Material-by-design has been a long-standing aspiration that has recently become a reality. Such designer materials have been repeatedly demonstrated using the top-down approach of mechanical exfoliation and stacking, leading to a variety of artificial 2D heterostructures with new properties that are otherwise unattainable. Consequently, tremendous research frontiers in physics, chemistry, engineering, and life science have been created. While thousands of layered crystals exist in nature, only a few dozen of them with manageable chemical-stability have been made into heterostructures using this method. Moreover, experimental investigations of materials that have received limited exploration in the 2D realm, such as cuprates, halides, and perovskites, along with their heterostructures, have been fundamentally hindered by their rapid chemical degradation. Another critical challenge imposed by exfoliating and stacking 2D layers in ambient environment is the absorption of itinerant gas molecules that further contaminate sensitive 2D interfaces in the heterostructures. Such contamination and compromised material properties significantly hinder surface-sensitive local probes—scanning probe microscopy (SPM)—that often require nanometer to atomic scale surface cleanliness. In this article, we aim to provide a technical review of recent development toward 2D materials and heterostructure fabrication in more controlled environments that are suitable for SPM characterizations. These include the development of more efficient mechanical exfoliation and dry-transfer techniques, as well as the incorporation of 2D material exfoliation and transfer in inert gas, low vacuum, and, eventually, ultra-high vacuum environments. Finally, we provide an outlook on the remaining challenges and opportunities in ultra-clean 2D material fabrication techniques.

2D heterostructures have been a major focus of recent physics, chemistry, mechanics, and electronics research. Novel quantum states of matter1 including unconventional2 and possibly topological3 superconductivity have been evidenced in these heterostructures, and 2D heterostructures are expected to be instrumental in a wide variety of applications from excitonic devices4 to chemical sensing.5 2D vertical heterostructures consist of layer-by-layer stacked or synthesized 2D materials, with degrees of freedom in the number of layers, the material composition of those layers, and the relative twist angle θ between each layer. The latter degree of freedom led to a recent surge in the research of Moiré heterostructures6 and their associated physics and chemistry.

Early 2D heterostructure research has primarily focused on materials that are chemically insert in air—particularly graphene and hexagonal boron nitride (hBN). At the same time, most 2D materials discovered are chemically reactive and readily degrade under ambient conditions. This includes most of transition metal dichalcogenides (TMDCs);7 elemental 2D materials8 of boron, silicon, phosphorous, gallium, germanium, arsenic, selenium, tin, antimony, tellurium, thallium, lead, and bismuth; Janus TMDCs such as MoSSe9 or MoSH;10 ferromagnets such as CrI311 or CrBr3,3 as well as a wide variety of layered compounds that are just beginning to be explored in the 2D realm, such as cuprate superconductors.12 2D materials have shown a wide variety of properties, including massless Dirac fermions,13 non-abelian statistics,14 unconventional superconductivity,12 and quantum spin hall effect.15 Undoubtedly, 2D heterostructures offer the potential to achieve designer materials with tailored properties, discover new physics, and develop new functional devices.

There are two general methods for fabricating such heterostructures: bottom-up synthesis and top-down fabrication. Bottom-up synthesis16 involves growing heterostructures layer by layer from its constituent parts through processes such as molecular beam epitaxy17 (MBE), chemical vapor deposition18 (CVD), and solution phase growth.19 This synthesis is beneficial for creating large continuous films20 and so far has been the only approach for creating lateral heterostructures19,21–25 consisting of dissimilar layers in a side-by-side rather than stacked geometry. However, bottom-up synthesis is less flexible as interfacial lattice and/or symmetry matching is often desirable for best growth results. In contrast, top-down approaches, which consist of obtaining (typically through exfoliation) and identifying 2D flakes and then stacking those flakes one at a time, offer much more flexibility,26 especially over the twist angle between each layer and the material species of each layer. As a consequence of the mechanical operations involved in the exfoliation and stacking, a grand challenge in top-down fabrication is preserving the material integrity and preventing surface and interfacial contamination—discoveries of new physics in 2D heterostructures rely on the use of high-quality and pristine materials. Figure 1 shows the chronology of representative development of 2D vertical heterostructures, beginning with the mechanical exfoliation of graphene through to ultra-high vacuum (UHV) assembly. Much of this development has been motivated by the fact that 2D materials are uniquely susceptible to contamination.27 In a single atomic layer, the entire material is exposed to the surface and there is no bulk. If any environmental contamination occurs, the entire material is contaminated. This obviously impacts samples that are chemically reactive in air, but it also can impact samples that are not chemically reactive in air through particle physisorption. For example, after exposure to atmospheric conditions for several days, graphene can be contaminated by environmental hydrocarbons28 [Figs. 2(a) and 2(b)]. The impacts on oxygen and moisture sensitive materials are even more dramatic. Figures 2(c) and 2(d) show black phosphorous samples degrading under ambient conditions over the course of a few days,29, Figs. 2(e)2(h) show CrI3 degrading under humid argon conditions over the course of an hour,30 and Figs. 2(i) and 2(j) show Fe3GeTe2 degrading under ambient conditions over the course of just 5 min.31 Therefore, many top-down fabrication techniques, which are performed under ambient conditions, succeed only with inert materials such as graphene and hBN. Air sensitive materials are typically only accessible using all-in-vacuum processing techniques, such as bottom-up MBE, or glovebox processing. However, there is increasing interest in searching for new exotic quantum states in 2D heterostructures performed by surface-sensitive characterization techniques such as scanning probe microscopy (SPM).32 At the spatiotemporal resolution limit, surface and interface cleanliness is critical. Because both lateral and vertical heterostructures grown via bottom-up synthesis in vacuum environments are broadly compatible with SPM techniques, we limit the scope of our review to advanced top-down 2D heterostructure fabrication techniques given the enormous challenges in achieving atomically pristine surfaces and interfaces. In addition, SPM techniques impose further requirements for the 2D heterostructures; specifically, the heterostructures often need to be electronically contactable with atomically clean and accessible surfaces and interfaces. To achieve such heterostructures, there are four principles a fabrication process should observe: (1) Dry transfer is preferred, although sometimes the use of solvents is unavoidable; (2) fabrication should be performed under inert or vacuum environments; (3) traditional lithography for making electrical contact should be avoided after the 2D heterostructure fabrication; and (4) ideally, there should be no encapsulating layer on the top of the heterostructure.

FIG. 1.

Chronological development of pristine 2D heterostructures. Reprinted with permission from Novoselov et al, “Electric field effect in atomically thin carbon films,” Science 306, 5696 (2004)127; Wang et al, “One-dimensional electrical contact to a two-dimensional material,” Science 342, 6158 (2013)128; Kim et al. Nano Lett 16, 1989 (2016). Copyright 2016 American Chemical Society; Cao et al., Nature 556, 43–50 (2018). Copyright 2024 Springer Nature Limited; Yu et al., Nature 575, 156–163 (2019). Copyright 2024 Springer Nature Limited; Mannix et al., Nat. Nanotechnol. 17, 361–366 (2022). Copyright 2024 Springer Nature Limited; and Wang et al., Nat. Electron. 6, 981–990 (2023). Copyright 2024 Springer Nature Limited.

FIG. 1.

Chronological development of pristine 2D heterostructures. Reprinted with permission from Novoselov et al, “Electric field effect in atomically thin carbon films,” Science 306, 5696 (2004)127; Wang et al, “One-dimensional electrical contact to a two-dimensional material,” Science 342, 6158 (2013)128; Kim et al. Nano Lett 16, 1989 (2016). Copyright 2016 American Chemical Society; Cao et al., Nature 556, 43–50 (2018). Copyright 2024 Springer Nature Limited; Yu et al., Nature 575, 156–163 (2019). Copyright 2024 Springer Nature Limited; Mannix et al., Nat. Nanotechnol. 17, 361–366 (2022). Copyright 2024 Springer Nature Limited; and Wang et al., Nat. Electron. 6, 981–990 (2023). Copyright 2024 Springer Nature Limited.

Close modal
FIG. 2.

(a) STM image of hydrocarbon contaminations on graphene, scale bar = 10 nm. (b) Schematic of the hydrocarbons shown in (a) on top of the graphene lattice. (c) AFM image of black phosphorus taken immediately after exfoliation. (d) AFM image of black phosphorus left under ambient conditions for a few days. (e)–(h) Time series optical micrographs of few-layer CrI3 flakes in humid argon. (i) STM image of Fe3GeTe2 with no exposure to ambient conditions. Scale bar = 180 nm. (j) STM image of the Fe3GeTe2 sample exposed to ambient conditions for 5 min. Scale bar = 40 nm. (a) and (b) Adapted from Pálinkás et al., Nat. Commun. 13, 6770 (2022). Copyright 2015 Author(s), licensed under a Creative Commons Attribution 4.0 License. (c) and (d) Reprinted with permission from Favron et al., Nat. Mater. 14, 826–832 (2015). Copyright 2024 Springer Nature Limited. (e)–(h) Reprinted with permission from Shcherbakov et al., Nano Lett. 18, 4214–4219 (2018). Copyright 2018 American Chemical Society. (i) and (j) Reprinted with permission from Kim et al., J. Korean Phys. Soc. 82, 204–208 (2023). Copyright 2024 Springer Nature Limited.

FIG. 2.

