Significance
Manipulating materials with atomic-scale precision is essential for the development of a next-generation material design toolbox. Tremendous efforts have been made to advance the compositional, structural, and spatial accuracy of material deposition and patterning. Here, we presented a new reaction pathway to implement the conversions of two-dimensional materials within the atomic-layer thickness at room temperature for electrical dipole manipulation. Not only could various Janus monolayer transition metal dichalcogenides with vertical dipole be realized, but also some heterostructures, including the dipole-nondipole heterostructures (MoS2-MoSSe) and multiheterostructures (MoS2-MoSSe-MoSeS-MoSe2) within the same monolayer host structure are developed, in which the dipoles can be selectively patterned to be zero (MoS2, MoSe2), positive (MoSSe), and negative (MoSeS).
Keywords: 2D materials, Janus transition-metal dichalcogenides, atomic-layer substitution, room temperature, heterostructures
Abstract
Technology advancements in history have often been propelled by material innovations. In recent years, two-dimensional (2D) materials have attracted substantial interest as an ideal platform to construct atomic-level material architectures. In this work, we design a reaction pathway steered in a very different energy landscape, in contrast to typical thermal chemical vapor deposition method in high temperature, to enable room-temperature atomic-layer substitution (RT-ALS). First-principle calculations elucidate how the RT-ALS process is overall exothermic in energy and only has a small reaction barrier, facilitating the reaction to occur at room temperature. As a result, a variety of Janus monolayer transition metal dichalcogenides with vertical dipole could be universally realized. In particular, the RT-ALS strategy can be combined with lithography and flip-transfer to enable programmable in-plane multiheterostructures with different out-of-plane crystal symmetry and electric polarization. Various characterizations have confirmed the fidelity of the precise single atomic layer conversion. Our approach for designing an artificial 2D landscape at selective locations of a single layer of atoms can lead to unique electronic, photonic, and mechanical properties previously not found in nature. This opens a new paradigm for future material design, enabling structures and properties for unexplored territories.
The wide and rich physics of two-dimensional (2D) materials have led to the fabrication of heterostructures (1–5), superlattices (6–8), and twisted structures (9–11) with breakthrough discoveries and applications. During the past 3 y, a new class of atomic layer structures, in which the top layer chalcogen atoms within the three-atom-thick transition-metal dichalcogenides (TMDs) are replaced by a different type of chalcogen atoms, has been developed, resulting in a Janus monolayer with broken out-of-plane symmetry, inherit strain, and a variety of attractive properties (12–16). However, such a transition has been subjected to thermal energy input because the top chalcogen–metal bonds need to be broken in order for the substitution to occur. Previous strategies (12, 15, 16) usually break off the original chalcogen atoms by purely harsh energetic means and then form new bonds with the incoming replacements. Since the displacement energies of these chalcogen atoms in 2D TMDs range from 5 to 7 eV (17, 18), the process requires either knocking off the top chalcogens via the kinetic energy of atoms in a hydrogen (12) or Se plasma (16) or purely high temperature (to form sulfur vacancies) (15) to happen.
It is well known that thermal energy has been widely used to enable chemical reactions. However, high reaction temperatures incur severe limitations (for example, the types of substrate on which materials can be synthesized) and undesirable energy consumptions. There has been a continuing pursuit to reduce material synthesis temperatures while maintaining the quality of as-synthesized material (19–21). Recently, a room-temperature method to synthesis monolayer Janus WSSe and MoSSe by sulfurization from monolayer WSe2 and MoSe2 was reported (14). Nevertheless, the mechanism for such conversion was only hypothesized and has not been completely understood.
This paper reports our concurrent work for room-temperature atomic-layer substitution (RT-ALS), including the conversion of WSe2 and MoSe2 (2H phase) being sulfurized into Janus WSeS and MoSeS, WS2 and MoS2 (2H phase) being selenized into Janus WSSe and MoSSe, and the conversion of 1T’ MoS2 being selenized into Janus 1T’ MoSSe. Using density functional theory (DFT) calculations, we were able to elucidate the evolution of free energies in each reaction step and present a clear understanding of why such a reaction can occur at room temperature. More significantly, by combining RT-ALS with lithography patterning and flip 2transfer, a completely new class of monolayer lateral heterostructures and multiheterostructures are designed and fabricated.
Fig. 1A illustrates our RT-ALS process. The presence of hydrogen radicals (22) produced by a remote plasma strip the chalcogenide atoms on the top layer in a gentle and chemical way. Meanwhile, the low-pressure system facilitates the supply of vaporized chalcogen substitutes (here illustrated with Se) to take the place of missing atoms, resulting in an asymmetric Janus structure of MXY (M = Mo or W, X = S or Se, and Y = Se or S) at room temperature.
Fig. 1.
