3a). The extracted activation energy decreases monotonically from ~7 meV to near zero as \({V}_{{{{{{\rm{tg}}}}}}{-{{{{{\rm{SiP}}}}}}}_{2}}\) increases from 2.5 V to 3.8 V (Supplementary Fig. 3b). As a result, \({R}_{{{{{{\rm{s}}}}}}}\) starts to decrease with cooling temperature and the positive \({{{{{\rm{d}}}}}}{R}_{{{{{{\rm{s}}}}}}}/{{{{{\rm{d}}}}}}T\) shows the typical metallic behavior when \({V}_{{{{{{\rm{tg}}}}}}{-{{{{{\rm{SiP}}}}}}}_{2}}\) \( > \) 3.8 V, directly indicating an insulator–metal transition23 in SiP2-gated 1L-MoS2. Such a transition in SiP2-gated 1L-MoS2 transistor directly manifests the excellent dielectric property of layered SiP2 as a gate medium to modulate the electronic states of ultrathin semiconductors.

To experimentally evaluate the dielectric constant of SiP2, we measured the sheet carrier density (\({n}_{2{{{{{\rm{D}}}}}}}\)) of MoS2 (5 nm) as a function of \({V}_{{{{{{\rm{tg}}}}}}{-{{{{{\rm{SiP}}}}}}}_{2}}\) based on Hall effect measurements. The \({n}_{2{{{{{\rm{D}}}}}}}\) values of top-gated MoS2 remain nearly unchanged below the threshold voltage of 1.7 V and can be continually modulated to 8 × 1012 cm–2 by increasing \({V}_{{{{{{\rm{tg}}}}}}{-{{{{{\rm{SiP}}}}}}}_{2}}\) to 5 V (Supplementary Fig. 4a). Note that the dual-gate-modulated \({n}_{2{{{{{\rm{D}}}}}}}\) can reach a maximum value close to 1013 cm–2 (Supplementary Fig. 4b). The relative dielectric constant \({\varepsilon }_{{{{{{\rm{r}}}}}}}\) is evaluated to be 8.1 for SiP2 by fitting the linear part of the \({n}_{2{{{{{\rm{D}}}}}}}\)\({V}_{{{{{{\rm{tg}}}}}}{-{{{{{\rm{SiP}}}}}}}_{2}}\) data using \({n}_{2{{{{{\rm{D}}}}}}}={\varepsilon }_{0}{\varepsilon }_{{{{{{\rm{r}}}}}}}{V}_{{{{{{\rm{tg}}}}}}{-{{{{{\rm{SiP}}}}}}}_{2}}/(e{t}_{{{{{{{\rm{SiP}}}}}}}_{2}})\), where \(e\) is the electron charge, \({\varepsilon }_{0}\) is the vacuum permittivity, and \({t}_{{{{{{{\rm{SiP}}}}}}}_{2}}\) = 20 nm is the thickness of SiP2 (more details in “Methods”). Such a dielectric constant of 8.1 in layered SiP2 is larger than those of SiO2 and h-BN dielectrics24,25 and comparable to that of Al2O3 dielectric24 (Fig. 1g, a detailed comparison is given in Supplementary Tables 3 and 4), well consistent with the theoretical estimation from first-principles calculations26.

As a typical vdW dielectric with excellent performance, another distinctive nature of SiP2 is the anisotropic lattice structure with non-symmorphic C2 symmetry. In sharp contrast to the highly symmetric crystal structure of those widely used dielectrics (oxides and h-BN), such an anisotropic in-plane lattice structure of SiP2 leads to a highly anisotropic ratio of electron effective masses (~16, Fig. 1g) for the band edge states in its electronic band structure26, and provides an opportunity to engineer the interfacial symmetry of vdW heterostructure combined monolayer TMDCs with SiP2. For example, the 1L-MoS2/SiP2 heterostructure shows no rotational symmetry and can exhibit in-plane anisotropic optical and electronic properties (details discussed below). In particular, if the zigzag direction of MoS2 and the PB–PB chain of SiP2 (the direction parallel to the PB–PB chain of SiP2 is defined as the y direction, while the perpendicular direction is defined as the x direction) are aligned in parallel (Fig. 2a, b), the mirror symmetry along the x direction can remain in the 1L-MoS2/SiP2 heterostructure; otherwise, all crystal symmetries in MoS2 are broken. The perturbation for the electronic structures of stacked heterostructures can be used to generate in-plane polarization and Berry curvature dipole at the interface27, realizing emergent interfacial phenomena such as directional quantum shift current5, nonlinear Hall effect1 and circular photo-galvanic effect (Supplementary Fig. 5).

