Introduction

The epitaxial growth of two-dimensional (2D) materials is a promising route to wafer-scale, single-crystalline, atomically thin films, showing great potential in integrated electronic and optoelectronic devices beyond silicon1,2,3,4. Substrate engineering plays a crucial role in realizing this goal. For monolayer growth, methods like introducing step edges5,6,7,8,9, defects10,11, and interfacial reconstructed layers12,13, as well as selecting appropriate substrate surface symmetry14,15,16, have been utilized to prepare graphene, hexagonal boron nitride, 2H-MoSe2, and 2H-WSe2. For multilayers, techniques like bevel-edge epitaxy17, interfacial epitaxy18, and remote epitaxy19 have been applied to control interfacial stacking order and thickness for rhombohedral boron nitride, 2H-MoS2, and 2H-WS2. However, most of these 2D materials are high-symmetry crystals (e.g., graphene: C6v; h-BN: C3v, 2H-MoS2, −MoS2, −WS2, and −WSe2: D3h), while low-symmetry ones such as triclinic ReX2 (X = S, Se) and monoclinic 1T′-MTe2 (M = Mo, W) receive limited attention in epitaxial 2D fabrication studies. These low-symmetry 2D layers require more stringent control of the lattice orientation than the high-symmetry crystals even for monolayers due to the following reasons.

Firstly, low-symmetry crystal structures lead to abundant in-plane anisotropic properties, which can only be effectively manifested at the macroscopic scale when the 2D material adopts a single lateral lattice orientation across a wide range. The presence of different orientated domains results in grain boundaries (GBs) that degrade carrier mobility and spatially averages out the anisotropic properties of low-symmetry crystals, ultimately impacting mechanical, electronic, and optical applications. Efforts have been devoted to the in-plane orientation-controlled epitaxial growth via substrate engineering20,21,22,23,24. Take ReS2 \(\left({{{\rm{P}}}}\bar{1}\right)\) as an example, which is a typical low-symmetry semiconducting 2D material possessing only an inversion center as the symmetric element, monolayer single-crystalline ReS2 domains with unidirectionally aligned b-axis (orientation of Re4 chains) were grown on Au(101) due to strong interfacial coupling and low-symmetry of the substrate surface (C2v)23. More recently, continuous 2D ReS2 films with a consistent b-axis orientation were achieved on Au(111) by introducing step edges along \(\left[01\bar{1}\right]\) to decrease the surface symmetry to Cs (having a mirror plane)24. However, these approaches suffer from high substrate cost (noble metal), limited crystal quality, non-uniform thickness, and, more importantly, the inability to control the out-of-plane lattice orientation of monolayers, which will result in the achievement of racemic mixture rather than a single structure (discussed in the next paragraph).

The necessity to control orientations of low-symmetry 2D crystals comes from an even more important reason, i.e., monolayers with opposite vertical lattice directions have different chirality in two dimensions when they are confined to a plane, meaning that they cannot be interconverted by any rotation about the axis perpendicular to the substrate. This is illustrated by monolayer ReS2 in Fig. 1a. R-ReS2 and S-ReS2 are enantiomers, despite having the same in-plane b-axis orientation (Re4 chain direction marked in red), and they display exotic physical properties like chiral Raman scattering25,26. In other words, when monolayer ReS2, which is achiral in three dimensions, binds to a substrate, its original inversion center disappears, thus generating a chiral ReS2/substrate interface in space. This phenomenon introduces an additional degree of freedom for structural control of low-symmetry 2D crystals, expanding the possibility for performance tuning but also increasing the difficulty of material fabrication. Currently, the construction of chiral 2D structures predominantly focuses on twisted-stacked 2D crystals via sequential mechanical transfer and screw dislocation-triggered growth27,28,29,30, layered 2D crystals intercalated by chiral molecules31, and chiral perovskites via direct synthesis32,33, among which diversified chirality-dependent exotic properties like spin selectivity, magnetic phases, and photoelectric responses were discovered33,34,35. Yet chirality-selected growth of low-symmetry monolayer single crystals by complete control over both in-plane and out-of-plane lattice orientations is blank and urgently needed.

Fig. 1: Terrace-step-kink strategy of substrate design.
figure 1

a Schematic illustrating chirality of monolayer ReS2 on the substrate, where R and S refer to rectus and sinister in Latin, respectively, representing two types of enantiomers. The structural differences between R- and S-ReS2 on the substrate surface are marked by red symbols. b–d Schematics showing the substrate surface reconstruction from MgO(11l) to terrace facets (001) with bunched steps along \(\left[\bar{1}10\right]\) (b) and reconstruction from chiral surface R-(12l)/S-(21l) to terrace facets (001) with bunched steps along \(\left[\bar{1}10\right]\) and kinks along \(\left[\bar{1}00\right]\)/[010] (c and d). e–g Scheme of the designed terrace-step-kink strategy, which gradually increases the structural selectivity of the as-grown ReS2 on the MgO substrate by sequentially introducing specific terraces, steps, and kinks to the reconstructed substrate surfaces to reduce their structural symmetry, thus eliminating energy degeneracy of monolayer ReS2 single crystals with different orientations and chirality. Blue and red blocks represent ReS2 domains in different chirality. The gray lines represent microstructures on the reconstructed substrate surfaces with the zoomed-in inset highlighting steps and kinks. h Binding energies (Eb) calculated using density functional theory (DFT) for ReS2 single-crystalline domains with different rotation angles (θ) on MgO(001), θ represents the angle between the b-axis ([010]) of ReS2 and MgO\(\left[\bar{1}10\right]\). i Formation energies (Ef) of three grain boundaries with different inter-grain twist angles (insets). j Binding energies (Eb) of monolayer single-crystalline ReS2 domains oriented perpendicularly on reconstructed MgO(11l) composed of terrace(001) and step\(\left[\bar{1}10\right]\). The red lines in the top panel of the insets represent step\(\left[\bar{1}10\right]\) on the reconstructed MgO surface, while the blue rhombuses represent Re4 chains of monolayer ReS2. k Binding energies (Eb) of R- and S-ReS2 monolayer single crystals with the same b-axis orientation on the chiral reconstructed MgO(12l)R surface. The red curves (top panels in the insets) represent the “step + kink” configuration on MgO.