RT-ALS with DFT calculation and comparison. (A) Schematic illustration of the RT-ALS process from monolayer MoS2. (B) Schematics of the five key reaction steps for RT-ALS process (cartons from left to right): before H adsorption, 2 H adsorption and diffusion to the same S, formation of H2S, desorption of H2S, and Se occupation of the S vacancy. Purple, yellow, red, and green balls are Mo, S, Se, and H atoms, respectively. (C) Free energy of each step in B, relative to that of the first step. (D) Comparison of activation energy barrier between the RT-ALS strategy (red) and the conventional high-temperature substitution (blue). The axes of the two pathways correspond to distances (d) between the substituted S atom to Mo plane in each case. The vertical dashed lines (blue and red) indicate equilibrium S to Mo plane distances (1.54 Å for high T pathway, 1.64 Å for RT-ALS pathway). In RT-ALS process, the largest energy barrier occurs at a critical distance where H2S starts to form. In the high-temperature pathway, the largest energy barrier corresponds to when S and Se are both disconnect with Mo.
Results and Discussion
In sharp contrast to previous severe strategies (12, 15, 16), the RT-ALS steers the reaction via a different pathway. We use DFT (within the generalized gradient approximation, including van der Waals corrections [PBE-D2] on a supercell containing 4 MoS2 unit cells [larger supercells are included in the SI Appendix]) to examine two reaction pathways to illustrate the difference, as shown in Fig. 1 B–D and SI Appendix, Fig. S1D. One pathway is the conventional high-temperature substitution (replacing S by Se via thermal activation), and the other is our RT-ALS process. Due to the active roles of H radicals in the RT-ALS process, the energy barrier becomes much lower, and the process is overall exothermic. Fig. 1B exhibits that H radicals firstly adsorb on the upper surface of monolayer MoS2 with adsorption energy of ∼0.5 eV/H (which further increases with the H coverage, as shown in SI Appendix, Fig. S1A). As the H coverage increases, two H atoms form bonds with one S atom (SI Appendix, Fig. S1B), and the H2S desorbs, leaving an S vacancy. Surprisingly, this desorption process only involves a small activation barrier <0.5 eV (i.e., 12 kcal/mol), an order of magnitude smaller than the total bond energies of 3 Mo-S bonds (∼6 eV). In such a scenario, the removal of S is unlikely to be caused by collisions with hydrogen radicals because the kinetic energy of an H radical in the remote hydrogen plasma is about two orders of magnitude smaller than the displacement energy of a chalcogen atom (16), and the masses are very different between H and S atoms. The S vacancy is then occupied by a nearby Se atom. In contrast, the high-temperature pathway (SI Appendix, Fig. S1D) involves Se adsorption on surface, Mo-S bond breaking, and Mo-Se bond reforming, requiring an energy barrier of ∼2.5 eV that occurs when the Mo-S bond breaks (Fig. 1D). This calculated energy barrier matches well with the experimental findings, wherein the high-temperature substitution happens at ∼1,000 K (23). Such a calculation comparison clearly reveals that RT-ALS is made possible via reducing the reaction energy barrier, and the reaction kinetics is controlled by the active H radicals.
For RT-ALS experiments, we start with chemical vapor deposition (CVD)-synthesized monolayer MoS2 and perform selenization for the top sulfur layer to obtain MoSSe. Afterward, we use flip-transfer to expose the bottom surface and perform RT-ALS a second time to achieve MoSSe conversion to MoSe2 (SI Appendix, Fig. S2). From the optical microscopy (OM) images (SI Appendix, Fig. S2 B–D), the optical contrast of the MoS2, MoSSe, and converted MoSe2 triangular single crystals remain uniform and intact after RT-ALS, suggesting the process makes negligible disruption to the 2D lattice. This is also confirmed by atomic force microscopy (AFM) images (SI Appendix, Fig. S2 E–G), as their surfaces remain pristine and flat after each RT-ALS step. In addition, it is straightforward to extend the single/double-sided conversions to continuous large-area films (SI Appendix, Fig. S2 H–J). From a material synthesis perspective, here MoSe2 is obtained from a MoS2 template by room-temperature reaction, without the need of high-temperature conversions as in contrast to earlier approaches (21, 24), and thus avoids subsequent lattice or substrate damage. This provides an alternative route to obtain Se-based TMDs at room temperature. More importantly, this could be used to synthesize TMD materials that are hard to be directly grown by CVD or other methods.
To confirm the substitution of the top chalcogenide layer, measurements on Raman and photoluminescence (PL) spectroscopy are performed due to their sensitivity to the crystal structure. For TMDs, the A1g and E2g mode correspond to the out-of-plane and in-plane lattice vibrations, respectively. Therefore, changing the top layer of chalcogens produces different phonon frequencies for the same type of lattice vibrations. As shown in Fig. 2A, the MoS2 A1g (404 cm−1) and E2g (383 cm−1) modes shift to 288 cm−1 and 355 cm−1, respectively, in Janus MoSSe12 due to the change of atomic mass and the broken symmetry in the vertical direction (12). Further flip-transfer and RT-ALS on the other sulfur layer yield full selenization, with a sharp peak at ∼239 cm−1 and a broad one at ∼284 cm−1, which are consistent with the A1g and E2g modes in monolayer MoSe2 (25). These conversions are further evidenced by the strong PL emission shown in Fig. 2B, in which the PL peak energy shifts from 1.85 eV (pristine MoS2) to 1.72 eV (Janus MoSSe) and then to 1.60 eV (converted MoSe2). All the observed PL peak energies are consistent with reported values of previous literature (12, 25). Besides, Raman mappings (SI Appendix, Fig. S3 A–C) and PL intensity mappings (Fig. 2B, Insets) also confirm that crystals produced by RT-ALS have high spatial homogeneity.
