Two-dimensional transition metal dichalcogenides (2D TMDs) is one of the promising materials for future electronics since they have, not only superior characteristics, but also a versatility that conventional materials do not have with a few nanometer thickness. One of the prerequisites for applying these materials to device fabrication is to deposit an ultrathin film below 10nm with excellent uniformity. However, TMD has quite a different surface chemistry and is fragile to external conditions compared to conventional materials. Thus, thin film deposition on 2D TMD with excellent uniformity using conventional deposition techniques is quite challenging. Currently, the most adequate deposition technique for sub-10nm-thick film growth is atomic layer deposition (ALD). A thin film is formed on the surface by the reaction between chemical and surface species based on the self-limiting growth manner. Owing to its unique and superior growth characteristics, such as excellent uniformity and conformality, ALD is an essential deposition technique for nanoscale device fabrication. However, since 2D TMD has a lack of reaction sites on the surface, various studies have reported that ALD on 2D TMDs surfaces without any treatment showed an island growth mode or formation of clusters rather than continuous films. For this reason, recent studies have been focused on the deposition of an ultrathin film on 2D TMDs with excellent uniformity. For a decade, there have been various approaches to obtain uniform films on 2D TMDs using ALD. Among them, the authors focus on the most frequently researched methods and adsorption control of chemical species by modifying the process parameters or functionalization of new chemical species that can assist adsorption on the chemically inert 2D TMD surface. In this review, the overall research progress of ALD on 2D TMD will be discussed which would, in turn, open up new horizons in future nanoelectronics fabrication using 2D TMDs.

Two-dimensional (2D) materials have been highlighted as promising materials that can exceed the current limitation of the three-dimensional (3D) bulk materials. Unlike bulk materials, 2D materials have been known to possess superior electrical, optical, and mechanical properties that cannot be obtained from conventional 3D materials.1–4 For instance, graphene, the first reported 2D material, had been highlighted as a promising material that can reach a breakthrough in current electronics.5 Owing to its hexagonal-honeycomb lattice structure, carbon atoms contain sp2-hybridized bonded layer; thus, it exhibits extremely high conductivity and carrier mobility with only a nanometer-thick layer.6 Despite these unprecedented characteristics, however, it has not been widely used as a material for electronic devices since it has no intrinsic bandgap, which is essential for a switching characteristic. Many researchers have tried to open the bandgap of graphene but had some problems obtaining a certain level of bandgap using graphene, retaining its pristine characteristics.7–9 For this reason, 2D transition metal dichalcogenide (TMD) has been considered a promising material for the channel material of electronics instead of graphene. Characteristics of TMDs are not only superior compared to those of conventional materials but also can be varied from a metal to a semiconductor and even to an insulator by a variety of combinations of metals and chalcogen elements.10 Molybdenum diselenide (MoS2), one of the representative TMD materials, has been highlighted as a promising channel material of the transistor, showing remarkable electrical characteristics that include high mobility and an excellent on/off ratio.11 In addition, direct to indirect bandgap transition of ultrathin 2D TMDs depending on the number of layers12,13 and excellent mechanical durability14–16 is used for the purpose of high-performance optical or mechanical device fabrication. Likewise, because of their superior properties, most 2D TMD research studies have focused their attention on the investigation of their own properties and expansion to the various applications using these peculiar characteristics over a decade.

To utilize 2D TMDs, deposition on the TMDs is frequently required in order to fabricate various devices17,18 or synthesize a hybrid material for energy applications with inorganic/organic materials.19,20 To date, there are numerous deposition techniques, including vapor deposition (physical or chemical), solution process, etc. Among them, atomic layer deposition (ALD), one of the specialized chemical vapor depositions that consists of sequential exposure of chemical species, is the most appropriate deposition method for nanoscale device fabrication.21,22 Based on self-limiting growth characteristics, precursor molecules chemisorb onto the surface reaction sites and subsequent reactant molecules react with the adsorbed precursor species forming a subnanometer thick film within a cycle. Owing to its unique deposition manner, various advantages, including atomic-scale thickness controllability and large area uniformity, can be obtained. Moreover, high conformal coating even on a 3D-complex structure like nanoflakes is also available.23–25 However, since the 2D material has quite a different surface chemistry compared to conventional 3D materials, the growth characteristics of ALD on the 2D material shows a significantly different aspect. Unlike conventional 3D materials, there is a lack of dangling bond on TMDs and this leads to a chemically inert surface on the basal plane of these materials.10 This indicates that no chemisorption of the chemical species is available on the surface. Even the exfoliated or synthesized 2D TMDs has inherent defects that are generated during the process, uneven distribution of reaction sites on TMDs induces nonuniform deposition on the surface after the ALD process,17 leading to a critical limitation on applying ALD on 2D materials. Therefore, many research studies have been focused on the uniform deposition of ALD-grown material without destroying or damaging a pristine TMD structure.

In this paper, ALD on 2D TMDs will be highlighted from the perspective of growth characteristics. There are numerous reports of ALD on 2D TMDs, but we will only focus on the growth mechanism and the methods for excellent uniformity. In terms of surface chemistry, dissimilar deposition mechanism of ALD on 2D TMDs will be discussed by referring to previous reports. Moreover, diverse approaches for uniform deposition of the ALD-grown material on 2D TMDs and its applications will be described in detail by suggesting the outlook and prospect of ALD on 2D materials.

The notable difference between conventional materials and 2D TMDs is the presence of the reaction sites on the surface. As mentioned in the Introduction briefly, a conventional material has a dangling bond on top of the surface that enables reactions with different chemical species. In the case of ideally infinite 2D materials, on the other hand, there was no dangling bond theoretically because of its peculiar crystal configuration26 that indicates 2D materials are chemically inert. The absence of the dangling bond on the basal plane is one of the major advantages of 2D TMDs, which can be avoided by electrical deterioration from interface traps.27 From the perspective of the ALD process, however, free of dangling bond suppresses a material growth on the surface. For instance, when the precursor or the reactant is exposed to the reactor on a dangling bond-rich substrate, a chemical reaction occurs on top of the surface by reacting with the dangling bonds (i.e., surface reaction sites), and, therefore, the film starts to grow. However, since there is lack of a reaction site on pristine 2D TMDs, film growth inhibition on the basal plane of 2D TMD is observed when ALD is conducted on the 2D materials.17,28

In the case of surface properties of nonideal finite 2D TMDs, it has slightly different surface chemistry. Initially, most of the 2D TMD research was conducted with a TMD flake that was mechanically29 or chemically exfoliated30,31 from the single-crystal TMD. Among them, a mechanical exfoliation process that uses an adhesive tape for peeling off the ultrathin 2D TMD flake from the bulk TMD crystal ensures TMD flakes with single crystalline, high purity, and immaculateness. In particular, since a mechanically exfoliated 2D TMD flake has a very low level of defect concentration, using this flake helps us to understand the intrinsic characteristics of 2D TMD.32,33 However, some defects or edge sites that exist on the exfoliated flake, generated during synthesis or transfer process (e.g., point defect or dislocations), also have a significant effect on its surface characteristics as well.34,35 Since these have dangling bonds that are chemically reactive,36–39 defects or edge sites of exfoliated TMD act as reaction sites for the gas molecules or chemicals.40,41 While the exfoliated 2D TMDs showed superior characteristics that are suitable for laboratory-scale device fabrication, it is not appropriate for scaling up of the large-scale fabrication and controlling thickness and size of the flakes that are essential for characteristic modulation. Accordingly, the development of fabrication methods for fabricating 2D TMD flakes with large scale and high uniformity have also been widely researched. In contrast to exfoliation, however, CVD-grown TMDs showed a relatively high level of defect densities due to imperfection of the synthesis process, which induces various structural defects in the material.42 From the perspective of surface chemistry, a high level of defect densities on 2D TMD imply that a reaction with the following chemical species would be more favorable on the CVD-grown TMD than on the exfoliated one.