(a) STM image of hydrocarbon contaminations on graphene, scale bar = 10 nm. (b) Schematic of the hydrocarbons shown in (a) on top of the graphene lattice. (c) AFM image of black phosphorus taken immediately after exfoliation. (d) AFM image of black phosphorus left under ambient conditions for a few days. (e)–(h) Time series optical micrographs of few-layer CrI3 flakes in humid argon. (i) STM image of Fe3GeTe2 with no exposure to ambient conditions. Scale bar = 180 nm. (j) STM image of the Fe3GeTe2 sample exposed to ambient conditions for 5 min. Scale bar = 40 nm. (a) and (b) Adapted from Pálinkás et al., Nat. Commun. 13, 6770 (2022). Copyright 2015 Author(s), licensed under a Creative Commons Attribution 4.0 License. (c) and (d) Reprinted with permission from Favron et al., Nat. Mater. 14, 826–832 (2015). Copyright 2024 Springer Nature Limited. (e)–(h) Reprinted with permission from Shcherbakov et al., Nano Lett. 18, 4214–4219 (2018). Copyright 2018 American Chemical Society. (i) and (j) Reprinted with permission from Kim et al., J. Korean Phys. Soc. 82, 204–208 (2023). Copyright 2024 Springer Nature Limited.

Close modal

The rest of this Review will proceed as follows: We will briefly review the commonly used wet transfer methods and their limited suitability for SPM characterization. We will then review the dry transfer methods currently being employed to fabricate 2D heterostructures for SPM. We will further discuss the technical requirements for performing dry-transfer in an inert-atmosphere glovebox and the techniques being used therein. Then, we will discuss exfoliation and transfer methods in vacuum, particularly UHV environments. Finally, we will review methods for making electrical contacts without the use of traditional wet lithography and provide some future outlook.

For any 2D material transfer method to work, two basic challenges must be addressed. The first is how to pick up the desired flake, and the second is how to deposit that flake. We here define wet-transfer to be methods that use liquids in solving one or both challenges. Several of the first methods developed for 2D heterostructure fabrication used a floating-flake wet-transfer scheme.33,34 These methods involve coating substrates with polymers that could either be dissolved in solvents33 or were hydrophobic,34 causing the desired flakes to float to the surface together with a supporting polymer layer where they could then be scooped and deposited. These methods also used solvents in later stages of the fabrication process for removing the supporting polymers. A similar method of injecting water into a salt-promoter layer has also been used to separate and isolate CVD-grown monolayers.35 

When preparing heterostructures for characterization with SPM, a more common wet-transfer approach36,37 involves using water or other solvents such as acetone38 to dissolve a sacrificial polymer layer depositing a flake in the desired location. Figures 3(a)3(h) show one such procedure37 where water is used to dissolve a polyvinyl alcohol (PVA) layer depositing the magic angle twisted bilayer graphene (MATBG)/hBN stack onto a handle [Fig. 3(f)]. This procedure fabricates twisted bilayer graphene devices for characterization using SPM, specifically scanning tunneling microscopy (STM), as shown in Fig. 3(i), but it has significant drawbacks. This procedure cannot be performed in a glovebox or vacuum, as the use of water would contaminate those environments, which means that only materials that are stable in air and water may be used. In addition, a significant amount of post-fabrication treatment is required in this procedure to make the device atomically clean because bubbles can become trapped between the layers,39 individual layers can become wrinkled,40,41 and environmental contaminants can adsorb to the surface.28 Such a treatment typically involves rinsing the device in heated dichloromethane, acetone, water, and isopropanol alcohol (IPA), in addition to annealing the device in UHV for over 10 h at temperatures above 100 °C.

FIG. 3.

A representative wet transfer method. (a) PVA/tape/PDMS stamp picks up hBN. (b) The hBN tears monolayer graphene, picking up half of the flake. (c) A twist angle is introduced and the remaining monolayer of graphene is picked up. (d) and (e) The heterostructure is pressed against a PDMS/tape/PMMA stamp. (f) Water is injected between the stamps, dissolving the PVA layer and leaving the heterostructure on the PDMS/tape/PMMA stamp. (g) and (h) The heterostructure is deposited onto a substrate with prepatterned contacts. Following step (h), the device is cleaned to remove residues. (i) STM image of 1.06° MATBG fabricated by this procedure. (a)–(i) Reprinted with permission from Wong et al., Nature 582, 198–202 (2020). Copyright 2024 Springer Nature Limited.

FIG. 3.

A representative wet transfer method. (a) PVA/tape/PDMS stamp picks up hBN. (b) The hBN tears monolayer graphene, picking up half of the flake. (c) A twist angle is introduced and the remaining monolayer of graphene is picked up. (d) and (e) The heterostructure is pressed against a PDMS/tape/PMMA stamp. (f) Water is injected between the stamps, dissolving the PVA layer and leaving the heterostructure on the PDMS/tape/PMMA stamp. (g) and (h) The heterostructure is deposited onto a substrate with prepatterned contacts. Following step (h), the device is cleaned to remove residues. (i) STM image of 1.06° MATBG fabricated by this procedure. (a)–(i) Reprinted with permission from Wong et al., Nature 582, 198–202 (2020). Copyright 2024 Springer Nature Limited.

Close modal

Dry-transfer techniques must employ different methods to pick up and deposit flakes that do not rely on solvents.42 This is typically done by using polymer stamps [made of polydimethylsiloxane (PDMS), polycarbonate (PC), etc.] with relatively low glass transition temperatures. Flakes are picked up and deposited by heating or cooling polymer stamps through their glass transition temperatures, which gives control over the malleability and adhesiveness of the stamp. Many43–52 2D heterostructures have been prepared using such methods for SPM measurements, particularly STM, and although the details often differ, the overall approach and major challenges are common.

Figure 4 shows a representative dry-transfer tear-and-stack technique43 for creating twisted bilayer graphene to be characterized by STM.45 This technique begins by exfoliating monolayer graphene from bulk graphite [Figs. 4(a)4(c)] using a silicon/PVA/polymethyl methacrylate (PMMA) stamp. Tape, with a window cut to access the graphene [Fig. 4(d)], is then used to peel away the PMMA/graphene layer. Once the monolayer of graphene to be used is identified, half of the flake is brought into contact with a piece of hBN, and the graphene is slowly lifted away tearing at the hBN edge [Figs. 4(e) and 4(f)]. This is known as the tear-and-stack technique, which is a popular52–57 method for controlling the twist angle between layers in heterostructures where both layers are the same material. A twist angle is then introduced, and the remaining half graphene piece is placed on top of the initial graphene [Figs. 4(g) and 4(h)]. One downside to the tear-and-stack method is that it can introduce unintended strain58 when the material is torn. This can be mitigated by using an atomic force microscope (AFM) tip to cut the material in half before lifting off [a step just before Figure 4(e)], a method known as the cut-and-stack technique.59–62 For 2D heterostructure fabrication involving layers that are not the same material, the same stacking process can be used, but to determine the relative orientation between layers, other methods are required. Second harmonic generation (SHG) provides one method for non-centrosymmetric63 materials such as odd-layer TMDCs.47,64,65 For materials with anisotropic crystal structures, such as MoO3, a combination of optical microscopy and AFM has been used to determine the crystallographic axes.66 

FIG. 4.

Schematic showing a representative dry-transfer tear-and-stack technique. (a) A silicon chip, with a layer of spin-coated PVA and a layer of spin-coated PMMA, approaches the bulk graphite. (b) and (c) Graphene is exfoliated from the bulk graphite onto the stack. After lifting the stamp, monolayers of graphene adhere to the PMMA. (d) A piece of tape with a window (dashed line) cut out so the tape that does not contact the graphene is then used to peel away the graphene/PMMA (shown top down) separating it from the silicon/PVA. (e) Separately, hBN is exfoliated and annealed in forming gas. The target graphene and hBN flakes are identified and brought together such that about half the target graphene flake contacts the hBN, and the rest of the graphene extends past the edge. (f) PMMA/graphene is slowly lifted, tearing the graphene flake in half but maintaining the same crystal orientation. (g) A twist angle θ is introduced and the two graphene halves are aligned vertically. (h) The two graphene flakes are brought together creating twisted-bilayer graphene. After the PMMA is removed, the sample is annealed in forming gas at 260 °C overnight to remove fabrication residues. (i) STM image of twisted (θ = 1.07°) bilayer graphene fabricated using the process described in (a)–(h). (i) Reprinted with permission from Jiang et al., Nature 573, 91–95 (2019). Copyright 2024 Springer Nature Limited.

FIG. 4.

Schematic showing a representative dry-transfer tear-and-stack technique. (a) A silicon chip, with a layer of spin-coated PVA and a layer of spin-coated PMMA, approaches the bulk graphite. (b) and (c) Graphene is exfoliated from the bulk graphite onto the stack. After lifting the stamp, monolayers of graphene adhere to the PMMA. (d) A piece of tape with a window (dashed line) cut out so the tape that does not contact the graphene is then used to peel away the graphene/PMMA (shown top down) separating it from the silicon/PVA. (e) Separately, hBN is exfoliated and annealed in forming gas. The target graphene and hBN flakes are identified and brought together such that about half the target graphene flake contacts the hBN, and the rest of the graphene extends past the edge. (f) PMMA/graphene is slowly lifted, tearing the graphene flake in half but maintaining the same crystal orientation. (g) A twist angle θ is introduced and the two graphene halves are aligned vertically. (h) The two graphene flakes are brought together creating twisted-bilayer graphene. After the PMMA is removed, the sample is annealed in forming gas at 260 °C overnight to remove fabrication residues. (i) STM image of twisted (θ = 1.07°) bilayer graphene fabricated using the process described in (a)–(h). (i) Reprinted with permission from Jiang et al., Nature 573, 91–95 (2019). Copyright 2024 Springer Nature Limited.