Herein, we demonstrate the chemical vapor deposition (CVD) growth of chirality-selected unidirectionally aligned monolayer ReS2 single crystals on a cost-effective and industry-compatible insulating substrate via a terrace-step-kink strategy. The acquisition of the target material is decomposed into three steps with correspondingly developed substrate engineering methods (Fig. 1e–g). Step 1: Realizing the single-crystallinity of each ReS2 domain, fulfilled by choosing the substrate terrace facet with an appropriate rotation symmetry, which can prohibit ReS2 from generating thermodynamically favorable GBs during the epitaxial growth (Fig. 1e). Step 2: Realizing the in-plane unidirectional alignment of every single-crystalline domain, reached by introducing specific steps on the terrace (Fig. 1f), thus breaking the binding energy degeneracy of differently oriented ReS2 domains. Step 3: Realizing the chirality selectivity of unidirectionally aligned single crystals. A chiral substrate surface was constructed by further introducing selected kinks to the “terrace + step” microstructures (Fig. 1g) so that a chirality transfer mechanism functions, which enables breaking both interfacial configuration and energy degeneracy of monolayer single-crystalline ReS2 enantiomers (definitions of steps and kinks are included in Supplementary Fig. 1).

Results

Terrace-step-kink strategy of substrate design

Considering crystal structure, physical properties, and material costs, MgO, a binary cubic system with high chemical and thermal stability, was selected as the substrate. Its insulating characteristic also facilitates better compatibility of the 2D semiconducting materials grown on it with the current integrated circuit technology. Firstly, we performed DFT calculations to verify the theoretical feasibility of the above substrate design strategy. The key to achieving single-crystalline ReS2 domains (Step 1) is to prevent Re4 chains from deflecting during growth and forming GBs. Previous literature has demonstrated that, distinct from 2D materials like MoS2 and WS2 which can spontaneously grow into triangular/hexagonal single crystals under most conditions, Re4 chains in ReS2 are prone to deflect integer multiples of  60° during domain growth, thereby generating a series of highly coherent GBs with low formation energies and leading to polycrystalline products36,37,38. Therefore, we chose MgO(001) with C4v as the substrate terrace facet rather than α-Al2O3(0001) (C3v) and mica(0001) (C6v) in order to prohibit ReS2 from generating thermodynamically advantageous GBs when epitaxially grown on the single-crystalline substrate. Theoretical calculations exhibit that monolayer ReS2 single crystals have two optimal epitaxial growth orientations orthogonal to each other on MgO(001), whose intersectional angles (θ) between the b-axis of ReS2 and MgO\(\left[\bar{1}10\right]\) is 0° and 90°, respectively. The suboptimal epitaxial orientation is θ = 45° (Fig. 1h and Supplementary Fig. 2). On the premise of ensuring a superior epitaxial growth orientation on the substrate, monolayer ReS2 can only construct GBs with inter-grain angles of 45° and 90°, whose calculated formation energies are significantly larger than those highly coherent GBs commonly observed in previous work38 (e.g., the GB with an inter-grain angle of 120°, Fig. 1i). It indicates thermodynamic prohibition to form GBs for epitaxially grown ReS2 single layers on MgO(001) with C4v symmetry.

To achieve unidirectional alignment of single-crystalline domains (Step 2), we introduced steps along the crystallographic orientation of \(\left[\bar{1}10\right]\) (named step\(\left[\bar{1}10\right]\)) on the terrace(001) of MgO. Such surface microstructures, in principle, can be obtained by spontaneous reconstruction of the vicinal MgO(001), referred to as MgO(11l) (\(l\)\(1\)), after annealing (Fig. 1b). It is noteworthy that the step orientation was deliberately designed to be parallel with and perpendicular to the b-axis orientations of two orthogonally oriented ReS2 single crystals epitaxially grown on the terrace(001). This design can maximize the interfacial structural disparity of two oriented ReS2 single crystals on the substrate and break energy degeneracy remarkably. DFT calculations show a prominently lower interfacial binding energy when the b-axis direction of ReS2 is parallel with step\(\left[\bar{1}10\right]\) on MgO than when they are perpendicular (Fig. 1j), implying that introducing step\(\left[\bar{1}10\right]\) can thermodynamically drive uniform orientation of monolayer ReS2 single crystals. It is worth noting that since the reconstructed MgO(11l) still has mirror symmetry (Cs), the structures of ReS2 enantiomers on the substrate surface mirror each other (Supplementary Fig. 3), displaying the same manner of interfacial atomic alignment and thus degenerated binding energies. Therefore, chirality selectivity is incapable of being realized solely by the “terrace(001) + step\(\left[\bar{1}10\right]\)” substrate design.

Finally, we introduced kinks along lattice directions of \(\left[\bar{1}00\right]\) and [010] (named kink\(\left[\bar{1}00\right]\) and kink[010]) on the MgO surface involving terrace(001) and step\(\left[\bar{1}10\right]\). Such design eliminates the mirror symmetry of the substrate, thus constructing chiral surfaces which may realize chirality-selective growth (Step 3, Fig. 1c, d and Supplementary Fig. 4). The chiral surfaces can commonly be obtained by thermal treatment of MgO(12l) and MgO(21l) (\(l\)\(2\)) because unstable high-index facets are inclined to reconstruct to thermodynamically favorable low-index counterparts. By definition, (12l) and (21l) can be assigned to rectus (R) and sinister (S) configurations, which are termed (12l)R and (12l)S (\(l\gg 2\)), respectively (Supplementary Fig. 5). It is noteworthy that the construction of chiral surfaces only requires (hkl) to satisfy h ≠ k ≠ l and h × k × l ≠ 0 geometrically. Here, we designed chiral surfaces of (12l)R and (12l)S (\(l\gg 2\)) for the following reasons: (i) Selection of high-index planes with \(l\gg 2\) ensures a sufficiently large area of each terrace layer and a low height of every step/kink after surface reconstruction, which facilitate monolayer nucleation and rapid domain expansion across steps/kinks. (ii) Reconstructed surfaces of (12l)R and (12l)S have a similar density of step\(\left[\bar{1}10\right]\) to that of the reconstructed (11l) designed in Step 2. Moreover, the length of the newly introduced kink\(\left[\bar{1}10\right]\)/kink[010] is 1/√2 ≈ 70% of the step\(\left[\bar{1}10\right]\) length. Such design allows the substrate to effectively control both the in-plane alignment and chirality of ReS2 single crystals. DFT calculations display that the binding energy of R-ReS2 on MgO(12l)R is lower than that of S-ReS2 on MgO(12l)R (Fig. 1k), suggesting chirality transfer from the substrate surface to the as-grown monolayer products.