Since 2D TMDs consist of two or more different elements, transition metal and chalcogen, it is much more possible to have various types of defects than graphene which only consists a single carbon element. A previous study showed various point defects of 2D MoS2 such as a monosulfur vacancy, a disulfur vacancy, a vacancy complex with Mo and three neighboring sulfur atoms, a vacancy complex with Mo and three disulfur pairs, antisite defects with a Mo atom, etc.42 In the case of graphene, the adsorption tendencies of chemical species on graphene depends on the type of defects.43 Likewise, the adsorption of chemical species on 2D TMDs differs from what the chemical species or a kind of defect sites are. A previous report investigated adsorption of various molecules on 2D MoS2.44 The authors calculated adsorption energies between various molecules (CO2, CO, O2, H2O, NO, N2, and H2) and basal plane or defects (S vacancy, S divacancy, Mo vacancy, one or two Mo atoms on S mono- and divacancy, and one or two S atoms on Mo mono- and divacancy) of 2D MoS2 by using density functional theory (DFT) calculation. They revealed that all the molecules have larger adsorption energy on defect sites than that of on the basal plane, and the adsorption energy varied depending on the surface species and the type of chemical species. Consequently, defects on 2D TMDs can play a role as highly reactive sites and trap sites for different molecules by exposure to chemical functional groups.45 

Based on the classical ALD theory, a film is formed on reaction sites of the surface by the chemical reaction between chemical species (i.e., precursor and reactant).21 This chemical reaction, named as dissociative chemisorption, occurs with the help of external energy, forming a chemical bond between adsorbent (solid) and adsorbate (gas). During the ALD process, since chemical species could only react with surface reaction sites of the substrate and residual molecules that are not reacted would be purged out, this self-limiting growth based on the chemisorption is irreversible in general.46 

In contrast to this chemisorption-based reaction, another adsorption mechanism, physisorption, becomes more and more important for the modern ALD process, especially on 2D materials. Unlike the chemisorption-based process, when chemical species approaches on top of the 2D surface, these are physically attached on the surface with weak intermolecular force (e.g., van der Waals force) without making bond modification between an adsorbent and adsorbate. Although physisorption is one of the fundamental bonding theories, this had been seldom considered in the ALD process until carbon-based materials (i.e., carbon nanotube47 or graphene40,48,49) came out to the world since most of the ALD reaction were conducted on the substrate with full of dangling bonds that enable a favorable chemical reaction on the surface. However, since graphene itself lacks dangling bonds on the top of the surface, it showed a chemically inert surface, resulting in growth inhibition or abnormal nucleation growth on top of the graphene surface.40,50,51 To overcome this limitation, many researchers have tried to physically attach molecules for the purpose of artificial adsorption sites for the reaction with subsequent chemical species.

Figure 1(a) shows a plot of Lennard–Jones potential model, which is generally used for molecular adsorption model. In this figure, the depth of the potential well indicates the attraction force between solid (substrate) and gas (precursor or reactant), and the x axis indicates a distance between them. As shown in the figure, the attraction force of physisorption is relatively low (∼40 meV) with a relatively large distance (>3 Å) compared to that of chemisorption (∼1 eV) with a small distance (<3 Å).52,53 Because of its weak bond strength and relatively long distance, physically attached molecules can be easily detached by external energy source, for instance, thermal heating54 or long purging time.55 That is, since physisorption does not induce bond modification between the adsorbate and the adsorbent, this could be reversible. Figure 1(b) shows a simple schematic of ALD on the 2D material with and without physisorption. As we mentioned in the previous paragraph briefly, ALD on the 2D substrate is favorable on the defect or edge sites, not on the basal plane. However, if some chemical species are physically attached to the basal plane of the 2D material, it may act as a reaction site for the following precursor, resulting in nucleation on the basal plane. In other words, if molecules can be physically attached on the entire basal plane, uniform thin film deposition on dangling bond-free 2D TMDs would be also available by ALD.

FIG. 1.

(a) Potential energy for physisorption and chemisorption, based on the Lennard–Jones potential model and (b) schematic of the ALD growth on the defect-free 2D TMD surface with and without physisorption.

FIG. 1.

(a) Potential energy for physisorption and chemisorption, based on the Lennard–Jones potential model and (b) schematic of the ALD growth on the defect-free 2D TMD surface with and without physisorption.

Close modal

In this paragraph, we will discuss the recent progress of ALD on 2D TMDs to obtain high uniform films even with a sub-10 nm thickness level that is essential for nanoscale device integration with high performance. Over a decade, numerous reports of ALD on 2D TMDs have been reported. In this report, however, we will mainly focus on the growth mechanism of ALD on 2D TMDs from a viewpoint of surface chemistry. Also, various methods for uniform thin film deposition will be described in detail. All reports of ALD on 2D TMDs hitherto are summarized in Table I as a function of the synthesis method of TMD.

TABLE I.

ALD processes on 2D TMDCs.