Fig. 2.
Characterizations of the monolayer (Janus) TMDs converted using RT-ALS and the conversion process. (A) Raman spectra of starting monolayer MoS2, Janus MoSSe, and converted MoSe2. (Insets) Crystal structures of MoS2, Janus MoSSe, and MoSe2, respectively. (B) PL spectra of starting monolayer MoS2, Janus MoSSe, and converted MoSe2. (Insets) Spatially resolved PL mappings of MoS2, Janus MoSSe, and converted MoSe2 flakes at 1.85 eV, 1.72 eV, and 1.60 eV, respectively. (C) Raman spectra of starting monolayer WSe2, Janus WSeS, and converted WS2. (D) PL spectra of starting monolayer WS2, Janus WSSe, and converted WSe2. (Insets) Crystal structures of WS2, Janus WSSe, and WSe2, respectively. (E) Tilted ADF-STEM image of a MoSSe sample to visualize the asymmetric atom structure in the vertical direction. The corresponding Mo, Se, and S atoms are schematically shown with blue, red, and yellow circles, respectively. (F) Intensity profile for the atomic chain highlighted in blue in E shows the intensity of individual Mo, Se, and S atoms. (G) Raman spectra collected at several monolayer Janus MoSSe samples treated with hydrogen plasma for 10 s, 30 s, 1 min, 3 min, 5 min, 10 min, and 15 min, respectively. (H) The relationship between the conversion time (10 s, 30 s, 1 min, 3 min, 5 min, and 10 min, respectively) and the corresponding intensity IA1g(MoSSe)/IA1g(MoS2) ratio over the conversion process. (Inset) Spatially resolved Raman mappings of two coalesced Janus MoSSe flakes converted within 30 s and 5 min, respectively, which show uniform Raman intensity over the whole MoSSe regions.
This RT-ALS is found to be a general strategy to obtain various Janus materials (including different transition metal atoms, chalcogen atoms, for converting only top layer or both top and bottom layers, and different crystal phases) through our experimental investigations. The conversions from WS2 to Janus WSSe and then to WSe2 were successfully obtained (SI Appendix, Figs. S2 K–M and S3 G–I), with the corresponding PL emissions shift from 2.01 eV (WS2) to 1.83 eV (WSSe) and to 1.63 eV (WSe2) (Fig. 2D). Compared with MoSSe, the larger elastic modulus and smaller carrier effective mass of WSSe may give rise to higher carrier mobility, especially for holes, due to the stronger spin-orbit coupling and overall cleaner optical signatures of W (16, 26). In addition, the conversion from monolayer MSe2 (M = Mo and W) to Janus MSeS and then to MS2 are realized by sulfurization (Fig. 2C and SI Appendix, Fig. S3 D–F) as well as the conversion from 1T’ MoS2 to Janus 1T’ MoSSe (SI Appendix, Fig. S3 J and K). The evolutions of Raman peaks are all in consistent with literature (12, 16, 27) and our theoretical predictions, suggesting the universality of RT-ALS strategy.
In order to verify the atomic structure of Janus MoSSe, we used annular dark-field scanning transmission electron microscopy (ADF-STEM) to obtain the tilted images of MoSSe. The image was obtained by rotating the MoSSe sample 10° from the [001] crystallographic orientation to a new orientation that resulted in the Mo, Se, and S atoms projected in a straight line. In that way, the Mo, Se, and S atoms can be individually identified. As shown in Fig. 2E, the tilted ADF-STEM image enables to separate the Se and S atoms. The results show that the Se atoms are located on one side of the monolayer MoSSe, and S atoms on the opposite side, which is direct evidence of the Janus structure. The corresponding intensity profile in Fig. 2F clearly show the individual Mo, Se and S atoms with their total peak intensities proportional to their atomic numbers.
To achieve a better understanding of the conversion process, we investigated the evolution of the Raman peaks during the synthesis. The results indicated in Fig. 2 G and H, and SI Appendix, Fig. S4A show that within 10 s, the A1g peak of Janus MoSSe begins to emerge [at a frequency ∼10 cm−1 lower than the final point, most likely due to the modulation of phonon frequency from the local phonon vibrations at the reaction onset (26)]. The intensity of the Janus MoSSe A1g mode increases with longer conversion time; meanwhile, the intensity of the MoS2 A1g peak decreases and eventually vanishes in 15 min. Based on the variation of Raman peaks at different time, the conversion yield (roughly defined by Y% = Intensity A1g of MoSSe/[Intensity A1g of MoSSe + Intensity A1g of MoS2]) from S to Se after RT-ALS treatment are estimated (SI Appendix, Fig. S4B). For material conversion through high-temperature substitution, previous reports (21) showed that the atomic replacement begins at energetically favorable sites (crystal edges, defect sites, and grain boundaries) and gradually progresses to the other regions of the flakes, indicating it is controlled by thermodynamics. In contrast, the room temperature operation here suggests it is a kinetic process, in which the energy needed for the atomic substitution is supplied by the highly reactive hydrogen radicals. This leads to the spatially homogeneous substitution of atoms over the whole flake since the reaction onset (Fig. 2H, Insets and SI Appendix, Fig. S4C) and a fast progression of the reaction with a time scale as short as 30 s (Fig. 2G and SI Appendix, Fig. S4A).