Growth of chirality-selected, unidirectional ReS2 single crystals

Experimentally, the chiral substrate (e.g., (12l)S) was achieved by custom-designing MgO(001) with a cutting angle of ~1° towards [210] (Supplementary Figs. 6 and 7). The surface was reconstructed into terraces(001) with a bunch of parallelly aligned steps along \(\left[\bar{1}20\right]\) (Supplementary Fig. 8) after annealing in oxygen. These steps are, in principle, composed of step\(\left[\bar{1}10\right]\) and kink[010] at the nanometer scale (Fig. 1d). Figure 2a shows spindle-shaped ReS2 crystals with a length of 15–20 µm grown on MgO(12l)S by hydrogen-free atmospheric-pressure CVD. They were unidirectionally aligned along MgO\(\left[\bar{1}10\right]\) and can merge together with atomically stitched interfaces (Supplementary Fig. 9). The angle-dependent Raman intensities of the vibrational mode at 212 cm−1 were measured in the parallel polarization configuration for different crystals and various locations in the same crystal of the as-grown ReS2 (inset of Fig. 2a and Supplementary Fig. 10). They exhibited almost overlapped double-lobed patterns with the maximized intensity corresponding to an angle of 353° ± 4° (0° in the polar plot corresponded to the major axis direction of the spindle-like domain), indicating that the as-prepared products have consistent crystallographic orientation within and between crystals and the major axis direction of the spindle-shaped crystals corresponds to the b-axis orientation of ReS2. The second harmonic generation (SHG) spectra and mapping further verified single-crystallinity without GBs of ReS2 grown on MgO(12l)S because no obvious signal was detected, which was in stark contrast to the strong SHG intensity observed in polycrystalline ReS2 prepared on α-Al2O3(0001) (Supplementary Figs. 11 and 12). Annular dark-field scanning transmission electron microscopy (ADF-STEM) showed well-organized Re4 chains with neither detectable lattice deformation nor atom missing, demonstrating monolayer thickness of the 2D products with superior crystal quality at the atomic scale (Fig. 2b). Selected area electron diffraction (SAED) captured at different domains displayed only one set of quasi-hexagonal patterns aligning in the same orientation without high-order diffraction spots splitting (Fig. 2c), manifesting consistent lattice orientation of disparate domains (Supplementary Fig. 13).

Fig. 2: Growth and characterization of chirality-selected, unidirectional, single-crystalline monolayer ReS2.
figure 2

a Scanning electron microscopy (SEM) image of unidirectionally aligned spindle-like ReS2 domains grown on MgO(12l)S. The inset shows the angle-resolved Raman intensities of the vibrational mode of ReS2 at 212 cm−1 captured at three different locations in one domain. Scale bar: 20 µm. b Annular dark-field scanning transmission electron microscopy (ADF-STEM) image of monolayer ReS2. The Re4 chains are represented by white rhombuses with crystallographic directions a [100] and b [010] labeled. Scale bar: 2 nm. c Selected area electron diffraction (SAED) patterns obtained at four different domains. The yellow hexagons mark the innermost diffraction spots, while the insets are the zoomed-in views of the red- and blue-boxed high-index diffraction spots. Scale bar: 2 nm−1. d Circularly polarized Raman (CPR) spectra of as-grown ReS2 in (a) (bottom) and the difference between the Raman intensities excited by right-handed (RCP) and left-handed (LCP) circularly polarized light (top). The wavelength of the polarized light is 532 nm. e Histogram showing the circular intensity difference (CID) distribution of Raman modes III (153 cm−1), V (212 cm−1), and VI (235 cm−1) captured from ≈50 ReS2 domains located at different regions on MgO(12l)S. f Atomic force microscopy (AFM) image displaying a ReS2 domain (left to the white dashed line) grown on bare MgO(12l)S (right to the white dashed line). The reconstructed steps along MgO\(\left[\bar{1}20\right]\) (which are composed of step\(\left[\bar{1}10\right]\) and kink[010] at the atomic scale) are marked by the red arrow, while the b-axis of ReS2 is labeled by the white arrow. Scale bar: 200 nm. g, Zoomed-in atomically resolved AFM image of the ReS2 domain in (f). The Re rhombic lattice along the b-axis is marked by white lines. Scale bar: 2 nm. h Cross-sectional ADF-STEM image of the ReS2/MgO interface viewed along MgO\(\left[\bar{1}10\right]\). Scale bar: 1 nm. i Schematic illustrating the epitaxial relationship of monolayer ReS2 on reconstructed MgO(12l)S. The orange and white spheres represent the first and second layers of O atoms of the MgO surface, respectively.

We then applied CPR spectroscopy to determine the 2D chirality of ReS2 on MgO(12l)S. Remarkable differences in relative intensities were observed in CPR spectra of as-grown domains under RCP and LCP due to phase differences among multiple phonon-mediated electron scattering pathways25 (Fig. 2d). The CID values, defined as (IRCPILCP) / (IRCP + ILCP), for vibrational mode III (151 cm−1), V (212 cm−1), and VI (235 cm−1) were measured to be 0.124, 0.523, and −0.214, respectively, agreeing well with those for S-ReS2. We characterized CPR spectra from 50 randomly selected ReS2 domains grown at various positions on MgO(12l)S. The statistical results demonstrated that the CID feature exhibited in Fig. 2d was universal, implying S chiral configuration enrichment of our sample (Fig. 2e).