SubstrateALD materialsMethod for improvement uniformityPrecursorReactantGrowth temperature (°C)Thickness of ALD materials (nm)EOT of ALD materials (nm)Growth typeApplicationsCharacteristicsYear
Exfoliated
MoS2 
Al2O3 None No information No information No information 30 13 — FET — 2013 
None TMA H2O/O3 200 <5 <2.2 Island/film — Conformally deposited Al2O3 using O3 2014 
Al seed layer TMA H2200 — Dual-gate charge-trap memory Estimated stable retention of ∼28% charge loss after 10 years 2015 
None TMA H2200 10 4.3 — FET Improvement of device stability 2016 
UV-O3 TMA H2— 50 21.7 — FET Enhancement of carrier mobility from 24.3 to 41.2 cm2 V−1 s−1 2016 
UV-O3 TMA H2— 10 4.3 Film MOSCAP – Electron density: ∼1017 cm−3
–Minimum Dit: ∼1011 eV−1 cm−2 
2016 
x seed layer — — 200 60 26 — Tunnel FET  2018 
None TMA H2— 50 21.7 — — Negative shift and broadening of PL emission 2018 
None TDMAH H2— 30 4.7 — FET – Electron mobility >200 cm2 V−1 s−1
–Current on/off ratios: 108 
2011 
None TDMAH H2170 20 3.1 — Direct-coupled FET logic First demonstration of integrated multistage system by using 2D materials 2012 
None TEMAH H2— 30 4.7 — FET Charge carrier density >n2D ∼3.6 × 1013 cm−2 2013 
None TDMAH H2200 15 2.3 Island — — 2013 
MgO, Al2O3, Y2O3 buffer layer TDMAH H295 2.8 0.4 Island/film FET – Conformally deposited HfO2 on Al2O3 and Y2O3
–Electron mobility of 63.7 cm2/V s
–On/off current ratio: >108 
2014 
HfO2 UV-O3 TDMAH H2200 ∼20 ∼3.1 Island — Conformally deposited HfO2 after UV-O3 2014 
None — — — 30 4.7 — FET biosensor — 2015 
None — — 250 16 2.5 — FET – Electron mobility of 25.3 cm2/V s
–SS: 86.0 mV/decade
–On/off ratio: 107 
2015 
None — — 90 6.5 1.0 Island Floating gate memory-based FET – On/off ratio: >106
–Memory window: 10 V
–Stable program /erase ratio: ∼105 
2015 
Al seeding layer — — 95 6.5 1.0 — Phototransistors – Photoresponsivity: 59.2 A W−1
–Detectivity: 3.8 × 1010 Jones 
2016 
UV-O3 TDMAH H2200 1.2 Film FET — 2017 
TiO2 None TTIP H2200 2.7 0.1 — FET Reduction of contact resistance from 34 to 28 kΩ 2016 
ZrO2 None — — 120 17.5 2.7 — FET Electron sheet densities: ∼1.0 × 1013 cm−2 2013 
Exfoliated
WS2 
Al2O3 None TMA H2— 50 21.7 — — Negative shift and broadening of PL emission 2018 
HfO2 Al seeding layer — — 95 6.5 1.0 — Phototransistors – Photoresponsivity: 59.2 AW−1
–Detectivity: 3.8 × 1010 Jones 
2016 
Exfoliated
WS2 
TiO2 None TTIP — 200 2.7 0.1 — FET Low contact resistance (24 and 34.8 kΩ μιη for Ti–TiO2–WS2 and Pd–TiO2–WS22016 
Exfoliated
MoSe2s 
HfO2 UV-O3 TDMAH H2200 0.6 Film — Improvement of dielectric constant 2015 
Al2O3 None TMA H2— 50 21.7 — — Negative shift and broadening of PL emission 2018 
Al2O3 None TMA O3 30–300 2.4 1.0 Island/film — Improvement of uniformity 2017 
Exfoliated
WSe2 
HfO2 None TMA H2— 50 21.7 — — Negative shift and broadening of PL emission 2018 
UV-O3 TDMAH H2200 0.6 Film — Improvement of dielectric constant 2015 
ZrO2 None No information No information 120 17.5 2.7 — FET – Electron sheet densities: ∼2.5 × 1012 cm−2 2013 
Exfoliated
WS2 
Al2O3 None TMA H2— 15 6.5 — FET Improvement of oxidation stability 2016 
None TMA H2300 30 13 — FET — 2013 
None — — 170 15 6.5 Island FET – Improvement the channels on-state conductance
–Increase the back-gated mobility by a factor of ∼3 
2013 
Native ΜοΟx TMA H2200 16 6.9 Flim FET Maximum drain current of 62.5 mA/mm at 2 V drain bias 2013 
None TMA O2 plasma — 20 8.7 — FET Temperature dependency of resistance passivated with ALD deposited Al2O3 2014 
Al2O3 Perylene bisimide TMA H280 1.3 Film — — 2015 
None TMA H2200 10 4.3 — Solar cell Improved power conversion fficiency:5.6% 2016 
CVD
MoS2 
None TMA H2150 1.3 Film FET Improved mobility with decreased H2O dosing time: 40 2016 
None TMA H2500 54 23.4 Film Superabsorber Strong light absorption (>70%) in atomically thin MoS2 films (≥4 layers) 2016 
None TMA H2— 4–20 1.7–8.7 Nanosphere — The electrons transfer is remarkable suppressed by Al2O3 2017 
None — — — 30 4.7 — FET First demonstration of common-source amplifier and active frequency mixer based on CVD MoS2 2015 
HfO2 None — — <90 16 2.5 — FET High carrier mobility: ∼36.4 cm2/V s 2016 
Al seeding layer — — — 30 4.7 — FET On/off ratio >106, cutoff current <1 pA μm−1 2018 
ZrO2 None TEMAZ H2200 27 4.2 Rough film FET High dielectric breakdown field of 4.9 MV cm−1 2018 
CVD
WS2 
Al2O3 Perylene bisimide TMA H280 1.3 Film — — 2015 
MBE
WSe2 
Al2O3 TiOPc TMA H2120 27 11.7 Film FET – Leakage current of 0.046 pA/μm2 at 1 V gate bias 2016 
SubstrateALD materialsMethod for improvement uniformityPrecursorReactantGrowth temperature (°C)Thickness of ALD materials (nm)EOT of ALD materials (nm)Growth typeApplicationsCharacteristicsYear
Exfoliated
MoS2 
Al2O3 None No information No information No information 30 13 — FET — 2013 
None TMA H2O/O3 200 <5 <2.2 Island/film — Conformally deposited Al2O3 using O3 2014 
Al seed layer TMA H2200 — Dual-gate charge-trap memory Estimated stable retention of ∼28% charge loss after 10 years 2015 
None TMA H2200 10 4.3 — FET Improvement of device stability 2016 
UV-O3 TMA H2— 50 21.7 — FET Enhancement of carrier mobility from 24.3 to 41.2 cm2 V−1 s−1 2016 
UV-O3 TMA H2— 10 4.3 Film MOSCAP – Electron density: ∼1017 cm−3
–Minimum Dit: ∼1011 eV−1 cm−2 
2016 
x seed layer — — 200 60 26 — Tunnel FET  2018 
None TMA H2— 50 21.7 — — Negative shift and broadening of PL emission 2018 
None TDMAH H2— 30 4.7 — FET – Electron mobility >200 cm2 V−1 s−1
–Current on/off ratios: 108 
2011 
None TDMAH H2170 20 3.1 — Direct-coupled FET logic First demonstration of integrated multistage system by using 2D materials 2012 
None TEMAH H2— 30 4.7 — FET Charge carrier density >n2D ∼3.6 × 1013 cm−2 2013 
None TDMAH H2200 15 2.3 Island — — 2013 
MgO, Al2O3, Y2O3 buffer layer TDMAH H295 2.8 0.4 Island/film FET – Conformally deposited HfO2 on Al2O3 and Y2O3
–Electron mobility of 63.7 cm2/V s
–On/off current ratio: >108 
2014 
HfO2 UV-O3 TDMAH H2200 ∼20 ∼3.1 Island — Conformally deposited HfO2 after UV-O3 2014 
None — — — 30 4.7 — FET biosensor — 2015 
None — — 250 16 2.5 — FET – Electron mobility of 25.3 cm2/V s
–SS: 86.0 mV/decade
–On/off ratio: 107 
2015 
None — — 90 6.5 1.0 Island Floating gate memory-based FET – On/off ratio: >106
–Memory window: 10 V
–Stable program /erase ratio: ∼105 
2015 
Al seeding layer — — 95 6.5 1.0 — Phototransistors – Photoresponsivity: 59.2 A W−1
–Detectivity: 3.8 × 1010 Jones 
2016 
UV-O3 TDMAH H2200 1.2 Film FET — 2017 
TiO2 None TTIP H2200 2.7 0.1 — FET Reduction of contact resistance from 34 to 28 kΩ 2016 
ZrO2 None — — 120 17.5 2.7 — FET Electron sheet densities: ∼1.0 × 1013 cm−2 2013 
Exfoliated
WS2 
Al2O3 None TMA H2— 50 21.7 — — Negative shift and broadening of PL emission 2018 
HfO2 Al seeding layer — — 95 6.5 1.0 — Phototransistors – Photoresponsivity: 59.2 AW−1
–Detectivity: 3.8 × 1010 Jones 
2016 
Exfoliated
WS2 
TiO2 None TTIP — 200 2.7 0.1 — FET Low contact resistance (24 and 34.8 kΩ μιη for Ti–TiO2–WS2 and Pd–TiO2–WS22016 
Exfoliated
MoSe2s 
HfO2 UV-O3 TDMAH H2200 0.6 Film — Improvement of dielectric constant 2015 
Al2O3 None TMA H2— 50 21.7 — — Negative shift and broadening of PL emission 2018 
Al2O3 None TMA O3 30–300 2.4 1.0 Island/film — Improvement of uniformity 2017 
Exfoliated
WSe2 
HfO2 None TMA H2— 50 21.7 — — Negative shift and broadening of PL emission 2018 
UV-O3 TDMAH H2200 0.6 Film — Improvement of dielectric constant 2015 
ZrO2 None No information No information 120 17.5 2.7 — FET – Electron sheet densities: ∼2.5 × 1012 cm−2 2013 
Exfoliated
WS2 
Al2O3 None TMA H2— 15 6.5 — FET Improvement of oxidation stability 2016 
None TMA H2300 30 13 — FET — 2013 
None — — 170 15 6.5 Island FET – Improvement the channels on-state conductance
–Increase the back-gated mobility by a factor of ∼3 
2013 
Native ΜοΟx TMA H2200 16 6.9 Flim FET Maximum drain current of 62.5 mA/mm at 2 V drain bias 2013 
None TMA O2 plasma — 20 8.7 — FET Temperature dependency of resistance passivated with ALD deposited Al2O3 2014 
Al2O3 Perylene bisimide TMA H280 1.3 Film — — 2015 
None TMA H2200 10 4.3 — Solar cell Improved power conversion fficiency:5.6% 2016 
CVD
MoS2 
None TMA H2150 1.3 Film FET Improved mobility with decreased H2O dosing time: 40 2016 
None TMA H2500 54 23.4 Film Superabsorber Strong light absorption (>70%) in atomically thin MoS2 films (≥4 layers) 2016 
None TMA H2— 4–20 1.7–8.7 Nanosphere — The electrons transfer is remarkable suppressed by Al2O3 2017 
None — — — 30 4.7 — FET First demonstration of common-source amplifier and active frequency mixer based on CVD MoS2 2015 
HfO2 None — — <90 16 2.5 — FET High carrier mobility: ∼36.4 cm2/V s 2016 
Al seeding layer — — — 30 4.7 — FET On/off ratio >106, cutoff current <1 pA μm−1 2018 
ZrO2 None TEMAZ H2200 27 4.2 Rough film FET High dielectric breakdown field of 4.9 MV cm−1 2018 
CVD
WS2 
Al2O3 Perylene bisimide TMA H280 1.3 Film — — 2015 
MBE
WSe2 
Al2O3 TiOPc TMA H2120 27 11.7 Film FET – Leakage current of 0.046 pA/μm2 at 1 V gate bias 2016 

