Selective molecular transport through nanoporous membranes forms the basis of many important separation and purification processes. In principle, two-dimensional (2D) materials such as graphene are ideal for preparing high-performance separation membranes. This is because they are extremely thin, thereby maximizing the molecular transport efficiency in the presence of nanopores1-5. Theoretical studies have suggested that membranes based on 2D materials could be considerably superior to the conventional polymeric membranes in terms of permeability and selectivity in water desalination and gas separation6-8. However, it is difficult to generate nanopores with precise molecular dimensions on 2D materials, especially over large areas as required for practical applications.
Various methods, including ion or electron irradiation3,4,9 and chemical or plasma etching2,10,11, have been used to create nanopores on 2D materials. However, these methods offer poor control of pore size and pore density. The obtained membranes typically have a nonuniform distribution of pore size, resulting in low permeability or poor selectivity when used for molecular separation9,12,13.
Compared with these posttreatment methods, direct generation of nanopores during the synthesis of 2D materials is more promising for achieving precise pore size control and high pore density. The synthesis of 2D materials via chemical vapor deposition (CVD) typically involves the simultaneous growth of multiple grains having different orientations. As the growth progresses, the grains are connected through grain boundaries (GBs) to form an extensive polycrystalline thin film. The formation of GBs in 2D materials induces structural inhomogeneity, which usually adversely affects the properties that single-crystalline structures are expected to possess. However, the presence of local structures at GBs may also create new opportunities for specific applications. For example, eight-membered rings (8-MRs) formed at graphene GBs can act as pores that allow Li+ to pass through14, whereas the perfect graphene lattice comprising 6-MRs is impermeable to Li+. Although 8-MRs at graphene GBs have demonstrated considerable potential for ionic transport, such pores are randomly produced by the uncontrolled growth of graphene, thereby limiting the pore density and, in turn, the permeability of the membrane. The GB structure is primarily determined by the orientation relation of adjacent grains, with the number of pores being directly proportional to the GB density. Hence, controlling the grain orientation and size during the growth of 2D materials is crucial for optimizing the desired pore structure on the GB.
Herein, we intentionally generated a large number of GBs with predesigned porous structures in monolayer molybdenum sulfide (MoS2) for efficient molecular separation. This was achieved by modifying our recently developed CVD method15 to precisely control the orientation and size of MoS2 grains. The MoS2 grains grown by this method typically exhibit anti-parallel orientations with a 60 degree relationship, resulting in several predictable GB structures. Among these structures, the 8-MR is dominant. The size of 8-MR in monolayer MoS2 exactly matches the requirement to enable rapid water transport while blocking various hydrated ions. Hence, the prepared films exhibit high water flux and water/ion selectivity when used as separation membranes. The number of 8-MRs can be increased drastically by reducing the average grain size, affording optimized membrane performance that exceeds that of state-of-the-art membranes for osmotically driven water/ion separation. Conversely, monolayer MoS2 prepared without controlling the grain orientation has less well-defined GB structures with fewer 8-MRs (mainly comprising smaller impermeable rings), thus exhibiting lower water flux and water/ion selectivity.
Grain boundary engineering in monolayer MoS2
A single-crystalline defect-free monolayer MoS2 comprising 6-MRs is impermeable to any atoms or molecules16. Structural defects with ring sizes exceeding the intrinsic 6-MR can provide channels for selective molecular transport. However, to achieve high selectivity, the generated structural defects should have appropriate channel diameters (i.e., ring sizes), depending on the requirements of specific applications. Molecular dynamics (MD) simulations were performed to explore appropriate MoS2 ring sizes for water/ion separation based on the density functional theory (DFT)–optimized structural models (Supplementary Fig. 1). The results revealed that 7-MR is still too small to transport water molecules or ions (Fig. 1a), whereas 8-MR exhibits the desired pore size (~4.2 × 2.4 Å), allowing fast water permeation while completely rejecting hydrated Na+ and Cl– ions (Fig. 1b). The MD simulations also revealed the single-file transport of water in MoS2 8-MRs. The average number of hydrogen bonds (H-bonds) per water molecule in 8-MRs was considerably lower than that of bulk water (1.6 vs. 3.3, Fig. 1c), similar to water transport through protein aquaporin channels17,18.