Taking advantage of the room temperature procedure of the whole process, the programmable design in monolayer TMD can be realized using lithography combined with flip-transfer techniques, as illustrated in Fig. 3 A–C. Arbitrary patterns in a 2D plane can be implemented by using a physical mask (here, lithographic patterns is just an example), enabling locally programmed out-of-plane atomic structures (Fig. 3A, where electron-beam lithography [EBL] can define arbitrary patterns using a polymethyl methacrylate [PMMA] mask). RT-ALS will convert the exposed areas to MSSe, allowing the formation of lateral heterostructure between MS2 and Janus MSSe within the same host monolayer (Fig. 3B). Furthermore, the chalcogen on the other side can be replaced by using a flip-over transfer (Methods) and another selective RT-ALS conversion, resulting in lateral multiheterostructures of MS2-MSSe-MSeS-MSe2 (Fig. 3C) with tailored functionalities. Fig. 3D shows the scanning electron microscopy (SEM) image of monolayer Janus MoSSe-MoS2 stripe patterns with different widths. Even though both MoSSe and MoS2 are semiconductors with comparable bandgaps, MoSSe is brighter than MoS2, possibly due to the intrinsic vertical dipole in the monolayer. Fig. 3E shows the spatially resolved PL mapping of Janus MoSSe with the pattern of the Massachusetts Institute of Technology mascot “Tim the beaver” made on a continuous monolayer CVD-MoS2 canvas (RT-ALS converted the exposed regions to Janus MoSSe, while the masked regions remained as pristine MoS2). The same pattern was characterized using various methods: OM image (SI Appendix, Fig. S5A), SEM image (SI Appendix, Fig. S5B), and Raman and PL mappings (SI Appendix, Fig. S5 C and D), which all clearly verify that MoS2 and MoSSe are well-located as designed. ADF-STEM is used to characterize the junction, which shows that the MoS2-MoSSe interface is seamlessly connected at the boundary defined by lithography (Fig. 3F and SI Appendix, Fig. S6A). Elemental diffusion (2, 4) is not anticipated for the RT-ALS strategy, as the substitution occurs at room temperature within a short time. As ADF-STEM intensity scales with the atomic number, the MoS2 region with weaker intensity from the S-S column on the left side of Fig. 3F is clearly distinguished from the stronger intensity from the S-Se column of MoSSe on the right. Additionally, the different structures of MoS2 and MoSSe are further verified by the intensity profile of electron scattering, where the atomic positions of Mo, S/Se, and 2S are corresponding to their gradually decreased intensities (Fig. 3 G and H). The energy-dispersive X-ray spectroscopy results (SI Appendix, Fig. S6 C–H) indicate that Se and S atoms are well-confined in the selective regions. As the selected area electron diffraction patterns show nonseparable patterns from these two lattices (SI Appendix, Fig. S7 A–D), real space strain mappings were further performed (SI Appendix, Fig. S7 G and H), in which the slight lattice difference in the MoSSe region can be observed compared to the reference MoS2 region.
Fig. 3.
RT-ALS with programmable design in dipole/nondipole lateral heterostructures and multiheterostructure. (A) Schematic representation of programmable ALS process by lithography patterning and ALS process. (B) Programmable monolayer MoS2-Janus MoSSe heterostructure on SiO2/Si substrate. (C) A multiheterostructure composed of monolayer MoS2-Janus MoSSe-Janus MoSeS-MoSe2 regions. (D) SEM images of monolayer Janus MoSSe-MoS2 stripe patterns with different widths. (E) Spatially resolved PL mapping for a Massachusetts Institute of Technology mascot “Tim the beaver” pattern (Janus MoSSe) on a monolayer MoS2 canvas. (F) Pseudocolor ADF-STEM image of a MoS2-MoSSe interface (Left), the zoomed-in views of a MoS2 region (Right Top, highlighted in yellow) and a Janus MoSSe region (Right Bottom, highlighted in blue), and their respective intensity profiles in G and H. (I) OM image of monolayer multijunction composed with MoS2-Janus MoSSe-Janus MoSeS-MoSe2. (J) AFM topography image for the monolayer multiheterostructure of MoS2-Janus MoSSe-Janus MoSeS-MoSe2. (K) SEM image of monolayer multiheterostructure composed with MoS2-Janus MoSSe-Janus MoSeS-MoSe2. (L) Spatially resolved Raman mapping for A1g mode intensity of monolayer multiheterostructures based on MoS2-Janus MoSSe-Janus MoSeS-MoSe2. The red area represents MoS2 region, the blue area represents Janus region, and the green area represents MoSe2 region.