Atomic force microscopy (AFM) and cross-sectional ADF-STEM were employed to unveil the epitaxial relationship between ReS2 and MgO(12l)S. The b-axis direction of monolayer ReS2 was found to deflect  18.5° from parallelly aligned edges along \(\left[\bar{1}20\right]\) on reconstructed MgO(12l)S in the AFM image (Fig. 2f, g and Supplementary Fig. 14), proving again that the major axis (b-axis) of the spindle-shaped ReS2 crystals is parallel to MgO\(\left[\bar{1}10\right]\) rather than MgO\(\left[\bar{1}20\right]\). The cross-sectional ADF-STEM image displayed the atomic arrangement of monolayer S-ReS2 on MgO along MgO\(\left[\bar{1}10\right]\) (Fig. 2h). The lattice spacing of ReS2 in the lateral direction was measured as 5.60 Å, well matching with \({a}_{{{{{\rm{ReS}}}}}_{2}}\times \sin {\alpha }_{{{{{\rm{ReS}}}}}_{2}}=5.61\,{{{\text{\AA }}}}\,\left(\right.{a}_{{{{{\rm{ReS}}}}}_{2}}=6.58\,{{{\text{\AA}}}},{\alpha }_{{{{{\rm{ReS}}}}}_{2}}=121.5^\circ\) representing the angle between a-axis and b-axis of ReS2), which corresponds to the lattice spacing of ReS2 perpendicular to its b-axis direction. The lattice spacing of MgO was measured as 2.98 Å, consistent with \({\sqrt{2}a}_{{{{\rm{MgO}}}}}/2=2.98\,{{{\text{\AA }}}}\,({a}_{{{{\rm{MgO}}}}}=4.21\,{{{\text{\AA }}}})\). These results directly confirmed that the (001) lattice plane of ReS2 is parallel to the (001) terrace facet of MgO and the b-axis of ReS2 is parallel to the \(\left[\bar{1}10\right]\) lattice direction of MgO, which can be expressed as [010]ReS2(001)// \(\left[\bar{1}10\right]\)MgO(001) (Fig. 2i). In addition, the atomic step on the reconstructed substrate showed a height of 2.14 Å corresponding to half the lattice constant of MgO (aMgO/2 = 2.16 Å), and ReS2 was observed to grow across steps (Supplementary Fig. 15). The distance between ReS2 and the MgO step edge (2.75 Å) is close to the van der Waals gap between ReS2 and the MgO terrace (3.10 Å), suggesting van der Waals coupling at the ReS2/MgO step edge interface. These characteristics indicate that the terrace facets of MgO determine the epitaxial orientation of monolayer ReS2, while the steps/kinks are used to break the energy degeneracy of ReS2 with different lattice directions to achieve full structural control. The terrace-step-kink strategy can also be extended to the chirality-selective growth of other unidirectional low-symmetry 2D single crystals, such as monoclinic MoO2 (Supplementary Fig. 16).

Structural selectivity improvement by substrate engineering

To achieve an in-depth comprehension of how terraces, steps, and kinks on MgO control the structure of monolayer ReS2 and validate the theoretical calculations in Fig. 1, we compared the growth results of ReS2 on MgO(001), MgO(11l), and chiral surfaces of MgO(12l)R and MgO(12l)S, respectively. On the flat MgO(001) without steps and kinks (Fig. 3a), the contrast under a polarized optical microscope and the angle-resolved Raman intensity mapping at 212 cm−1 displayed homogeneous brightness in each crystal (inset of Fig. 3e and Supplementary Figs. 17 and 18), indicating the achievement of single-crystalline ReS2 domains on MgO(001) with C4v symmetry. However, two epitaxial growth orientations of ReS2 crystals along [110] and \(\left[\bar{1}10\right]\) of MgO, respectively, were observed, which were evidenced by the optical microscopy (OM) image and the sharp intensity disparity between two orthogonally oriented ReS2 crystals in the angle-resolved Raman mapping (Fig. 3e). Moreover, twelve spots which can be assigned to two sets of quasi-hexagonal patterns perpendicular to each other were obtained when conducting low-energy electron diffraction (LEED) on the as-grown ReS2 (Fig. 3f), further demonstrating the incapability of realizing unidirectional alignment of ReS2 single crystals on MgO(001).

Fig. 3: Step-by-step structural selectivity improvement of as-grown ReS2 by substrate engineering.
figure 3

a–d AFM images showing bare MgO(001) (a), and monolayer ReS2 grown on reconstructed MgO(11l) (b), MgO(12l)R (c), and MgO(12l)S (d), respectively. Regions to the left of the white dashed lines in (b–d) are monolayer ReS2-covered MgO surfaces with steps along \(\left[\bar{1}10\right],\left[\bar{2}10\right]\), and \(\left[\bar{1}20\right]\) marked by white, blue, and red arrows, respectively. Areas to the right of the white lines are bare substrates. The atomic models corresponding to each surface are displayed below the AFM images. Large spheres in different shades of orange represent O atoms in different layers of the MgO substrate, while small gray spheres represent Mg atoms. Scale bars: 200 nm. Insets in (b–d) are the nanometer-scale AFM images showing the reconstructed steps and kinks on the corresponding substrates after annealing. The white and yellow dashed lines mark the steps and kinks, respectively. Scale bar: 30 nm. e Optical image showing ReS2 domains grown on MgO(001) with two preferred orientations of 0° and 90°. Scale bar: 50 µm. The inset is an angle-resolved Raman intensity mapping of mode V for ReS2 domains grown on MgO(001). Scale bar: 5 μm. f Low-energy electron diffraction (LEED) patterns of ReS2 grown on MgO(001). g In-plane orientation histograms of ReS2 domains grown on MgO(001) and MgO(11l), respectively. 100 domains were counted on each substrate. h Linearly polarized Raman spectra of R-ReS2 and S-ReS2 for a laser wavelength of 532 nm. Inset is an optical image showing that the b-axis of ReS2 (black arrow) is rotated 120° with respect to the polarization direction of the incident laser (green arrow) during characterization. Scale bar: 5 μm. i, j Polarization dependence of the intensities of Raman modes III (red) and V (blue) at 151 cm−1 and 212 cm−1 for R-ReS2 (i) and S-ReS2 (j), respectively. The b-axis of ReS2 is marked by the black arrow. k, l Statistical distributions of the normalized intensity of Raman mode III (IIII/IV) for ReS2 domains grown on MgO(12l)R and MgO(12l)S (insets), respectively. The measurement condition was the same as that in (h). Each substrate counted ≈ 200 domains.

