Recent progress in atomic-level manufacturing of two-dimensional transition metal dichalcogenides beyond exfoliation and restacking

Huihui Lin *a and Yang Meng b
aInstitute of Sustainability for Chemicals, Energy and Environment (ISCE2), Agency for Science, Technology and Research (A*STAR), 1 Pesek Road, Singapore 627833, Singapore. E-mail: linhuihui86@gmail.com
bNational University of Singapore (Suzhou) Research Institute, Suzhou, 215123, China

Received 11th February 2025 , Accepted 2nd April 2025

First published on 3rd April 2025


Abstract

Two-dimensional transition metal dichalcogenides (2DTMDCs) are promising in quantum computing, flexible electronics, spintronics, sustainable energy systems, and advanced healthcare. To transition 2DTMDCs from the lab to industry, it is crucial to develop scalable and industrially compatible strategies for precisely modulating novel quantum states. In this review, we provide a new classification of atomic-level manufacturing strategies for quantum state manipulation of 2DTMDCs beyond conventional exfoliation and restacking. We begin by summarizing emerging synthesis strategies for high-quality intrinsic 2DTMDCs and approaches for atomic-level engineering. We then explore the novel quantum phenomena that arise from these modifications, examining their underlying mechanisms in three key aspects: (a) quantum state manipulation in intrinsic 2DTMDCs, (b) quantum state engineering through intrinsic atomic engineering, and (c) quantum state modulation via extrinsic heteroatom incorporation. Finally, we discuss the challenges and future prospects of atomic-scale manufacturing in 2DTMDCs, providing insights into potential research directions.


image file: d5ta01124h-p1.tif

Huihui Lin

Dr Huihui Lin is a Staff Scientist in Energy and Environment (ISCE2) at the Agency for Science, Technology and Research (A*STAR) and the National University of Singapore. He received his PhD from the School of Physics at Nanjing University in 2019 under the supervision of Prof. Libo Gao, and subsequently served as a Research Fellow with Prof. Jiong Lu at the National University of Singapore. His current research focuses on the atomic-level engineering of novel 2D materials and the development of new applications, including 2D electronic devices, sustainable energy solutions, and biomedical technologies.

image file: d5ta01124h-p2.tif

Yang Meng

Dr Yang Meng is an Associate Investigator at the National University of Singapore (Suzhou) Research Institute (NUSRI), working under the supervision of Prof. Jiong Lu. She received her PhD from Guizhou University in 2024. Her current research focuses on the theoretical exploration of two-dimensional (2D) materials on catalytic mechanisms and practical applications.


1. Introduction

Two-dimensional transition metal dichalcogenides (2DTMDCs) are a class of materials that can be reduced to few-layer or monolayer structures, characterized by the MX2 configuration, where M represents a transition metal atom (such as Mo, W, or Nb) and X denotes a chalcogen atom (such as S, Se, or Te).1–5 Due to their quantum confinement, broken inversion symmetry, and strong spin–orbit coupling effects, 2DTMDCs exhibit a range of unique physical phenomena, including Ising superconductivity,6–9 charge density waves (CDWs),10–13 single-photon emission,14–17 and plasmonic effects.18–20 These remarkable properties make 2DTMDCs highly promising for applications in low-power, high-performance, and flexible electronic devices, as well as in energy-related technologies.

At the forefront of quantum nanotechnology lies the design and fabrication of artificial quantum materials with atomic precision, tailored to exhibit distinct quantum states. Although significant progress has been made in developing preparation techniques to achieve various quantum states in 2DTMDCs, a fundamental understanding of the atomic-level mechanisms governing these states remains incomplete. A simple exfoliation and restacking approach1,21–25 has been widely used to assemble a variety of 2D heterostructures,26–28 superlattices,26,29–31 and moiré superlattices.32–36 However, this method faces significant technical bottlenecks, rendering it unsuitable for practical industrial applications. A crucial direction for future research involves devising scalable and efficient strategies for mass-producing functionalized 2DTMDCs beyond conventional exfoliation and restacking, facilitating their integration into high-performance electronic and energy devices. Achieving this goal is essential to unlocking the full potential of 2DTMDCs in next-generation technologies. Therefore, it is crucial to systematically compile the latest research advancements and establish a methodological framework, providing valuable reference and guidance for future research.

In this review, we introduce a new classification of atomic engineering (or atomic-level manufacturing) strategies for quantum state manipulation in 2DTMDCs beyond conventional exfoliation and restacking, focusing on recently reported cutting-edge research and providing new insights into recent advancements and their applications. We begin by discussing strategies for inducing novel quantum states in 2DTMDC materials, along with the mechanisms underlying their formation. We then explore the modulation and potential applications of distinct quantum states in TMDCs, categorized into three primary areas: (a) quantum state manipulation in intrinsic 2DTMDCs, (b) quantum state engineering through intrinsic atomic modifications, and (c) quantum state modulation via extrinsic heteroatom incorporation. Finally, we highlight the challenges and opportunities associated with industrially compatible fabrication methods and the practical implementation of novel quantum states in 2DTMDCs.

2. Atomic engineering of 2DTMDCs

Designing and realizing diverse quantum states in 2DTMDCs is crucial for their integration into next-generation high-performance devices. Various strategies have been developed to synthesize 2DTMDCs, including mechanical exfoliation, which yields high-quality atomic layers;1,21–25 chemical vapor deposition (CVD), enabling large-scale production;37–40 and electrochemical intercalation–exfoliation.41–48 Atomic-scale manufacturing of 2DTMDCs can be further refined through modular stacking, annealing, and chemical modification (Fig. 1).
image file: d5ta01124h-f1.tif
Fig. 1 Some cases of atomic engineering of 2DTMDCs. (a) Stacking of intrinsic 2DTMDCs. (b) Quantum states manipulation by intrinsic atomic engineering. (c) Quantum states manipulation by atomic engineering of extrinsic heteroatoms.

2DTMDCs obtained through conventional exfoliation and restacking exhibit high crystalline quality, providing a solid foundation for studying their intrinsic properties.1,21–25 However, the limited sample size and scalability hinder their practical applications and industrial scalability.49–51 Additionally, due to their two-dimensional confinement and high surface energy, the physical properties of 2DTMDCs are highly sensitive to external conditions.52 Therefore, it is crucial to develop strategies that not only enhance their environmental stability but also establish universal and highly reproducible modification techniques beyond conventional exfoliation and restacking. In this section, we review recently developed strategies for the atomic engineering of 2DTMDCs crucial for achieving precise quantum state control.

2.1 Recently developed preparation strategies of high quality intrinsic 2DTMDCs

2.1.1 Newly developed ion intercalation–exfoliation strategy. Ion intercalation–exfoliation, especially with alkali metal ions, has been widely used in the exfoliation of atomically thin sheets. One of the most widely accepted methods, chemical lithiation, involves organic lithium reagents such as n-butyllithium (n-BuLi), first introduced in 1986.53,54 The solvent-based chemical ion intercalation–exfoliation process consists of two main steps: first, intercalating ions weaken the interlayer adhesion forces in layered materials, leading to an increased interlayer distance and atomic-layer separation. Then, protonic solvents (e.g., water) often facilitate gas release (e.g., hydrogen gas), which further expands the layers, ultimately achieving exfoliation.54–60

However, solvent-based chemical ion intercalation–exfoliation using n-BuLi as the intercalant involves charge transfer between the inserted ions and the layered crystal, inducing the formation of charged layers. Despite its effectiveness, this method has several limitations: n-BuLi is highly explosive in ambient air, requires long intercalation times, and demands stringent reaction conditions (e.g., 100 °C in an argon-filled glovebox).55,61–63

To address the safety and scalability challenges of solvent-based methods, solid-state ion intercalation–exfoliation strategies have been developed in recent years.62,64 One promising approach utilizes solid-state lithiation and exfoliation for the scalable synthesis of transition metal telluride (TMT) nanosheets, achieving ultra-fast, gram-scale production (Fig. 2a–c).62,66 This method involves mixing bulk TMT crystals with a safer lithium source, such as lithium tetrahydroborate (LiBH4), followed by a brief 10-minute heat treatment in an argon atmosphere to complete solid-state lithiation. The LixMTe2 (where x is the molar ratio of LiBH4 to MTe2) is then hydrolyzed within seconds to achieve exfoliation.


image file: d5ta01124h-f2.tif
Fig. 2 Recently developed preparation strategies of high quality intrinsic 2DTMDCs. (a–c) Solid-state lithiation and exfoliation strategy for TMT nanosheets: (a) schematic illustration of the solid-state lithiation and exfoliation process for TMT nanosheets. (b) Photograph of exfoliated NbTe2 nanosheets. (c) SEM image of high quality and uniform TMT nanosheets. (d–f) Electrically driven ammonium ion co-intercalation–exfoliation strategy: (d) photograph of tetrabutyl ammonium (TBA)-exfoliated 2DTMDCs dispersed in propylene carbonate (PC) solvent. (e) Optical image of a monolayer NbSe2, with the inset showing exfoliated NbSe2 flakes in PC. (f) Atomically resolved high-angle annular dark-field scanning transmission electron microscopy (ADF-STEM) image of the NbSe2 monolayer, with the inset demonstrating a high yield (>75%) of large-area single-crystal monolayers. (g and h) Two-step vapor deposition strategy: (g) schematic representation of the two-step vapor deposition growth process, involving magnetron sputtering of an Nb film, followed by selenylation into NC-NbSe2, enabling wafer-scale production. (h) Photograph of a Nb film and a selenized Nb film deposited on 2-inch sapphire wafers. Panels reproduced with permission from: (a–c) ref. 62, Copyright 2024, Springer Nature Limited; (d–f) ref. 32, Copyright 2021, Springer Nature Limited; (g) ref. 65, Copyright 2025, Science Press and EDP Sciences; (h) ref. 49, Copyright 2019, Springer Nature Limited.

This strategy demonstrates broad applicability, enabling the scalable production of various TMT nanosheets, including MoTe2, WTe2, NbTe2, TaTe2, and TiTe2. These nanosheets exhibit excellent processability and can be integrated with advanced atomic manufacturing techniques to form various inks for fabricating membranes, thin films, nanocomposites, and heterostructures.67

Another innovative strategy leverages liquid gallium (Ga) intercalation and surface tension to disrupt the van der Waals (vdW) forces in layered materials, enabling the exfoliation of bulk layered crystals into 2D nanosheets (2D-NS) at near-room temperature.68 Utilizing Ga's fluidic and metallic properties, this method gently weakens the vdW interactions, producing high-quality, large-area, surfactant-free 2D-NS.68–70 The exfoliation process involves mild mixing of molten Ga with layered material powders using a magnetic stirrer. The fluid behavior of Ga and the adhesive nature of Ga2O3 enable the shear forces generated by the stirring process to separate the 2D layers from the bulk crystal. This approach has been successfully applied to exfoliate over ten different 2D materials, including MoTe2 and MoSe2, yielding high-aspect-ratio, surfactant-free 2D-NS. With significant advantages in scalability and cost-effectiveness, this strategy holds immense technological potential for applications in optics, energy conversion, and beyond.