Notably, 8-MR is a typical defect structure of polycrystalline monolayer MoS2, which is commonly found at the boundaries between two grains with a 60° orientation relationship19,20. To gain insights into this phenomenon, we performed DFT calculations to compare the formation energies of the GB structures that were previously observed between grains with a 60° orientation relationship, including 4, 8-4, 8-4-8-4, 12-4, 12-4-12-4, and 4-8-4-4-12 structures (the numbers refer to ring sizes; see Fig. 1d). The calculation results showed that 8-4 (i.e., 8-MR combined with 4-MR) is the most stable configuration with the lowest formation energy, regardless of the S concentration; moreover, structures involving 12-MR generally exhibited high formation energies, suggesting that the formation of 12-MR is energetically less favorable than the formation of 8-MR (Fig. 1e). Therefore, polycrystalline monolayer MoS2 films exclusively comprising grains with a fixed orientation relationship (0° or 60°) are expected to have abundant 8-MRs suitable for water/ion separation.
We successfully fabricated such films on sapphire (α-Al2O3 (0001)) substrates using our recently developed CVD-based method15. The key to this method is a two-step process that facilitates the alignment of MoS2 seed crystals before further growth (see experimental section for details). For comparison, we also prepared monolayer MoS2 films comprising randomly oriented grains (denoted as R-MoS2) by omitting the prior alignment step. The films with controlled and random grain orientations are denoted as C-MoS2 and R-MoS2, respectively. The difference in grain orientation between C-MoS2 and R-MoS2 can be clearly identified from optical microscopy images showing triangular MoS2 flakes recorded at an early stage of the CVD process (Figs. 2a, 2d). As the CVD process proceeded, the triangular MoS2 flakes continued to grow and connected to each other through GBs, eventually forming continuous films (Figs. 2b, 2e). The as-prepared MoS2 films exhibited excellent uniformity throughout the substrate (diameter: 2 inch), as seen from the Raman and photoluminescence spectra as well as atomic force microscopy (AFM) images (Supplementary Figs. 2-4).
The GB structures on C-MoS2 and R-MoS2 films were probed using atomic-resolution annular dark-field scanning transmission electron microscopy (ADF-STEM). The images revealed that 7-MR is the main defect structure of R-MoS2, prevalent at GBs between grains with non-60° orientation relationships. Figure 2c shows a typical example, where many 7-MRs are observed along the boundary between two grains misaligned by 19°. This observation is consistent with the conclusions of previous theoretical studies that the formation of 7-5 rings in non-mirror-twin boundaries (or tilt boundaries) is energetically favorable21-23. Interestingly, when the two interconnected grains have a 60° orientation relationship, as in the case of C-MoS2, a string of 8-MRs is observed along the mirror-twin boundaries (Fig. 2f and Supplementary Fig. 5). This observation is consistent with the expectation that building GBs with MoS2 grains with a 60° orientation relationship can yield abundant 8-MRs in the film. Statistics based on ADF-STEM imaging revealed that 8-MRs accounted for ~97% of all the pores observed in the GBs of C-MoS2 (4-MRs were not included in the statistics because of their impermeability), while pores of sizes >1 nm were rare, accounting for only ~2.3% of all the observed pores (Supplementary Fig. 6).
The grain sizes and grain boundary numbers in C-MoS2 and R-MoS2 films can be roughly determined by visualizing GBs using an established method. Specifically, heating the films at 100°C for 1 h in humid air (60% relative humidity) caused the accumulation of contaminants (water, oxygen, and hydrocarbon molecules) along the GB, allowing the easy identification of GBs using scanning electron microscopy (SEM)24. The SEM results show that R-MoS2 and C-MoS2 films grown under standard conditions exhibit similar average grain sizes of ~7.4 and ~6.8 µm2, respectively; however, the intersection of multiple GBs was observed in R-MoS2 but not in C-MoS2, confirming the difference in grain orientation (i.e., random vs. fixed) between the two films (Supplementary Fig. 7).