The initial research of ALD on 2D TMDs began on the exfoliated TMD flakes from single-crystal TMD. In 2011, Radisavljevic et al. investigated device properties of MoS2 top-gate transistors with or without high-k dielectrics.11 For a fabrication monolayer MoS2 field-effect transistor (FET) without high-k dielectrics, the MoS2 monolayer (∼6.5 Å) was mechanically exfoliated onto a 270-nm-thick SiO2/degenerately doped silicon substrate. Since the mobility of 2D materials combined with high-k dielectrics could be improved owing to dielectric screening,56–59 the authors deposited ALD HfO2 on monolayer MoS2 without any treatment. HfO2 was chosen high-k dielectrics due to its high dielectric constant (k ∼ 25) and large bandgap (∼5.7 eV).60 For the deposition of HfO2, tetrakis(dimethylamino)hafnium (TDMAHf) was used as an Hf precursor, and H2O was used as a reactant, and 30 nm of HfO2 was deposited on the MoS2 flake for top-gate dielectric. After deposition, the electrical properties of FET were greatly enhanced owing to the screening effect, as mentioned before. However, even though direct ALD on an untreated MoS2 surface might induce numerous nucleation, their result showed insignificant change on surface morphology before and after ALD HfO2. In general, nucleation and growth of ALD films favorably occur on surfaces that have a high concentration of hydroxyl (–OH)-terminated groups, which is considered as adsorption sites for the precursors.61 However, although the mechanically exfoliated TMD flake has no inherent –OH species, relatively flat thin film was observed. This result was verified later that, when the thickness of HfO2 was quite thick, the nucleated HfO2 started to be connected eventually, forming a relatively flat thin film.62 

Likewise, there is no specific obstacle to deposit relatively thick dielectric (>15 nm) on dangling bond-free TMD substrates using ALD. However, the thickness of the dielectric should be decreased for the integration of complementary metal oxide semiconductor (CMOS).41 In addition, the electrical properties of TMD-based FET would be remarkably improved when the thickness of the gate dielectric is small enough, especially at the sub-10 nm region, resulting in better control of the channel and larger drive current.59 Therefore, deposition of ultrathin high-k dielectrics with excellent uniformity on 2D TMD layers is a prerequisite for high-performance device fabrication, and after that time, many studies have focused on the deposition that satisfies these requirements.

The initial approach to obtain a uniform thin film (<10 nm) on 2D TMDs was controlling of process temperature. In 2012, Liu et al. reported Al2O3 ALD on exfoliated MoS2 and boron nitride (BN) with an experimental and theoretical method.28 The deposition of ALD Al2O3 was conducted on mechanically exfoliated 2D MoS2 and BN films using trimethylaluminum (TMA) and H2O by differing process temperatures to investigate the effect of temperature on the uniform deposition. Figure 2(a) shows atomic force microscopy (AFM) images of MoS2 and BN substrates after 111 cycles of ALD Al2O3 depending on the process temperature (200, 300, and 400 °C, respectively). The nominal thickness of Al2O3 that was calculated based on the growth rates of Al2O3 on the SiO2 substrate was about 10 nm since there was no temperature dependence on growth rates of ALD Al2O3 on SiO2 from 200 to 400 °C, indicating chemical adsorption of TMA on –OH groups was not affected by temperature from 200 to 400 °C. However, the morphology of ALD Al2O3 on BN and MoS2 was strongly affected as a function of growth temperature. At 200 °C, ALD Al2O3 that was deposited on BN and MoS2 showed excellent uniformity without voids or pinholes. As the process temperature of ALD Al2O3 increased, however, many voids and pinholes started to be formed. When the process temperature was increased up to 400 °C, numerous islands of Al2O3 were observed on the surface rather than on a thin film, regardless of the substrate. Surface coverage of ALD Al2O3 on MoS2 that is calculated from these AFM images was slightly higher than that on BN, regardless of the process temperature. The aspect of the film growth depending on the process temperature is closely related to the physisorption of precursor on MoS2 and BN. As discussed in Sec. III, precursor (or reactant) molecule would be physically adsorbed on the dangling bond-free substrate, forming a weak van der Waals interaction. As the temperature increases, however, some physisorbed molecules could be desorbed by breaking a weak bond and nonuniform reaction sites on TMDs remain. While the physisorbed molecules on the basal plane gradually detached as the process temperature increased, edge sites of TMDs still remained and acted as reaction sites because of the presence of dangling bonds.40,50 Accordingly, the ALD process could not deposit uniform Al2O3 films on BN and MoS2 surfaces as the process temperature increased.