2.1.2 Electrically driven intercalation–exfoliation strategy. Lithium-ion (Li+) intercalation–exfoliation is a widely used strategy for exfoliating layered crystals. During the lithiation process, each Li+ insertion is accompanied by the injection of an electron into the host crystal.54,61 However, excessive Li+ intercalation introduces a large number of electrons into bulk TMDC crystals, which can trigger an undesirable phase transition from the semiconducting 2H phase to the metallic 1T phase. To prevent such uncontrolled phase changes, it is essential to minimize electron injection into the host 2D crystal.32,57

One approach to mitigating unwanted phase transitions is replacing small Li+ ions (d ≈ 2 Å) with larger cations, such as quaternary ammonium ions. For instance, Lin et al. demonstrated that electrochemical intercalation using bulky ammonium molecules—such as tetraheptyl ammonium bromide (THAB; d ≈ 20 Å), followed by mild ultrasonic treatment enables the exfoliation of pure-phase 2H-MoS2 nanosheets.61,71,72 This method offers a reliable route for large-scale production of high-quality nanosheets, particularly for applications in electronics, optoelectronics, and thermoelectric devices. However, the increased cation size inevitably raises the intercalation barrier, reducing both intercalation efficiency and exfoliation rates.

To address the challenges of phase transition and slow exfoliation rates, a versatile and mild electrochemical intercalation–exfoliation strategy has been developed for the large-scale production of highly crystalline monolayer 2DTMDCs by Lu's group (Fig. 2d). This method employs a flexible co-intercalation agent, where ammonium cations are solvated with a large number of neutral solvent molecules. This unique combination facilitates complete penetration into bulk crystals, leading to the efficient exfoliation of high-quality, large-area single-layer TMDCs.32,73 For example, using this approach, they achieved a high yield (>75%) of large-sized single-crystal monolayer NbSe2 with lateral dimensions reaching up to 300 μm (Fig. 2e and f). The resulting monolayer TMDCs exhibit pristine interfaces, making them highly promising for twisted van der Waals heterostructures (vdWH) and wafer-scale printed films with novel superconducting properties.

Despite its immense potential for producing solution-processable 2D semiconductors, electrochemical molecular intercalation–exfoliation typically requires large-sized bulk single crystals (e.g., crystal lengths > 3 mm) as the starting material, making it costly and impractical for large-scale synthesis.71,74 A recently reported electrically driven intercalation–exfoliation strategy utilizing a liquid-metal-assisted method offers a scalable and cost-effective alternative. This approach enables the electrochemical intercalation of low-cost, readily available crystal powders, significantly broadening the range of exfoliable materials.75,76 By incorporating liquid-metal-based soft frameworks, this method accommodates the substantial volume expansion of layered crystals during molecular intercalation while maintaining strong electrical contact with microcrystals.77,78 The resulting solution-processable MoS2 nanosheets exhibit comparable quality to those exfoliated from bulk single crystals. As a high-throughput exfoliation technique using low-cost powdered materials, this strategy enables the large-scale, cost-effective production of high-performance electronic inks (>50 different types) and facilitates the fabrication of large-area electronic devices with significantly reduced manufacturing costs.

2.1.3 Recently developed CVD growth strategy. Scalable top-down synthesis methods have enabled the fabrication of 2DTMDCs with physical properties comparable to those of intrinsic 2DTMDCs.60,62,72 Meanwhile, bottom-up fabrication techniques, particularly CVD, offer advantages such as large-scale production, scalability, and compatibility with industrial manufacturing processes.79–81 As a result, extensive research has been devoted to developing CVD-based growth strategies.

For graphene synthesis using gas-phase precursors, high-crystalline-quality, large-area single-crystal graphene can be obtained by employing techniques such as growth on single-crystal metal substrates.82–84 However, the growth of high-quality intrinsic 2DTMDCs is more complex, as it involves at least two solid precursors and a gas–liquid–solid three-phase reaction process. Among the various approaches, salt-assisted chemical vapor deposition (SA-CVD) has emerged as a highly effective method for synthesizing 2DTMDCs.85–88

In 2015, Li et al. first demonstrated the growth of large monolayer WSe2 and WS2 using alkali metal halides as growth promoters.87 By mixing tungsten oxide (WO3 ∼ 1473 °C) with halide salts (AX, where A = Na, K and X = Cl, Br, I), monolayer WSe2 and WS2 could be synthesized at significantly lower temperatures (∼700 °C), more than 100 °C lower than the same process using only tungsten oxide. Subsequent research by the same group reported the vapor–liquid–solid (VLS) growth of monolayer MoS2, yielding highly crystalline ribbon structures with widths ranging from tens to thousands of nanometers.86,89 This growth process is mediated by molten Na–Mo–O droplets, which, upon saturation with sulfur, guide MoS2 ribbon formation in a “crawling mode”. This technique provides an alternative approach for controllable atomic-scale nanostructure arrays and hybrid-dimensional heterostructures for nanoelectronics. Furthermore, Zhou et al. elucidated how halide salts lower the melting point of reactants and facilitate the formation of intermediate species, the overall reaction rate.88 The scalability of SA-CVD was also demonstrated in this study, enabling the growth of a wide range of 2DTMDCs, including 32 binary compounds and 13 alloyed structures.

In a very recent breakthrough, He et al. proposed the 2D Czochralski (2DCZ) method, a reliable approach for growing centimeter-scale single-crystal MoS2 domains.81,90,91 This method introduces a solid–liquid–solid (SLS) process that facilitates the transformation of polycrystalline MoS2 into single-crystal MoS2.81 Firstly, a large-area 2D liquid precursor film is formed on a molten glass substrate. Then, rapid sulfurization under an ultrafast process yields atomically smooth MoS2 single crystals. This technique successfully produced 1.5 cm-scale MoS2 single-crystal domains with ultra-low defect densities (∼2.9 × 1012 cm−2). The 2D Czochralski method represents a major advancement in scalable, high-quality 2D semiconductor material fabrication, offering significant potential for the production of next-generation electronic and optoelectronic devices.

A novel two-step vapor deposition strategy for the scalable fabrication of large-area 2DTMDCs has been proposed by our group (Fig. 2g and h).49 Theoretical analysis indicates that oxygen readily bonds with 2DTMDCs during growth, leading to the formation of atomic vacancies and a reduction in environmental stability.49,65 Therefore, achieving a completely oxygen-free growth environment is essential for producing stable 2DTMDC films. As a result, a two-step vapor deposition method to synthesize environmentally stable 2DTMDC thin films is developed. To achieve high-crystallinity 2DTMDC films, deeply degassed magnetron sputtering under O2- and H2O-free conditions is employed to deposit thin metal films onto heated sapphire substrates. Subsequently, a typical CVD process was used to convert the crystalline metal films into various 2DTMDCs or carbides by introducing different source materials. Through a series of studies, the universality of the two-step vapor deposition model was extended and its effectiveness demonstrated in achieving precise atomic-scale manufacturing of 2DTMDCs. This method successfully enabled the synthesis of single-crystal MoSe2, single-crystal NbSe2, wafer-scale polycrystalline NbSe2 and TiSe2, and wafer-scale nanocrystalline NbSe2.49,65,92 Moreover, the programmable and scalable fabrication capability of this two-step vapor deposition process was validated by achieving batch-scale production of wafer-scale nanocrystalline NbSe2 (NC-NbSe2).65 The industrially compatible, atomically customizable fabrication of 2DTMDC thin films lays the foundation for developing multifunctional 2DTMDC-based devices with complex integrated structures, offering significant technological implications for future applications. Following the work by Wang et al., an innovative topological transformation strategy was reported for fabricating transition-metal dichalcogenide superconductor (TMDSC) nanostructures using pre-patterned metal precursors. By leveraging pre-patterned metal templates, this method allows for the direct transformation of precursor structures into superconducting TMDSCs, paving the way for their integration into next-generation electronic and quantum computing applications.80,93

2.2 Recently developed strategies for engineering of 2DTMDCs

Due to their atomic-layer thickness, 2DTMDCs inherently exhibit structural disorder during synthesis.94,95 This disorder has a dual impact on their properties: In high-performance devices that rely on the intrinsic properties of 2DTMDCs, structural disorder can degrade their electronic, optical, and mechanical characteristics, ultimately hindering device performance. Therefore, minimizing disorder is essential for achieving optimal material functionality.96 On the other hand, at the atomic scale, engineered disorder can be deliberately introduced to tailor the quantum states of 2DTMDCs, unlocking novel electronic, optical, chemical, and magnetic functionalities.96 This opens exciting opportunities for next-generation quantum devices and emerging applications in photonics, electronics, and catalysis. In this section, we highlight recently developed atomic engineering strategies that leverage disorder for precise quantum state modulation, enabling new functionalities in 2DTMDCs.97,98

One effective strategy for intrinsic atomic engineering in 2DTMDCs is the deliberate introduction of atomic vacancies, which can serve as quantum-scale functional units. As shown in Fig. 3a–c, Li et al. demonstrated that single chalcogen vacancies in PtTe2 can act as atomic-scale building blocks, enabling the bottom-up fabrication of quantum antidots (QADs)—periodic vacancy structures that modulate quantum states at the atomic level.97,98 The relatively low diffusion barrier of Te single vacancies (SVs) in PtTe2 allows for vacancy migration, leading to the formation of a superlattice of vacancies.


image file: d5ta01124h-f3.tif
Fig. 3 Recently developed strategies for atomic engineering of 2DTMDCs. (a) Annealing strategy for intrinsic atomic engineering: (a) schematic illustration of the self-assembly process leading to the ordered clustering of Te vacancies in atomically thin PtTe2. (b and c) STEM-ADF and Scanning Tunneling Microscopy (STM) images of few-layer PtTe2, highlighting atomic-scale vacancy ordering. (d–g) Phase control strategy for atomic engineering of extrinsic heteroatoms: (d) schematic representation of the electrochemical intercalation–exfoliation to synthesize 1T′-MoS2 NSs. (e) TEM image and (f) AFM image of exfoliated 1T′-MoS2 NSs, confirming their nanosheet morphology. (g) Thickness distribution histogram of 1T′-MoS2 NSs, obtained through AFM measurements. (h–j) Direct preparation strategy for atomic engineering of extrinsic heteroatoms: (h) schematic illustration of the hydrothermal synthesis process used to fabricate Ni1/MoS2 SASCs. (i) TEM image of Ni1/MoS2. (j) Atomically resolved ADF-STEM image of a single NiMo site (marked by a red circle). Panels reproduced with permission from: (a) ref. 97, copyright 2023, Springer Nature Limited; (b and c) ref. 98, copyright 2021, Springer Nature Limited; (d–g) ref. 99, copyright 2023, Springer Nature Limited; (h–j) ref. 100, copyright 2023, Springer Nature Limited.

This engineered PtTe2 structure, with intentionally introduced vacancy defects, exhibits multi-level quantum hole states, a property of immense significance for electrocatalytic hydrogen evolution reactions (HER), where vacancy sites enhance catalytic activity.98 And quantum antidots-based quantum devices, where atomic-precision vacancy engineering can define electronic and optical quantum states.97 By precisely controlling vacancy distribution, this approach provides a pathway toward functional atomic engineering, opening new frontiers in quantum materials design and nanoscale device fabrication.