Close modal

When flakes are exposed to polymers (e.g., PMMA) in such dry-transfer processes, polymer residues are left behind on the flakes.67 These residues interfere with taking high quality SPM data. To obtain the STM image shown in Fig. 4(i), the sample was annealed at 260 °C in forming gas (10% H2 and 90% Ar) overnight, which removed the fabrication residues.45 Another option to prevent residue contamination entirely involves using clever polymer geometries68 to ensure that no polymer has direct contact with a targeted flake surface, as illustrated in Figs. 5(a) and 5(b), in a polymer-contact free method. The goal of the polymer-contact free method is to fabricate a 2D heterostructure without allowing any polymer to contact a surface that is part of the 2D heterostructure, thus ensuring that no polymer residue is left behind. To accomplish this, a base flake is used to build the heterostructure and contact polymer stamps in a very specific geometry. The two stamps consist of PDMS with a hole (stamp 1) and a trench (stamp 2) partially covered by a poly(vinyl chloride) (PVC) thin film [Fig. 5(a)]. The procedure starts by picking up the base flake with stamp 1 such that the PVC film contacts roughly half of the base flake [Fig. 5(b-i)]. A target flake [green in Fig. 5(b-ii)] is then made in contact with and picked up by the part of the base flake that is in contact with PVC. Once this target layer has been picked up, the other half of the base flake is contacted by the suspended PVC film on stamp 2 such that the target flake/base flake heterostructure extends out into the vacant trench of stamp 2 as shown in Fig. 5(b-iii). Stamp 1 is then slowly removed horizontally [Fig. 5(b-iv)], and stamp 2 is flipped over and pressed against a final substrate [yellow in Fig. 5(b-v)]. Finally, by pulling away stamp 2, the target flake/base flake heterostructure is deposited on this substrate and ready to be characterized [Fig. 5(b-vi)]. This approach creates a heterostructure where the interface and the top surface of the target flake have not had any contact with polymers. Optical and STM images, as shown in Figs. 5(b)5(e), demonstrate the cleanliness of a variety of heterostructures this technique can produce.

FIG. 5.

Transfer technique showing a polymer-contact free method for ultra-clean surface 2D heterostructures. (a) Two stamps are prepared with a hole (on stamp 1) or a trench (on stamp 2) with a PVC thin film partially covering the opening. (b) Transfer steps for picking up a base flake and a target flake that ensure that the PVC film never touches the target flake. (c) Optical image showing a bilayer graphene/hBN heterostructure fabricated with this technique. (d) Optical image showing a bilayer WTe2/NbSe2 heterostructure fabricated using this technique. (e) STM image of bilayer WTe2 on bulk NbSe2. Only a few single-atomic defects are observed, and their abundance is close to the material’s bulk defect concentration. (a)–(e) Reprinted with permission from Jin et al., Adv. Mater. Interfaces 11, 2300658 (2024) Copyright 1999–2024 John Wiley & Sons.

FIG. 5.

Transfer technique showing a polymer-contact free method for ultra-clean surface 2D heterostructures. (a) Two stamps are prepared with a hole (on stamp 1) or a trench (on stamp 2) with a PVC thin film partially covering the opening. (b) Transfer steps for picking up a base flake and a target flake that ensure that the PVC film never touches the target flake. (c) Optical image showing a bilayer graphene/hBN heterostructure fabricated with this technique. (d) Optical image showing a bilayer WTe2/NbSe2 heterostructure fabricated using this technique. (e) STM image of bilayer WTe2 on bulk NbSe2. Only a few single-atomic defects are observed, and their abundance is close to the material’s bulk defect concentration. (a)–(e) Reprinted with permission from Jin et al., Adv. Mater. Interfaces 11, 2300658 (2024) Copyright 1999–2024 John Wiley & Sons.

Close modal

In addition to contamination from polymer residues, 2D heterostructures also face substantial contamination risk from the ambient environment, as discussed in the Introduction. The most common method of reducing ambient environment contamination involves moving 2D heterostructure fabrication to an inert atmosphere glovebox, typically filled with either nitrogen or argon (both nitrogen and argon are inert to most materials, although notably nitrogen does react with lithium compounds69 commonly used in energy applications). Gloveboxes are primarily used to mitigate water and oxygen induced sample degradation, and, as such, any wet technique that relies on water is inherently incompatible. Volatile solvents such as chloroform65,70 used in other wet-transfer techniques, while perhaps not being inherently incompatible, are extremely bad for the catalysts in a glovebox, and so techniques relying on these solvents can only be compatible with gloveboxes with solvent traps/scrubbers.71 To fabricate 2D heterostructures in a glovebox, a transfer station,71–74 like the one shown in Fig. 6(a), is typically set up inside a glovebox and dry-transfer techniques are employed. By motorizing each of the stage positioners, the fine movements associated with the transfer process can be controlled remotely without being encumbered by the glovebox. Dry transfer techniques are broadly glovebox compatible, and a wide variety of polymers and tapes have been used for glovebox exfoliation and stacking. The technique shown in Fig. 5 is entirely glovebox compatible, and Fig. 6(b) shows four stamps used in gloveboxes utilizing a wide variety of polymers: (i) PMMA,73 (ii) PC,75 (iii) Elvacite,76 and (iv) PDMS, tape, and polycaprolactone (PCL),77 for a variety of 2D materials, including (i) black phosphorous (BP),73 (ii) MoS2,75 (iii) hBN,76 and (iv) ZnPS3.77 Fabricating in a glovebox provides two major advantages: first, with fewer contaminants, the surfaces are cleaner, which is necessary for SPM; and second, materials sensitive to oxygen and/or moisture can be incorporated into 2D heterostructures. Indeed, the WTe2/NbSe2 heterostructure68 shown in Fig. 5(d) was fabricated in a glovebox because both materials are air sensitive.

FIG. 6.

Transfer in inert environment. (a) Photo of a typical transfer station in a glovebox. The positioners and rotation knob, along with the sample stage x, y can be attached to motors and controlled remotely. (b) Examples of various polymer stamps and materials used in glovebox setups. Polymers include (i) PMMA73 (ii) PC,78 (iii) Elvacite,75 and (iv) PDMS, tape, and PCL.76 (c) shows a vacuum suitcase that interfaces with both a glovebox and vacuum characterization systems. (a) Reprinted with permission from Zheng et al., “Emerging van der Waals junctions based on TMDs materials for advanced gas sensors,” Coordination Chem. Rev. 15, 214151 (2021).129 Copyright 2021 Elsevier. (c) Reprinted with permission from Gray et al., Rev. Sci. Instrum. 91, 073909 (2020). Copyright 2020 AIP Publishing.

FIG. 6.

Transfer in inert environment. (a) Photo of a typical transfer station in a glovebox. The positioners and rotation knob, along with the sample stage x, y can be attached to motors and controlled remotely. (b) Examples of various polymer stamps and materials used in glovebox setups. Polymers include (i) PMMA73 (ii) PC,78 (iii) Elvacite,75 and (iv) PDMS, tape, and PCL.76 (c) shows a vacuum suitcase that interfaces with both a glovebox and vacuum characterization systems. (a) Reprinted with permission from Zheng et al., “Emerging van der Waals junctions based on TMDs materials for advanced gas sensors,” Coordination Chem. Rev. 15, 214151 (2021).129 Copyright 2021 Elsevier. (c) Reprinted with permission from Gray et al., Rev. Sci. Instrum. 91, 073909 (2020). Copyright 2020 AIP Publishing.

Close modal

For air sensitive samples prepared in a glovebox, transferring the samples to external characterization tools presents another set of challenges. There are two strategies that we will discuss briefly here: (1) encapsulating air sensitive samples with other materials, and (2) using a protective suitcase to transfer the sample. Encapsulation is a particularly common method for bottom-up MBE grown samples, but there are also examples of this method in a top-down fabrication regime.78–81 Inevitably, encapsulation adds significant challenges to SPM measurements, which typically require the sample surface to be exposed. There are a few examples of researchers using STM to scan through a top encapsulating layer of graphene,79,80 or a layer of MoS2.78 However, this adds additional complexity to data interpretation as signal deconvolution is required. For encapsulation layers in the form of deposited chalcogen (e.g., Te and Se), in situ thermal desorption combined with ion sputtering before SPM measurements can effectively recover a pristine sample surface.82 Occasionally, top layers of van der Waals crystals can be locally removed with an STM tip as with graphene81 or Bi2Sr2CaCu2O8+x (BSCCO),83 exposing subsurface layers. Another option is to use a protective suitcase to transfer samples. These suitcases can be either evacuated or filled with inert gas, and an example71 is shown in Fig. 6(c). The biggest challenge with this method is ensuring that the glovebox and the characterization equipment can interface with the protective suitcase. Vacuum versions of these suitcases are also often bulky.

All-in-vacuum systems build on the advantages of inert atmosphere gloveboxes in two important ways. First, all-in-vacuum systems achieve lower levels of contamination and thus cleaner materials and interfaces. As Guo et al.84 demonstrated, one way to compare protection methods is to assume that a sample surface has a sticking coefficient of unity and use the partial pressure to compare various sample environments. Using this method, under air, a pristine sample surface becomes fully covered in about 1 ns; under typical inert glovebox conditions [oxygen and water content at 0.1 parts per million (ppm)], the pristine surface still becomes covered in 1–10 ms; and UHV conditions (better than 10−9 mbar) extend that time to 103 s and beyond. Even sub-0.1 ppm contaminants present in gloveboxes can be detrimental to 2D material properties and interfacial qualities.85 Therefore, the development of methods and instrumentation for exfoliating van der Waals crystals and fabricating 2D heterostructures under vacuum conditions is highly desirable. Second, UHV conditions can enhance exfoliation yields to the extent that Sun et al.85 classified using ultra-flat and clean substrates to exfoliate materials in UHV as a universal exfoliation method. Gold substrates86 and thin films87 have been shown to be particularly effective non-polymer stamps for exfoliating many 2D crystals. Thus, all-in-vacuum systems enable both cleaner samples and access to universal exfoliation methods.