We fabricated three other C-MoS2 films exhibiting different average grain sizes by varying the CVD conditions to tune the nucleation density (see experimental section). Their average grain sizes, which were determined using the aforementioned method, were ~443.4, ~87.3, and ~0.26 µm2, respectively (Supplementary Fig. 8). Thus, a total of five films were fabricated, hereinafter denoted as “C-MoS2-X” or “R-MoS2-X,” where “X” indicates the average grain size. The 8-MR density of the film increases with decreasing grain size. From the ADF-STEM imaging results, the density of 8-MR pores in C-MoS2-0.26 was estimated to be ~5.5–7.6 × 1011 cm−2 (Supplementary Note 1).
MoS2 membranes for water/ion separation
To fabricate separation membranes, the CVD-grown monolayer MoS2 films were transferred from sapphire to porous polycarbonate (PC) substrates using the polydimethylsiloxane-assisted technique24,25 (Fig. 3a and Supplementary Fig. 9). PC was chosen as the substrate because its smooth surface and high surface energy (~40 mJ m−2) can provide a strong adhesion with the supported MoS2 film26,27. The top-view SEM image shows that the MoS2 film is continuous and devoid of cracks (Fig. 3b), while the cross-sectional ADF-STEM image confirms the monolayer nature and homogeneity of the film (Fig. 3c).
The water and ion transport properties of the fabricated MoS2 membranes were first verified using a forward osmosis (FO) configuration. When pure water was used as the feed solution and 2 M NaCl was used as the draw solution, the C-MoS2-6.8 membrane exhibited a water flux of ~1.2 × 105 mol m−2 h−1 (Supplementary Fig. 10a). Its water/ion separation capability was evaluated using various salt solutions (K+, Na+, Ca2+, Mg2+, and Al3+; 0.1 M) as the feed solution and a sucrose solution (2 M) as the draw solution. In all the tests, the ion flux of the C-MoS2-6.8 membrane was ultralow; it was more than five orders of magnitude lower than that of the bare PC substrate (Supplementary Fig.10b), corresponding to extremely high water/ion selectivity. For example, when tested with 0.1 M NaCl, the membrane had a water permeance of ~118.9 mol m−2 h−1 bar−1 and a water/NaCl selectivity of ~4.6 × 104. The nonzero ion fluxes (<0.16 mol m−2 h−1) can be attributed to the presence of a few large pores (Supplementary Fig. 6), which are only occasionally observed along the GBs (Supplementary Fig.11) but are absent inside the MoS2 grains (Supplementary Fig.12). Under identical test conditions, the R-MoS2-7.4 membrane showed considerably lower water permeance (~10.5 mol m−2 h−1 bar−1) and water/NaCl selectivity (~4418) than the C-MoS2-6.8 membrane despite having similar average MoS2 grain sizes (Fig. 3d). This result is consistent with the expectations and is attributed to the presence of the impermeable 7-MR as the primary defect structure on the randomly grown MoS2 film.
The C-MoS2-87.3 and C-MoS2-443.4 membranes showed lower permeability and water/NaCl selectivity compared to the C-MoS2-6.8 membrane, while the C-MoS2-0.26 membrane outperformed all the tested membranes, exhibiting a water permeance of ~232 mol m−1 h−1 bar-1 and a water/NaCl selectivity of ~6.5 × 104 (Fig. 3d). These results demonstrate that the water/ion separation performance of the C-MoS2 membranes is dependent on the number of 8-MRs, which increases as the grain size decreases. Unlike the conventional water desalination membranes that have a tradeoff between water permeance and water/ion selectivity, C-MoS2 membranes show positively correlated water permeance and water/ion selectivity, both of which increase with the number of 8-MRs, suggesting a molecular sieving–based separation mechanism.