Fig. 2: Anisotropic optical response at the TMDC/SiP2 interface.
figure 2

a, b Optical image (a) and schematic illustration of the top view (b) of a 1L-MoS2/SiP2 heterostructure, in which the armchair direction of MoS2 and the x direction of SiP2 are parallel. The red and white dashed lines in (a) highlight the 1L-MoS2 and SiP2 sample areas. Scale bar is 20 μm. The red rectangle and diamond in (b) represent the unit cells of SiP2 and MoS2, respectively. The black dashed line represents the mirror plane m. c Polar plot of polarization-resolved second harmonic generation (SHG) intensities of bare 1L-MoS2 (blue shadow) and 1L-MoS2/SiP2 (red) under the parallel configuration (the detection polarization is parallel to the excitation polarization). The red solid line represents the fitting with Eq. (2) in the “Methods”. d Photoluminescence (PL) spectra of SiP2 (green curve), 1L-MoS2 (blue curve), and 1L-MoS2/SiP2 heterostructure (red curve) at 77 K. Insets are the corresponding schematics of the measured sample geometry. The exciton emission of 1L-MoS2 is highlighted with shadows. e Corresponding color plot of the PL intensity in (d) as a function of emission photon energy at different detection polarization angles θ. Here θ denotes the angle between the analyzer polarization direction and the x direction, as defined in (b). f Polar plots of polarization-resolved PL integrated intensities of 1L-MoS2 (blue curve) and 1L-MoS2/SiP2 (red curve and red shaded area). The circles represent the experimental data and the solid curves represent the sine fitting results.

SiP2-induced anisotropic optical response in 1L-MoS2

To understand the engineered symmetry of heterointerfaces, we performed second harmonic generation (SHG) measurements on 1L-MoS2/SiP2 heterostructures under a parallel measurement geometry (Fig. 2c). The SHG signal in pristine 1L-MoS2 shows a sixfold-rotational symmetric pattern with maxima of SHG intensity along its armchair direction and can be well fitted with Eq. (1), implying the C3-rotational symmetry of 1L-MoS2 samples28. While in 1L-MoS2/SiP2 heterostructure, the SHG signal presents an additional twofold component imposed to the six symmetric petals (detailed analysis in “Methods”). Such a twofold component does not originate from the C3-symmetric lattice of 1L-MoS2 itself, but results from the reduced symmetry at the 1L-MoS2/SiP2 heterointerface. The SHG intensities can be well fitted with Eq. (2) in the “Methods”, similar to those distorted SHG scenarios in uniaxially-strained 1L-MoS2 samples29,30. Note that SiP2 itself has no SHG signal due to the existing inversion symmetry in its crystal lattice (Supplementary Fig. 9c), ensuring that the observed SHG signal of 1L-MoS2/SiP2 heterostructure mainly originates from 1L-MoS2 whose band structure is effectively modified by the potential on the heterointerface created by the bottom C2 symmetric SiP2. More importantly, the anisotropic SHG response at 1L-MoS2/SiP2 changes little with increasing SiP2 thickness (Supplementary Fig. 11), further confirming that this is an interfacial phenomenon induced by symmetry breaking at the 1L-MoS2/SiP2 heterointerface between the 1L-MoS2 and the topmost SiP2 layer.

To investigate the effect of symmetry engineering on the optical properties of such an interface, we performed polarization-dependent photoluminescence (PL) measurements in the 1L-MoS2/SiP2 heterostructure at 77 K (Fig. 2d, e and Supplementary Fig. 6). For SiP2, the PL signal with an excitonic emission energy of 2.06 eV at 77 K shows a linear polarization along the x direction of the lattice26. In contrast, for 1L-MoS2 without SiP2 stacking layer, the excitonic emission of 1L-MoS2 at 1.91 eV at 77 K remains unchanged with the detection polarization angle and shows no linear polarization. While, for the 1L-MoS2/SiP2 heterostructure, the excitonic state of 1L-MoS2 with an emission energy of 1.91 eV becomes linearly polarized along the x direction of SiP2 (Fig. 2f and Supplementary Fig. 7). The similar results can be observed at 77 K and 300 K in 1L-WS2/SiP2 heterostructure (Supplementary Fig. 8), indicating the anisotropic PL response in TMDC/SiP2 is robust with temperatures. Similar anisotropic SHG responses are also observed in 1L-WS2/SiP2 heterostructure (details in Supplementary Fig. 9), indicating that SiP2 dielectric can effectively engineer the symmetry of its neighboring monolayer TMDC through tunable interlayer interactions. More interestingly, the anisotropic PL and SHG responses strongly depended on the twist angle between MoS2 and SiP2, and the corresponding anisotropy dramatically decreased when the mirror symmetry of moiré superlattice varnishes with changing the twist angle (Supplementary Figs. 12 and 13). This result indicates that the mirror symmetry of MoS2/SiP2 moiré superlattice plays an important role in controlling the magnitude of anisotropic optical responses at the heterointerface. The symmetry breaking induced anisotropic behavior can exist at those interfaces stacked with the C3-symmetric monolayer TMDCs and C2-symmetric dielectrics, enabling a strategy to explore applications such as polarization-sensitive photodetectors31.