The second strategy is atomic engineering of tuned extrinsic heteroatom density via different 2DTMDC Phases. For example, Shi et al. investigated the influence of the 2H and 1T′ phases of MoS2 on the distribution and catalytic behavior of Pt atoms.99,101,102 Their study revealed that the 2H phase template facilitates the epitaxial growth of Pt nanoparticles, whereas the 1T′ phase supports the formation of dispersed single-atom Pt (s-Pt) sites, achieving a high Pt loading of up to 10 wt%. To achieve this, tetraheptyl ammonium bromide molecules were intercalated into KxMoS2 crystals within an electrochemical cell (Fig. 3d). After exfoliation, ultrathin MoS2 nanosheets with a thickness of 1.4 ± 0.4 nm and lateral sizes reaching several micrometers were obtained (Fig. 3e–g).

In the s-Pt/1T′-MoS2 system, the hydrogen adsorption free energy of Pt atoms anchored above Mo sites approaches 0 eV, making these sites highly active for efficient HER in acidic media. When used as a cathode catalyst in a prototype proton exchange membrane (PEM) electrolyzer, the s-Pt/1T′-MoS2 system demonstrated remarkable performance: To maintain the 1T′ phase at room temperature, the required cell voltages for 500 mA cm−2 and 1000 mA cm−2 were only 1.73 V and 1.82 V, respectively. The catalyst also exhibited long-term stability, operating continuously for 500 hours at 100 mA cm−2. These findings highlight the phase-dependent heteroatom engineering of 2DTMDCs as a powerful tool for optimizing electrocatalytic activity and durability in energy conversion applications.103–105

The third strategy is a direct preparation strategy for atomic engineering of extrinsic heteroatoms. For example, Sun et al. developed a versatile and scalable hydrothermal technique to synthesize a comprehensive collection of single-atom spin catalysts (SASCs) using MoS2 as a base matrix for substituting magnetic atoms (M1), as shown in Fig. 3h.100 During the hydrothermal process, the decomposition of thioacetamide serves a dual purpose: supplying sulfur for the formation of 2H-phase MoS2 and creating an acidic environment that prevents nanoparticle aggregation and favors atomic-level dispersion of doped metals. The resulting Ni1/MoS2, adopts a twisted tetragonal structure (Fig. 3i and j), which promotes ferromagnetic coupling between adjacent S atoms and neighboring Ni1 sites, leading to widespread room-temperature ferromagnetism. In a magnetic field, these heterogeneous single-atom spin catalysts reduce the potential barrier for the oxygen evolution reaction (OER), enhance metal utilization, and boost reaction efficiency.

3. Quantum state engineering of 2DTMDCs

3.1 Quantum state engineering through stacking of intrinsic 2DTMDCs

2DTMDCs exhibit a diverse range of properties depending on their composition and phase, making them highly promising for various applications. The exfoliation and restacking approaches have been widely utilized to assemble nearly arbitrary heterostructures, providing a versatile material platform for fundamental studies of exotic physical phenomena and proof-of-concept device demonstrations. Beyond conventional exfoliation and restacking, stacking of intrinsic 2DTMDCs via intercalation–exfoliation or epitaxial growth represents key fabrication strategies. These approaches enable the production of large-area, high-quality 2DTMDCs that are compatible with advanced integrated circuits, laying the foundation for next-generation chip materials and energy conversion catalysts. For example, twisted vdWHs and wafer-scale printed films via intercalation–exfoliation can be fabricated by stacking superconducting monolayers.32 These stacked structures exhibit an interesting critical current that can be modulated by a magnetic field when one flux quantum aligns with an integer number of moiré cells.
3.1.1 LEGO block of 2DTMDCs. A newly developed high-to-low temperature strategy enables the controlled wafer-scale growth of multi-layered vdW superconductor heterostructure (vdWSH) films based on our developed two-step vapor deposition process.29,49 In this approach, “high” and “low” refer to the different growth temperatures required for various 2D materials. Using excessively high temperatures can lead to decomposition, etching, or alloying of the pre-grown bottom-layer 2D material, making precise control of the deposition process essential. To address this challenge, a multi-cycle two-step vapor deposition process was designed, transitioning from low to high temperatures to fabricate stacked vdWSH films with clean interfaces and preserved superconducting properties. Based on this strategy, Zhou et al. successfully synthesized 27 double-block, 15 triple-block, 5 four-block and 3 five-block vdWSH films, where each block represents a different 2D material.29 The resulting continuous vdW interface ensures proximity-induced superconductivity across centimeter-scale regions and enables the realization of superconducting Josephson junctions (Fig. 4a and b). Due to the inevitable contamination of transferred vdWSH interfaces, directly stacked NbSe2/PtTe2 heterostructures exhibit significantly stronger superconducting behavior than transferred NbSe2/PtTe2 (Fig. 4b). This indicates that 2D superconductors can be seamlessly integrated into vdWSHs without degradation, preserving their intrinsic superconducting properties. Given its scalability and reproducibility, this multi-layer vdWSH fabrication strategy offers a promising pathway for accelerating the design and application of next-generation functional superconducting devices.108,109
image file: d5ta01124h-f4.tif
Fig. 4 Quantum state engineering through stacking of intrinsic 2DTMDCs. (a and b) vdW stacking of 2DTMDCs. (a) Schematic illustration and cross-sectional ADF-STEM image depicting the stacked growth of multi-block vdWHs by the high-to-low temperature strategy. (b) Variable temperature resistance (R) measurement of individual NbSe2 films vertically grown on sapphire, demonstrating the temperature-dependent electronic properties of vertically stacked 2DTMDCs. (c) Twisted Stacked 2DTMDCs. Atomic force microscopy (AFM) images showing aligned WS2 layers with various twist angles, grown around WOx particles on SiO2/Si substrates, providing insight into the structural modulation and moiré superlattice effects in twisted 2DTMDCs. (d) 3D printing of 2DTMDCs: photographs of 3D-printed NbSe2 atomic models and Merlion statues, demonstrating the feasibility of additive manufacturing techniques for fabricating customized 2D material architectures. (e) 3D integration of 2DTMDCs. Cross-sectional ADF-STEM image showcasing the structure of vertically integrated CMOS devices, highlighting the potential of 3D-stacked 2DTMDCs for next-generation nanoelectronic applications. Panels reproduced with permission from: (a and b) ref. 29, copyright 2024, Springer Nature Limited; (c) ref. 106, copyright 2020, AAAS; (d) ref. 32, copyright 2021, Springer Nature Limited; (e) ref. 107, copyright 2025, Springer Nature Limited.

Another innovative strategy to stacked 2DTMDCs involves the formation of twisted superstructures through the controlled growth of spiral dislocation helices on non-Euclidean surfaces.106 This model predicts continuously twisted superstructures, where the twist angles are fully determined by the geometry of the non-Euclidean surface. Zhao et al. demonstrated this model by achieving the growth of highly twisted WS2 and WSe2 helices using water vapor-assisted chemical vapor transport (CVT).106 In this process, MX2 (M = Mo, W; X = S, Se) was used as the precursor, while water vapor served as the transport agent. The twisted superstructures were grown on SiO2/Si substrates with a tunable twist angle (Fig. 4c). The key to achieving this non-Euclidean twisting was the introduction of SiO2 (200 nm diameter) or WO3 (<100 nm diameter) nanoparticles onto the SiO2/Si substrate before the growth reaction. These nanoparticles disrupted the conventional layered growth mechanism, promoting the formation of continuously twisted van der Waals superstructures. This strategy provides an innovative method for integrating spiral dislocation helices with non-Euclidean surface engineering, enabling the controlled synthesis of twisted van der Waals materials.110,111 Future research could further refine nanoscale substrate modifications to achieve even more precise control over twist angles. This approach opens new avenues for studying the properties of twisted 2DTMDCs and exploring their applications in twistronics and next-generation quantum devices.

3.1.2 3D block of 2DTMDCs. The development of 3D architectures incorporating 2DTMDCs brings these materials closer to practical applications. In 2020, Li et al. demonstrated monolayer NbSe2 exhibits excellent solubility in various solvents, making it highly compatible with modern thin-film manufacturing and printing technologies.32 The printed NbSe2 films demonstrate several key advantages over conventional thin films, including an increased upper critical field, higher superconducting transition temperature, and improved environmental stability. Furthermore, exfoliated NbSe2 monolayers can be seamlessly integrated with commercial Genesis resin, enabling their application in both 3D and 2D printing technologies. In these 3D-printed structures, NbSe2 atomic layers remain uniformly dispersed, while maintaining a controlled topology (Fig. 4d). The temperature-dependent magnetization measurements of these printed structures, conducted under a 200 Oe magnetic field, reveal a diamagnetic signal at approximately 6.8 K, confirming the superconducting transition. This successful integration of 2D superconductors into 3D-printed architectures represents a significant step toward scalable fabrication of next-generation superconducting materials and functional electronic devices.112,113

In a recent study, CVD-based 3D integration of 2DTMDCs has been successfully demonstrated. Kim et al. developed a method for growing single-crystal channel materials (composed of 2DTMDCs) on both amorphous and polycrystalline surfaces.107,114 Notably, the growth temperature was sufficiently low to preserve the integrity of the underlying electronic components, making it a viable approach for integrated circuits. The study showcased the growth of single-crystal channel materials on silicon wafers coated with an amorphous oxide layer at a remarkably low temperature of 385 °C. This breakthrough enabled the seamless monolithic integration of vertical single-crystal logic transistor arrays, as illustrated in Fig. 4e.

This technique achieved n-type metal oxide semiconductor (nMOS) and p-type metal oxide semiconductor (pMOS) vertical monolithic integration, facilitating the fabrication of functional vertical inverters. The seamless stacking approach substantially reduces interconnect distances, thereby mitigating resistive–capacitive (RC) delays and effectively doubling the transistor density within a given wafer space. The average voltage gain (Again) and noise margin (NM) of these vertical inverters surpass those of previously reported TMD-based inverters fabricated via conventional stacking methods, owing to the seamless nature of the in situ growth process. Furthermore, the study successfully demonstrated NAND and NOR functionalities using these vertical inverters. This strategy marks the first realization of a vertical complementary metal oxide semiconductor (CMOS) array composed of grown single-crystal channels, offering a promising avenue for monolithic 3D (M3D) integration of various electronic hardware using single-crystal materials.

3.2 Quantum state engineering through structure tailoring of intrinsic 2DTMDCs

Structural tailoring of intrinsic 2DTMDCs enables precise modulation of their quantum states, unlocking new functionalities for quantum electronics, optoelectronics, and energy applications. In this section, we summarize several representative strategies for structure tailoring of intrinsic 2DTMDCs, which provide fundamental insights into how controlled structural modifications can be leveraged to enhance the quantum characteristics of 2DTMDC materials.