Notably, the optimized C-MoS2 (i.e., C-MoS2-0.26) membrane outperforms advanced membranes made of various nanomaterials (Fig. 3f). The water/NaCl selectivity and water permeance of C-MoS2-0.26 are approximately five times and more than 10 times the corresponding values of commercial polyamide thin-film composite membranes (Fig. 3f and Supplementary Table 1)28,29. Seven membranes fabricated using C-MoS2-0.26 collected from different regions of a sapphire substrate (diameter: 2 inch) showed consistent water/ion separation properties (Supplementary Fig.13), demonstrating the structural homogeneity of monolayer MoS2 and reliability of the water separation tests. Because C-MoS2-0.26 membranes delivered the best performance, these membranes were used in subsequent studies.
During 30 days of continuous testing with 0.6 M NaCl in the feed (similar to the ionic strength of seawater) and 2 M sucrose as the draw solution, the C-MoS2-0.26 membrane showed a steady NaCl rejection of >99% and only slightly reduced water permeance, demonstrating long-term mechanical and chemical stability. The excellent mechanical strength of C-MoS2-0.26 is also evidenced by its high Young's modulus of 0.251 TPa (Supplementary Fig. 14). When synthetic seawater (containing KCl, NaCl, Na2SO4, CaCl2, and MgCl2; ionic strength: ~1.15 M) was used as the feed, the C-MoS2-0.26 membrane exhibited >99.9% ion rejection and a water permeance of ~214 mol m−2 h−1 bar−1 (Supplementary Fig. 15).
Exploration of the separation mechanism
The excellent water/ion separation ability of C-MoS2 membranes is believed to originate from the unique size-exclusion effect of 8-MR as hydrated ions encounter a considerably higher energy barrier than water molecules when entering the subnanometer-sized pores30, 31,32. To gain more insight into the separation mechanism, we calculated the Arrhenius activation energy (EA) for water transport through the C-MoS2-0.26 membrane based on the water fluxes measured at different temperatures. The results showed that EA is barely affected by the system pH (e.g., EA = 15.1 and 15.3 kJ mol−1 at pH 3.0 and 7.8, respectively, Fig. 4a); this finding indicates a lack of proton concentration–governed water–pore interactions. Therefore, it is more likely that the energy barrier is associated with the drastic change in the H-bond configuration when water molecules are transported from the bulk solution into the 8-MR pores. Likewise, different anion species did not have considerable impact on the water/Na+ separation performance of the C-MoS2-0.26 membrane (Fig. 4b), confirming the irrelevance of solution chemistry to the separation.
The C-MoS2-0.26 membrane was also evaluated using a reverse osmosis (RO) configuration. For these tests, membranes were fabricated by supporting monolayer C-MoS2-0.26 films on highly porous anodized aluminum oxide (AAO) substrates. The use of AAO substrates occasionally resulted in the generation of pinholes in the membrane; these pinholes could be sealed subsequently via an interfacial polymerization strategy4,12 (see details in the experimental section). At a feed salt concentration of 0.01 M and an operating pressure of 2 bar, the C-MoS2-0.26 membrane exhibited the following ion diameter–dependent rejection (R) sequence: R (AlCl3) > R (MgCl2) > R (Na2SO4) > R (NaCl) > R (KCl) (Fig. 4c). The fact that MgCl2 is rejected to a greater extent than Na2SO4 rules out the Donnan exclusion as the main separation mechanism. When tested with NaCl, the membrane exhibited a high water permeance of ~46 L m−2 h−1 bar−1 with an NaCl rejection of ~97%, which is considerably higher than that of the reported 2D material–based membranes (<90%)2,33-35 (Supplementary Table 2). Its separation performance remained mostly constant with increasing applied pressure up to 13 bar, feed concentration up to 6000 ppm (confirming that the Donnan exclusion effect is negligible), and under continuous long-term operation up to 450 h (Supplementary Fig. 16). In addition to common cations, C-MoS2 membranes can effectively exclude boron, which exists as B(OH)3 in neutral solutions with a low molecular weight of 61.8 g mol−1. Compared with the commercial polyamide thin-film composite membrane SW30, the C-MoS2-0.26 membrane exhibited a higher boron rejection (~90.5% vs. ~70%; at pH = 7) and an approximately 40-fold increase in water permeance (Fig. 4d). Furthermore, unlike SW30 whose boron rejection is pH-sensitive, the high boron rejection of the C-MoS2-0.26 membrane remained almost constant in the pH range of 6–10. These results collectively indicated that the strong molecular-sieving effect of 8-MR pores, rather than the surface charge, is the primary mechanism responsible for the exceptional water/ion selectivity of C-MoS2 membranes, regardless of the type of driving force (FO or RO) for the water transport.