Giant anisotropic conductance in SiP2-gated MoS2 transistors

To investigate the interfacial symmetry modulation on the electronic transport properties of MoS2, we measured the conductance \({G}_{x}\) (\({G}_{y}\)) along the x (y) directions of SiP2-gated 1L-MoS2 transistors (Fig. 3a, Supplementary Figs. 14 and 15). Figure 3b shows a comparison between \({G}_{x}\) and \({G}_{y}\) under different \({V}_{{{{{{\rm{tg}}}}}}{-{{{{{\rm{SiP}}}}}}}_{2}}\). One can see that the anisotropy index \({G}_{y}\)/\({G}_{x}\) can be as high as 103 at the off-state with \({V}_{{{{{{\rm{tg}}}}}}{-{{{{{\rm{SiP}}}}}}}_{2}}\) < 1 V (Fig. 3c), implying that the symmetry engineering using SiP2 dielectric can drive the isotropic conductivity of C3-symmetric 1L-MoS2 into highly-anisotropic electronic states. With further increasing \({V}_{{{{{{\rm{tg}}}}}}{-{{{{{\rm{SiP}}}}}}}_{2}}\), the anisotropy index gradually approaches the value of 1, suggesting that 1L-MoS2 recovers back to isotopically conducting states at the on-state. Such continuous modulation of \({G}_{x}\), \({G}_{y}\), and \({G}_{y}\)/\({G}_{x}\) index can also be achieved at a wide range of \({V}_{{{{{{\rm{bg}}}}}}{-{{{{{\rm{SiO}}}}}}}_{2}}\) (Fig. 3d–f). The tunable conductance from anisotropic to isotropic characteristics suggests that SiP2 with in-plane anisotropy is anticipated to stimulate device functionality exploration for anisotropic digital inverters32, anisotropic memorizers33, or artificial synaptic devices34.

Fig. 3: Gate tunable anisotropic transfer characteristics of MoS2/SiP2 heterointerfaces.
figure 3

a Schematic diagram of a SiP2-gated MoS2 transistor. The top gate medium is 20-nm-thick SiP2, and the bottom gate medium is 300-nm-thick SiO2. \({G}_{x}\) and \({G}_{y}\) are the sheet conductance along the x and y directions of the heterostructure, corresponding to the measurement geometries described in Supplementary Fig. 10. b Transfer characteristics of \({G}_{x}\) and \({G}_{y}\) for 1L-MoS2 at \({V}_{{{{{{\rm{bg}}}}}}{-{{{{{\rm{SiO}}}}}}}_{2}}\) = 20 V and 2 K. Note that \({V}_{{{{{{\rm{bg}}}}}}{-{{{{{\rm{SiO}}}}}}}_{2}}\) is necessarily applied to reduce contact resistance and activate top-gated channel for achieving measurable four-terminal conductance and a good signal-to-noise ratio. c Anisotropy index \({G}_{y}\)/\({G}_{x}\) as a function of \({V}_{{{{{{\rm{tg}}}}}}{-{{{{{\rm{SiP}}}}}}}_{2}}\) at \({V}_{{{{{{\rm{bg}}}}}}{-{{{{{\rm{SiO}}}}}}}_{2}}\) = 20 V. d, e \({G}_{y}\) (d) and \({G}_{x}\) (e) as a function of \({V}_{{{{{{\rm{tg}}}}}}{-{{{{{\rm{SiP}}}}}}}_{2}}\) and \({V}_{{{{{{\rm{bg}}}}}}{-{{{{{\rm{SiO}}}}}}}_{2}}\). f Color plot of anisotropy index \({G}_{y}\)/\({G}_{x}\) as a function of \({V}_{{{{{{\rm{tg}}}}}}{-{{{{{\rm{SiP}}}}}}}_{2}}\) and \({V}_{{{{{{\rm{bg}}}}}}{-{{{{{\rm{SiO}}}}}}}_{2}}\). The dashed lines highlight where \({G}_{y}\)/\({G}_{x}\) equals 1, 10, and 100. \({G}_{y}\)/\({G}_{x}\) values for \({V}_{{{{{{\rm{tg}}}}}}{-{{{{{\rm{SiP}}}}}}}_{2}}\) < 1.4 V and \({V}_{{{{{{\rm{bg}}}}}}{-{{{{{\rm{SiO}}}}}}}_{2}}\) < 20 V are not given since \({G}_{x}\) is too small to achieve measurable four-terminal conductance. The inset is a schematic illustration of the top view of a 1L-MoS2/SiP2 heterostructure. g The maximum \({G}_{y}\)/\({G}_{x}\) value of the MoS2/SiP2 device as a function of MoS2 thickness. h The anisotropic conductance in layered materials as a function of their bandgap values. Note that all other anisotropic materials host intrinsically-anisotropic conductance in nature while in our case we can drive the intrinsically-isotropic conductance in MoS2 into the anisotropic state. The values of anisotropic conductance and bandgaps for other anisotropic materials are generated from previous reports18,35.