Rolling up 2D vdWHs into high-order vdW superlattices extends the functionality of conventional 2D vdWHs, offering a rich materials platform for both fundamental studies and technological applications.31,71,115 Cui et al. (2018) demonstrated that monolayer TMDC sheets grown via CVD can be spontaneously rolled into 2D TMDC nanoscrolls (NSs) within 5 seconds using an ethanol droplet-based approach, achieving an almost 100% yield. The mobility of MoS2 NSs was measured in the range of 200–700 cm2 V−1 s−1, with the on/off ratio exceeding 1.0 × 105.115 Compared to monolayer MoS2 flakes (10–20 cm2 V−1 s−1), MoS2-NSs exhibit a nearly 30-fold enhancement in carrier mobility. Due to their internally open topology, 2D TMDC NSs allow for tunable interlayer spacing, which facilitates the incorporation of organic molecules, polymers, nanoparticles, 2D materials, and biomolecules, offering broad potential applications in nanoelectronics, catalysis, and sensing.71,116,117

In 2021, Zhao et al. reported a straightforward approach for realizing high-order vdW superlattices by rolling up vdW heterostructures (Fig. 5a). By exposing CVD-grown 2D/2D vdW heterostructures (e.g., SnS2/WSe2) to an ethanol–water–ammonia solution, capillary forces drive spontaneous delamination and rolling, forming high-order 2D/2D vdW superlattices without requiring multiple transfer and restacking steps (Fig. 5b–d).31 This rolling-induced stacking significantly alters the optical and electronic properties. The photoluminescence (PL) peak of WSe2 in the SnS2/WSe2 heterobilayer redshifts from 771 nm to 815 nm in the SnS2/WSe2 vdW superlattice, accompanied by a PL quenching effect. Due to the substantial bandgap reduction, charge transport in the rolled-up vdW superlattice is significantly enhanced, with the total current in the SnS2/WSe2 vdW superlattice-based FET being 2 to 6 orders of magnitude higher than that of a bilayer heterostructure-based FET at zero gate voltage. At 3 K, the magnetoresistance of the 1D rolled-up structure exhibits a linear dependence on the magnetic field, in contrast to the quadratic dependence observed in the 2D vdW heterostructure, further demonstrating the impact of rolling-induced quantum confinement effects.116,120


image file: d5ta01124h-f5.tif
Fig. 5 Quantum state engineering through structure tailoring of intrinsic 2DTMDCs. (a–d) High-Order vdW superlattice formation through rotational stacking: (a) schematic representation of the high-order vdW superlattice. (b) Cross-sectional ADF-STEM image of a representative SnS2/WSe2 roll-up. (c) Atomically resolved ADF-STEM image of a SnS2/MoS2/WS2 vdW superlattice. (d) Energy-dispersive X-ray spectroscopy (EDS) mapping of W (blue), Mo (green), and Sn (red), visualizing the element distribution within the superlattice. (e–g) Unconventional superconductivity in chiral 2DTMDC superlattices: (e) schematic illustration of chiral molecule intercalated 2H-TaS2. (f) Circular Dichroism (CD) spectra of right-handed methylbenzylamine (R-MBA) and left-handed methylbenzylamine (S-MBA) intercalated 2H-TaS2. (g) Resistance vs. temperature curves for pristine, R-MBA, and S-MBA intercalated 2H-TaS2. (h–k) Humidity sensing applications of 2DTMDCs. (h) Schematic of the phase-switchable synthesis process for 1T′-WS2 NSs. (i) Mechanical property measurements of the WS2 humidity sensor, with the inset showcasing the flexibility of the device. (j) Nasal breath monitoring curve. (k) Quantitative relationship between finger distance and device current under a 10 V bias, with the inset illustrating a finger approaching the sensor. Panels reproduced with permission from: (a–d) ref. 31, copyright 2023, Springer Nature Limited; (e–g) ref. 118, copyright 2024, Springer Nature Limited; (h–k) ref. 119, copyright 2024, Springer Nature Limited.

Surface functionalization provides another powerful strategy to modify intrinsic 2DTMDCs, introducing new quantum states and topological properties. Qian et al. explored the interaction between atomic layers in crystalline materials and self-assembled chiral molecular layers, revealing the potential to create exotic topological superconductors.30,118 By incorporating chiral molecules into conventional superconducting lattices, non-centrosymmetry is introduced, facilitating the realization of chiral superconductivity (Fig. 5e–g). This system exhibits several key characteristics: an exceptionally large in-plane upper critical field (Bc2,‖) exceeding the Pauli paramagnetic limit, a robust π-phase shift observed in Little–Parks measurements under different cooling sequences, and a field-free superconducting diode effect (SDE). These phenomena, which are absent in pristine 2H-TaS2, underscore the unique interplay between the crystalline atomic layers and the self-assembled chiral molecular layers, potentially leading to the emergence of novel topological quantum materials. The π-phase shift observed in TaS2 with chiral molecular insertion is not observed in pristine 2H-TaS2 or TaS2 with non-chiral molecular insertions, confirming the critical role of molecular chirality in the emergence of unconventional superconductivity.121

Another effective strategy for tailoring the quantum properties of intrinsic 2DTMDCs is phase engineering during material synthesis, which enables the discovery of new functionalities and applications. Liang et al. demonstrated that discharge current density during material synthesis can selectively tune the phase composition of WS2.119,122 At a low discharge current density (0.005 A g−1, cutoff voltage 0.9 V), a pure 2H-phase WS2 bilayer is formed. In contrast, at a higher discharge current density (0.02 A g−1, cutoff voltage 0.7 V), a predominantly semimetallic 1T′-phase WS2 monolayer is obtained (Fig. 5h).

This phase-engineered WS2 demonstrates promising applications in humidity sensing. The 2H-WS2 humidity sensor exhibits excellent performance at 65% RH, maintaining a stable response even after 1300 bending cycles (Fig. 5i). The device is also applied for breath monitoring, a crucial diagnostic tool for slowly progressing diseases such as cancer, diabetes, and sleep apnea-hypoventilation syndrome, where early indicators are limited. The sensor exhibits high sensitivity to breathing rate variations, as demonstrated in Fig. 5j, where the sensor current value dynamically responds to breathing rate changes. Additionally, the WS2 humidity sensor meets the requirements for real-time, non-contact spatial interface detection. As shown in Fig. 5k, the sensor begins to register a slight response when a human finger, acting as a moisture source, is positioned 6 mm above the sensor, with the signal increasing as the finger approaches. This study highlights the potential of solution-processable phase engineering of 2DTMDCs for next-generation sensing technologies. Furthermore, Lim et al. developed a light-driven strategy for phase engineering, overcoming the slow charge transfer limitation encountered under dark conditions.123,124 By leveraging a photo-redox activation reaction pathway under low-power 455 nm illumination, the 2H-to-1T phase transition in few-layer TMDs was accelerated by up to six orders of magnitude. This strategy provides a sustainable and scalable approach for phase engineering, paving the way for large-scale material and device fabrication using light-driven oxidation–reduction reactions.

3.3 Quantum state engineering through atom tailoring of intrinsic 2DTMDCs

Quantum state engineering via atomic tailoring of intrinsic 2DTMDCs enables the modulation of diverse quantum properties, paving the way for their applications in next-generation energy and electronic devices. In this section, we summarize several representative strategies for atomically tailoring intrinsic 2DTMDCs to achieve the desired quantum functionalities. Specifically, we discuss three key strategies: atomic reconstruction in intrinsic 2DTMDCs, atomic vacancy design in 2DTMDCs, and grain boundary reconstruction in 2DTMDCs.

Atomic reconstruction in intrinsic 2DTMDCs plays a crucial role in tailoring their electronic and catalytic properties, offering new opportunities for quantum state engineering and energy applications. A notable example of this strategy is demonstrated by Yang et al., who discovered that Nb-terminated 2H Nb1+xS2 surfaces exhibit a hydrogen adsorption free energy close to thermoneutral, significantly enhancing HER activity.125 Cross-sectional ADF-STEM images provide direct evidence of the 2H phase structure in Nb1+xS2 (Fig. 6a). The measured interlayer spacing d (0002) = 6.35 Å (corresponding to c = 12.60 Å) is in good agreement with the theoretical values for Nb1.35S2, confirming the excess Nb incorporation and atomic reconstruction. The synthesis of metallic 2H-phase NbS2 with additional Nb (2H Nb1+xS2, where x ∼ 0.35) exhibits remarkable catalytic performance for the HER, achieving current densities exceeding 5000 mA cm−2 at approximately −420 mV versus a reversible hydrogen electrode (RHE) (Fig. 6b). Furthermore, this material demonstrates exceptional electrochemical stability, with negligible degradation in polarization curves and overpotential even after 10[thin space (1/6-em)]000 cycles (Fig. 6c). These findings highlight atomic reconstruction as a powerful strategy for tuning the electronic, structural, and catalytic properties of 2DTMDCs.


image file: d5ta01124h-f6.tif
Fig. 6 Quantum state engineering through atom tailoring of intrinsic 2DTMDCs. (a–c) Enhanced catalytic performance via a transition metal-terminated activation strategy. (a) Cross-sectional ADF-STEM image of 2H-phase Nb1.35S2 and 3R phase Nb1+xS2. (b) Polarization curves comparing 2H MoS2, 1T MoS2 and WS2, 2H Nb1.35S2, 3R Nb1+xS2, 2H NbS2 and 3R NbS2, and Pt in 0.5 M H2SO4. (c) Stability test showing polarization curves of 2H Nb1.35S2 before (black) and after (red) 10[thin space (1/6-em)]000 cycles, with the inset displaying a two-electrode electrolyzer using a Pt anode and a 2H Nb1.35S2 cathode. (d–f) Enhanced catalytic performance via an atomic vacancy activation strategy. (d) Partial charge distribution of bilayer PtTe2-3Te and PtTe2–1Te, with an iso-surface level of 0.002 e Å−3. (e) LSV curves of bulk PtTe2 crystals, PtTe2 NSs, PtTe2-200 NSs, PtTe2-400 NSs, PtTe2-600 NSs, and Pt/C, highlighting improved catalytic activity. (f) Overpotential comparison at 10 mA cm−2 and exchange current density (J0 ECSA), demonstrating enhanced activity in vacancy-engineered catalysts. (g–i) Enhanced catalytic performance via a GB activation strategy. (g) ADF-STEM image of MoS2 grains, showing the GBs between three MoS2 domains. (h) Polarization curves of MoS2-based catalysts. (i) Tafel plots, confirming improved HER kinetics in GB-activated MoS2. Panels reproduced with permission from: (a–c) ref. 125, copyright 2019, Springer Nature Limited; (d–f) ref.98, copyright 2021, Springer Nature Limited; (g–i) ref.126, copyright 2020, Springer Nature Limited.