FIG. 2.

(a) Atomic force microscope image of BN and MoS2 after 111 cycles of ALD Al2O3 at 200, 300, and 400 °C (scan size: 2 × 2 μm2 with a scale bar of 500 nm). (b) An illustrative Lennard–Jones potential model for physical adsorption at 2D crystal surfaces. Reprinted with permission from Liu et al., Appl. Phys. Lett. 100, 152115 (2012). Copyright 2012, AIP Publishing LLC.

FIG. 2.

(a) Atomic force microscope image of BN and MoS2 after 111 cycles of ALD Al2O3 at 200, 300, and 400 °C (scan size: 2 × 2 μm2 with a scale bar of 500 nm). (b) An illustrative Lennard–Jones potential model for physical adsorption at 2D crystal surfaces. Reprinted with permission from Liu et al., Appl. Phys. Lett. 100, 152115 (2012). Copyright 2012, AIP Publishing LLC.

Close modal

In order to investigate the interaction between precursors and the surface of 2D materials and understand the reaction mechanism of initial ALD cycles on 2D materials, adsorption energy between TMA and MoS2 or BN were calculated using DFT calculation. The physical adsorption energy of TMA on the BN surface was 8.7 kcal/mol, which was larger than that of H2O on BN. The physisorption energy of TMA on MoS2 was 21.9 kcal/mol, which is larger than that of H2O on MoS2. This indicates that TMA is more strongly physisorbed on 2D materials than H2O. In addition, TMA has the largest interaction energy with MoS2 because of the following reasons; positively charged Al can more favorably interact with negatively charged N and S than negatively charged O, and polarizabilities of TMA and MoS2 are larger than those of H2O and BN, respectively. Based on these results, Lennard–Jones potential is depicted in Fig. 2(b). These calculation results confirmed the higher coverage of Al2O3 on MoS2 than on BN in the high temperature range (300–400 °C), which corresponds to the coverage difference. The authors revealed that temperature dependent growth indicates that ALD Al2O3 on the 2D substrate is mainly controlled by physisorption of the precursors, instead of chemisorption.

From both experimental and calculation results, the authors deduced that the growth of ALD Al2O3 on 2D materials is critically affected by physisorption, not chemisorption. Since physical adsorption on 2D TMDs is highly affected by not only adsorption energies but also polarizabilities of chemical species and TMD, a proper selection of the precursor and the reactant depending on the 2D material is highly demanded to enhance physisorption on a dangling bond-free substrate.

Even though lowering process temperature is one of the possible methods to obtain a uniform thin film on 2D TMDs, however, it also brings undesirable effects as well. For instance, low temperature ALD leads to poor film density63 and much incorporation of impurities from precursor (e.g., C, N, O, etc.)64 due to insufficient activation energy for the reaction between precursor and surface reaction sites. For these reasons, it is highly required to obtain uniform film using ALD without degrading the film quality.

In this section, various surface treatment methods will be discussed to acquire a highly uniform thin film on the 2D TMD substrate. Diverse methods including not only attaching molecules but also generating –OH species or defects on the TMD surface will be described in detail.

1. Molecule functionalization on 2D TMDs

In 2013, McDonnell et al. reported ALD HfO2 on a mechanically exfoliated MoS2 substrate. TDMAHf and H2O were used for HfO2 deposition at 200 °C.17 In this report, 30 nm of HfO2 deposited on MoS2 covered the entire MoS2 layer but not uniformly, unlike previous reports.11,28,65 The different thing from those and this report was presumed by the residue removal process. In order to fabricate TMD-based FET, subsequent coating of organic materials followed by removal with solvents is typically carried out after the transfer or lithography process.65,66 However, residues that are generated during these processes are hardly removed even after UHV annealing,67 thus, these remained surface contaminants may wact as reaction sites. To investigate whether intentionally attached molecules (i.e., organic or solvent contaminants) would promote nucleation of ALD HfO2 on MoS2 flakes, they were soaked into acetone or N-methyl-2-pyrrolidone (NMP) or spin-coated with poly(methyl methacrylate) (PMMA) prior to ALD HfO2. As shown in Fig. 3(a), even though there was no significant change after ALD when MoS2 was submerged in acetone, the clear difference in phase images of AFM was observed after PMMA coating or the NMP treatment. From the AFM images, the author deduced that the organic residue or solvent, which remained on the MoS2 substrate acted as a nucleation promotion site that improved uniformity of ALD HfO2. They also confirmed the effect of purge time of the ALD process on the physisorption of chemical species. With extremely long purge time between precursor and reactant exposure (over 6 h), no HfO2-related peaks were observed from x-ray photoelectron spectroscopy (XPS) [Fig. 3(b)], while there were clear Hf4f and O1s peaks when the precursor and reactant purge time were 5 and 10 s, respectively. This could be a piece of evidence that physisorbed precursor molecules can be detached by long purge time. Therefore, physisorption or molecule attachment (functionalization) would enhance the nucleation of ALD on the low number of reaction sites on the TMD surface.

FIG. 3.

AFM topography and phase images of ALD HfO2 on (a) untreated MoS2 after mechanical exfoliation, (b) MoS2 submerged in acetone for 2.5 h, (c) MoS2 submerged in NMP for 2.5 h, (d) MoS2 coated with PMMA and submerged in acetone for 2.5 h, and (e) coated with PMMA and submerged in NMP for 2.5 h. (b) XPS spectra of ALD HfO2 on untreated MoS2 with long purge time (>6 h) and short purge time (5 and 10 s for precursor and reactant purge time, respectively). (c) Three-dimensional STM image of the TiOPc monolayer with a molecular resolution, (d) schematic of the TiOPc monolayer, including the WSe2 monolayer and the HOPG, and (e) AFM images of Al2O3/bulk WSe2 without and (f) with the TiOPc layer. (a) and (b) Reprinted with permission from McDonnell et al., ACS Nano 7, 10354 (2013). Copyright 2013, American Chemical Society. (c)–(f) Reprinted with permission from Park et al., ACS Nano 10, 6888 (2016). Copyright 2016, American Chemical Society.

FIG. 3.

AFM topography and phase images of ALD HfO2 on (a) untreated MoS2 after mechanical exfoliation, (b) MoS2 submerged in acetone for 2.5 h, (c) MoS2 submerged in NMP for 2.5 h, (d) MoS2 coated with PMMA and submerged in acetone for 2.5 h, and (e) coated with PMMA and submerged in NMP for 2.5 h. (b) XPS spectra of ALD HfO2 on untreated MoS2 with long purge time (>6 h) and short purge time (5 and 10 s for precursor and reactant purge time, respectively). (c) Three-dimensional STM image of the TiOPc monolayer with a molecular resolution, (d) schematic of the TiOPc monolayer, including the WSe2 monolayer and the HOPG, and (e) AFM images of Al2O3/bulk WSe2 without and (f) with the TiOPc layer. (a) and (b) Reprinted with permission from McDonnell et al., ACS Nano 7, 10354 (2013). Copyright 2013, American Chemical Society. (c)–(f) Reprinted with permission from Park et al., ACS Nano 10, 6888 (2016). Copyright 2016, American Chemical Society.