To confirm that such anisotropic conductance originates from the MoS2/SiP2 heterointerface, we compared the \({G}_{x}\) and \({G}_{y}\) values of SiP2-gated MoS2 transistors by increasing the thickness of MoS2 from monolayer to 20 nm. As a result, the observed anisotropy index \({G}_{y}\)/\({G}_{x}\) at the off-state decreases rapidly to ~1 when the thickness of MoS2 is increased to 20 nm (Fig. 3g, Supplementary Figs. 16 and 17), suggesting a nearly isotropic conductance in thicker samples. Such a thickness-dependent behavior is proposed to be attributed to the competition between the surface and bulk conductance, as qualitatively described in Supplementary Fig. 18. Specifically, the anisotropy index of MoS2 can be written as \(\frac{{G}_{y}}{{G}_{x}}=\frac{{G}_{y}^{{{{{{\rm{surface}}}}}}}+{G}_{y}^{{{{{{\rm{bulk}}}}}}}}{{G}_{x}^{{{{{{\rm{surface}}}}}}}+{G}_{x}^{{{{{{\rm{bulk}}}}}}}}\), where bulk conductance (proportional to the sample thickness) is isotropic \({G}_{y}^{{{{{{\rm{bulk}}}}}}}\) ≈ \({G}_{x}^{{{{{{\rm{bulk}}}}}}}\) while surface conductance is anisotropic since only the surface layer of MoS2 with certain thickness (within the Thomas-Fermi screening length) can be tuned for carrier accumulation with SiP2 dielectric based on our numerical Poisson-Schrödinger calculations (Supplementary Figs. 19 and 20). For SiP2-gated 1L-MoS2 case, \({G}_{y}^{{{{{{\rm{bulk}}}}}}}\) = \({G}_{x}^{{{{{{\rm{bulk}}}}}}}\) = 0 and only the MoS2 layer on the interface (namely the whole monolayer) contributes to the conductance, so the anisotropy index can be written as \(\frac{{G}_{y}}{{G}_{x}}=\frac{{G}_{y}^{{{{{{\rm{surface}}}}}}}}{{G}_{x}^{{{{{{\rm{surface}}}}}}}}\), whose value can be as high as 1000. When increasing the thickness of MoS2, \({G}_{y}^{{{{{{\rm{bulk}}}}}}}\) and \({G}_{x}^{{{{{{\rm{bulk}}}}}}}\) begin to increase and gradually dominate the total conductance with \({G}^{{{{{{\rm{surface}}}}}}}\)\({G}^{{{{{{\rm{bulk}}}}}}}\) at the thick limit. As a result, the anisotropy index is reduced to \(\frac{{G}_{y}}{{G}_{x}}=\frac{{G}_{y}^{{{{{{\rm{bulk}}}}}}}}{{G}_{x}^{{{{{{\rm{bulk}}}}}}}}=1\), which is consistent with our experimental observation of less anisotropy in SiP2-gated MoS2 with a thickness of 20 nm. This result indicates that the anisotropic conductance behavior is contributed exactly from the interface of the MoS2/SiP2 heterostructures (details in Supplementary Figs. 1820). Compared to those vdW materials with in-plane anisotropic crystal lattices and electronic structures18,35, our SiP2-gated 1L-MoS2 not only has the largest anisotropy index but also hosts the greatest capability to tune such an anisotropy index (Fig. 3h).