Atomic vacancy engineering is a powerful approach for modulating the electronic structure and catalytic activity of 2DTMDCs. Utilizing a straightforward exfoliation and thermal annealing method, Li et al. introduced vacancy defects into PtTe2, elucidating the relationship between atomically defined Pt sites, adsorption energy, and enhanced HER activity.98 During the thermal annealing process, Te single-atom vacancies (SAVs) in exfoliated PtTe2 NSs migrate and self-assemble into ordered triangular Te-SAV clusters (Fig. 6d), effectively lowering the hydrogen adsorption energy and improving HER kinetics.98 Benefiting from the atomic-scale thickness of PtTe2, the coordinatively unsaturated Pt sites remain highly exposed and stable, leading to superior HER activity and durability compared to commercial Pt/C catalysts. Among these structures, PtTe2-600 NSs demonstrate outstanding HER performance, exhibiting an ultra-low onset potential (∼0 mV) and an overpotential of 22 mV at η = 10 mA cm−2, surpassing the performance of the best 20 wt% Pt/C catalyst (26 mV at η = 10 mA cm−2, Fig. 6e and f). At the same potential, the mass activity of Pt in PtTe2-600 NSs significantly exceeds that of Pt/C. Specifically, at −0.2 V vs. RHE, the mass activity of Pt in PtTe2-600 NSs reaches 1.55 A per mg Pt, whereas the value for Pt in Pt/C is only 1.13 A per mg Pt. This approach offers an efficient pathway for developing high-performance, atomically engineered electrocatalysts with enhanced activity and stability.127

Grain boundaries (GBs) are intrinsic structural defects in atomically thin or so-called 2D polycrystalline materials and can be described as line defects. Grain boundary reconstruction strategies play a crucial role in shaping the mechanical strength, optoelectronic properties, and catalytic performance of 2D materials for various applications.38,94 He et al. utilized a climb-and-driven 0D/2D interaction growth mechanism, employing gold quantum dot-assisted vapor-phase growth to fabricate wafer-scale, atomically thin TMD films with sub-10 nm grain sizes, exhibiting an exceptionally high GB density of up to ∼1012 cm−2.128 Due to its high GB density and superior catalytic activity, the nanocrystalline thin film exhibits outstanding HER performance, with an onset potential of −25 mV and a Tafel slope of 54 mV dec−1, indicating efficient catalytic kinetics. Beyond electrocatalysis, these nanograined films hold promising applications in resistive memory devices and molecular sieve membranes, demonstrating their potential for next-generation nanoelectronic and energy-related technologies.129,130

3.4 Quantum state engineering via extrinsic heteroatom incorporation

Extrinsic heteroatom incorporation represents a powerful strategy for tailoring the electronic, magnetic, and catalytic properties of 2DTMDCs. By integrating foreign atoms into the 2D lattice, novel quantum states can be engineered, enabling applications in quantum materials, energy conversion, and advanced electronics. This approach complements intrinsic atomic engineering strategies by introducing additional degrees of freedom for modulating material properties. In this section, we summarize several representative strategies for extrinsic heteroatom incorporation in 2DTMDCs, focusing on three key approaches: extrinsic zero-dimensional (0D) heteroatoms, extrinsic one-dimensional (1D) heteroatoms, and extrinsic 2D heteroatoms.

The incorporation of 0D heteroatoms into 2DTMDCs enables the tuning of their magnetic, electronic, and catalytic properties, offering promising applications in quantum materials and electrocatalysis. Sun et al. developed a room-temperature ferromagnetic Ni1/MoS2 system, where the observed ferromagnetic behavior is attributed to the ferromagnetic coupling between adjacent NiS4 sites embedded within the MoS2 lattice (Fig. 7a).100 Due to the high Ni doping concentration, the ferromagnetic interaction between the two NiMo sites is 30.67 meV/Ni atom stronger than the competing antiferromagnetic coupling, leading to stable ferromagnetism at room temperature. Beyond its magnetic properties, Ni incorporation also enhances the electrocatalytic performance of MoS2. During the OER, strong Ni–S hybridization effectively modulates the spin density on sulfur sites, optimizing the adsorption of O-related intermediate species, thereby promoting O2 formation. Experimental results in 1 M KOH demonstrate a strong magnetoelectric effect in the Ni1/MoS2 system. As the magnetic field increases from 0 to 502 mT, the OER current density increases by 5 to 28 times at 1.7 V and 1.55 V vs. RHE, respectively (Fig. 7b and c). These findings highlight that SASCs exhibit remarkable magnetoelectric coupling effects, significantly accelerating water and seawater electrolysis, and providing new insights into the design of spin-controlled electrocatalysts.133,134


image file: d5ta01124h-f7.tif
Fig. 7 Quantum state engineering via extrinsic heteroatom incorporation. (a–c) Catalytic enhancement via 0D heteroatom engineering: (a) three distinct S sites (S1, S2, and S3) in the NiMo region. (b) LSV curves of Ni1/MoS2 under varying magnetic fields. The inset displays Ni1/MoS2 as an OER catalyst in water-splitting cells at 1.65 V, without (left) and with (right) an applied magnetic field. (c) Giant magnetoelectric effect observed in Ni1/MoS2. (d–f) Enhanced conductivity via 1D heteroatom engineering: (d) schematic illustration of the fabrication process for single-metal-atom chains. (e) False-colored ADF-STEM image of platinum SMACs embedded in monolayer MoS2, with an inset showing a fast Fourier transform of the highlighted region. (f) Room-temperature resistivity vs. channel length, extracted from a two-terminal transmission-line model device, confirming the superior electrical transport properties of 1D heteroatom-engineered MoS2. (g–i) Tunable bandgap via 2D heteroatom engineering: (g) schematic representation of epitaxially aligned Au on MoS2, enabling bandgap modulation. (h) Virtual bright-field image, reconstructed using MoS2 (100) diffraction spots, showing encapsulated Au (dark area) in bilayer MoS2, with a bilayer Moiré pattern (5.5-nm period). (i) DFT-calculated interaction energy of Au with the twisted bilayer MoS2, revealing stacking-dependent bandgap tuning. Panels reproduced with permission from: (a–c) ref.100, copyright 2023, Springer Nature Limited; (d–f) copyright 2022, Springer Nature Limited; (d–g) ref. 131, copyright 2023, Springer Nature Limited; (g–i) ref. 132, copyright 2024, SAAA.

The incorporation of 1D heteroatoms into 2DTMDCs introduces unique quantum transport phenomena and enhances electrical conductivity. Guo et al. developed a CVD method to construct wafer-scale single-metal-atom chain (SMAC) networks within atomically thin films (Fig. 7d and e).131 The resulting atomic chains exhibit an average length of ∼17 nm and a high density exceeding 10 wt%, demonstrating a scalable approach for integrating 1D atomic networks into 2D materials. The electronic delocalization along the platinum atomic chains leads to a coherent metallic 1D channel, forming an interconnected conduction network within the 2D thin-film matrix. This results in a substantial improvement in electrical transport properties, as the resistivity of the SMAC-embedded MoS2 film is approximately two orders of magnitude lower than that of conventional CVD-grown MoS2 (Fig. 7f). The successful large-scale fabrication of air-stable SMAC networks provides a minimal 1D platform for studying fundamental quantum phenomena such as Luttinger liquid behavior, Fermi liquid microstructures, and other theoretical 1D quantum models.135–137 These findings highlight the potential of 1D atomic engineering in 2DTMDCs for advancing quantum transport research and next-generation nanoelectronic devices.

The incorporation of 2D heteroatoms into 2DTMDCs provides an innovative strategy for modulating interlayer interactions, tuning electronic structures, and stabilizing novel 2D materials. Cui et al. demonstrated a method in which nanoscale Au nanoparticles were annealed between two exfoliated hexagonal MoS2 substrates (Fig. 7h).132 The basal planes of these MoS2 layers were misoriented relative to each other, resulting in twist angles ranging from 0° to 60°. Transmission electron microscopy (TEM) studies revealed a twist-dependent atomic configuration of Au between the two MoS2 layers (Fig. 7i and j). The chemical interactions between Au and S atoms were found to be significantly stronger than conventional van der Waals interactions, yet weaker than covalent bonds.138,139 This unique intermediate bonding state not only stabilizes 2D Au structures but also enables precise control over their orientation. This twisted epitaxy strategy leverages the van der Waals spacing in stacked bilayer 2D materials as a nanoscale reaction chamber, providing an effective means to constrain material growth and tune crystal morphology, orientation, and lattice strain.140,141 Further investigations are necessary to elucidate the functional characteristics of these materials and assess their potential applications in areas such as quantum electronics, catalysis, and strain-engineered 2D devices.142

4. Perspective

Two-dimensional materials exhibit unique physical properties, and while many intriguing quantum phenomena have been observed in 2D systems, the synthesis of 2D materials and the precise modulation of their quantum states—particularly for large-area 2DTMDCs—remain active areas of research. The exploration of novel 2DTMDC-based materials, along with their potential applications in next-generation semiconductor chips, topological superconductors for quantum computing, room-temperature magnetic materials for high-density data storage, and high-performance materials for energy conversion, represents a critical direction for future research in the post-Moore's Law era. Based on these insights, we outline the following research perspectives.

4.1 Exploring novel quantum states for targeted applications

A wide range of 2DTMDCs with novel quantum states have been successfully synthesized. A deeper investigation into their physical properties and potential applications will be essential for advancing the field and unlocking their economic value. Identifying the most suitable applications for these materials will be keys to their further development and commercialization.

4.2 Scalable and industry-compatible synthesis of functionalized 2DTMDCs

Developing industrially viable methods for the large-scale synthesis and functionalization of 2DTMDCs is of paramount importance. Approaches such as low-temperature growth and simplified fabrication routes offer promising strategies for making 2DTMDCs more accessible for practical applications. Establishing scalable manufacturing processes will be crucial for their integration into next-generation electronic and energy devices.

4.3 Advancing 3D stacked growth for next-generation chip materials

Initial progress has been achieved in the three-dimensional (3D) stacked growth of 2DTMDCs.106,107 Moving forward, the development of more complex 3D integration strategies for 2DTMDCs represents a viable pathway toward next-generation chip materials in the post-Moore's Law era. The exploration of advanced 3D stacking techniques will be critical for realizing high-performance electronic and computing devices.

By addressing these key challenges, the field of 2DTMDCs can move closer to practical and large-scale applications, paving the way for breakthroughs in quantum technology, high-performance computing, and sustainable energy solutions.

Data availability

The original contributions presented in this study are included in the article; further inquiries can be directed to the corresponding authors.

Author contributions

H. Lin. proposed the topic of the Review. H. Lin and Y. Meng wrote the manuscript. All the authors contributed to the scientific discussion.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work was supported by Low Carbon Energy Research grant LCERFI01-0033.