Flipping over the MoS2-MoSSe heterostructures and repeating selective RT-ALS on the other side yielded lateral multiheterostructures with MoS2, Janus MoSSe and MoSeS, and MoSe2 (Fig. 3 I–L and SI Appendix, Fig. S8). From the OM image in Fig. 3I, it can be seen that Janus MoSSe and Janus MoSeS has very similar optical contrast, while the AFM image Fig. 3J indicates the slight height increase after each RT-ALS step, consistent with the observation from SI Appendix, Fig. S2 E–G. Fig. 3K shows the SEM of the multiheterostructure regions. The two types of Janus regions (MoSSe [having vertical dipole pointing up] and MoSeS [having vertical dipole point down]) demonstrates clear intriguing contrast under SEM. These two types of regions should have very similar properties (e.g., material composition, bandgap, etc.) except the dipole direction, such observation indicates the role played by the dipole moment in material functionalities. In Fig. 3L, spatially resolved Raman mapping for A1g mode intensity of such monolayer multiheterostructures is presented, with the red region being MoS2, blue region being Janus (either MoSSe or MoSeS), and green region being converted MoSe2.
In principle, the minimum feature size of RT-ALS should be determined by the resolution of lithography (Fig. 3D and SI Appendix, Fig. S5 E–H). With current EBL, features of hundreds of nanometers can be achieved, with potential for scaling down to tens of nanometers using extreme ultraviolet, helium-beam, or AFM-based lithography (28). The interface revealed by ADF-STEM (Fig. 3F) indicates edge roughness due to EBL limitations. However, sharp edges are anticipated if atomically sharp physical masks such as aligned carbon nanotubes (29, 30) can be used in the future. Nevertheless, even with the current EBL, many new device structures with unusual physical properties and potential applications can already be envisioned. For example, periodic arrays of these planar heterostructures with 100-nm-scale pitch can be defined as an atomically thin optical grating, and metasurfaces for efficient modulation or steering of visible/near-infrared light can be designed. In these structures, the extremely strong light-matter interactions in 2D materials and the optical nonlinearity in Janus TMD (31, 32) could play critical roles in expanding the design space of these atomically thin optical metasurfaces with unprecedented efficiencies or novel functionalities. The ability to create these artificial TMD multiheterostructures within a continuous atomic-layer “canvas” has never been achieved before and presents extremely exciting potentials.
A major application of 2D materials and their heterostructures lies in novel electronic or optoelectronic devices. With RT-ALS, we found that high-quality materials are obtained, and the energy misalignment of band edges exist at both the MoSSe-MoS2 and MoSe2-MoSeS lateral heterojunctions with a type-II band alignment. Multiple back-gate transistors with long and short electrodes were fabricated and four-probe measurements were used to eliminate the contact resistance (OM images in Fig. 4 B and C, Insets). Typical transfer characteristics (channel current ID versus back-gate voltage VG) are plotted in Fig. 4A and used to extract the threshold voltage VT, defined as the intersection of the linear fit of the on-current and the x-axis, and the field effect mobility µFE, defined as µFE = (L/W)gm/(VDCox), where L and W are the channel length and channel width, gm is the transconductance, VD is the source-drain voltage (bias across the channel), and Cox = 12.1 nF/cm2 (capacitance of the 285 nm SiO2 gate oxide). The VT and µFE of tens of devices were measured for the starting CVD MoS2, single-side-converted Janus MoSSe, as well as double-side-converted MoSe2, summarized in Fig. 4A, Inset. Average µFE for MoS2, Janus MoSSe, and ALS MoSe2 are 8.94 cm2/Vs, 2.11 cm2/Vs, and 0.17 cm2/Vs, respectively. Average VT for them are −1.9 V, 36.0 V, and 51.6 V, respectively. The gradual degradation of mobility from MoS2 to MoSe2 may arise from the introduction of defects through the RT-ALS or transfer processes. Note that the mobility value measured in our Janus MoSSe devices is two orders of magnitude higher than through the route of conventional high-temperature substitution (15), which is indicative of the high quality of materials from RT-ALS. Furthermore, the increasing of VT values indicates that MoS2 is the most n-type doped, whereas Janus MoSSe and MoSe2 are decreasingly n-doped, in agreement with DFT calculated band alignment (Fig. 4F). This band alignment configuration further determines the electrical polarity of Janus MoSSe-MoS2 and MoSe2-Janus MoSeS lateral heterojunctions. The Janus MoSSe and the converted MoSe2 regions are the anodes of these two n-n+ junctions, respectively, and confirmed with output characteristics (ID versus VM with different VG, where VM is the voltage drop across the two inner short electrodes [M1 and M2] as ID is applied across the two long electrodes [D and S]), as shown in Fig. 4 B and C. The weak rectification behaviors observed on both lateral junctions suggest that the barrier heights (denoted as ΦB) at the lateral heterojunctions are very low. We performed temperature-dependent four-probe transport measurement and used the thermionic emission model to extract the gate-dependent barrier heights (see Methods for details, and SI Appendix, Figs. S9 and S10), plotted in Fig. 4D, for ΦB = 30 meV and 50 meV for MoSSe-MoS2 and MoSe2-MoSeS lateral heterojunctions, respectively, when the channels are turned on completely (VG > 50 V).
Fig. 4.