Anisotropic moiré potential at 1L-MoS2/SiP2 interface

To further understand the influence of interface structures on the anisotropic optical and transport behavior, we explore the structural and electronic properties of the 1L-MoS2/SiP2 heterostructure by using density functional theory (DFT) calculations. Note that the mirror symmetry of the constructed MoS2/SiP2 heterointerface originates from the parallel or antiparallel alignment of the zigzag chain in MoS2 along the y direction of SiP2 (see details in Supplementary Fig. 21). Thus, two kinds of moiré patterns can be obtained (labeled as case-I and case-II, details in Section 13 in Supplementary Information). Taking the moiré pattern of case-I as an example (Fig. 4a, b), due to the lattice mismatch between MoS2 and SiP2, there are three typical stacking structures, labeled as I-AA, I-AB, and I-BA, as indicated by the colored rectangular areas in Fig. 4a. The details about atomic stacked configurations and stacked structures with II-AA, II-AB, and II-BA in the moiré pattern of case-II are shown in Supplementary Fig. 21d.

Fig. 4: Structural and electronic properties of the MoS2/SiP2 heterointerface.
figure 4

a Moiré pattern of case-I formed by stacking MoS2 monolayer on SiP2 monolayer. The colored rectangles mark different stacked configurations: I-AA (red area), I-AB (green area), and I-BA (blue area). The vertical dashed line represents the mirror plane m. b The side view of the case-I heterostructure model. The interlayer distance is marked by the black lines, which corresponds to the difference between the z coordinates of adjacent S atoms and P atoms in the vdW gap. c The real space distribution of the interlayer distance in the moiré pattern of case-I. The dashed rectangular area corresponds to the moiré superlattice shown in (a). The effective hopping paths between the trapped states along the x and y directions are shown by the black curve arrows. d The band alignment between unstrained monolayer MoS2 and unstrained monolayer SiP2. CBM and VBM represent the conduction band minimum and valence band maximum, respectively. The work function and bandgap for each layer are calculated with the HSE-06 functional. The conduction band edge contributed by monolayer MoS2 is marked by “‡”. e, f The illustrations of the plane-averaged charge density on the lowest conduction band edge contributed by 1L-MoS2 along the z direction (e) and y direction (f) of the case-I-ABBA heterostructure. g, h Top view (g) and side view (h) for the calculated charge density for the lowest conduction band edge contributed by monolayer MoS2 in case-I-ABBA heterostructure. The isosurface is set as 0.0002 e Å–3.

The structural corrugations in 1L-MoS2/SiP2 heterostructure with two kinds of moiré patterns are simulated via DFT calculations (Fig. 4c and Supplementary Fig. 27b, e). After being placed on the SiP2 lattice, the atomic flat structure of 1L-MoS2 will be deformed due to the interface coupling, resulting in the formation of moiré potential on 1L-MoS2 that effectively breaks C3-rotational symmetry of the pristine MoS2. In contrast, the structural corrugations in 1L-SiP2 are much smaller compared with that in 1L-MoS2 (Supplementary Fig. 25). Furthermore, we plot the distribution of the interlayer distance (marked in Fig. 4b) for the moiré pattern of case-I in Fig. 4c to demonstrate the moiré potential in this heterostructure. In the moiré pattern of case-I, the stacked region of I-BA hosts the smallest interlayer distance between MoS2 and SiP2, indicating the largest structural deformation on 1L-MoS2 and interlayer coupling. However, in the moiré pattern of case-II such region with the largest moiré potential and lattice deformation becomes II-AA (Supplementary Fig. 27e). On the other hand, the mirror symmetry parallel to the armchair direction of 1L-MoS2 (also the x direction of the heterostructure) is always observed in both moiré patterns of 1L-MoS2/SiP2 heterostructures that is determined by the specific stacked regulations in the fabrication of experimental devices. Such characters with reduced symmetry in corrugated 1L-MoS2 are consistent with the symmetry analysis to stacking structures with moiré patterns. Our simulated structural deformation of corrugated 1L-MoS2 with moiré patterns gives a consistent interpretation on the change of the symmetric shape of the experimental SHG spectra from sixfold (1L-MoS2) to twofold (1L-MoS2/SiP2).