References

  1. M. Chhowalla, H. S. Shin, G. Eda, L.-J. Li, K. P. Loh and H. Zhang, Nat. Chem., 2013, 5, 263 CrossRef PubMed .
  2. S. Manzeli, D. Ovchinnikov, D. Pasquier, O. V. Yazyev and A. Kis, Nat. Rev. Mater., 2017, 2, 17033 CrossRef CAS .
  3. S. B. Desai, S. R. Madhvapathy, A. B. Sachid, J. P. Llinas, Q. Wang, G. H. Ahn, G. Pitner, M. J. Kim, J. Bokor, C. Hu, H.-S. P. Wong and A. Javey, Science, 2016, 354, 99–102 CrossRef CAS PubMed .
  4. W. Cao, H. Bu, M. Vinet, M. Cao, S. Takagi, S. Hwang, T. Ghani and K. Banerjee, Nature, 2023, 620, 501–515 CrossRef CAS PubMed .
  5. C. Liu, H. Chen, S. Wang, Q. Liu, Y.-G. Jiang, D. W. Zhang, M. Liu and P. Zhou, Nat. Nanotechnol., 2020, 15, 545–557 CrossRef CAS PubMed .
  6. J. M. Lu, O. Zheliuk, I. Leermakers, N. F. Q. Yuan, U. Zeitler, K. T. Law and J. T. Ye, Science, 2015, 350, 1353–1357 CrossRef CAS PubMed .
  7. H. Ji, Y. Liu, C. Ji and J. Wang, Acc. Mater. Res., 2024, 5, 1146–1157 CrossRef CAS .
  8. J. Lu, O. Zheliuk, Q. Chen, I. Leermakers, N. E. Hussey, U. Zeitler and J. Ye, Proc. Natl. Acad. Sci. U. S. A., 2018, 115, 3551–3556 CrossRef CAS PubMed .
  9. W. Li, J. Huang, X. Li, S. Zhao, J. Lu, Z. V. Han and H. Wang, Mater. Today Phys., 2021, 21, 100504 CrossRef CAS .
  10. C.-S. Lian, C. Si and W. Duan, Nano Lett., 2018, 18, 2924–2929 CrossRef CAS PubMed .
  11. A. A. Balandin, S. V. Zaitsev-Zotov and G. Grüner, Appl. Phys. Lett., 2021, 119, 170401 CrossRef CAS .
  12. D. Lin, S. Li, J. Wen, H. Berger, L. Forró, H. Zhou, S. Jia, T. Taniguchi, K. Watanabe, X. Xi and M. S. Bahramy, Nat. Commun., 2020, 11, 2406 CrossRef CAS PubMed .
  13. R. Bianco, I. Errea, L. Monacelli, M. Calandra and F. Mauri, Nano Lett., 2019, 19, 3098–3103 CrossRef CAS PubMed .
  14. J.-P. So, H.-R. Kim, H. Baek, K.-Y. Jeong, H.-C. Lee, W. Huh, Y. S. Kim, K. Watanabe, T. Taniguchi, J. Kim, C.-H. Lee and H.-G. Park, Sci. Adv., 2021, 7, eabj3176 CrossRef CAS PubMed .
  15. J.-P. So, K.-Y. Jeong, J. M. Lee, K.-H. Kim, S.-J. Lee, W. Huh, H.-R. Kim, J.-H. Choi, J. M. Kim, Y. S. Kim, C.-H. Lee, S. Nam and H.-G. Park, Nano Lett., 2021, 21, 1546–1554 CrossRef CAS PubMed .
  16. S. Gupta, W. Wu, S. Huang and B. I. Yakobson, J. Phys. Chem. Lett., 2023, 14, 3274–3284 CrossRef CAS PubMed .
  17. H. Zhao, M. T. Pettes, Y. Zheng and H. Htoon, Nat. Commun., 2021, 12, 6753 CrossRef CAS PubMed .
  18. A. Elbanna, H. Jiang, Q. Fu, J.-F. Zhu, Y. Liu, M. Zhao, D. Liu, S. Lai, X. W. Chua, J. Pan, Z. X. Shen, L. Wu, Z. Liu, C.-W. Qiu and J. Teng, ACS Nano, 2023, 17, 4134–4179 CrossRef CAS PubMed .
  19. M.-G. Lee, S. Yoo, T. Kim and Q. H. Park, Nanoscale, 2017, 9, 16244–16248 RSC .
  20. S. Bang, N. T. Duong, J. Lee, Y. H. Cho, H. M. Oh, H. Kim, S. J. Yun, C. Park, M.-K. Kwon, J.-Y. Kim, J. Kim and M. S. Jeong, Nano Lett., 2018, 18, 2316–2323 CrossRef CAS PubMed .
  21. K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang, S. V. Dubonos, I. V. Grigorieva and A. A. Firsov, Science, 2004, 306, 666–669 CrossRef CAS PubMed .
  22. X. Xi, Z. Wang, W. Zhao, J.-H. Park, K. T. Law, H. Berger, L. Forró, J. Shan and K. F. Mak, Nat. Phys., 2016, 12, 139 Search PubMed .
  23. B. Radisavljevic, A. Radenovic, J. Brivio, V. Giacometti and A. Kis, Nat. Nanotechnol., 2011, 6, 147–150 CrossRef CAS PubMed .
  24. A. Splendiani, L. Sun, Y. Zhang, T. Li, J. Kim, C.-Y. Chim, G. Galli and F. Wang, Nano Lett., 2010, 10, 1271–1275 CrossRef CAS PubMed .
  25. K. F. Mak, C. Lee, J. Hone, J. Shan and T. F. Heinz, Phys. Rev. Lett., 2010, 105, 136805 CrossRef PubMed .
  26. S. J. Haigh, A. Gholinia, R. Jalil, S. Romani, L. Britnell, D. C. Elias, K. S. Novoselov, L. A. Ponomarenko, A. K. Geim and R. Gorbachev, Nat. Mater., 2012, 11, 764–767 CrossRef CAS PubMed .
  27. A. K. Geim and I. V. Grigorieva, Nature, 2013, 499, 419–425 CrossRef CAS PubMed .
  28. H. Fang, C. Battaglia, C. Carraro, S. Nemsak, B. Ozdol, J. S. Kang, H. A. Bechtel, S. B. Desai, F. Kronast, A. A. Unal, G. Conti, C. Conlon, G. K. Palsson, M. C. Martin, A. M. Minor, C. S. Fadley, E. Yablonovitch, R. Maboudian and A. Javey, Proc. Natl. Acad. Sci. U. S. A., 2014, 111, 6198–6202 CrossRef CAS PubMed .
  29. Z. Zhou, F. Hou, X. Huang, G. Wang, Z. Fu, W. Liu, G. Yuan, X. Xi, J. Xu, J. Lin and L. Gao, Nature, 2023, 621, 499–505 CrossRef CAS PubMed .
  30. Q. Qian, H. Ren, J. Zhou, Z. Wan, J. Zhou, X. Yan, J. Cai, P. Wang, B. Li and Z. Sofer, Nature, 2022, 606, 902–908 CrossRef CAS PubMed .
  31. B. Zhao, Z. Wan, Y. Liu, J. Xu, X. Yang, D. Shen, Z. Zhang, C. Guo, Q. Qian and J. Li, Nature, 2021, 591, 385–390 CrossRef CAS PubMed .
  32. J. Li, P. Song, J. Zhao, K. Vaklinova, X. Zhao, Z. Li, Z. Qiu, Z. Wang, L. Lin, M. Zhao, T. S. Herng, Y. Zuo, W. Jonhson, W. Yu, X. Hai, P. Lyu, H. Xu, H. Yang, C. Chen, S. J. Pennycook, J. Ding, J. Teng, A. H. Castro Neto, K. S. Novoselov and J. Lu, Nat. Mater., 2021, 20, 181–187 CrossRef CAS PubMed .
  33. Y. Liu, N. O. Weiss, X. Duan, H.-C. Cheng, Y. Huang and X. Duan, Nat. Rev. Mater., 2016, 1, 16042 CrossRef CAS .
  34. K. Kang, K.-H. Lee, Y. Han, H. Gao, S. Xie, D. A. Muller and J. Park, Nature, 2017, 550, 229–233 CrossRef PubMed .
  35. D. R. Klein, D. MacNeill, J. L. Lado, D. Soriano, E. Navarro-Moratalla, K. Watanabe, T. Taniguchi, S. Manni, P. Canfield, J. Fernández-Rossier and P. Jarillo-Herrero, Science, 2018, 360, 1218–1222 CrossRef CAS PubMed .
  36. Y. Cao, V. Fatemi, S. Fang, K. Watanabe, T. Taniguchi, E. Kaxiras and P. Jarillo-Herrero, Nature, 2018, 556, 43–50 CrossRef CAS PubMed .
  37. H. Wang, X. Huang, J. Lin, J. Cui, Y. Chen, C. Zhu, F. Liu, Q. Zeng, J. Zhou, P. Yu, X. Wang, H. He, S. H. Tsang, W. Gao, K. Suenaga, F. Ma, C. Yang, L. Lu, T. Yu, E. H. T. Teo, G. Liu and Z. Liu, Nat. Commun., 2017, 8, 394 CrossRef PubMed .
  38. S. Najmaei, Z. Liu, W. Zhou, X. Zou, G. Shi, S. Lei, B. I. Yakobson, J.-C. Idrobo, P. M. Ajayan and J. Lou, Nat. Mater., 2013, 12, 754–759 CrossRef CAS PubMed .
  39. Y.-H. Lee, X.-Q. Zhang, W. Zhang, M.-T. Chang, C.-T. Lin, K.-D. Chang, Y.-C. Yu, J. T.-W. Wang, C.-S. Chang, L.-J. Li and T.-W. Lin, Adv. Mater., 2012, 24, 2320–2325 CrossRef CAS PubMed .
  40. J. Shi, X. Wang, S. Zhang, L. Xiao, Y. Huan, Y. Gong, Z. Zhang, Y. Li, X. Zhou, M. Hong, Q. Fang, Q. Zhang, X. Liu, L. Gu, Z. Liu and Y. Zhang, Nat. Commun., 2017, 8, 958 CrossRef PubMed .
  41. R. Yang, L. Mei, Q. Zhang, Y. Fan, H. S. Shin, D. Voiry and Z. Zeng, Nat. Protoc., 2022, 17, 358–377 CrossRef CAS PubMed .
  42. Y. Gong, H. Yuan, C.-L. Wu, P. Tang, S.-Z. Yang, A. Yang, G. Li, B. Liu, J. van de Groep, M. L. Brongersma, M. F. Chisholm, S.-C. Zhang, W. Zhou and Y. Cui, Nat. Nanotechnol., 2018, 13, 294–299 CrossRef CAS PubMed .
  43. Z. Zeng, Z. Yin, X. Huang, H. Li, Q. He, G. Lu, F. Boey and H. Zhang, Angew. Chem., Int. Ed., 2011, 50, 11093–11097 CrossRef CAS PubMed .
  44. Z. Zeng, T. Sun, J. Zhu, X. Huang, Z. Yin, G. Lu, Z. Fan, Q. Yan, H. H. Hng and H. Zhang, Angew. Chem., Int. Ed., 2012, 51, 9052–9056 CrossRef CAS PubMed .
  45. W. Yu, J. Li, T. S. Herng, Z. Wang, X. Zhao, X. Chi, W. Fu, I. Abdelwahab, J. Zhou and J. Dan, Adv. Mater., 2019, 31, 1903779 CrossRef CAS PubMed .