Electrical properties of MoS2, Janus MoSSe, MoSe2, and their lateral heterostructures produced by ALS. (A) Transfer characteristics (current density ID/W versus back-gate voltage VG, VD = 1 V) of typical MoS2, single-side-converted Janus MoSSe, and double-side-converted MoSe2. (Inset) Field effect mobility µFE versus threshold voltage VT for tens of devices. The average mobility values for MoS2, Janus MoSSe, and MoSe2 are 8.94, 2.11, and 0.17 cm2/Vs, and the average VT for them are −1.9, 36.0, and 51.6 V, respectively. (B and C) Four-probe output characteristics (current density ID/W versus bias voltage VM with various back-gate voltage VG) for (A) MoSSe-MoS2 and (B) MoSeS-MoSe2 lateral heterojunctions. (Insets) Diagrams and OM images of these devices. (D) Barrier height ΦB as a function of VG for the MoSSe-MoS2 (red) and the MoSeS-MoSe2 (orange) lateral heterojunctions. (E) Kelvin probe force microscope image of the MoS2-MoSSe-MoSeS-MoSe2 multiheterostructure. The Janus MoSSe region with the Se layer on top and the MoSeS region with the S layer on top display the highest and the lowest surface potential, respectively, whereas the surface potentials for the MoS2 and MoSe2 regions are in between. (F) Conduction band minima and valence band maxima energies of MoS2, Janus MoSSe, and MoSe2 relative to vacuum, obtained from DFT calculations. For Janus MoSSe, the vacuum energy levels from upper and lower planes are averaged.
More uniquely, our RT-ALS approach brings about a programmable control of the out-of-plane electrical dipole in TMD materials (induced by the electronegativity difference between S and Se atoms in Janus MoSSe). We performed Kelvin probe force microscope measurements on the MoS2-MoSSe-MoSeS-MoSe2 grid samples (Fig. 4E) to observe clear surface potential differences among these four distinct materials. As expected, the Janus MoSSe region with the Se layers on top and the flipped MoSeS region with the S layers on top display the highest and the lowest surface potentials, respectively, whereas the surface potentials for the MoS2 and MoSe2 regions are in between. The measured potential difference between the top Se layer of the MoSSe region and the top S layer of the MoSeS region is around 100 meV. Such potential difference is even larger than that between the MoSe2 and MoS2 regions (which is around 15 meV), suggesting that the Janus MoSSe (or MoSeS) holds a large intrinsic dipole that is locked to the 2D surface normal. Note that this dipole can be selectively patterned to be zero (MoS2, MoSe2), positive (MoSSe), and negative (MoSeS) on such a 2D material “canvas,” which would enable many nanostructures and devices with intriguing electrical and optoelectronic properties. For example, with proper choice of physical masks, such as e-beam resist, or carbon nanotubes (28, 29) with atomically sharp edges, it is possible to make in-plane quantum wells, superlattices, or photonic devices by generating periodic arrays of such dipole/nondipole lateral heterostructures with 1- to 100-nm periodicity. Not only can the transport and optical properties of these MoXY multiheterostrutures be altered, any materials in close vicinity would also be modulated by such an electrostatic “canvas.” Unusual physical properties and unprecedented functionality may be possible on such a platform, ranging from nonlinear optics, electronic band engineering, and nanoscale origami to electrochemical catalysis, all of which can be modulated by different electrostatic forces.
In summary, the RT-ALS combined with patterning and flip transfer present a powerful yet universal strategy to program material properties at the atomic-layer limit. It allows for creating diverse artificial low-symmetry 2D materials and their heterostructures. Such a process generates seamless, high-quality interfaces between different structures and can complement existing vdW heterostructure fabrication techniques by adding another intriguing class of materials (Janus and lateral Janus heterostructures and multiheterostructures). Furthermore, by designing these heterostructure patterns and surface charge distributions, great potential is anticipated in new physical discoveries and future applications.
Methods
Synthesis of MoS2.
A target SiO2 substrate was suspended between two other SiO2/Si substrates with predeposited perylene-3, 4, 9, 10-tetracarboxylic acid tetrapotassium solution. All of these substrates were placed facedown on a crucible containing MoO3 precursor in a 1-inch quartz tube. This crucible was placed in the middle of the heating zone with another sulfur crucible on the upstream. Before heating, the whole CVD system was purged with 1,000 sccm Ar (99.999% purity) for 3 min. Then, 20 sccm Ar was introduced into the system as a carrier gas. The growth system was heated to 625 °C for 15 min. The MoS2 growth was carried out around 620 to 630 °C for 3 min under atmospheric pressure. After growth, the whole system was naturally cooled down to room temperature.
Plasma-Assisted ALS.
We use a remote commercial inductively coupled plasma system to substitute the top-layer sulfur atoms of monolayer MS2 (M = Mo, W) with selenium. The CVD grown monolayer MoS2/WS2 were placed in the middle of a quartz tube. The plasma coil made by a cylindrical copper tube placed at the upstream of CVD furnace. The distance between the sample and the plasma coil is around 8 to 10 cm, with the selenium powder placed around 5 cm away from the plasma generator on the other side. Whenever masks were needed, the sample was covered with PMMA patterns defined by EBL. At the beginning of the process, the whole system was pumped down to a low pressure to remove air in the chamber. Then, hydrogen was introduced into the system and the plasma generator was ignited. The hydrogen atoms assist the removal of the sulfur atoms on the top layer of MoS2. At the same time, the vaporized selenium filled in the vacancy of the sulfur atoms, resulting in the asymmetric Janus structure of MoSSe. The whole process was performed at room temperature. After the reaction, the whole system was purged with Ar gas (99.999% purity) to remove the residual reaction gas, and the pressure was recovered to atmospheric.