To realize these advantages, all-in-vacuum systems have recently been demonstrated that can exfoliate85,88–93 and stack73,84,87,94–96 flakes into the desired heterostructures. While ideally both exfoliation and deterministic stacking should be completed in vacuum for a complete top-down fabrication process, the process of exfoliating bulk crystals and stacking 2D layers is challenging in vacuum due to limited degrees of freedom for sample manipulation. As such, several systems have been demonstrated that focus only on in-vacuum exfoliation or stacking, and to our knowledge, only two systems have been demonstrated to completely perform both exfoliation and deterministic stacking in an entirely top-down regime in UHV.

In-vacuum exfoliation has been developed by several research groups, beginning with a setup for graphene.91 One such method involves a transfer substrate consisting of a wheel88 covered with an insulating, low outgassing, and residue-free clean-room adhesive tape (UltraTape product) that was used to pick up flakes under ambient conditions. The wheel was then introduced to a vacuum chamber (10−8 mbar) and rolled over a sample substrate, depositing monolayers of VSe2. Another method consists of cleaving a bulk crystal in high vacuum89 or UHV85 and preparing a sample substrate through sputtering, annealing, and plasma cleaning. The substrate material can be flexible, depending on the flakes to be exfoliated, including Ag, Au, Ge, MgO(100), SrTiO3(100), Al2O3(0001), and Si(111). By pressing the substrate against the freshly cleaved material at variable temperatures (between room temperature and 500 K), a variety of materials including MoS2, graphene, FeSe, and Bi-2212 were exfoliated. Preparation in vacuum can also enhance the size of exfoliated monolayers up to millimeters as demonstrated with graphene.93 

Monolayers grown using bottom-up synthesis techniques are occasionally stacked using methods compatible with top-down fabrication stacking processes for STM characterizations.35 Imamura et al.94 used a high vacuum (10−6 mbar) system to stack two layers of CVD grown graphene together. The stacking process consists of pressing the two layers together at 200 °C for 1 h. This setup faced the drawback that the twist angle between the two layers is difficult to maintain, and the setup is incapable of deterministic stacking. A more controlled system was developed by Mannix et al.95 consisting of an assembly robot in high vacuum (10−6 Torr), see Figs. 7(a) and 7(b), that robotically stacks layers together, assembling a heterostructure that contains as many as 80 layers [16 layers of MoS2 are shown in Fig. 7(c)]. The assembly uses a stamp made from several polymers, beginning with PDMS, a support layer of MicroChem lift-off resist, a release layer of poly(cyclohexene propylene carbonate) (PCPC), and an adhesion layer of poly(benzyl methacrylate) (PBzMA), see Fig. 7(d), which it manipulates via heating and cooling through glass transition temperatures to pick up and place each layer. The system has been used for a variety of materials, including metal–organic CVD grown TMDCs, Au thin films, exfoliated flakes of graphite, and hBN, and is expected to be applicable to a variety of materials compatible with van der Waals stacking. One downside to this approach is that since polymers were used, the completed heterostructures have to be cleaned using acetone or chloroform, a process that is incompatible with vacuum.

FIG. 7.

(a) Photograph of an in-vacuum (10−6 Torr) assembly robot. (b) Diagram of the vacuum assembly robot. The stamp lifts up target flakes at 145 °C and deposits at 175 °C. (c) White-light optical micrograph of the 1–16 layer (bottom right = 1 layer, top left = 16 layers) MoS2 grid structure fabricated with this setup. (d) Schematic of the adhesive stamp structure in (b). The multilayer structure mediates programmed adhesion and deposition by thermal and/or ultraviolet light activation of the release layer to generate rapid decomposition. (e) Photograph of an instrument for UHV (10−10 mbar) polymer-free fabrication. (f) Close-up of the optical lens, cantilever, and sample stage therein. (g) Scanning electron microscopy (SEM) image of several cantilevers protruding from a silicon chip. (h) Optical image of a cantilever with a 2D flake adhered. The inset shows a schematic of flake deposition. (i) Example of a UHV fabricated 2D heterostructure of hBN-encapsulated graphene on graphite. The graphene is outlined with a black dashed line. The inset shows a 25 μm AFM image on the area highlighted by the blue dashed box. (j) Schematic showing the multilayer metallic coating of the cantilever holding a 2D material specimen. (a)–(d) Reprinted with permission from Mannix et al., Nat. Nanotechnol. 17, 361–366 (2022). Copyright 2024 Springer Nature. (h) Reprinted with permission from Wang, “Polymer-free assembly of ultraclean van der Waals heterostructures,” Nat. Rev. Phys. 4, 504–504 (2022).130 Copyright 2024 Springer Nature Limited. (e)–(g) and (i) and (j) Reprinted from Wang et al., Nat. Electron. 6, 981–990 (2023). Copyright 2024 Springer Nature Limited.

FIG. 7.

(a) Photograph of an in-vacuum (10−6 Torr) assembly robot. (b) Diagram of the vacuum assembly robot. The stamp lifts up target flakes at 145 °C and deposits at 175 °C. (c) White-light optical micrograph of the 1–16 layer (bottom right = 1 layer, top left = 16 layers) MoS2 grid structure fabricated with this setup. (d) Schematic of the adhesive stamp structure in (b). The multilayer structure mediates programmed adhesion and deposition by thermal and/or ultraviolet light activation of the release layer to generate rapid decomposition. (e) Photograph of an instrument for UHV (10−10 mbar) polymer-free fabrication. (f) Close-up of the optical lens, cantilever, and sample stage therein. (g) Scanning electron microscopy (SEM) image of several cantilevers protruding from a silicon chip. (h) Optical image of a cantilever with a 2D flake adhered. The inset shows a schematic of flake deposition. (i) Example of a UHV fabricated 2D heterostructure of hBN-encapsulated graphene on graphite. The graphene is outlined with a black dashed line. The inset shows a 25 μm AFM image on the area highlighted by the blue dashed box. (j) Schematic showing the multilayer metallic coating of the cantilever holding a 2D material specimen. (a)–(d) Reprinted with permission from Mannix et al., Nat. Nanotechnol. 17, 361–366 (2022). Copyright 2024 Springer Nature. (h) Reprinted with permission from Wang, “Polymer-free assembly of ultraclean van der Waals heterostructures,” Nat. Rev. Phys. 4, 504–504 (2022).130 Copyright 2024 Springer Nature Limited. (e)–(g) and (i) and (j) Reprinted from Wang et al., Nat. Electron. 6, 981–990 (2023). Copyright 2024 Springer Nature Limited.

Close modal

Two systems have been demonstrated to perform complete exfoliation and deterministic stacking in UHV. Wang et al.87 demonstrated a UHV (10−10 mbar) polymer-free transfer setup [Figs. 7(e) and 7(f)] using a SiNx cantilever [Figs. 7(g) and 7(h)] coated with a thin metal trilayer film for better adhesion to 2D materials. While a PMMA polymer was initially attempted, the PMMA outgassed and released hydrocarbons in UHV as evidenced by bubbles in the fabricated hBN/graphene/hBN heterostructures. To solve this problem, inorganic SiNx was used, leading to heterostructures with minimal contaminants [Fig. 7(i)]. Within the trilayer metal film, a 1 nm layer of Ta and a 5 nm layer of Pt were used to smooth out the surface and to give a flat place for the Au to bond without clumping [Fig. 7(j)]. A 0.65 nm thick layer of Au was found to be optimal for 2D material transfer at 150 °C.

Guo et al.84 demonstrated a UHV (5 × 10−10 mbar) system that is also directly connected to an MBE growth chamber. The assembly chamber [Fig. 8(a)] consists of two stages, a tool stage and a sample stage [Fig. 8(b)]. Samples are brought in on the sample stage, picked up by the tool stage, and then placed in their desired location on the substrate [which is also held on the sample stage as shown in Fig. 8(c)]. The whole assembly is optically accessible via a long working distance objective. Exfoliation is performed using UHV-compatible Kapton tapes onto a silicon oxide substrate (or Au coated silicon), as shown in Fig. 8(d). A sapphire/PDMS/PMMA/polypropylene carbonate (PPC) stamp, which is assembled outside and degassed before being brought into the UHV chamber, is then used to pick up exfoliated monolayers for stacking into heterostructures. With all-in-vacuum operations, this system has produced heterostructures with precise rotational alignment, as shown in the BP/hBN heterostructure in Fig. 8(e), although due to the use of polymers there are still some residues visible in the AFM image as bright specks [Fig. 8(f)].

FIG. 8.

(a) Schematic of an all-in-UHV (5 × 10−10 mbar) 2D material fabrication setup. A microscope offers optical access to the chamber, which is equipped with a wobble stick to move samples and tools. (b) Schematic of the sample and tool stage, which are driven by high-precision piezoelectric positioners. (c) Schematic showing exfoliation occurring in the sample stage, with a prepared substrate in the tool stage. (d) Step-by-step schematic of the complete 2D material exfoliation process. (e) Optical image of a completed BP/hBN heterostructure on a SiO2/Si substrate. (f) AFM image of the heterostructure shown in (e) (the region in the dashed square). Bright dots are residue on the surface. (a)–(f) Reprinted with permission from Guo et al., Rev. Sci. Instrum. 94, 013903 (2023). Copyright 2024 AIP Publishing.