The out-of-plane structural corrugations of 1L-MoS2 with moiré patterns can strongly modulate its electronic structures and thus influence the optical properties. Since the direct simulation of the large-size moiré lattice by using DFT are too expensive to afford, to overcome this issue, we use a strained MoS2 and SiP2 to construct a heterostructure guaranteeing the conduction band offset between MoS2 and SiP2 same with that in the moiré heterostructure, then simulate the influence of the moiré potential on the electronic states of 1L-MoS2 (details in Sections 14–17 in Supplementary Information). The exemplified results are presented in the heterostructure model with stacking regions of I-AB and I-BA (named as case-I-ABBA). Figure 4g and h demonstrates the charge density distribution for conduction band edge in case-I-ABBA, and Fig. 4e and f shows the plane-averaged charge density along the z and y directions (Section 17 in Supplementary Information). In the heterostructure model with fully relaxation, out-of-plane corrugation can be clearly observed with the retaining of the mirror symmetry in the moiré pattern. The conduction band edge in 1L-MoS2/SiP2 heterostructure is dominated by the state from the MoS2 layer, while its charge density distribution has been strongly modified by the lattice deformation. The calculated results for other heterostructure models containing different stacked regions are shown in Section 17 in Supplementary Information. The conduction band edge in 1L-MoS2/SiP2 heterostructure is always strongly modified by the moiré potential. For the case-II moiré pattern, we build similar models and obtain the same conclusions (details in Sections 16 and 17 in Supplementary Information). Compared with pristine 1L-MoS2 with C3-rotational symmetry, the symmetry engineering on the conduction band edge can be clearly observed in 1L-MoS2/SiP2 heterostructure. With breaking C3-rotational symmetry by SiP2, the conduction band edge on deformed 1L-MoS2 only keeps the mirror symmetry (Fig. 4f). Therefore, when one electron is excited on conduction band edge and couples with hole states, the optical matrix elements in formed exciton should be strongly modified by the moiré potential with lower symmetry and the lowest bright exciton absorption becomes highly anisotropic, which is consistent with the observation from our PL experiments.

The formation of moiré potential with symmetry engineering can also explain the experimentally observed giant anisotropic conductance in SiP2-gated MoS2 transistors. At the off-state with low carrier density (~5 × 109 cm–2) and low temperature (2 K), the charge carriers in 1L-MoS2/SiP2 heterostructure are mainly trapped by charged impurities22,36,37,38 and the moiré potentials, thus the giant anisotropic conductance of 1L-MoS2 is mainly contributed by the effective hopping between trapped charge states in the moiré potentials22,36,37,38. Since the charged defects in 1L-MoS2 are distributed randomly without anisotropy, the anisotropic moiré potential should be a critical factor for the anisotropic conductance in the SiP2-gated MoS2 transistor. Similar to previous discussions, we also take the moiré pattern of case-I as an example (the discussion of the moiré pattern of case-II draws the same conclusion), the distribution of interlayer distance between SiP2 and MoS2 in real space (Fig. 4c) shows that the smallest interlayer distance is located in the I-BA stacking region, which corresponds to the largest interlayer potential and can effectively trap the charge carriers inside. On the other hand, the anisotropic moiré potential results in highly anisotropic hopping between trapped states. For example, the effective hopping along the parallel direction (||) to the mirror plane is naturally smaller than that along the perpendicular direction (), indicating that, at the off state, the effective mass of these trapped states is highly anisotropic (\({m}_{\parallel }\gg {m}_{\perp }\)). The large ratio of effective mass (\({m}_{\parallel }\)/\({m}_{\perp }\)) leads to the highly-anisotropic conductance in 1L-MoS2/SiP2 heterostructure. With increasing the electron density to reach the on-state, the moiré potential cannot fully trap these charge carriers and its influence on transport becomes unimportant, delocalizing the 2D electron gas formed on the 1L-MoS2 layer. Thus, the conductance turns out to be isotropic at the on-state.

Discussion

In conclusion, we demonstrate an anisotropic van der Waals dielectric SiP2 that can simultaneously tune the electronic states of channel semiconductors and induce symmetry engineering at TMDC/SiP2 interfaces. Our first-principles calculations reveal that these anisotropic characteristics originate from the formation of the anisotropic-symmetric moiré potential in the MoS2/SiP2 heterostructure, which strongly modulates structural and electronic properties of 1L-MoS2 with tunable anisotropic symmetry. The tunable interfacial symmetry in the TMDC/SiP2 heterostructure can provide a unique platform for investigating symmetry-related interfacial physics and corresponding phenomena, including the generation of the in-plane polarization5 (bulk photovoltaic effect and quantum shift current) and the Berry curvature dipole39,40,41 (circular photo-galvanic effect and nonlinear Hall effect). The giant anisotropy generated in the TMDC/SiP2 heterostructure, which is absent in the pristine TMDC material, sheds light on the moiré physics of the engineered interface with reduced symmetry, and provides an effective way to control the degree of freedom of electrons in condensed matter systems.