We then cut MgO(001) towards [110] with an angle of ~1° to achieve MgO(11l). The substrate surface was reconstructed to straight, low-index atomic steps arranged parallel to \(\left[\bar{1}10\right]\) after annealing (Fig. 3b). OM and angle-resolved polarization Raman spectroscopy showed that the introduction of step\(\left[\bar{1}10\right]\) on MgO(001) led to unidirectional alignment of spindle-shaped ReS2 single crystals with an orientation selectivity of >99% (red columns in Fig. 3g and Supplementary Fig. 19). In contrast, on MgO(001), the ratio of ReS2 domains aligning along [110] and \(\left[\bar{1}10\right]\) was 49:51 (blue columns in Fig. 3g), demonstrating the decisive role of step \(\left[\bar{1}10\right]\) in breaking the symmetry of the orthogonal orientations of ReS2 single crystals. Next, we applied a method to rapidly identify the chiral configuration of the 2D products by comparing the Raman peak intensity between the chirality-sensitive mode III (151 cm−1) and chirality-insensitive mode V (212 cm−1), which was measured by setting an angle (φ = 120°) between the linear polarization direction of the incident laser and the b-axis of ReS2 (Fig. 3h, details discussed in the next paragraph and Supplementary Fig. 20). The statistics captured from 100 showed that the quantity ratio of S- and R-ReS2 was 43:57 (Supplementary Fig. 21), implying that solely introducing step\(\left[\bar{1}10\right]\) on terrace(001) cannot reach chirality selectivity of monolayer ReS2.

Finally, the chiral surfaces of MgO(12l)R and MgO(12l)S were applied to prepare monolayer ReS2. Well-organized step edges along \(\left[\bar{2}10\right]\) and \(\left[\bar{1}20\right]\) were displayed on the reconstructed chiral surfaces after annealing (top panels in Fig. 3c, d), which exhibited jagged edges composed of low-index step\(\left[\bar{1}10\right]\) and kink\(\left[\bar{1}00\right]\) /kink[010] at the nanometer scale (bottom panels in Fig. 3c, d). Since Fig. 2 has included comprehensive characterization of the single-crystallinity, orientation consistency, and epitaxial relationship of monolayer ReS2 on MgO(12l)S and utilized CID to demonstrate the single chirality of the 2D products qualitatively, here we focused on the quantitative analysis of the chirality selectivity of monolayer ReS2 crystals grown on R and S chiral MgO surfaces from the statistics perspective. The prerequisite to realizing this goal is to find a fast, accurate, and non-destructive approach for chiral structure identification. It is worth noting that the CPR method used in Fig. 2d has several limitations, including the requirement of frequent optical path switching (each test point needs to switch the optical path once in order to achieve RCP and LCP light), low characterization efficiency, and difficulty in maintaining the test position precisely unchanged before and after optical path switching. Therefore, the CPR method is not suitable for statistical chirality characterization of a quantity of 2D crystals. We utilized the approach mentioned in Fig. 3h and used the peak intensity ratio between Raman mode III and V (IIII/IV), denoted as the normalized intensity of mode III, to determine the 2D chiral configuration of ReS2 on MgO. This strategy needs no optical path switching and can identify the chirality of each test point by only one spectrum. Figure 3i, j showed the angle-resolved Raman intensities of mode III and mode V under parallel configuration corresponding to one R-ReS2 domain and one S-ReS2 domain grown on MgO(12l)R and MgO(12l)S, respectively. It displayed that the double-lobed pattern of mode V (blue) did not change with the variation of the product chirality, manifesting that Raman mode V of ReS2 is insensitive to the product’s chirality. On the contrary, the patterns of mode III (red) corresponding to R- and S-ReS2 were mirror symmetric, indicating that the intensity of mode III had a significant dependence on the product’s chirality. We measured IIII/IV under φ = 120° to determine ReS2 chirality, which is due to the following reasons: (i) Selecting the normalized rather than the absolute intensity of Raman mode III as chirality criterion can avoid the influence of experimental variables unrelated to ReS2 chiral configurations, such as tiny fluctuation of incident laser power and subtle variation of laser focusing degree on different samples. (ii) When φ = 120°, the IIII/IV corresponding to S- and R-ReS2 reach the maximum and secondary maximum values in their polar plots, respectively (Supplementary Fig. 22), implying that the change rates of IIII/IV with φ are minimum at these two points. Therefore, selecting φ = 120° helps improve the robustness of the measured IIII/IV even when the true φ values in the experiments slightly deviated from 120°. The normalized mode III intensities of R- and S-ReS2 were measured to be 2.2 ± 0.6 and 6.3 ± 1.8, respectively. Polarization-resolved Raman mapping further shows homogeneous IIII/IV distributions within and across crystals, confirming single-crystallinity and uniform chirality (Supplementary Fig. 23). Statistics of IIII/IV corresponding to 200 ReS2 crystals grown on MgO(12l)R and MgO(12l)S, showed that R-ReS2 accounted for 97.5% of the total domains on MgO(12l)R (Fig. 3k), and S-ReS2 accounted for 98.7% on MgO(12l)S (Fig. 3l). Apart from the control of chirality, the nucleation density of monolayer ReS2 single crystals can also be modulated by adjusting the density of step edges and kinks on the chiral MgO substrate (Supplementary Figs. 2426).

Device performance of chiral ReS2 single crystals

We transferred the as-grown ReS2 on a SiO2/Si substrate to fabricate a back-gated field effect transistor (FET) with Cr/Au utilized as the contact electrode to achieve matched work function at the interface (Supplementary Fig. 27). The Ids-Vds curve displays linear characteristics, indicating a low-resistance ohmic contact between ReS2 and the electrode (Supplementary Fig. 28). The transfer curve in Fig. 4a manifests a typical n-type semiconducting behavior. We calculated the carrier mobility (\(\mu=\frac{d{I}_{{{{\rm{ds}}}}}}{d{V}_{{{{\rm{g}}}}}}\cdot \frac{L}{W}\cdot \frac{1}{{C}_{{{{\rm{i}}}}}\cdot {V}_{{{{\rm{d}}}}{{{\rm{s}}}}}}\), where L and W represent the length and width of the FET device, and Ci = 12.1 nF cm−2 is the measured capacitance per unit area for 285 nm SiO2 layer), yielding μ = 3.35 cm2 V−1 s−1 for the single-crystalline ReS2 grown on MgO(12l)R at room temperature, which is among the highest reported for CVD-grown ReS2 and is 30 times higher than that of polycrystalline ReS2 prepared on α-Al2O3(0001) (0.11 cm2 V−1 s−1). The transconductance (\({g}_{m}=\frac{{{dI}}_{{ds}}}{{{{\rm{d}}}}{V}_{g}}\)) extracted from the linearly region of the transfer curve is 0.35 μA/V, comparable to those of mechanically exfoliated ReS239,40.