FIG. 8.

(a) Schematic of an all-in-UHV (5 × 10−10 mbar) 2D material fabrication setup. A microscope offers optical access to the chamber, which is equipped with a wobble stick to move samples and tools. (b) Schematic of the sample and tool stage, which are driven by high-precision piezoelectric positioners. (c) Schematic showing exfoliation occurring in the sample stage, with a prepared substrate in the tool stage. (d) Step-by-step schematic of the complete 2D material exfoliation process. (e) Optical image of a completed BP/hBN heterostructure on a SiO2/Si substrate. (f) AFM image of the heterostructure shown in (e) (the region in the dashed square). Bright dots are residue on the surface. (a)–(f) Reprinted with permission from Guo et al., Rev. Sci. Instrum. 94, 013903 (2023). Copyright 2024 AIP Publishing.

Close modal

The characterization of 2D heterostructures, e.g., using STM, often requires them to be electrically contacted. Furthermore, the performance of 2D electronic, optoelectronic, and spintronic devices is greatly influenced by the intricate details of how these 2D materials are electrically interfaced with external circuits. For an in-depth discussion on achieving various types of flexible, ohmic, seamless, or impedance-matched contacts with 2D materials, we refer the readers to the relevant review articles.97–99 Defining metal contacts via lithography is an indispensable solution; however, the use of polymers inevitably leaves residues and the process of removing resists through lift-off requires the use of various chemicals that are detrimental to air sensitive materials. As a result, interfaces between substrates and 2D heterostructures are often contaminated by the accumulation of resists, polymers, desorbed gasses, and solvents. Most of the time, the contaminants significantly impact the performance and characterization as they tend to modify the electrochemical composition with the introduction of scattering centers, altered potential landscape, reduced charge carriers and mobility.100,101 Techniques such as annealing and vacuum treatment have shown some success in mitigating contamination issues. However, such a heat treatment introduces further complications to samples that are chemically reactive or heat sensitive.

For SPM characterizations that are sensitive to disorders and contamination at the nanometer to atomic length scale, deposition of contacts on substrates before 2D heterostructure fabrication and in situ patterning and contact formation on as-prepared 2D heterostructures in a vacuum environment are preferred methods as they preserve surface cleanliness and enables uncontaminated contact/sample interfaces. Toward this goal, several techniques have been demonstrated for the latter scheme. These include lithography-free techniques for simple patterning and shadow/stencil mask techniques for more complicated patterning and contacts producing high quality heterostructure devices.102–104 Most of the techniques involve the use of macroscopic and nanostructured stencil masks in UHV chambers through which the electrical contacts and patterns are formed by subsequent metal deposition using focused ion beams, thermal evaporation, and e-beam evaporation. Through the use of independent mask carriers, evaporation through multiple masks also becomes feasible. These methods are purely mechanical, resistless, and UHV compatible, thus minimizing the introduction of contaminants.

Masks with precise geometry and nanopatterns can be prepared by standard lithography processes.105–107 A typical silicon mask preparation process is schematically illustrated in Fig. 9(a). Double side polished Si(100) was used as a substrate, and 200 nm of Cr was deposited to serve as an etching mask layer, followed by subsequent deposition of PMMA, an e-beam resist. The other side of the wafer was coated with photoresist to etch away three-quarters of the wafer by using photolithography and KOH etching. The front side of the substrate was exposed to an e-beam to create the patterns for a desirable shadow mask structure. The e-beam resist was developed after exposure, and Cr etchant is used to remove exposed Cr patterns as openings on the mask. Once prepared, those masks can be transferred to UHV to form electrical contact by directly depositing contact material through the mask onto the 2D heterostructures at the desired location [Fig. 9(b)]. The masks so prepared can be used several times before they are destroyed or the apertures are covered with deposited metal. Bao et al.102 compared the cleanliness of the contacts made by e-beam lithography and shadow masks using AFM. The contacts made by e-beam lithography left a layer of residue, which needed additional sputtering and heat treatments for removal [left panel in Fig. 9(c)], whereas contacts made by shadow masks are much cleaner and need no post-deposition treatments, making the exposure of extended atomically cleaner areas suitable for STM imaging [Fig. 9(d)]. For making few local contacts with no complex patterning, nano-soldering with indium or other alloys of indium with low melting points are used to create ohmic contacts to nanostructures as demonstrated by Girit and Zettl;103 main components of their miniaturized soldering setup are shown in Fig. 9(e). A sample with a small bead of indium is placed on a holder, which can be heated up to melt the indium bead. The molten indium is precisely manipulated using a W tip under an optical microscope using XYZ micromanipulators forming a spike to create submicron sized contacts to the nanostructured samples.

FIG. 9.

Electrical contact fabrication. (a) Schematic illustration of the fabrication process of Si-shadow masks. (b) Direct deposition of metallic electrodes through the mask. (c) Optical image of a single-layer graphene device. The electrodes A, B, C, and D were deposited by evaporation through a shadow mask, and E, F, and G were fabricated using standard electron beam lithography. (d) STM images of an as-fabricated device using the shadow mask technique. The main panel displays an image of 85 × 85 nm2 area, and the inset shows the atomic lattice over an area of 2.5 × 2.5 nm2. (e) Schematic of the nano-soldering setup, consisting of an optical microscope, a micromanipulator, and a sample heater used to contact graphene and other nanostructures (center), SEM image of an indium solder spike ending in a 50 nm radius tip (upper left), and optical image of a contacted graphene device (upper right). (f) Schematic of fabrication of a graphene device. An ultra-thin quartz filament is used as a shadow mask. (g) AFM image of a four-layer graphene channel. (h) and (i) Schematics for planar tunnel junction fabrication by using the same technique but with angled evaporation. (a)–(d) Reprinted from Bao et al., Nano Res. 3, 98–102 (2010). Copyright 2024 Springer Nature. (e) Reprinted with permission from Girit and Zettl, Appl. Phys. Lett. 91, 193512 (2007). Copyright 2024 AIP Publishing. (f)–(i) Reprinted with permission from Staley et al., Appl. Phys. Lett. 90, 143518 (2007). Copyright 2024 AIP Publishing.

FIG. 9.

Electrical contact fabrication. (a) Schematic illustration of the fabrication process of Si-shadow masks. (b) Direct deposition of metallic electrodes through the mask. (c) Optical image of a single-layer graphene device. The electrodes A, B, C, and D were deposited by evaporation through a shadow mask, and E, F, and G were fabricated using standard electron beam lithography. (d) STM images of an as-fabricated device using the shadow mask technique. The main panel displays an image of 85 × 85 nm2 area, and the inset shows the atomic lattice over an area of 2.5 × 2.5 nm2. (e) Schematic of the nano-soldering setup, consisting of an optical microscope, a micromanipulator, and a sample heater used to contact graphene and other nanostructures (center), SEM image of an indium solder spike ending in a 50 nm radius tip (upper left), and optical image of a contacted graphene device (upper right). (f) Schematic of fabrication of a graphene device. An ultra-thin quartz filament is used as a shadow mask. (g) AFM image of a four-layer graphene channel. (h) and (i) Schematics for planar tunnel junction fabrication by using the same technique but with angled evaporation. (a)–(d) Reprinted from Bao et al., Nano Res. 3, 98–102 (2010). Copyright 2024 Springer Nature. (e) Reprinted with permission from Girit and Zettl, Appl. Phys. Lett. 91, 193512 (2007). Copyright 2024 AIP Publishing. (f)–(i) Reprinted with permission from Staley et al., Appl. Phys. Lett. 90, 143518 (2007). Copyright 2024 AIP Publishing.

Close modal

Staley et al.104 used an alternative graphene device fabrication technique by using ultra-thin quartz filaments up to submicron size in diameter as masks. Once the desired 2D material/heterostructure is prepared, a thin quartz filament is placed onto an area of the material as a mask and contact material such as gold can be evaporated on top normal to the surface [Fig. 9(f)], which resulted in exposure of micrometer sized graphene [Fig. 9(g)]. Similarly, graphene based tunnel junctions can also be fabricated by evaporating alumina and metal contacts at angles, taking advantage of the shadows created by the quartz filaments [Figs. 9(h) and 9(i)].

Several similar techniques have been employed in UHV using resistless methods to make electrical contacts for 2D material characterization in STM and other multifold experimental areas. These techniques allow for flexible, inexpensive, fast, and easy to integrate electrical contacts for heterostructure devices in the UHV system; various patterns and geometries can be designed simply by changing the masks in the mask carriers. However, the drawback of this fabrication process is the formation of penumbra during the evaporation process caused by mechanical instability and placement distance between the mask and the substrate, creating smudged patterns due to edge spreading.108,109 Nonetheless, methods such as SPM-assisted resistless lithography110 and well-designed masks with capacitively controlled alignment111 have enabled UHV lithography with sub-micron precision and, therefore, show great potential to be integrated with UHV exfoliation and stacking setups to create functional heterostructure devices all in vacuum.112 

We have reviewed the recent development of wet- and dry-transfer methods, especially emphasizing those that are UHV compatible for fabricating 2D heterostructures as well as strategies for making clean electrical contacts for SPM characterizations. Fabrication processes in protected environments, both in gloveboxes and in vacuum systems, create significant material advantages and research opportunities. Unconventional air-sensitive 2D materials such as cuprate superconductors, halides, and perovskites can now be integrated into novel and functional 2D heterostructures with preserved material properties and, critically, high reproducibility.63,113 SPM characterizations that are sensitive to atomic scale surface and interface contamination will significantly benefit from these techniques.