Close modal

In 2016, Park et al. successfully deposited Al2O3 on exfoliated WSe2 by organic functionalization using the titanyl phthalocyanine (TiOPc) monolayer.68 They were motivated by a previous report of uniform ALD on graphene using TiOPc.69 From the previous first-principle calculation, they found a strong binding energy between dimethyl aluminum, surface species that were generated by dissociation of TMA with surface hydroxyl group, and elements of TiOPc (C, N, or O), over 1.5 eV, which indicates that TMA would attach on the TiOPc layer favorably, resulting in reaction sites for subsequent reactants. In this report, WSe2 was deposited on highly ordered pyrolytic graphite (HOPG) using molecular beam epitaxy (MBE), then a monolayer of TiOPc was also deposited using MBE. From the scanning tunneling microscopy (STM) images after TiOPc deposition, shown in Fig. 3(c), highly ordered morphology without defect was observed, indicating that uniform and thin TiOPc monolayer was well deposited on MBE-grown WSe2. The energy level that is represented with color profiles, and this shows that every single TiOPc molecule has a single bright peak that is negatively charged, providing potential binding sites for polar ALD precursors [Fig. 3(d)]. To confirm that this TiOPc is effective for nucleation of following ALD, ALD Al2O3 was deposited on bare and TiOPc-deposited WSe2. As shown in Figs. 3(e) and 3(f), after TiOPc deposition, smooth and flat Al2O3 was obtained (RMS roughness ∼0.15 nm), which is comparable to the roughness of conventional ALD Al2O3 roughness,63 while ALD on bare WSe2 showed large pinholes (RMS roughness ∼3.6 nm).

2. Plasma treatment

Plasma treatment is one of the most frequently used techniques in the modern fabrication process that readily primes any surface for better acceptance of the following process.70,71 Owing to highly energetic species that are made during plasma ignition, functional groups would be generated on the substrate surface by the reaction between reactive ion/electron species and surface group.72 In particular, this kind of surface treatment has been widely used for carbon nanotubes (CNTs)73 or graphene,74 which have lack of dangling bond on them, to promote a nucleation of ALD by generating hydroxyl species on top of the surface. However, it has critical obstacle that pristine substrate may get a damage after plasma ignition because of ion bombardment from ion/electron species that have high kinetic energy.75 Since 2D TMDs are considerably thin, compared to bulk materials, the effect of plasma damage on 2D TMDs is much critical than that on conventional bulk materials. Therefore, it is highly desirable to functionalize the surface of the ultrathin 2D TMD surface with less damage.

In 2013, Kim et al. reported the effect of oxygen-plasma pretreatment on the growth of ALD Al2O3 and HfO2 on 2D MoS2.76 To minimize damage from plasma, they used remote inductively coupled plasma (ICP) which plasma glow and surface treatment region was in a separate reactor. Using this remote plasma surface treatment, the TMD surface was not subjected to intense ion and electron bombardment during the treatment process.77 Firstly, the authors investigated the growth characteristics of ALD Al2O3 and HfO2 on pristine MoS2 at 250 °C as a function of thickness. Multilayered 2D MoS2 was prepared by mechanical exfoliation, and TMA, tetrakis(ethylmethylamino)hafnium (TEMAHf), and H2O were used for ALD Al2O3 and HfO2. Before the ALD process, oxygen plasma was treated on the MoS2 flake to improve the following precursors’ adsorption. When nominally 1-nm-thick Al2O3 was deposited on pristine MoS2, the poor surface coverage of Al2O3 was observed (∼23%), while when the thickness of Al2O3 was increased to 10 and 30 nm, the surface coverage was also increased to 77 and 97%, indicating the connection of individual Al2O3 islands. However, there were still uncovered grain boundaries even after 30 nm Al2O3 deposition. In the case of HfO2 deposition, the more uniform film was deposited on MoS2. However, numerous pinholes were still observed because of different reactivity of basal plain and grain boundary with TEMAHf precursor. These results are consistent with previous reports that are mentioned in Sec. IV A.

In order to enhance surface coverage of the ALD film on MoS2, O2 plasma was treated prior to film deposition by varying treatment time (10, 20, and 30 s). The surface coverage of ALD Al2O3 films was significantly increased only after 10 s of plasma pretreatment (93%), compared to that without plasma treatment (77%). Nearly completely covered Al2O3 films would be formed with 30 s of plasma pretreatment, showing that the surface of reaction sites would be increased as the plasma treatment time increased. A similar tendency of coverage improvement with plasma pretreatment was also confirmed for the ALD HfO2 case. They confirmed optimal plasma treatment time for uniform ALD on 2D MoS2 as 30 s. They also deposited both Al2O3 and HfO2 with various thickness after 30 s plasma treatment on MoS2, and SEM images of them are shown in Fig. 4(a). From these figures, nearly complete deposition of 10-nm-thick Al2O3 and HfO2 were available on MoS2 after 30 s of O2 plasma treatment. Figure 4(b) shows root mean square (RMS) surface roughness of ALD Al2O3 and HfO2 that were extracted from AFM images as a function of various plasma pretreatment times. The RMS surface roughness gradually decreased as the plasma pretreatment time increased on both Al2O3 and HfO2, indicating that more surface reaction sites were generated by increasing treatment time. Thus, dielectrics were more uniformly deposited as the plasma-treated time increased.

FIG. 4.

(a) SEM images of the ALD-grown (a)–(c) Al2O3 and (d)–(f) HfO2 films on the MoS2 flakes pretreated with oxygen plasma for 30 s as a function of thickness. (b) RMS roughness values of ∼10-nm-thick ALD-grown Al2O3 and HfO2 films on MoS2 that are extracted from the AFM measurement as a function of the oxygen-plasma treatment time (scan size: 1 × 1 μm2). (c) XPS spectra of Mo3d and S2s core levels measured from pristine and oxygen-plasma treated (30 s) MoS2 surfaces. (d) The cross-sectional TEM image of multilayer MoSe2 after doping through O2 plasma treatment and Al2O3 layer deposition. A high-magnification cross-sectional TEM image of the top layer is shown in the right (e) XPS Mo3d, Se3d, and O1s spectra of multilayer MoSe2 before and after the O2 plasma treatment. (a)–(c) Reprinted with permission from Yang et al., ACS Appl. Mater. Interfaces 5, 4739 (2013). Copyright 2013, American Chemical Society. (d) and (e) Reprinted with permission from Adv. Electron. Mater. 4, 1800308 (2018). Copyright 2018, Wiley-VCH Verlag GmbH & Co.

FIG. 4.

(a) SEM images of the ALD-grown (a)–(c) Al2O3 and (d)–(f) HfO2 films on the MoS2 flakes pretreated with oxygen plasma for 30 s as a function of thickness. (b) RMS roughness values of ∼10-nm-thick ALD-grown Al2O3 and HfO2 films on MoS2 that are extracted from the AFM measurement as a function of the oxygen-plasma treatment time (scan size: 1 × 1 μm2). (c) XPS spectra of Mo3d and S2s core levels measured from pristine and oxygen-plasma treated (30 s) MoS2 surfaces. (d) The cross-sectional TEM image of multilayer MoSe2 after doping through O2 plasma treatment and Al2O3 layer deposition. A high-magnification cross-sectional TEM image of the top layer is shown in the right (e) XPS Mo3d, Se3d, and O1s spectra of multilayer MoSe2 before and after the O2 plasma treatment. (a)–(c) Reprinted with permission from Yang et al., ACS Appl. Mater. Interfaces 5, 4739 (2013). Copyright 2013, American Chemical Society. (d) and (e) Reprinted with permission from Adv. Electron. Mater. 4, 1800308 (2018). Copyright 2018, Wiley-VCH Verlag GmbH & Co.