Transfer.
For normal transfer, the samples were spin-coated with PMMA as a supporting layer. Then, they were put in the KOH solution and the PMMA/2D material was detached from the growth substrate and was floated on the surface of KOH solution. Afterward, it was taken to a deionized (DI) water bath by a glass slide and washed several times and then picked up with the target substrate. After baking on a hot plate at 100 °C for 15 min, the PMMA layer was removed with acetone and isopropanol (IPA). For flip-over transfer, the premade MoSSe sample was picked up by a polydimethylsiloxane stamp and then was released onto an intermediate substrate. Then, the MoSSe was released through etching the SiO2 in a KOH solution. The film was washed with DI water for several times and then put upside-down on the target SiO2/Si substrate, with the bottom-layer selenium atoms touching the substrate. At last, the PMMA film was removed by acetone and IPA.
Supplementary Material
Acknowledgments
The preliminary experiments of this work are supported by the Air Force Office of Scientific Research under the Multidisciplinary University Research Initiative (MURI)-FATE program, Grant No. FA9550-15-1-0514. The characterization of the Janus Materials at a later stage was supported by the US Department of Energy (DOE), Office of Science, Basic Energy Sciences under Award DE‐SC0020042. Y.L. and T.P. acknowledge the US Army Research Office through the Institute for Soldier Nanotechnologies under Cooperative Agreement No. W911NF-18-2-0048 and the Science-Technology Center (STC) for Integrated Quantum Materials, NSF Grant No. DMR 1231319. P.-C.S. and A.-Y.L. acknowledge the funding from the Center for Energy Efficient Electronics Science (NSF Award No. 0939514) and the US Army Research Office through the Institute for Soldier Nanotechnologies at Massachusetts Institute of Technology, under Cooperative Agreement No. W911NF-18-2-0048. K.X. and T.C. are partially supported by NSF through the University of Washington Materials Research Science and Engineering Center Grant No. DMR-1719797. K.X. acknowledges support by the state of Washington through the University of Washington Clean Energy Institute. B.Y. and Y.Y. acknowledge the funding from Natural Science Foundation of China (Grant No. 21805184), NSF Shanghai (Grant No. 18ZR1425200), and the Center for High-resolution Electron Microscopy at ShanghaiTech University (Grant No. EM02161943). C.S. and J.W. acknowledge support through US Army Research Office under Grant No. W911NF-18-1-0431. Q.J. acknowledges support from the STC Center for Integrated Quantum Materials, NSF Grant No. DMR 1231319. S.H. and K.Z. acknowledge financial support from NSF (ECCS-1943895). L.D. acknowledges support from the US Department of Defense, Office of Naval Research (Grant No. N00014-19-1-2296). E.S. acknowledges support from the Davidson School of Chemical Engineering of Purdue University. J.L. and C.S. acknowledge support from an Office of Naval Research MURI (Grant No. N00014-17-1-2661). The crystallographic tilted STEM image research was conducted at the Center for Nanophase Materials Sciences, which is a DOE Office of Science User Facility (J.-C.I.). We thank X. Zhang, Y. Han, G. Cheng, N. Yao, and N. Yan for helpful discussions.
Footnotes
The authors declare no competing interest.
This article is a PNAS Direct Submission. K.F.M. is a guest editor invited by the Editorial Board.
This article contains supporting information online at https://www.pnas.org/lookup/suppl/doi:10.1073/pnas.2106124118/-/DCSupplemental.
Data Availability
All study data are included in the article and/or SI Appendix. All materials are available upon request to J.K.
References
- 1.Gong Y., et al., Vertical and in-plane heterostructures from WS2/MoS2 monolayers. Nat. Mater. 13, 1135–1142 (2014). [DOI] [PubMed] [Google Scholar]
- 2.Huang C., et al., Lateral heterojunctions within monolayer MoSe2-WSe2 semiconductors. Nat. Mater. 13, 1096–1101 (2014). [DOI] [PubMed] [Google Scholar]
- 3.Li M. Y., et al., NANOELECTRONICS. Epitaxial growth of a monolayer WSe2-MoS2 lateral p-n junction with an atomically sharp interface. Science 349, 524–528 (2015). [DOI] [PubMed] [Google Scholar]
- 4.Sahoo P. K., Memaran S., Xin Y., Balicas L., Gutiérrez H. R., One-pot growth of two-dimensional lateral heterostructures via sequential edge-epitaxy. Nature 553, 63–67 (2018). [DOI] [PubMed] [Google Scholar]
- 5.Xiang R., et al., One-dimensional van der Waals heterostructures. Science 367, 537–542 (2020). [DOI] [PubMed] [Google Scholar]
- 6.Zhang Z., et al., Robust epitaxial growth of two-dimensional heterostructures, multiheterostructures, and superlattices. Science 357, 788–792 (2017). [DOI] [PubMed] [Google Scholar]