Going forward, we believe there exist significant opportunities for technical development that will address multiple challenges facing ultra-clean 2D material fabrication at present. For example, improving the efficiency and flexibility of in-vacuum fabrication processes will both increase research capabilities and develop scalability necessary for future commercial applications. We can imagine a future closed-cycle autonomous laboratory powered by artificial intelligence and robotic experimental agent that can take and exfoliate bulk materials, identify target flakes, assemble desired 2D heterostructures, perform material characterizations, and refine fabrication strategies using feedback obtained through data analysis. The robotic assembly95 system described above provides an example of what the assembly portion of such a system might look like. Substantial effort114 has already gone into developing machine learning algorithms to identify 2D flakes based on optical contrast115 and shape,116 and significant progress has recently been made in self-driven autonomous laboratories for materials discovery in chemistry, biology, and materials science.117–121 Therefore, we believe that it will be a matter of time before the realization of fully automated and closed-cycle discoveries of novel functional 2D heterostructures.

Another opportunity lies in the development of polymer-less stamps that are compatible with higher vacuum. Using polymers to create stamps is a common fabrication practice in air and glovebox environments, and there is some evidence that these polymers can be used under high vacuum88 and UHV84 conditions given careful selection and preparation. However, given the significantly lower outgassing rates of inorganic materials compared to polymers,122 it stands to reason that higher vacuum levels can be achieved using polymer-free setups like the SiNx discussed above.87 Transfer processes relying on metals or layers of metals, particularly Au86,87 or Ni,123 have successfully been used to transfer flakes on a variety of materials. Further development of these methods and identification of new stamp materials in UHV environments will further mitigate contamination and thus cleaner surfaces and interfaces.

Finally, there are opportunities to combine bottom-up and top-down fabrication processes in exciting ways. Remote epitaxy124 is a process for epitaxial growth that uses a van der Waals material (e.g., graphene) to allow the grown layer to be peeled off. The free layer could then be incorporated into a 2D heterostructure with conventional top-down exfoliated flakes. Such a scheme will become even more powerful when combined with micro-MBE that uses shadow masks to achieve growth resolution on the order of 100 nm,125,126 allowing for the flexibility of top-down fabrication with the material control of bottom-up synthesis.

Without question, technical advances toward fabricating 2D heterostructures of highest quality will unlock groundbreaking research opportunities that were previously inconceivable and introduce invaluable new research frontiers promised by 2D materials.

J.M., N.S., and X.L. acknowledge the support from the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences, under Award No. DE-SC0024291. X.L. acknowledges the support from the Oak Ridge Associated Universities Ralph E. Powe Junior Faculty Enhancement Award.

The authors have no conflicts to disclose.

James McKenzie: Investigation (lead); Visualization (lead); Writing – original draft (lead); Writing – review & editing (lead). Nileema Sharma: Investigation (supporting); Visualization (supporting); Writing – original draft (supporting); Writing – review & editing (supporting). Xiaolong Liu: Conceptualization (lead); Funding acquisition (lead); Investigation (equal); Project administration (lead); Resources (lead); Supervision (lead); Writing – review & editing (equal).