Close modal

After the optimization of the plasma treatment process, XPS and Raman spectroscopy were measured to investigate what makes the improved growth nature of ALD dielectric on MoS2. From the XPS spectra of MoS2 [Fig. 4(c)], Mo6+ peaks which correspond to MoO3 appeared after 30 s of the O2 plasma treatment, while intensities of Mo4+ and S2− was slightly decreased, indicating that some MoS2 layers were transformed to MoO3 after plasma treatment. MoO3 is known as hydrophilic material that contains numerous hydroxyl (–OH) species on the surface,78 thus the transformation to MoO3 could lead to an improvement of surface coverage of ALD films. Also, from Raman spectroscopy analysis, a decrease of peak intensities of two MoS2 Raman peaks (E12g and A1g) were observed, which indicates the oxidation of MoS2. However, there was no identifiable peak shift by the additional physical damage. From this paper, they concluded that the improvement of surface roughness after the plasma is mainly due to the oxidation of the surface, however, undesirable MoO3 formation followed as well.

In 2018, Hong et al. reported direct plasma treatment on the CVD-grown MoSe2 layer to obtain uniform and high quality of dielectric to passivate bulky MoSe2 from unintentional molecular adsorption which may induce deterioration of FET performance.79 MoSe2 is one of the 2D TMDs that has smaller bandgap (1.57 eV for monolayer, 1.09 for bulk) than MoS2 (1.88 eV for monolayer, 1.23 for bulk), therefore, better electrical properties could be anticipated in MoSe2 FET.80,81 Unlike previous reports that used remote plasma, they used direct plasma to oxidize the topmost MoSe2 layer intentionally. The oxygen-plasma treatment (100 W, 30 SCCM of O2, 10−2 Pa) was conducted on 50 nm of MoSe2 layer for 1 min prior to passivation. Then, 40 nm of Al2O3 was deposited on the MoSe2 using TMA and H2O at 200 °C. Cross-sectional and high-resolution transmission electron microscopy (TEM) images of multilayer MoSe2 is shown in Fig. 4(d). From these figures, around 3 nm of MoO3 was observed after direct plasma exposure, which corresponds to three to five uppermost MoSe2 layers were transformed. Notwithstanding the relatively strong plasma exposure, the rest of the MoSe2 layers seem to retain its crystallinity. Since the thickness of the MoSe2 layer was around 50 nm, the oxidation of the topmost MoSe2 layer might not be critical to the entire MoSe2, and the oxide layer (MoOx) may act as a nucleation promotion layer between MoSe2 and Al2O3. From XPS analysis, as shown in Fig. 4(e), clear peak shifts were observed from overall spectra, that indicates oxygen incorporation into MoSe2.82 Notable thing from these spectra is that Se–O, similar species to S–O that we mentioned in previous section, was also observed after direct plasma treatment. And these Se–O species would also assist chemical species to be adsorbed on top of the surface.

Recently, Huang et al reported H2O plasma treatment for the uniform dielectric deposition on the MoS2 substrate.83 They used H2O plasma to generate –OH species on top of TMD surface since H2O has hydrogen atom itself, unlike O3 or O2 plasma, hydroxyl radical (–OH*) can be generated by plasma ignition.84,85 As the generated hydroxyl groups adsorbed on the MoS2 substrate acting as reaction sites, uniform and high quality Al2O3 was obtained on the MoS2 flake. Using H2O plasma treatment, they successfully obtained uniform Al2O3 film of on exfoliated MoS2 down to 1.5 nm, which is a state-of-art result of thickness of uniform deposition on TMD material. Moreover, from the Raman and PL analysis, the peak intensities were not decreased after H2O plasma treatment, rather increased because the carbonaceous contaminants on the mechanically exfoliated MoS2 were removed during the plasma process (self-cleaning effect),86 suggesting a promising method for the TMD surface treatment.

3. UV-O3 treatment for surface hydroxylation

As described, although plasma treatment can promote nucleation of ALD on TMD surface by forming dangling bonds on top of the surface, it also damages ultrathin TMD itself because of highly reactive species that are generated during plasma ignition, resulting in a deterioration of its intrinsic and unique characteristics. Thus, forming a dangling bond on 2D TMDs without detriment should be carried out. For these reasons, researchers made efforts to develop a facile way that can generate reaction sites on the 2D TMD surface without damaging itself.

In 2014, Azcatl et al. investigated the effect of the room temperature ultraviolet-ozone (UV-O3) treatment on the growth of ALD Al2O3 on 2D MoS2.87 UV-O3 treatment is one of the traditional treatment methods which can remove surface contaminants by the reaction with highly reactive radicals that are generated from UV and O3 reaction.88 But nowadays, it becomes a more captivating technique to make a hydrophilic surface by forming a hydroxyl species on the surface.89 Likewise, using this surface treatment, it is anticipated that ALD on 2D TMD could obtain a smooth film retaining its original structure, and this was already confirmed on graphene.90,91 They proposed a nondestructive MoS2 functionalization method without breaking sulfur-molybdenum bonds by using UV-O3 exposure. To remove the top layers from bulk MoS2, the authors performed mechanical exfoliation by using adhesive tape. After the exfoliation, the authors loaded the sample into the ultrahigh vacuum (UHV) within 5 min and transferred to a chamber for the UV-O3 treatment.

in situ XPS was used to investigate chemical state changes of the MoS2 surface after UV-O3 exposure without affecting external contamination from ambient. Figures 5(a) and 5(b) show XPS spectra of S2p, Mo3d, O1s, and C1s before and after UV-O3 treatment. As shown in Fig. 5(a), a small peak was newly observed on Mo3d after the UV-O3 treatment, which corresponds to S–O, however, the oxygen-related peak (i.e., Mo-O) was not found. In S2p spectra, the additional doublet peak near 164.8 eV was observed after UV-O3 exposure. It is coincident with the appearance of an O1s peak [Fig. 5(b)], which corresponds to the S–O feature that was found in Mo3d. The newly generated S2p peaks were matched with S4+ which corresponds to the previous reports on MoS2 oxidation,92 instead of S2−. From these spectra changes, it could be deduced that S–O was formed on top of the MoS2 surface without Mo–S bond cleavage. They also calculated that the functionalized oxygen was formed with a monolayer coverage by calculating peak intensities of S2p (IS-O/Is), which confirms that functionalization using UV-O3 would make plenty of reaction sites for subsequent precursors on MoS2. Carbon residue on the surface was also eliminated as well after UV-O3 treatment [Fig. 5(b)] since UV-O3 is quite effective to remove carbonaceous contaminants on the surface, as mentioned earlier in this section. All the changes described above occurred when UV and O3 were exposed simultaneously since UV can not only greatly increase the reactivity of O3 (Ref. 88) but also highly reactive O radicals could be generated by O2 dissociation by UV absorption.92 From the DFT calculation, they showed that the formation of S–O bonds is energetically favorable without cleaving Mo–S bonds. Also, they claimed that O is hard to substitute S atom since S vacancy formation energy was quite large.