Methods

Crystal symmetry analyses of TMDC/SiP2 heterostructures

The orthorhombic SiP2 crystal exhibits an anisotropic layered structure (space group \(P{nma}\)) with an embedded quasi-one-dimensional PB–PB chains along the y direction of the crystal lattice. Specifically, three important spatial symmetry operations should be addressed in this atomic structure of the SiP2 crystal when stacking SiP2 with monolayer TMDCs. First, there is a non-symmorphic C2 symmetry about the z direction (the screw symmetry S2 that combines a twofold rotational symmetry with a translation along the z direction in the half-unit-cell, S2 = C2 + z/2) in the SiP2 crystal. This non-symmorphic C2 symmetry of bulk SiP2 is incompatible with C3 symmetry of TMDCs and will result in highly anisotropic nature (C1 symmetry) of the TMDC/SiP2 heterostructures. Second, the atomic structure of SiP2 is inversion symmetric, which means there is no SHG signal of SiP2 flakes, ensuring that the distorted SHG signals at the TMDC/SiP2 heterostructure mainly come from the symmetry breaking at the heterointerface. Third, there is a mirror symmetry perpendicular to the PB–PB chains (y direction) of the SiP2 crystal. This vertical mirror can remain in the TMDC/SiP2 heterostructures when stacking the TMDCs and SiP2 by aligning their mirror planes (the zigzag direction of TMDCs is parallel to the PB–PB chains of SiP2), generating a mirror symmetric anisotropic moiré potential at the heterointerface.

Optical measurements of TMDC/SiP2 heterointerfaces

SiP2 flakes and TMDC flakes were prepared by mechanical exfoliation onto polydimethylsiloxane (PDMS) stamps and SiO2/Si wafers (300-nm-thick SiO2 layer). TMDC/SiP2 heterointerfaces were fabricated using a dry-transfer method and stacked by parallelly aligning the zigzag direction of MoS2 and y direction of SiP2. The crystal axes of the 1L-MoS2 samples are confirmed by their polarized-SHG results. And the crystal axes of SiP2 are first identified by their optical image and then confirmed by their polarized PL results. The whole sample fabrication was processed in a glove box to avoid any degradation. Room temperature PL measurements were performed using a confocal Raman system (WITec Alpha 300) using a 50× objective lens with an incident laser (laser power of 1 mW) focused to a 1 μm spot size. Nitrogen-filled environments were established by protecting samples with continuous nitrogen gas flow. Low-temperature PL measurements were performed under vacuum conditions in cryostats (Cryo Instrument of America RC102–CFM Microscopy Cryostat). For polarized PL measurements, the excitation polarization is fixed along the x direction, and the detection polarization is changed from θ = 0° to 180° (θ is the angle between the detection polarization and the x direction).

The SHG measurements were performed using a Ti:sapphire oscillator with an excitation wavelength of 810 nm, pulse width of 70 fs, and repetition rate of 80 MHz. The laser pulse was focused to an ~1 μm spot size by a 40× objective lens. The SHG signals are obtained under a configuration with the detection polarization parallel to the excitation polarization. For the SHG signal of 1L-TMDCs on the SiO2 substrate, the sixfold symmetric SHG intensities are fitted by Eq. (1):

$${I}_{{{{{{\rm{SHG}}}}}}}^{\parallel }\propto {\cos }^{2}3\theta$$
(1)

Normally, for those uniaxially-strained TMDCs, the SHG intensities \({I}_{{{{{{\rm{SHG}}}}}}}^{\parallel }\) parallel to the incident laser polarization can be written as:

$${I}_{{{{{{\rm{SHG}}}}}}}^{\parallel }\propto {\left[\cos 3\theta+{\varepsilon }_{y}\left({k}_{1}{\cos }^{3}\theta -{k}_{2}{\sin }^{2}\theta \cos \theta \right)\right]}^{2}$$
(2)

where \({\varepsilon }_{y}\) is the strain along the y direction, \({k}_{1}\) and \({k}_{2}\) are parameters related to TMDCs. In our case of the 1L-MoS2/SiP2, the C3 symmetry of 1L-MoS2 is also reduced to low symmetry such as C1. Therefore, we fit our data by Eq. (2) to describe the anisotropic SHG response in our 1L-MoS2/SiP2.