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

1.
X.
Liu
and
M. C.
Hersam
, “
2D materials for quantum information science
,”
Nat. Rev. Mater.
4
,
669
684
(
2019
).
2.
Y.
Cao
et al, “
Unconventional superconductivity in magic-angle graphene superlattices
,”
Nature
556
,
43
50
(
2018
).
3.
S.
Kezilebieke
et al, “
Topological superconductivity in a van der Waals heterostructure
,”
Nature
588
,
424
428
(
2020
).
4.
A.
Ciarrocchi
,
F.
Tagarelli
,
A.
Avsar
, and
A.
Kis
, “
Excitonic devices with van der Waals heterostructures: Valleytronics meets twistronics
,”
Nat. Rev. Mater.
7
,
449
464
(
2022
).
5.
H.-L.
Hou
,
C.
Anichini
,
P.
Samorì
,
A.
Criado
, and
M.
Prato
, “
2D Van der Waals heterostructures for chemical sensing
,”
Adv. Funct. Mater.
32
,
2207065
(
2022
).
6.
F.
He
et al, “
Moiré patterns in 2D materials: A review
,”
ACS Nano
15
,
5944
5958
(
2021
).
7.
S.
Manzeli
,
D.
Ovchinnikov
,
D.
Pasquier
,
O. V.
Yazyev
, and
A.
Kis
, “
2D transition metal dichalcogenides
,”
Nat. Rev. Mater.
2
,
17033
17115
(
2017
).
8.
K. A.
Lozovoy
et al, “
Single-element 2D materials beyond graphene: Methods of epitaxial synthesis
,”
Nanomaterials
12
,
2221
(
2022
).
9.
A.-Y.
Lu
et al, “
Janus monolayers of transition metal dichalcogenides
,”
Nat. Nanotechnol.
12
,
744
749
(
2017
).
10.
X.
Wan
et al, “
Synthesis and characterization of metallic Janus MoSH monolayer
,”
ACS Nano
15
,
20319
20331
(
2021
).
11.
B.
Huang
et al, “
Layer-dependent ferromagnetism in a van der Waals crystal down to the monolayer limit
,”
Nature
546
,
270
273
(
2017
).
12.
Y.
Yu
et al, “
High-temperature superconductivity in monolayer Bi2Sr2CaCu2O8+δ
,”
Nature
575
,
156
163
(
2019
).
13.
K. S.
Novoselov
et al, “
Two-dimensional gas of massless Dirac fermions in graphene
,”
Nature
438
,
197
200
(
2005
).
14.
C.
Nayak
,
S. H.
Simon
,
A.
Stern
,
M.
Freedman
, and
S.
Das Sarma
, “
Non-Abelian anyons and topological quantum computation
,”
Rev. Mod. Phys.
80
,
1083
1159
(
2008
).
15.
X.
Qian
,
J.
Liu
,
L.
Fu
, and
J.
Li
, “
Quantum spin Hall effect in two-dimensional transition metal dichalcogenides
,”
Science
346
,
1344
1347
(
2014
).
16.
X.
Zhou
and
G.
Yu
, “
Preparation engineering of two-dimensional heterostructures via bottom-up growth for device applications
,”
ACS Nano
15
,
11040
11065
(
2021
).
17.
A.
Ohtake
and
Y.
Sakuma
, “
Two-dimensional WSe2/MoSe2 heterostructures grown by molecular-beam epitaxy
,”
J. Phys. Chem. C
125
,
11257
11261
(
2021
).
18.
Y.
Gong
et al, “
Vertical and in-plane heterostructures from WS2/MoS2 monolayers
,”
Nat. Mater.
13
,
1135
1142
(
2014
).
19.
E.
Shi
et al, “
Two-dimensional halide perovskite lateral epitaxial heterostructures
,”
Nature
580
,
614
620
(
2020
).
20.
L.
Liu
and
T.
Zhai
, “
Wafer-scale vertical van der Waals heterostructures
,”
InfoMat
3
,
3
21
(
2021
).
21.
P. K.
Sahoo
et al, “
Bilayer lateral heterostructures of transition-metal dichalcogenides and their optoelectronic response
,”
ACS Nano
13
,
12372
12384
(
2019
).
22.
A.
Patsha
,
K.
Ranganathan
,
M.
Kazes
,
D.
Oron
, and
A.
Ismach
, “
Halide chemical vapor deposition of 2D semiconducting atomically-thin crystals: From self-seeded to epitaxial growth
,”
Appl. Mater. Today
26
,
101379
(
2022
).
23.
S.
Xie
et al, “
Coherent, atomically thin transition-metal dichalcogenide superlattices with engineered strain
,”
Science
359
,
1131
1136
(
2018
).
24.
X.
Liu
and
M. C.
Hersam
, “
Borophene-graphene heterostructures
,”
Sci. Adv.
5
,
eaax6444
(
2019
).
25.
M. P.
Levendorf
et al, “
Graphene and boron nitride lateral heterostructures for atomically thin circuitry
,”
Nature
488
,
627
632
(
2012
).
26.
K. S.
Novoselov
,
A.
Mishchenko
,
A.
Carvalho
, and
A. H.
Castro Neto
, “
2D materials and van der Waals heterostructures
,”
Science
353
,
aac9439
(
2016
).
27.
X.
Wang
,
Y.
Sun
, and
K.
Liu
, “
Chemical and structural stability of 2D layered materials
,”
2D Mater.
6
,
042001
(
2019
).
28.
A.
Pálinkás
et al, “
The composition and structure of the ubiquitous hydrocarbon contamination on van der Waals materials
,”
Nat. Commun.
13
,
6770
(
2022
).
29.
A.
Favron
et al, “
Photooxidation and quantum confinement effects in exfoliated black phosphorus
,”
Nat. Mater.
14
,
826
832
(
2015
).
30.
D.
Shcherbakov
et al, “
Raman spectroscopy, photocatalytic degradation, and stabilization of atomically thin chromium tri-iodide
,”
Nano Lett.
18
,
4214
4219
(
2018
).
31.
H. K.
Kim
et al, “
In-situ scanning tunneling microscopy observation of thickness-dependent air-sensitive layered materials and heterodevices
,”
J. Korean Phys. Soc.
82
,
204
208
(
2023
).
32.
X.
Liu
and
M. C.
Hersam
, “
Interface characterization and control of 2D materials and heterostructures
,”
Adv. Mater.
30
,
1801586
(
2018
).
33.
C. R.
Dean
et al, “
Boron nitride substrates for high-quality graphene electronics
,”
Nat. Nanotechnol.
5
,
722
726
(
2010
).
34.
G. F.
Schneider
,
V. E.
Calado
,
H.
Zandbergen
,
L. M. K.
Vandersypen
, and
C.
Dekker
, “
Wedging transfer of nanostructures
,”
Nano Lett.
10
,
1912
1916
(
2010
).
35.
P.-Y.
Chen
et al, “
Tunable moiré superlattice of artificially twisted monolayers
,”
Adv. Mater.
31
,
1901077
(
2019
).
36.
Y.
Li
et al, “
Symmetry breaking and anomalous conductivity in a double-moiré superlattice
,”
Nano Lett.
22
,
6215
6222
(
2022
).
37.
D.
Wong
et al, “
Cascade of electronic transitions in magic-angle twisted bilayer graphene
,”
Nature
582
,
198
202
(
2020
).
38.
P. J.
Zomer
,
S. P.
Dash
,
N.
Tombros
, and
B. J.
van Wees
, “
A transfer technique for high mobility graphene devices on commercially available hexagonal boron nitride
,”
Appl. Phys. Lett.
99
,
232104
(
2011
).
39.
K.
Kang
et al, “
Layer-by-layer assembly of two-dimensional materials into wafer-scale heterostructures
,”
Nature
550
,
229
233
(
2017
).
40.
A.
Kushima
,
X.
Qian
,
P.
Zhao
,
S.
Zhang
, and
J.
Li
, “
Ripplocations in van der Waals layers
,”
Nano Lett.
15
,
1302
1308
(
2015
).
41.
M. C.
Wang
et al, “
Mechanical instability driven self-assembly and architecturing of 2D materials
,”
2D Mater.
4
,
022002
(
2017
).
42.
A.
Castellanos-Gomez
et al, “
Deterministic transfer of two-dimensional materials by all-dry viscoelastic stamping
,”
2D Mater.
1
,
011002
(
2014
).
43.
K.
Kim
et al, “
van der Waals heterostructures with high accuracy rotational alignment
,”
Nano Lett.
16
,
1989
1995
(
2016
).
44.
N.
Tilak
et al, “
Flat band carrier confinement in magic-angle twisted bilayer graphene
,”
Nat. Commun.
12
,
4180
(
2021
).
45.
Y.
Jiang
et al, “
Charge order and broken rotational symmetry in magic-angle twisted bilayer graphene
,”
Nature
573
,
91
95
(
2019
).
46.
C.
Rubio-Verdú
et al, “
Moiré nematic phase in twisted double bilayer graphene
,”
Nat. Phys.
18
,
196
202
(
2022
).
47.
S.
Shabani
et al, “
Deep moiré potentials in twisted transition metal dichalcogenide bilayers
,”
Nat. Phys.
17
,
720
725
(
2021
).
48.
J.
Lee
et al, “
Imaging electrostatically confined Dirac fermions in graphene quantum dots
,”
Nat. Phys.
12
,
1032
1036
(
2016
).
49.
K.
Yasuda
,
X.
Wang
,
K.
Watanabe
,
T.
Taniguchi
, and
P.
Jarillo-Herrero
, “
Stacking-engineered ferroelectricity in bilayer boron nitride
,”
Science
372
,
1458
1462
(
2021
).
50.
X.
Wang
et al, “
Light-induced ferromagnetism in moiré superlattices
,”
Nature
604
,
468
473
(
2022
).
51.
N. C. H.
Hesp
et al, “
Observation of interband collective excitations in twisted bilayer graphene
,”
Nat. Phys.
17
,
1162
1168
(
2021
).
52.
Y.
Choi
et al, “
Electronic correlations in twisted bilayer graphene near the magic angle
,”
Nat. Phys.
15
,
1174
1180
(
2019
).
53.
Y.
Zhou
et al, “
Bilayer Wigner crystals in a transition metal dichalcogenide heterostructure
,”
Nature
595
,
48
52
(
2021
).
54.
T. I.
Andersen
et al, “
Excitons in a reconstructed moiré potential in twisted WSe2/WSe2 homobilayers
,”
Nat. Mater.
20
,
480
487
(
2021
).
55.
M.
Oh
et al, “
Evidence for unconventional superconductivity in twisted bilayer graphene
,”
Nature
600
,
240
245
(
2021
).
56.
H. S.
Arora
et al, “
Superconductivity in metallic twisted bilayer graphene stabilized by WSe2
,”
Nature
583
,
379
384
(
2020
).
57.
Y.
Xie
et al, “
Fractional Chern insulators in magic-angle twisted bilayer graphene
,”
Nature
600
,
439
443
(
2021
).
58.
T.
Benschop
et al, “
Measuring local moiré lattice heterogeneity of twisted bilayer graphene
,”
Phys. Rev. Res.
3
,
013153
(
2021
).
59.
Y.
Saito
,
J.
Ge
,
K.
Watanabe
,
T.
Taniguchi
, and
A. F.
Young
, “
Independent superconductors and correlated insulators in twisted bilayer graphene
,”
Nat. Phys.
16
,
926
930
(
2020
).
60.
M.
He
et al, “
Competing correlated states and abundant orbital magnetism in twisted monolayer-bilayer graphene
,”
Nat. Commun.
12
,
4727
(
2021
).
61.
S.
Chen
et al, “
Electrically tunable correlated and topological states in twisted monolayer–bilayer graphene
,”
Nat. Phys.
17
,
374
380
(
2021
).
62.
A.
Jaoui
et al, “
Quantum critical behaviour in magic-angle twisted bilayer graphene
,”
Nat. Phys.
18
,
633
638
(
2022
).
63.
C. N.
Lau
,
M. W.
Bockrath
,
K. F.
Mak
, and
F.
Zhang
, “
Reproducibility in the fabrication and physics of moiré materials
,”
Nature
602
,
41
50
(
2022
).
64.
X.
Wang
et al, “
Moiré trions in MoSe2/WSe2 heterobilayers
,”
Nat. Nanotechnol.
16
,
1208
1213
(
2021
).
65.
R.
Nieken
et al, “
Direct STM measurements of R-type and H-type twisted MoSe2/WSe2
,”
APL Mater.
10
,
031107
(
2022
).
66.
M.
Chen
et al, “
Configurable phonon polaritons in twisted α-MoO3
,”
Nat. Mater.
19
,
1307
1311
(
2020
).
67.
J. J.
Schwartz
et al, “