  46. J. Peng, Y. Liu, X. Luo, J. Wu, Y. Lin, Y. Guo, J. Zhao, X. Wu, C. Wu and Y. Xie, Adv. Mater., 2019, 31, 1900568 CrossRef PubMed .
  47. Z. Lin, Z. Wan, F. Song, B. Huang, C. Jia, Q. Qian, J. S. Kang, Y. Wu, X. Yan, L. Peng, C. Wan, J. Zhou, Z. Sofer, I. Shakir, Z. Almutairi, S. Tolbert, X. Pan, Y. Hu, Y. Huang and X. Duan, Chem, 2021, 7, 1887–1902 CAS .
  48. W. Yu, Z. Wang, X. Zhao, J. Wang, T. S. Herng, T. Ma, Z. Zhu, J. Ding, G. Eda, S. J. Pennycook, Y. P. Feng and K. P. Loh, Adv. Funct. Mater., 2020, 30, 2003057 CrossRef CAS .
  49. H. Lin, Q. Zhu, D. Shu, D. Lin, J. Xu, X. Huang, W. Shi, X. Xi, J. Wang and L. Gao, Nat. Mater., 2019, 18, 602–607 CrossRef CAS PubMed .
  50. X. Meng, A. Pant, H. Cai, J. Kang, H. Sahin, B. Chen, K. Wu, S. Yang, A. Suslu, F. M. Peeters and S. Tongay, Nanoscale, 2015, 7, 17109–17115 RSC .
  51. J. Shi, Y. Huan, M. Xiao, M. Hong, X. Zhao, Y. Gao, F. Cui, P. Yang, S. J. Pennycook, J. Zhao and Y. Zhang, ACS Nano, 2020, 14, 9011–9020 CrossRef CAS PubMed .
  52. S. Vranic, R. Kurapati, K. Kostarelos and A. Bianco, Nat. Rev. Chem, 2025, 9, 173–184 CrossRef CAS PubMed .
  53. P. Joensen, R. Frindt and S. R. Morrison, Mater. Res. Bull., 1986, 21, 457–461 CrossRef CAS .
  54. Z. Zeng, Z. Yin, X. Huang, H. Li, Q. He, G. Lu, F. Boey and H. Zhang, Angew. Chem., Int. Ed., 2011, 47, 11093–11097 CrossRef PubMed .
  55. R. Yang, Y. Fan, L. Mei, H. S. Shin, D. Voiry, Q. Lu, J. Li and Z. Zeng, Nat. Synth., 2023, 2, 101–118 CrossRef CAS .
  56. R. Yang, L. Mei, Z. Lin, Y. Fan, J. Lim, J. Guo, Y. Liu, H. S. Shin, D. Voiry and Q. Lu, Nat. Rev. Chem, 2024, 1–23 Search PubMed .
  57. G. Eda, H. Yamaguchi, D. Voiry, T. Fujita, M. Chen and M. Chhowalla, Nano Lett., 2011, 11, 5111–5116 CrossRef CAS PubMed .
  58. L. Ries, E. Petit, T. Michel, C. C. Diogo, C. Gervais, C. Salameh, M. Bechelany, S. Balme, P. Miele and N. Onofrio, Nat. Mater., 2019, 18, 1112–1117 CrossRef CAS PubMed .
  59. J. Peng, J. Wu, X. Li, Y. Zhou, Z. Yu, Y. Guo, J. Wu, Y. Lin, Z. Li and X. Wu, J. Am. Chem. Soc., 2017, 139, 9019–9025 Search PubMed .
  60. L. Wu, H. Lin, X. Cao, Q. Tong, F. Yang, Y. Miao, D. Ye and Q. Fan, Angew. Chem., Int. Ed., 2024, e202405937 CAS .
  61. Z. Lin, Y. Liu, U. Halim, M. Ding, Y. Liu, Y. Wang, C. Jia, P. Chen, X. Duan, C. Wang, F. Song, M. Li, C. Wan, Y. Huang and X. Duan, Nature, 2018, 562, 254–258 CrossRef CAS PubMed .
  62. L. Zhang, Z. Yang, S. Feng, Z. Guo, Q. Jia, H. Zeng, Y. Ding, P. Das, Z. Bi and J. Ma, Nature, 2024, 628, 313–319 Search PubMed .
  63. J. Peng, Y. Liu, X. Luo, J. Wu, Y. Lin, Y. Guo, J. Zhao, X. Wu, C. Wu and Y. Xie, Adv. Mater., 2019, 31, 1900568 CrossRef PubMed .
  64. H.-L. Tsai, J. Heising, J. L. Schindler, C. R. Kannewurf and M. G. Kanatzidis, Chem. Mater., 1997, 9, 879–882 CrossRef CAS .
  65. L. Huihui, L. Yiwen, L. Yuxin, W. Luyan, C. Zhaolong and L. Jing, Natl. Sci. open, 2025, 4, 20240043 Search PubMed .
  66. S. Zhang, L. Zhang and Z.-S. Wu, ACS Nano, 2024, 18, 24617–24621 CrossRef CAS PubMed .
  67. Y. Zhai, Z. Shi, Q. Xia, W. Han, W. Li, X. Deng and X. Zhang, ACS Nano, 2024, 18, 26477–26502 CrossRef CAS PubMed .
  68. Y. Bai, Y. Xu, L. Sun, Z. Ward, H. Wang, G. Ratnayake, C. Wang, M. Zhao, H. He, J. Gao, M. Wu, S. Lu, G. Bepete, D. Peng, B. Liu, F. Kang, H. Terrones, M. Terrones and Y. Lei, Adv. Mater., 2025, 37, 2416375 CrossRef CAS PubMed .
  69. V. Nicolosi, M. Chhowalla, M. G. Kanatzidis, M. S. Strano and J. N. Coleman, Science, 2013, 340, 1226419 CrossRef .
  70. H. Wang, S. Chen, H. Li, X. Chen, J. Cheng, Y. Shao, C. Zhang, J. Zhang, L. Fan and H. Chang, Adv. Funct. Mater., 2021, 31, 2100274 CrossRef CAS .
  71. S. Wang, J. Xue, D. Xu, J. He, Y. Dai, T. Xia, Y. Huang, Q. He, X. Duan and Z. Lin, Nat. Protoc., 2023, 18, 2814–2837 CrossRef CAS PubMed .
  72. C. Wang, Q. He, U. Halim, Y. Liu, E. Zhu, Z. Lin, H. Xiao, X. Duan, Z. Feng and R. Cheng, Nature, 2018, 555, 231–236 CrossRef CAS PubMed .
  73. H. Ren, X. Xia, Y. Sun, Y. Zhai, Z. Zhang, J. Wu, J. Li and M. Liu, J. Mater. Chem. A, 2024, 12, 23416–23424 RSC .
  74. Q. Hao, J. Liu, G. Wang, J. Chen, H. Gan, J. Zhu, Y. Ke, Y. Chai, J. Lin and W. Zhang, ACS Nano, 2020, 14, 11373–11382 CrossRef CAS PubMed .
  75. S. Wang, W. Li, J. Xue, J. Ge, J. He, J. Hou, Y. Xie, Y. Li, H. Zhang and Z. Sofer, Nat. Commun., 2024, 15, 6388 CrossRef CAS PubMed .
  76. B. Tang, M. Sivan, J. F. Leong, Z. Xu, Y. Zhang, J. Li, R. Wan, Q. Wan, E. Zamburg and A. V. Thean, npj 2D Mater. Appl., 2024, 8, 74 CrossRef .
  77. W. Kong, Z. Wang, M. Wang, K. C. Manning, A. Uppal, M. D. Green, R. Y. Wang and K. Rykaczewski, Adv. Mater., 2019, 31, 1904309 CrossRef CAS PubMed .
  78. C. Wang, Y. Gong, B. V. Cunning, S. Lee, Q. Le, S. R. Joshi, O. Buyukcakir, H. Zhang, W. K. Seong, M. Huang, M. Wang, J. Lee, G.-H. Kim and R. S. Ruoff, Sci. Adv., 2021, 7, eabe3767 CrossRef CAS PubMed .
  79. T. Zhang, J. Wang, P. Wu, A.-Y. Lu and J. Kong, Nat. Rev. Mater., 2023, 8, 799–821 CrossRef CAS .
  80. X. Wang, H. Wang, L. Ma, L. Zhang, Z. Yang, D. Dong, X. Chen, H. Li, Y. Guan and B. Zhang, Nat. Commun., 2023, 14, 4282 CrossRef CAS PubMed .
  81. H. Jiang, X. Zhang, K. Chen, X. He, Y. Liu, H. Yu, L. Gao, M. Hong, Y. Wang and Z. Zhang, Nat. Mater., 2025, 24, 188–196 CrossRef CAS PubMed .
  82. Z. Yan, Z. Peng and J. M. Tour, Acc. Chem. Res., 2014, 47, 1327–1337 CrossRef CAS PubMed .
  83. T.-A. Chen, C.-P. Chuu, C.-C. Tseng, C.-K. Wen, H.-S. P. Wong, S. Pan, R. Li, T.-A. Chao, W.-C. Chueh and Y. Zhang, Nature, 2020, 579, 219–223 CrossRef CAS PubMed .
  84. J. Zhang, L. Lin, K. Jia, L. Sun, H. Peng and Z. Liu, Adv. Mater., 2020, 32, 1903266 CrossRef CAS PubMed .
  85. W. Zhang, J. K. Huang, C. H. Chen, Y. H. Chang, Y. J. Cheng and L. J. Li, Adv. Mater., 2013, 25, 3456–3461 CrossRef CAS PubMed .
  86. S. Li, Y.-C. Lin, W. Zhao, J. Wu, Z. Wang, Z. Hu, Y. Shen, D.-M. Tang, J. Wang, Q. Zhang, H. Zhu, L. Chu, W. Zhao, C. Liu, Z. Sun, T. Taniguchi, M. Osada, W. Chen, Q.-H. Xu, A. T. S. Wee, K. Suenaga, F. Ding and G. Eda, Nat. Mater., 2018, 17, 535–542 CrossRef CAS PubMed .
  87. S. Li, S. Wang, D.-M. Tang, W. Zhao, H. Xu, L. Chu, Y. Bando, D. Golberg and G. Eda, Appl. Mater. Today, 2015, 1, 60–66 CrossRef .
  88. J. Zhou, J. Lin, X. Huang, Y. Zhou, Y. Chen, J. Xia, H. Wang, Y. Xie, H. Yu and J. Lei, Nature, 2018, 556, 355–359 CrossRef CAS PubMed .
  89. S. Li, iScience, 2021, 24, 103229 CrossRef CAS PubMed .
  90. M.-C. Chang, P.-H. Ho, M.-F. Tseng, F.-Y. Lin, C.-H. Hou, I.-K. Lin, H. Wang, P.-P. Huang, C.-H. Chiang and Y.-C. Yang, Nat. Commun., 2020, 11, 3682 Search PubMed .
  91. H. Suzuki, R. Hashimoto, M. Misawa, Y. Liu, M. Kishibuchi, K. Ishimura, K. Tsuruta, Y. Miyata and Y. Hayashi, ACS Nano, 2022, 16, 11360–11373 CrossRef CAS PubMed .
  92. H. Lin, M. Chang, X. Fu, P. Li, M. Chen, L. Wu, F. Yang and Q. Zhang, Molecules, 2023, 28, 1059 CrossRef CAS PubMed .
  93. L. Ma, X. Wang, H. Wang, X. Wang, G. Zou, Y. Guan, S. Guo, H. Li, Q. Chen and L. Kang, Nano Lett., 2023, 23, 6892–6899 CrossRef CAS PubMed .
  94. D. Rhodes, S. H. Chae, R. Ribeiro-Palau and J. Hone, Nat. Mater., 2019, 18, 541–549 CrossRef CAS PubMed .