Direct seawater splitting and selective proton transport
Water/ion separation membranes enable the combination of the FO process with electrolysis to produce hydrogen directly from seawater, avoiding side reactions on the electrodes and corrosion issues36. An electrolytic cell was assembled for this application, in which the inner electrolyte (7.1 mL of 0.8 M NaH2PO4) and the outer solution (0.5 L of 0.6 M NaCl) were separated by the C-MoS2-0.26 membrane (Fig. 5a). The C-MoS2-0.26 membrane balanced the water influx and outflux at ~32.5 L m−2 h−1 under a constant current of 76 mA, with ~100% faradaic efficiencies for both hydrogen evolution and oxygen evolution reactions. Owing to its high water flux, the C-MoS2-0.26 membrane achieves a remarkable energy storage efficiency of 43.68 kJ h−1 cm−2 via H2 generation, which is nearly 40 times that of commercial cellulose acetate membranes36 (Fig. 5a and Supplementary Note 3). Notably, this performance remained constant over 80 h of continuous operation (Supplementary Fig. 17 and Table 3). Conversely, the electrochemical splitting of seawater without using a membrane yielded poor energy conversion efficiency even in a short operating period (Supplementary Fig. 18).
In addition to their water/ion separation capacity, C-MoS2 membranes also demonstrate an interesting ability for selective proton transport. The single-file water chains can adopt two configurations that differ in the manner of formation of H-bonds between water molecules (Fig. 5b). When a water molecule forms H-bonds with two adjacent water molecules via one O and one H, it can facilitate proton transport through the Grotthuss mechanism37. According to MD simulations, ~13% of the water molecules in MoS2 8-MRs adopt this configuration.
The proton transport properties of the C-MoS2-0.26 membrane were tested using a simple device fabricated by transferring the C-MoS2-0.26 film onto a silicon nitride substrate with an aperture of 5 µm diameter (Fig. 5c). Because the MoS2 grains were substantially smaller than the aperture area, abundant GBs could be used for ion transport. The assembled device exhibited low areal conductance (<100 S m−2) for various aqueous solutions, containing 0.1 M of KCl, NaCl, LiCl, CaCl2, or MgCl2. In contrast, a high areal conductance (~1.3 × 104 S m−2) was observed when 0.1 M HCl solution was used for the measurement, demonstrating the selective proton transport capability of C-MoS2-0.26 (Fig. 5d). The measured selectivity of H+ relative to other cations (e.g., H+/Li+ of ~170, H+/Na+ of ~160, and H+/K+ of ~127) is higher than that of various artificial proton channels (Supplementary Table 4).
The ion transport behaviors of the device hardly changed after soaking in 0.1 M HCl solution for 10 h, indicating that the GB structures of C-MoS2 have good acid resistance (Supplementary Fig. 19). Replacing the C-MoS2-0.26 film with a single-crystalline defect-free MoS2 film in the device caused loss of proton transport ability (Supplementary Fig. 20), confirming the critical role played by the defect structure in proton transport.
Proton transport through C-MoS2-0.26 was further investigated using several HCl concentration pairs, each with a four-fold concentration difference: 0.4 M / 0.1 M, 0.2 M / 0.05 M, and 0.1 M / 0.025 M (Supplementary Fig. 21). For all the tests, the zero-voltage current remained positive, and the zero-current voltage was nearly constant at ~−35 mV, indicating that H+ ions were the primary contributors to the net current. The diffusive proton permeability was calculated to be ~1 × 10−4 cm2 s−1, which is comparable to that achieved by single water chains confined in carbon nanotubes18,38,39 (Supplementary Note 4).