Fig. 4: Electrical and photoelectric properties of chiral monolayer ReS2 single crystals.
figure 4

a Ids-Vg curves of single-crystalline and polycrystalline monolayer ReS2-based field effect transistor (FET) devices. Here, Ids denotes the drain-source current, and Vg is the gate voltage. The channel length/width ratios (L/W) are 2.57 and 2.49 for the single- and poly-crystalline devices, respectively. Inset is the linear polarization-dependent photocurrent of the single-crystalline device under 532 nm laser illumination. b Carrier mobility vs. anisotropic ratio in photodetection of various 2D low-symmetry materials to compare our ReS2 device performance with the other state-of-the-art devices (Supplementary Table 1). c Statistical distributions of full width at half maximum (FWHM) and intensities corresponding to Raman mode I and V for ReS2 single crystals grown on chiral MgO and polycrystals grown on α-Al2O3. Each substrate counted 30 domains. d CPR spectra for R- and S-ReS2 transferred onto the MgO(001), α-Al2O3(0001), and SiO2/Si substrate, respectively. e CID values for R- and S-ReS2 on MgO(001), α-Al2O3(0001), and SiO2/Si. Each substrate counted 20 domains. Error bars are standard deviations. f, g Ids-Vds curves of R- and S-ReS2-based photodetectors under 532 nm circularly polarized light (CPL) illumination. Error bars represent the standard deviation from several measurements. h Comparison of the asymmetrical factor for photocurrent between our monolayer chiral single-crystalline ReS2-based photodetector for CPL detection and the other state-of-the-art low-symmetry 2D material-based devices (Supplementary Table 2).

Monolayer ReS2 can produce anisotropic photoresponse in the visible band due to its low lattice symmetry and direct bandgap of 1.43 eV41 (Supplementary Fig. 29). Under 532 nm laser illumination, the responsivity of the single-crystalline ReS2 device (\(R=\frac{{I}_{{{{\rm{ph}}}}}}{P},{{{\mbox{where }}}}\,{I}_{{{{\rm{ph}}}}}\) is the photocurrent and P is the light power) was 113.76 A W−1 at zero gate voltage (Vg = 0) and a low bias voltage (Vds = 0.1 V), whose performance is in the first echelon among the CVD-prepared 2D anisotropic materials under similar measurement conditions. More importantly, the photocurrent of the ReS2 device showed prominent dependence on the linear polarization direction of the incident light under the bias voltage ranging from 0 to 2 V (Supplementary Fig. 30), yielding a superior anisotropic ratio of 1.9 when Vds = 1.7 V (inset of Fig. 4a). The photocurrent reached maximum when the laser polarization direction was parallel to the b-axis of ReS2. Both the optical anisotropic ratio and the carrier mobility of our single-crystalline ReS2-fabricated device are among the highest values reported for CVD-grown ReS2 to date and are comparable to those of many mechanically exfoliated anisotropic 2D materials with larger thickness (Fig. 4b).

We further verified the crystal quality by the in-plane (mode I at 131 cm−1) and out-of-plane (mode V at 212 cm−1) Raman vibrational modes to unclose the excellence reason of our sample in some electrical and photoelectric performances. Statistics show that single-crystalline ReS2 grown on MgO(12l)R exhibits remarkably narrower distributions of full width at half maximum (FWHM) and more uniform Raman peak intensities for both mode I and V compared with the polycrystalline counterparts prepared on α-Al2O3(0001) (Fig. 4c). The FWHM and peak intensity variances of the single-crystalline samples are only ~30% and ~20% of those of the polycrystalline specimens (Supplementary Fig. 31), indicating high and homogeneous crystal quality.

Next, we investigated the 2D chiral structure-related physical properties and device performances of as-grown ReS2. We transferred monolayer R-and S-ReS2 grown on MgO(12l)R and MgO(12l)S to three achiral substrate surfaces with different chemical compositions and structures, which are MgO(001), α-Al2O3(0001) and the amorphous SiO2/Si. The ReS2 enantiomers displayed obvious helicity-dependent Raman intensities on all three substrates, and the R- and S-ReS2 showed opposite chiral Raman responses, indicating that the 2D chirality of monolayer ReS2 is predominantly determined by its out-of-plane orientation and is independent of whether the substrate surface is chiral or not (Fig. 4d, e). These results validate a universal and robust strategy to construct 2D chiral interfaces on diversified substrates using atomically thin low-symmetry materials. The CID values of R-ReS2 on the three substrates excited by LCP are higher than those by RCP for mode III (153 cm−1) and V (212 cm−1), while mode IV (162 cm−1) and VI (235 cm−1) display opposite chiral Raman responses. S-ReS2 behaved the opposite. In addition, the CID values for the same Raman mode of ReS2 on different substrates vary slightly, which may stem from different electron-photon and electron-phonon interactions in ReS2 due to disparate charge transfer between ReS2 and the substrate and the intrinsic optical property disparity of various substrates25,42,43.

The CPL detection ability of our chiral monolayer ReS2 was then studied by fabricating a photodetector directly on the MgO substrate in order to eliminate the negative impact on the device performance from the contamination introduced in the transfer procedure. The R-ReS2 device displayed a larger photocurrent under 532 nm RCP illumination than under the LCP illumination, while the S-ReS2 device behaved the opposite, demonstrating prominent CPL distinguishability (Fig. 4f, g). We quantified the distinguishability of CPL detection by an asymmetrical factor for photocurrent (\({g}_{{{{\rm{Iph}}}}}=\frac{2({I}_{{{{\rm{LCP}}}}}-{I}_{{{{\rm{RCP}}}}})}{{I}_{{LCP}}+{I}_{{{{\rm{RCP}}}}}}\), where ILCP and IRCP are the photocurrent under LCP and RCL irradiation) and yielded \({g}_{{{{\rm{Iph}}}}}\) of −0.31 and 0.20 for R-ReS2 and S-ReS2, respectively, which are comparable to those chiral perovskites and metal oxides with larger thickness (Fig. 4h and Supplementary Table 2).