Chemical identification of interlayer contaminants within van der Waals heterostructures
,”
ACS Appl. Mater. Interfaces
11
,
25578
25585
(
2019
).
68.
K.
Jin
et al, “
Assembly of arbitrary designer heterostructures with atomically clean interfaces
,”
Adv. Mater. Interfaces
11
,
2300658
(
2024
).
69.
S.
Yamaguchi
et al, “
Nitrogen dissociation via reaction with lithium alloys
,”
ACS Omega
2
,
1081
1088
(
2017
).
70.
T. A.
de Jong
et al, “
Imaging moiré deformation and dynamics in twisted bilayer graphene
,”
Nat. Commun.
13
,
70
(
2022
).
71.
M. J.
Gray
et al, “
A cleanroom in a glovebox
,”
Rev. Sci. Instrum.
91
,
073909
(
2020
).
72.
Y.
Cao
et al, “
Quality heterostructures from two-dimensional crystals unstable in air by their assembly in inert atmosphere
,”
Nano Lett.
15
,
4914
4921
(
2015
).
73.
G.
Long
et al, “
Achieving ultrahigh carrier mobility in two-dimensional hole gas of black phosphorus
,”
Nano Lett.
16
,
7768
7773
(
2016
).
74.
N.
Tilak
,
G.
Li
,
T.
Taniguchi
,
K.
Watanabe
, and
E. Y.
Andrei
, “
Moiré potential, lattice relaxation, and layer polarization in marginally twisted MoS2 bilayers
,”
Nano Lett.
23
,
73
81
(
2023
).
75.
J.
Martinez-Castro
et al, “
Scanning tunneling microscopy of an air sensitive dichalcogenide through an encapsulating layer
,”
Nano Lett.
18
,
6696
6702
(
2018
).
76.
S.
Masubuchi
et al, “
Dry pick-and-flip assembly of van der Waals heterostructures for microfocus angle-resolved photoemission spectroscopy
,”
Sci. Rep.
12
,
10936
(
2022
).
77.
S.
Son
et al, “
Strongly adhesive dry transfer technique for van der Waals heterostructure
,”
2D Mater.
7
,
041005
(
2020
).
78.
S.
Fan
,
Q. A.
Vu
,
M. D.
Tran
,
S.
Adhikari
, and
Y. H.
Lee
, “
Transfer assembly for two-dimensional van der Waals heterostructures
,”
2D Mater.
7
,
022005
(
2020
).
79.
Z.
Qiu
et al, “
Visualizing atomic structure and magnetism of 2D magnetic insulators via tunneling through graphene
,”
Nat. Commun.
12
,
70
(
2021
).
80.
M. A.
Altvater
et al, “
Charge density wave vortex lattice observed in graphene-passivated 1T–TaS2 by ambient scanning tunneling microscopy
,”
Nano Lett.
21
,
6132
6138
(
2021
).
81.
K.
Schouteden
,
A.
Volodin
,
T.
Moorkens
, and
C.
Van Haesendonck
, “
Peeling off graphene from co nanoparticle covered graphite in a scanning tunneling microscope
,”
Carbon
49
,
2258
2263
(
2011
).
82.
L.
Liu
et al, “
Scalable Van der Waals encapsulation by inorganic molecular crystals
,”
Adv. Mater.
34
,
2106041
(
2022
).
83.
M. H.
Hamidian
et al, “
Detection of a cooper-pair density wave in Bi2Sr2CaCu2O8+x
,”
Nature
532
,
343
347
(
2016
).
84.
S.
Guo
et al, “
An ultra-high vacuum system for fabricating clean two-dimensional material devices
,”
Rev. Sci. Instrum.
94
,
013903
(
2023
).
85.
Z.
Sun
et al, “
Exfoliation of 2D van der Waals crystals in ultrahigh vacuum for interface engineering
,”
Sci. Bull.
67
,
1345
1351
(
2022
).
86.
Y.
Huang
et al, “
Universal mechanical exfoliation of large-area 2D crystals
,”
Nat. Commun.
11
,
2453
(
2020
).
87.
W.
Wang
et al, “
Clean assembly of van der Waals heterostructures using silicon nitride membranes
,”
Nat. Electron.
6
,
981
990
(
2023
).
88.
Á.
Pásztor
,
A.
Scarfato
, and
C.
Renner
, “
Note: Mechanical in situ exfoliation of van der Waals materials
,”
Rev. Sci. Instrum.
88
,
076104
(
2017
).
89.
A.
Grubišić-Čabo
et al, “
In situ exfoliation method of large-area 2D materials
,”
Adv. Sci.
10
,
2301243
(
2023
).
90.
D.
Edelberg
,
H.
Kumar
,
V.
Shenoy
,
H.
Ochoa
, and
A. N.
Pasupathy
, “
Tunable strain soliton networks confine electrons in van der Waals materials
,”
Nat. Phys.
16
,
1097
1102
(
2020
).
91.
O.
Ochedowski
et al, “
Graphene on Si(111)7 × 7
,”
Nanotechnology
23
,
405708
(
2012
).
92.
M.
Mohan
et al, “
In-situ STS studies and first principles calculations on bare and Sn adsorbed UHV exfoliated WS2 layers
,”
IOP Conf. Ser. Mater. Sci. Eng.
1221
,
012046
(
2022
).
93.
H.
Brune
,
D.
Merk
, and
S.
Rusponi
, “
Large area monolayer graphene transfer in ultra-high vacuum
,”
SSRN Scholarly
(
published online, 2023
).
94.
H.
Imamura
et al, “
Twisted bilayer graphene fabricated by direct bonding in a high vacuum
,”
Appl. Phys. Express
13
,
075004
(
2020
).
95.
A. J.
Mannix
et al, “
Robotic four-dimensional pixel assembly of van der Waals solids
,”
Nat. Nanotechnol.
17
,
361
366
(
2022
).
96.
B.
El-Kerdi
et al, “
Evidence of strong Dzyaloshinskii–Moriya interaction at the cobalt/hexagonal boron nitride interface
,”
Nano Lett.
23
,
3202
3208
(
2023
).
97.
A.
Allain
,
J.
Kang
,
K.
Banerjee
, and
A.
Kis
, “
Electrical contacts to two-dimensional semiconductors
,”
Nat. Mater.
14
,
1195
1205
(
2015
).
98.
Y.
Zheng
,
J.
Gao
,
C.
Han
, and
W.
Chen
, “
Ohmic contact engineering for two-dimensional materials
,”
Cell Rep. Phys. Sci.
2
,
100298
(
2021
).
99.
S.
Batool
,
M.
Idrees
,
S.-T.
Han
,
V. A. L.
Roy
, and
Y.
Zhou
, “
Electrical contacts with 2D materials: Current developments and future prospects
,”
Small
19
,
2206550
(
2023
).
100.
J.
Chen
,
Introduction to Scanning Tunneling Microscopy
(
Oxford University Press
,
2021
), Vol.
69
.
101.
J.
Moser
,
A.
Barreiro
, and
A.
Bachtold
, “
Current-induced cleaning of graphene
,”
Appl. Phys. Lett.
91
,
163513
(
2007
).
102.
W.
Bao
et al, “
Lithography-free fabrication of high quality substrate-supported and freestanding graphene devices
,”
Nano Res.
3
,
98
102
(
2010
).
103.
Ç. Ö.
Girit
and
A.
Zettl
, “
Soldering to a single atomic layer
,”
Appl. Phys. Lett.
91
,
193512
(
2007
).
104.
N.
Staley
et al, “
Lithography-free fabrication of graphene devices
,”
Appl. Phys. Lett.
90
,
143518
(
2007
).
105.
K.
Ono
,
H.
Shimada
,
S.
Kobayashi
, and
Y. O. Y.
Ootuka
, “
A new fabrication method for ultra small tunnel junctions
,”
Jpn. J. Appl. Phys.
35
,
2369
(
1996
).
106.
S.
Matsui
and
Y.
Ochiai
, “
Focused ion beam applications to solid state devices
,”
Nanotechnology
7
,
247
(
1996
).
107.
M. A. F.
van den Boogaart
,
G. M.
Kim
,
R.
Pellens
,
J.-P.
van den Heuvel
, and
J.
Brugger
, “
Deep-ultraviolet–microelectromechanical systems stencils for high-throughput resistless patterning of mesoscopic structures
,”
J. Vac. Sci. Technol., B: Microelectron. Nanometer Struct.--Process., Meas., Phenom.
22
,
3174
3177
(
2004
).
108.
A.
Linklater
and
J.
Nogami
, “
Defining nanoscale metal features on an atomically clean silicon surface with a stencil
,”
Nanotechnology
19
,
285302
(
2008
).
109.
C.
Gärtner
et al, “
Fully ultrahigh-vacuum-compatible fabrication of submicrometer-spaced electrical contacts
,”
Rev. Sci. Instrum.
77
,
026101
(
2006
).
110.
R.
Lüthi
et al, “
Parallel nanodevice fabrication using a combination of shadow mask and scanning probe methods
,”
Appl. Phys. Lett.
75
,
1314
1316
(
1999
).
111.
P.
Bhaskar
et al, “
Mask aligner for ultrahigh vacuum with capacitive distance control
,”
Appl. Phys. Lett.
112
,
161602
(
2018
).
112.
F.
Ghahari
et al, “
An on/off Berry phase switch in circular graphene resonators
,”
Science
356
,
845
849
(
2017
).
113.
A.
Castellanos-Gomez
et al, “
Van der Waals heterostructures
,”
Nat. Rev. Methods Primers
2
,
58
(
2022
).
114.
Y.
Mao
,
L.
Wang
,
C.
Chen
,
Z.
Yang
, and
J.
Wang
, “
Thickness determination of ultrathin 2D materials empowered by machine learning algorithms
,”
Laser Photonics Rev.
17
,
2200357
(
2023
).
115.
J.
Sanchez-Juarez
,
M.
Granados-Baez
,
A. A.
Aguilar-Lasserre
, and
J.
Cardenas
, “
Automated system for the detection of 2D materials using digital image processing and deep learning
,”
Opt. Mater. Express
12
,
1856
1868
(
2022
).
116.
J.-L.
Uslu
et al, “
An open-source robust machine learning platform for real-time detection and classification of 2D material flakes
,”
Mach. Learn. Sci. Technol.
5
,
015027
(
2024
).
117.
M.
Abolhasani
and
E.
Kumacheva
, “
The rise of self-driving labs in chemical and materials sciences
,”
Nat. Synth.
2
,
483
492
(
2023
).
118.
B.
Burger
et al, “
A mobile robotic chemist
,”
Nature
583
,
237
241
(
2020
).
119.
N. J.
Szymanski
et al, “
An autonomous laboratory for the accelerated synthesis of novel materials
,”
Nature
624
,
86
91
(
2023
).
120.
D. A.
Boiko
,
R.
MacKnight
,
B.
Kline
, and
G.
Gomes
, “
Autonomous chemical research with large language models
,”
Nature
624
,
570
578
(
2023
).
121.
J. T.
Rapp
,
B. J.
Bremer
, and
P. A.
Romero
, “
Self-driving laboratories to autonomously navigate the protein fitness landscape
,”
Nat. Chem. Eng.
1
,
97
107
(
2024
).
122.
B.
Halliday
, “
An introduction to materials for use in vacuum
,”
Vacuum
37
,
583
585
(
1987
).
123.
J.
Shim
et al, “
Controlled crack propagation for atomic precision handling of wafer-scale two-dimensional materials
,”
Science
362
,
665
670
(
2018
).
124.
H.
Kim
et al, “
Remote epitaxy
,”
Nat. Rev. Methods Primers
2
,
40
(
2022
).
125.
C.
Yan
et al, “
An integrated quantum material testbed with multi-resolution photoemission spectroscopy
,”
Rev. Sci. Instrum.
92
,
113907
(
2021
).
126.
Y. X.
Zhou
,
A. T.
Johnson
,
J.
Hone
, and
W. F.
Smith
, “
Simple fabrication of molecular circuits by shadow mask evaporation
,”
Nano Lett.
3
,
1371
1374
(
2003
).
127.
K. S.
Novoselov
et al, “
Electric field effect in atomically thin carbon films
,”
Science
306
,
5696
(
2004
).
128.
L.
Wang
et al, “
One-dimensional electrical contact to a two-dimensional material
,”
Science
342
,
6158
(
2013
).
129.
W.
Zheng
et al, “
Emerging van der Waals junctions based on TMDs materials for advanced gas sensors
,”
Coord. Chem. Rev.
15
,
214151
(
2021
).
130.
W.
Wang
, “
Polymer-free assembly of ultraclean van der Waals heterostructures
,”
Nat. Rev. Phys.
4
,
504
(
2022
).