- 7.Xie S., et al., Coherent, atomically thin transition-metal dichalcogenide superlattices with engineered strain. Science 359, 1131–1136 (2018). [DOI] [PubMed] [Google Scholar]
- 8.Wang C., et al., Monolayer atomic crystal molecular superlattices. Nature 555, 231–236 (2018). [DOI] [PubMed] [Google Scholar]
- 9.Cao Y., et al., Unconventional superconductivity in magic-angle graphene superlattices. Nature 556, 43–50 (2018). [DOI] [PubMed] [Google Scholar]
- 10.Cao Y., et al., Correlated insulator behaviour at half-filling in magic-angle graphene superlattices. Nature 556, 80–84 (2018). [DOI] [PubMed] [Google Scholar]
- 11.Ribeiro-Palau R., et al., Twistable electronics with dynamically rotatable heterostructures. Science 361, 690–693 (2018). [Google Scholar]
- 12.Lu A. Y., et al., Janus monolayers of transition metal dichalcogenides. Nat. Nanotechnol. 12, 744–749 (2017). [DOI] [PubMed] [Google Scholar]
- 13.Berry J., Ristić S., Zhou S., Park J., Srolovitz D. J., The MoSeS dynamic omnigami paradigm for smart shape and composition programmable 2D materials. Nat. Commun. 10, 5210 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Trivedi D. B., et al., Room-temperature synthesis of 2D Janus crystals and their heterostructures. Adv. Mater. 32, e2006320 (2020). [DOI] [PubMed] [Google Scholar]
- 15.Zhang J., et al., Janus monolayer transition-metal dichalcogenides. ACS Nano. 11, 8192–8198 (2017). [DOI] [PubMed] [Google Scholar]
- 16.Lin Y.-C., et al., Low energy implantation into transition-metal dichalcogenide monolayers to form Janus structures. ACS Nano. 14, 3896–3906 (2020). [DOI] [PubMed] [Google Scholar]
- 17.Komsa H.-P., et al., Two-dimensional transition metal dichalcogenides under electron irradiation: Defect production and doping. Phys. Rev. Lett. 109, 035503 (2012). [DOI] [PubMed] [Google Scholar]
- 18.Ghorbani-Asl M., Kretschmer S., Spearot D. E., Krasheninnikov A. V., Two-dimensional MoS2 under ion irradiation: From controlled defect production to electronic structure engineering. 2D Mater. 4, 025078 (2017). [Google Scholar]
- 19.Li Z., et al., Low-temperature growth of graphene by chemical vapor deposition using solid and liquid carbon sources. ACS Nano. 5, 3385–3390 (2011). [DOI] [PubMed] [Google Scholar]
- 20.Chiang W. H., Sankaran R. M., Synergistic effects in bimetallic nanoparticles for low temperature carbon nanotube growth. Adv. Mater. 20, 4857–4861 (2008). [Google Scholar]
- 21.Yun S. J., et al., Telluriding monolayer MoS2 and WS2 via alkali metal scooter. Nat. Commun. 8, 2163 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Phillips J. A., Charge equilibrium ratios for hydrogen ions from solids. Phys. Rev. 97, 404 (1955). [Google Scholar]
- 23.Su S.-H., et al., Controllable synthesis of band-gap-tunable and monolayer transition-metal dichalcogenide alloys. Front. Energy Res. 2, 27 (2014). [Google Scholar]
- 24.Mahjouri-Samani M., et al., Patterned arrays of lateral heterojunctions within monolayer two-dimensional semiconductors. Nat. Commun. 6, 7749 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Wang X., et al., Chemical vapor deposition growth of crystalline monolayer MoSe2. ACS Nano. 8, 5125–5131 (2014). [DOI] [PubMed] [Google Scholar]
- 26.Xia C. X., et al., Universality of electronic characteristics and photocatalyst applications in the two-dimensional Janus transition metal dichalcogenides. Phys. Rev. B 98, 165424 (2018). [Google Scholar]
- 27.Liu L., et al., Phase-selective synthesis of 1T’ MoS2 monolayers and heterophase bilayers. Nat. Mater. 17, 1108–1114 (2018). [DOI] [PubMed] [Google Scholar]
- 28.Balčytis A., et al., “High precision fabrication of antennas and sensors” in Proceedings of the Ninth International Symposium on Precision Engineering Measurement and Instrumentation Vol. 9446, Cui J., Tan J., Wen X., Eds. (International Society for Optics and Photonics, Changsha/Zhangjiajie, China, 2014), 94461G-1 to 94461G-11. [Google Scholar]
- 29.Liu L., et al., Aligned, high-density semiconducting carbon nanotube arrays for high-performance electronics. Science 368, 850–856 (2020). [DOI] [PubMed] [Google Scholar]
- 30.He X., et al., Wafer-scale monodomain films of spontaneously aligned single-walled carbon nanotubes. Nat. Nanotechnol. 11, 633–638 (2016). [DOI] [PubMed] [Google Scholar]
- 31.Riis-Jensen A. C., Deilmann T., Olsen T., Thygesen K. S., Classifying the electronic and optical properties of Janus monolayers. ACS Nano 13, 13354–13364 (2019). [DOI] [PubMed] [Google Scholar]
- 32.Palsgaard M., Gunst T., Markussen T., Thygesen K. S., Brandbyge M., Stacked Janus device concepts: Abrupt pn-junctions and cross-plane channels. Nano Lett. 18, 7275–7281 (2018). [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
All study data are included in the article and/or SI Appendix. All materials are available upon request to J.K.