Discussion

We demonstrate a terrace-step-kink strategy for the direct growth of single-chirality, unidirectional, semiconducting, anisotropic monolayer ReS2 single crystals on the custom-designed chiral surface of insulating MgO, yielding >99% in-plane orientation consistency and >97.5% chirality selectivity. Our combined experimental and theoretical results show that terrace facets determine the epitaxial direction of ReS2 while steps and kinks break the configuration and energy degeneracy of ReS2 in the lateral orientation and chirality. The as-grown products can be facilely transferred to various substrates and robustly construct chiral interfaces. The high quality of monolayer ReS2 is further evidenced by the large carrier mobility, superior anisotropic ratio in photoresponse (1.9) comparable to mechanically exfoliated samples, and high distinguishability of CPL comparable to those typical chiral materials with larger thickness. The terrace-step-kink strategy is applicable to the chiral epitaxy of other low-symmetry 2D single crystals, like monoclinic MoO2. This work extends the range of control over 2D material growth to chirality, demonstrating full manipulation of both in-plane and out-of-plane lattice orientations of 2D crystals via the synergy of terraces, steps, and kinks on the substrate and enabling chirality transfer from the substrate to the 2D materials. Our findings offer the opportunity to produce wafer-scale, chirality-selective, single-crystalline ReS2 membranes and have the potential to be generalized to other low-symmetry 2D crystals.

Methods

Annealing of MgO substrate

Single-side polished MgO substrates were purchased from Hefei Ruijing Photoelectric Technology Co. To reconstruct terraces, steps, and kinks on the surface, the MgO substrates were annealed in an oxygen atmosphere at 1050 °C for approximately 2 h.

Growth and transfer of ReS2 domains

A 0.0025 mol/L aqueous solution of NaReO4 was prepared by dissolving NaReO4 in deionized water. Single-side polished c-plane sapphire were treated with oxygen plasma to create a hydrophilic surface. The prepared salt solution was spin-coated onto the plasma-treated sapphire surface at 5000 rpm for 30 s, which was then used to provide precursor vapor to the target substrate for ReS2 growth. The growth process was carried out in a 1-inch diameter atmospheric pressure CVD system. During growth, sulfur powder was placed in an Al2O3 crucible and positioned at the center of the low-temperature zone, with the temperature set to 140–180 °C. The MgO substrate or other substrates were placed face-to-face on the sapphire substrate spin-coated by NaReO4 solution to form a spatially confined reaction microchamber, which was positioned at the center of the high-temperature zone (820–880 °C). For the growth of monolayer ReS2 domains, 50 sccm of high-purity Ar was used as a carrier gas, and growth was conducted for 10–20 min (Supplementary Fig. 32).

The transfer of ReS2 domains was conducted using a polymer-assisted method. A thin film of poly (methyl methacrylate) (PMMA) was spin-coated onto the ReS2/substrate surface. The sample was then floated on a 2 mol/L potassium hydroxide (KOH) solution, allowing the PMMA/ReS2 film to gently detach from the substrate. The detached film was rinsed in deionized water three times to remove residual etchants. The PMMA/ReS2 film was subsequently transferred onto a TEM grid or other substrates, naturally dried in air, and baked on a hotplate at 180 °C for 15 min. To remove the PMMA, the sample was immersed in acetone for 8 h.

Characterization

  1. (i).

    Optical characterization. Optical images were captured using an Olympus BX53M microscope. Raman spectra were acquired using a Renishaw Qontor confocal Raman spectrometer with a 532 nm excitation laser. The laser was focused, and scattered light was collected through either a 100× or 50× Leica objective lens. A grating of 1800 lines/mm was used for spectral acquisition. All linearly polarized Raman measurements were conducted in a parallel polarization configuration, where the analyzer was aligned parallel to the incident polarization direction. The CPR spectra were obtained by inserting a quarter-wave plate into the light path. SHG measurements were carried out using a microspectral scanning system (MStarter 100) with a 1064 nm laser.

  2. (ii).

    Top-view STEM characterization. ADF-STEM imaging in Fig. 2b was performed at room temperature using an aberration-corrected Titan Cubed Themis G2 300 microscope. The STEM was operated under an accelerating voltage of 300 kV, with a condenser lens aperture of 50 mm, a convergence semi-angle of 21.3 mrad, and a collection angle of 39–200 mrad. The dwell time for a single frame was set to 2 µs per pixel, with a pixel size of 0.012 nm px−1 and a beam current of 30 pA. Image processing was conducted using ImageJ software. A Gaussian blur filter (≈ 2 pixels) was applied to high-magnification ADF-STEM images for smoothing. To enhance visualization, false color LUT was applied to some grayscale ADF-STEM images. Atomic models were constructed using the Accelrys Discovery Studio Visualizer and VESTA44 software.

  3. (iii).

    Cross-sectional STEM characterization. A dual-beam focused ion beam (FIB) system (Thermo Scientific Scios 2) was employed for sample preparation. Initially, a 200-nm carbon protective layer was deposited using e-beam deposition within the FIB system. The sample was then tilted perpendicular to the ion beam, after which a tungsten strap layer (2 μm) was deposited to shield the region of interest during the milling process. A Ga+ ion beam was utilized to thin and purge the sample to a final thickness of 50 nm for subsequent analysis. For in-plane STEM investigations, the prepared samples were transferred onto TEM grids via a wet-transfer method, and the experiments were conducted using a JEOL GrandARM300F at an accelerating voltage of 300 kV. In STEM imaging, the electron beam was focused to <0.1 nm, using a semi-convergence angle of approximately 24 mrad and a 30 μm condenser lens aperture.

  4. (iv).