  95. A. W. Tsen, L. Brown, M. P. Levendorf, F. Ghahari, P. Y. Huang, R. W. Havener, C. S. Ruiz-Vargas, D. A. Muller, P. Kim and J. Park, Science, 2012, 336, 1143–1146 CrossRef CAS PubMed .
  96. D. Wang, H. Yu, L. Tao, W. Xiao, P. Fan, T. Zhang, M. Liao, W. Guo, D. Shi and S. Du, Nano Res., 2018, 11, 6102–6109 CrossRef CAS .
  97. H. Fang, H. Mahalingam, X. Li, X. Han, Z. Qiu, Y. Han, K. Noori, D. Dulal, H. Chen and P. Lyu, Nat. Nanotechnol., 2023, 18, 1401–1408 CrossRef CAS PubMed .
  98. X. Li, Y. Fang, J. Wang, H. Fang, S. Xi, X. Zhao, D. Xu, H. Xu, W. Yu and X. Hai, Nat. Commun., 2021, 12, 2351 CrossRef CAS PubMed .
  99. Z. Shi, X. Zhang, X. Lin, G. Liu, C. Ling, S. Xi, B. Chen, Y. Ge, C. Tan and Z. Lai, Nature, 2023, 621, 300–305 CrossRef CAS PubMed .
  100. T. Sun, Z. Tang, W. Zang, Z. Li, J. Li, Z. Li, L. Cao, J. S. Dominic Rodriguez, C. O. M. Mariano and H. Xu, Nat. Nanotechnol., 2023, 18, 763–771 CrossRef CAS PubMed .
  101. Y. Yu, G.-H. Nam, Q. He, X.-J. Wu, K. Zhang, Z. Yang, J. Chen, Q. Ma, M. Zhao and Z. Liu, Nat. Chem., 2018, 10, 638–643 CrossRef CAS PubMed .
  102. L. Liu, J. Wu, L. Wu, M. Ye, X. Liu, Q. Wang, S. Hou, P. Lu, L. Sun and J. Zheng, Nat. Mater., 2018, 17, 1108–1114 CrossRef CAS PubMed .
  103. H. Li, Y. Lin, J. Duan, Q. Wen, Y. Liu and T. Zhai, Chem. Soc. Rev., 2024, 53, 10709 RSC .
  104. Q. Yun, Y. Ge, Z. Shi, J. Liu, X. Wang, A. Zhang, B. Huang, Y. Yao, Q. Luo and L. Zhai, Chem. Rev., 2023, 123, 13489–13692 CrossRef CAS PubMed .
  105. W.-G. Cui, F. Gao, G. Na, X. Wang, Z. Li, Y. Yang, Z. Niu, Y. Qu, D. Wang and H. Pan, Chem. Soc. Rev., 2024, 53, 10253 RSC .
  106. Y. Zhao, C. Zhang, D. D. Kohler, J. M. Scheeler, J. C. Wright, P. M. Voyles and S. Jin, Science, 2020, 370, 442–445 CrossRef CAS PubMed .
  107. K. S. Kim, S. Seo, J. Kwon, D. Lee, C. Kim, J.-E. Ryu, J. Kim, J. M. Suh, H.-G. Jung, Y. Jo, J.-C. Shin, M.-K. Song, J. Feng, H. Ahn, S. Lee, K. Cho, J. Jeon, M. Seol, J.-H. Park, S. W. Kim and J. Kim, Nature, 2024, 636, 615–621 CrossRef CAS PubMed .
  108. R. Wu, H. Zhang, H. Ma, B. Zhao, W. Li, Y. Chen, J. Liu, J. Liang, Q. Qin and W. Qi, Chem. Rev., 2024, 124, 10112–10191 CrossRef CAS PubMed .
  109. G. Xue, B. Qin, C. Ma, P. Yin, C. Liu and K. Liu, Chem. Rev., 2024, 124, 9785–9865 CrossRef CAS PubMed .
  110. L. Zhang, J. Dong and F. Ding, Chem. Rev., 2021, 121, 6321–6372 CrossRef CAS PubMed .
  111. R. Bian, C. Li, Q. Liu, G. Cao, Q. Fu, P. Meng, J. Zhou, F. Liu and Z. Liu, Natl. Sci. Rev., 2022, 9, nwab164 CrossRef CAS PubMed .
  112. J. Dong, L. Liu, C. Tan, Q. Xu, J. Zhang, Z. Qiao, D. Chu, Y. Liu, Q. Zhang, J. Jiang, Y. Han, A. P. Davis and Y. Cui, Nature, 2022, 602, 606–611 CrossRef CAS PubMed .
  113. M. Zhao, C. Casiraghi and K. Parvez, Chem. Soc. Rev., 2024, 53, 3036–3064 RSC .
  114. R. Pendurthi, N. U. Sakib, M. U. K. Sadaf, Z. Zhang, Y. Sun, C. Chen, D. Jayachandran, A. Oberoi, S. Ghosh, S. Kumari, S. P. Stepanoff, D. Somvanshi, Y. Yang, J. M. Redwing, D. E. Wolfe and S. Das, Nat. Nanotechnol., 2024, 19, 970–977 CrossRef CAS PubMed .
  115. X. Cui, Z. Kong, E. Gao, D. Huang, Y. Hao, H. Shen, C.-a. Di, Z. Xu, J. Zheng and D. Zhu, Nat. Commun., 2018, 9, 1301 CrossRef PubMed .
  116. Y. Wu, D. Li, C.-L. Wu, H. Y. Hwang and Y. Cui, Nat. Rev. Mater., 2023, 8, 41–53 CrossRef .
  117. Y. Liu, Y. Huang and X. Duan, Nature, 2019, 567, 323–333 CrossRef CAS PubMed .
  118. Z. Wan, G. Qiu, H. Ren, Q. Qian, Y. Li, D. Xu, J. Zhou, J. Zhou, B. Zhou and L. Wang, Nature, 2024, 632, 69–74 CrossRef CAS PubMed .
  119. L. Mei, Z. Gao, R. Yang, Z. Zhang, M. Sun, X. Liang, Y. Zhang, T. Ying, H. Hu and D. Li, Nat. Synth., 2024, 4, 303–313 CrossRef .
  120. A. Castellanos-Gomez, X. Duan, Z. Fei, H. R. Gutierrez, Y. Huang, X. Huang, J. Quereda, Q. Qian, E. Sutter and P. Sutter, Nat. Rev. Methods Primers, 2022, 2, 58 CrossRef CAS .
  121. S. Aftab, M. Z. Iqbal and Y. S. Rim, Small, 2023, 19, 2205418 CrossRef CAS PubMed .
  122. D. Voiry, H. Yamaguchi, J. Li, R. Silva, D. C. Alves, T. Fujita, M. Chen, T. Asefa, V. B. Shenoy and G. Eda, Nat. Mater., 2013, 12, 850–855 CrossRef CAS PubMed .
  123. J. Lim, J.-I. Lee, Y. Wang, N. Gauriot, E. Sebastian, M. Chhowalla, C. Schnedermann and A. Rao, Nature, 2024, 633, 83–89 Search PubMed .
  124. X. Yin, C. S. Tang, Y. Zheng, J. Gao, J. Wu, H. Zhang, M. Chhowalla, W. Chen and A. T. Wee, Chem. Soc. Rev., 2021, 50, 10087–10115 Search PubMed .
  125. J. Yang, A. R. Mohmad, Y. Wang, R. Fullon, X. Song, F. Zhao, I. Bozkurt, M. Augustin, E. J. Santos and H. S. Shin, Nat. Mater., 2019, 18, 1309–1314 Search PubMed .
  126. Y. He, P. Tang, Z. Hu, Q. He, C. Zhu, L. Wang, Q. Zeng, P. Golani, G. Gao, W. Fu, Z. Huang, C. Gao, J. Xia, X. Wang, X. Wang, C. Zhu, Q. M. Ramasse, A. Zhang, B. An, Y. Zhang, S. Martí-Sánchez, J. R. Morante, L. Wang, B. K. Tay, B. I. Yakobson, A. Trampert, H. Zhang, M. Wu, Q. J. Wang, J. Arbiol and Z. Liu, Nat. Commun., 2020, 11, 57 Search PubMed .
  127. X. Zhao, Z. H. Levell, S. Yu and Y. Liu, Chem. Rev., 2022, 122, 10675–10709 CrossRef CAS PubMed .
  128. Y. He, P. Tang, Z. Hu, Q. He, C. Zhu, L. Wang, Q. Zeng, P. Golani, G. Gao and W. Fu, Nat. Commun., 2020, 11, 57 CrossRef CAS PubMed .
  129. Y. Ouyang, C. Ling, Q. Chen, Z. Wang, L. Shi and J. Wang, Chem. Mater., 2016, 28, 4390–4396 CrossRef CAS .
  130. F. Liu and Z. Fan, Chem. Soc. Rev., 2023, 52, 1723–1772 RSC .
  131. S. Guo, J. Fu, P. Zhang, C. Zhu, H. Yao, M. Xu, B. An, X. Wang, B. Tang and Y. Deng, Nat. Synth., 2022, 1, 245–253 CrossRef CAS .
  132. Y. Cui, J. Wang, Y. Li, Y. Wu, E. Been, Z. Zhang, J. Zhou, W. Zhang, H. Y. Hwang and R. Sinclair, Science, 2024, 383, 212–219 Search PubMed .
  133. Y. Hao, S.-F. Hung, L. Wang, L. Deng, W.-J. Zeng, C. Zhang, Z.-Y. Lin, C.-H. Kuo, Y. Wang and Y. Zhang, Nat. Commun., 2024, 15, 8015 Search PubMed .
  134. H. T. Tang, H. Y. Zhou, Y. M. Pan, J. L. Zhang, F. H. Cui, W. H. Li and D. Wang, Angew. Chem., Int. Ed., 2024, 136, e202315032 CrossRef .
  135. R. Yang, Y. Fan, J. Hu, Z. Chen, H. S. Shin, D. Voiry, Q. Wang, Q. Lu, C. Y. Jimmy and Z. Zeng, Chem. Soc. Rev., 2023, 52, 7687–7706 RSC .
  136. L. Liu, J. Hu, Z. Ma, Z. Zhu, B. He, F. Chen, Y. Lu, R. Xu, Y. Zhang and T. Ma, Nat. Commun., 2024, 15, 305 CrossRef CAS PubMed .
  137. T. Zhang, D. Wang and J. Liu, Adv. Mater., 2024, 36, 2408259 CrossRef CAS PubMed .
  138. A. Allain, J. Kang, K. Banerjee and A. Kis, Nat. Mater., 2015, 14, 1195–1205 CrossRef CAS PubMed .
  139. K. Reidy, J. D. Thomsen, H. Y. Lee, V. Zarubin, Y. Yu, B. Wang, T. Pham, P. Periwal and F. M. Ross, Nano Lett., 2022, 22, 5849–5858 CrossRef CAS PubMed .
  140. Y.-C. Li, X.-L. Zhang, X.-L. Tai, X.-P. Yang, P.-C. Yu, S.-C. Dong, L.-P. Chi, Z.-Z. Wu and Y.-C. Zhang, Angew. Chem., Int. Ed., 2025, e202422054 CAS .
  141. Y. Bao, J. Shao, H. Xu, J. Yan, P.-T. Jing, J. Xu, D. Zhan, B. Li, K. Liu and L. Liu, ACS Nano, 2024, 18, 27411–27419 CrossRef CAS PubMed .
  142. G. Awiaz, J. Lin and A. Wu, Exploration, 2023, 3, 20220072 CrossRef PubMed .

Footnote

These authors contributed equally to this work.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.