Electrical transport measurements

The 1L-MoS2 (or few-layer MoS2) and SiP2 flakes for electronic transport measurements were exfoliated onto a PDMS stamp and transferred onto a silicon substrate with prepatterned electrodes (Ti/Au with a thickness of 3/9 nm) in sequence. A top gate on the SiP2 flake (Ti/Au with a thickness of 5/45 nm) was then made using electron-beam lithography and electron-beam evaporation. Electrical transport measurements were performed in a cryo-free superconducting magnet system (Oxford Instruments TeslatronPT). Four-terminal resistance \({R}_{{xx}}\) was acquired using a Keithley 2182 voltmeter with a DC current supplied by a Keithley 2400 sourcemeter. The gate voltage is supplied by a Keithley 2400 sourcemeter. The sheet carrier density (\({n}_{2{{{{{\rm{D}}}}}}}\)) is obtained based on Hall effect measurements on Au/SiP2/MoS2 sandwiched devices22. The Au/SiP2/MoS2 device can be considered as a parallel plate capacitor, and the amount of charge per unit area can be written as:

$$e{n}_{2{{{{{\rm{D}}}}}}}=\frac{{\varepsilon }_{0}{\varepsilon }_{{{{{{\rm{r}}}}}}-{{{{{\rm{Si}}}}}}{{{{{{\rm{P}}}}}}}_{2}}}{{t}_{{{{{{\rm{Si}}}}}}{{{{{{\rm{P}}}}}}}_{2}}}{V}_{{{{{{\rm{tg}}}}}}-{{{{{\rm{Si}}}}}}{{{{{{\rm{P}}}}}}}_{2}}$$
(3)

where \(e\) is the electron charge, \({\varepsilon }_{0}\) is the vacuum permittivity, \({t}_{{{{{{\rm{Si}}}}}}{{{{{{\rm{P}}}}}}}_{2}}\) = 20 nm is the thickness of SiP2, and \({\varepsilon }_{{{{{{\rm{r}}}}}}-{{{{{\rm{Si}}}}}}{{{{{{\rm{P}}}}}}}_{2}}\) is the relative dielectric constant of SiP2 within the Au/SiP2/MoS2 sandwiched structure. The \({\varepsilon }_{{{{{{\rm{r}}}}}}-{{{{{\rm{Si}}}}}}{{{{{{\rm{P}}}}}}}_{2}}\) is obtained by linear fitting of Eq. (3).

DFT calculations

The Vienna Ab initio Simulation package (VASP)42 was used for the first-principles calculations. The generalized-gradient approximation (GGA) of the Perdew-Burke-Ernzerhof-type functional43 was used with the projected-augmented-wave method44,45 and an energy cutoff of 500 eV. For the electronic self-consistent calculations, the convergence criterion was set as 10–6 eV. Considering the van der Waals interaction, the DFT-D3 method46 was applied as the correction. The force criterion was chosen to be 0.01 eV Å–1. After fully relaxing the 2H-MoS2 bulk structure, we obtained the lattice constants of 3.16 Å for a (and b) and 12.39 Å for c, which were consistent with previous studies47,48. The k-point mesh of 15 × 15 × 3 was used to sample the Brillouin zone (BZ). After the full relaxation of the SiP2 bulk structure, the lattice constants were 10.11 Å for a, 3.44 Å for b, and 14.18 Å for c, which were consistent with the experimental results49. The k-point mesh of 5 × 15 × 4 was used to sample the BZ of bulk SiP2. Moreover, except for the structural relaxation, the spin-orbit coupling was considered in all DFT calculations. For all slab models used in this work, a vacuum layer with 20 Å was added along the z direction. Moreover, the calculations with the HSE06 functional50,51 were performed to obtain the accurate values of the work function and bandgap of the unstrained and strained MoS2 and SiP2 monolayers.

For these heterostructure models with strained MoS2 and SiP2 which were used to simulate the electronic properties of 1L-MoS2/SiP2 heterostructure directly, the \(2\sqrt{3}\times 1\) rectangle supercell for monolayer MoS2 was used and its lattice was tensed to 3.25 Å for a, and was compressed to 10.69 Å for b. Correspondingly, the lattice of monolayer SiP2 was enlarged to 3.50 Å for a and 10.69 Å for b. In each strained heterostructure model, it contains 13 × 1 monolayer SiP2 unitcells and 14 × 1 rectangle supercells for monolayer MoS2. The k-point mesh of 5 × 1 × 1 was used to sample the BZ of heterostructure model. For strained case-I-ABBA and case-II-AABA lattices, we fully relaxed the whole structures. While, for strained case-I-AA and case-II-AB lattices, we fixed the x and y coordinates of the two central S atoms in MoS2 layer and one central P atom in SiP2 layer in the AA/AB stacking region. Then we fully relaxed other atoms in the strained heterostructure model.