    SEM and AFM characterization. SEM images were obtained using Regulus 8100 system at 3 kV, 10 μA. AFM measurements were performed by an Asylum Cypher S system. Oxford (AC200TS) Al(100)-coated silicon cantilevers were used. The resonance frequency was 150 kHz. The spring constant was 9 N m–1 and the tip curvature radius was 7 nm.

  5. (v).

    XRD, SAED, and LEED characterization. XRD measurements were conducted using multipurpose X-ray diffractometer (SmaterLab SE) with a copper target. SAED patterns were obtained using transmission electron microscopy (JEOL F200). LEED (OCI, LEED 450) were performed under an ultrahigh vacuum at room temperature.

  6. (vi).

    Device fabrications and measurements. The ReS₂ domains grown on MgO and α-Al2O3 substrates were transferred onto SiO2/Si (285 nm) substrates for the fabrication of FETs (Fig. 4a), while photodetectors (Fig. 4f, g) were fabricated directly on the chiral MgO substrates. Both FETs and photodetectors were constructed using a photolithographic pattern transfer technique. Electrodes (10 nm Cr/40 nm Au) were deposited on the substrate through standard photolithography, followed by thermal evaporation and a lift-off process. The prepared electrodes were then transferred onto the target sample using the aligned dry-transfer method. Electronic and optoelectronic measurements were performed on a probe station using a semiconductor device analyzer (Keithley 2612B). A linearly polarized beam (532 nm) with a defined polarization direction was focused onto the device through a 50× objective, with its intensity calibrated using a standard optical power meter (Thorlabs PM100D). CPL was generated using a quarter-wave plate (KungXuan K0.A225.532). All measurements were conducted in the air at room temperature.

DFT calculations

DFT calculations were performed using the plane-wave method implemented in the Vienna Ab initio Simulation Package45,46. The projected augmented wave pseudo-potentials47, the Perdew-Burke-Ernzerh exchange-correlation functional were adopted in all the calculations48. The energy cutoff was set as 400 eV for the plane-wave basis set for the sake of high accuracy. A vacuum layer of 20 Å was used to avoid the interference of the adjacent periodic images. For supercells featuring distinct geometries or sizes, Monkhorst–Pack k-point sampling was implemented with a grid spacing of ~ 0.03 Å−1 along each reciprocal lattice vector within the Brillouin zone. The energy and force convergence criteria were set to 10−4 eV and 0.01 eV/Å, respectively.

We constructed the ReS2-MgO(001) heterojunction with orientations varying by intervals of 15°(Supplementary Fig. 33 and Supplementary Table 3). The 0° orientation corresponds to the configuration where the b-axis of ReS2 is aligned parallel to the MgO[110] direction. The binding energies were calculated as

$${E}_{b}=\left({E}_{{{{{\rm{ReS}}}}}_{2}-{{{\rm{MgO}}}}(001)}-{E}_{{{{{\rm{ReS}}}}}_{2}}-{E}_{{{{\rm{MgO}}}}(001)}\right)/S$$
(1)

where the \({E}_{{{{{\rm{ReS}}}}}_{2}-{{{\rm{MgO}}}}(001)},{E}_{{{{{\rm{ReS}}}}}_{2}},{E}_{{{{\rm{MgO}}}}(001)}\) represent the total energy of ReS2 on MgO(001) substrate and the energy of isolated ReS2 and MgO(001) substrate, respectively. S indicates the contacting area between ReS2 and MgO(001) substrate. Then, we compared the grain boundary formation energies of 90° and 45° GBs using the following formula:

$${E}_{f}={E}_{{tot}}-{xE}({{{\mathrm{Re}}}{{{\rm{S}}}}}_{2})$$
(2)

where \({E}_{{tot}}\) is the total energy of the ReS2 with GBs, \(E({{{\mathrm{Re}}}{{{\rm{S}}}}}_{2})\) is the energy of ReS2 per unit (each unit here has one Re and two S atoms), x is the total number of Re atoms (number of units) in the structure of ReS2 with GBs.

Next, we built a ReS2-MgO(11l) heterojunction featuring step \(\left[\bar{1}10\right]\) (Supplementary Fig. 34 and Supplementary Table 4). We compared the binding energies between two orientations where the b-axis in ReS2 aligns parallel or perpendicular to the step \(\left[\bar{1}10\right]\). The binding energies were calculated as

$${E}_{b}=\left({E}_{{{{{\rm{ReS}}}}}_{2}-{{{\rm{MgO}}}}\left(11l\right)}-{E}_{{{{{\rm{ReS}}}}}_{2}}-{E}_{{{{\rm{MgO}}}}\left(11l\right)}\right)/2L$$
(3)

where \({E}_{{{{{\rm{ReS}}}}}_{2}-{{{\rm{MgO}}}}\left(11l\right)},{E}_{{{{{\rm{ReS}}}}}_{2}},{{{\mbox{and }}}}\,{E}_{{{{\rm{MgO}}}}\left(11l\right)}\) represent the total energy of ReS2 on the MgO(11l) substrate, the energy of isolated ReS2, and the energy of the MgO(11l) substrate, respectively. L indicates the length along the step \(\left[\bar{1}10\right]\).

Finally, we constructed ReS2-chiral MgO heterojunction, introducing both step \(\left[\bar{1}10\right]\) and kink \(\left[\bar{1}00\right]\), (Supplementary Fig. 35 and Supplementary Table 5). We compared the binding energies of ReS2 with different chirality on the chiral MgO surface, using the formula:

$${E}_{b}=\left({E}_{{{{{\rm{ReS}}}}}_{2}-{{{\rm{chiral\; MgO}}}}}-{E}_{{{{{\rm{ReS}}}}}_{2}}-{E}_{{{{\rm{chiral\; MgO}}}}}\right)/S$$
(4)

where the \({E}_{{{{{\rm{ReS}}}}}_{2}-{{{\rm{chiral\; MgO}}}}},{E}_{{{{{\rm{ReS}}}}}_{2}},{E}_{{{{\rm{chiral\; MgO}}}}}\) represent the total energy of ReS2 on chiral MgO substrate, the energy of isolated ReS2 and the energy of chiral MgO substrate, respectively. S denotes the contact area between ReS2 and chiral MgO substrate.