FIG. 5.

XPS spectra of (a) Mo3d and S2p and (b) O1s and C1s of MoS2 before and after the UV-O3 treatment. (c) AFM images of ALD Al2O3 before (top) and after (middle) the UV-O3 treatment, and cross-sectional TEM images of ALD Al2O3 on MoS2 after the UV-O3 treatment (bottom). XPS spectra of ALD HfO2 on (d) MoSe2 and (e) WSe2 as a function of ALD cycles and calculated peak ratio of oxygen-related species to (f) MoSe2 and (g) WSe2. (a)–(c) Reprinted with the permission from Azcatl et al., Appl. Phys. Lett. 104, 111601 (2014). Copyright 2014, AIP Publishing LLC. (d)–(g) Reprinted with permission from Angelica et al., 2D Mater. 2, 014004 (2015). Copyright 2015, A. Azcatl et al., licensed under a Creative Commons Attribution 3.0 License.

FIG. 5.

XPS spectra of (a) Mo3d and S2p and (b) O1s and C1s of MoS2 before and after the UV-O3 treatment. (c) AFM images of ALD Al2O3 before (top) and after (middle) the UV-O3 treatment, and cross-sectional TEM images of ALD Al2O3 on MoS2 after the UV-O3 treatment (bottom). XPS spectra of ALD HfO2 on (d) MoSe2 and (e) WSe2 as a function of ALD cycles and calculated peak ratio of oxygen-related species to (f) MoSe2 and (g) WSe2. (a)–(c) Reprinted with the permission from Azcatl et al., Appl. Phys. Lett. 104, 111601 (2014). Copyright 2014, AIP Publishing LLC. (d)–(g) Reprinted with permission from Angelica et al., 2D Mater. 2, 014004 (2015). Copyright 2015, A. Azcatl et al., licensed under a Creative Commons Attribution 3.0 License.

Close modal

To investigate growth characteristics and uniformity of film growth on treated MoS2, ALD Al2O3 on UV-O3 functionalized MoS2 was conducted, and surface morphology was measured by AFM and high-resolution TEM (HR-TEM). Figure 5(c) showed AFM images of Al2O3-deposited MoS2 at 200, 250, and 300 °C before and after the UV-O3 treatment and cross-sectional TEM images after the treatment. Without the UV-O3 treatment, nucleation occurred selectively on edges sites of MoS2, regardless of temperature. At 200 °C, the density of Al2O3 clusters on the MoS2 is higher than those of higher temperature conditions because the thermal desorption of chemisorbed oxygen species was suppressed at relatively low temperature. In contrast, the coverage is significantly increased on UV-O3 functionalized MoS2. Al2O3 films showed very small RMS roughness values of 0.14, 0.17, and 0.30 nm, with ALD process temperatures of 200, 250, and 300 °C, respectively. From the HR-TEM cross-section images of Al2O3 on UV-O3 functionalized MoS2, Al2O3 films with excellent uniformity were observed at 200° and 250 °C. However, Al2O3 deposited at 300 °C exhibited island-type growth, suggesting that S–O species that were generated after UV-O3 treatment would stand up to 250 °C, and started to be decomposed or desorbed at 300 °C. From the XPS analysis after ALD Al2O3 on functionalized MoS2, the S–O species was disappeared after ALD Al2O3 from the XPS analysis, which indicates surface S–O species acted as a sacrificial layer for ALD nucleation. Moreover, there was no Mo–Al or S–Al peak, indicating a noncovalent bond of Al2O3 on MoS2, similar to the previous result of HfO2 on MoS2.17 

The authors also showed the surface chemistry of WSe2 and MoSe2 with the UV-O3 treatment.93 The study showed that reactivities of TMDs diverse with their molecular species, and the different oxidation with the UV-O3 treatment. Unlike MoS2, both MoSe2 and WSe2 showed oxidation of transition metal (i.e., Mo and W) after 6 min of UV-O3 exposure, and this might due to the reactivity difference toward oxidation. From the previous report, WSe2 is the highest oxidation tendency, and MoS2 is the lowest among the three TMDs.94 This infers that MoSe2 or WSe2 are more sensitive to form artificial nucleation sites without damaging the original materials, compared with MoS2. However, it was found that the surface oxides generated on MoSe2 and WSe2 by UV-O3 exposure were reduced or eliminated by the ALD process. As shown in Figs. 7(d) and 7(f), only after one cycle of ALD HfO2 on UV-O3 treated MoSe2, the oxidized MoSe2 species (MoSexOy) started to be diminished and a new Mo5+ peak that might be induced by reduction of Mo6+ was observed. This phenomenon, referred to as “self-cleaning effect,”95 was originated from the ligand exchange between surface oxide species and Hf precursor since the formation of HfO2 is thermodynamically favorable than retaining MoOx.96,97 On the other hand, when HfO2 was deposited on WSe2, WOx species were not completed removed after 30 cycles of ALD HfO2 even though HfO2 is more favorable to be formed than retaining WOx. This might because of the following reasons: further reaction such as dissociation of W–O might be preceded before ligand exchange with Hf precursor,98 or high stability of WOx hindered the reduction of WOx.99 This implies that even UV-O3 treatment has been shown as promising method for surface functionalization, the effect might be dissimilar to the different 2D TMDs.

FIG. 6.

(a) Schematic of buffer layer insertion between MoS2 and ALD HfO2; (b) SEM images of ALD HfO2 on MoS2 without buffer layer, with MgO, Y2O3, and Al2O3 buffer layer; and (c) Raman spectra of pristine MoS2 before and after Y2O3 deposition, followed by ALD HfO2. (d) AFM images of 3-nm-thick ALD Al2O3 without (left) and with (right) buffer layer. (a)–(c) Reprinted with permission from Zou et al., Adv. Mater. 26, 6255 (2014). Copyright 2014, Wiley-VCH Verlag GmbH & Co. (d) Reprinted with permission from Son et al., Appl. Phys. Lett. 106, 021601 (2015). Copyright 2015, AIP Publishing LLC.

FIG. 6.

(a) Schematic of buffer layer insertion between MoS2 and ALD HfO2; (b) SEM images of ALD HfO2 on MoS2 without buffer layer, with MgO, Y2O3, and Al2O3 buffer layer; and (c) Raman spectra of pristine MoS2 before and after Y2O3 deposition, followed by ALD HfO2. (d) AFM images of 3-nm-thick ALD Al2O3 without (left) and with (right) buffer layer. (a)–(c) Reprinted with permission from Zou et al., Adv. Mater. 26, 6255 (2014). Copyright 2014, Wiley-VCH Verlag GmbH & Co. (d) Reprinted with permission from Son et al., Appl. Phys. Lett. 106, 021601 (2015). Copyright 2015, AIP Publishing LLC.

Close modal
FIG. 7.

(a) AFM images and height profiles of ALD Al2O3 on MoS2 using H2O (30 cycles) (left), O3 (30 cycles) at 200 °C (middle), and (c) O3 (five cycles of seeding layer at 30 °C, followed by 45 cycles of ALD at 200 °C) (right). (b) AFM images of ALD Al2O3 on MoS2 using water (left) and O2 plasma (right). (c) Cross-sectional TEM images of thermal ALD (left) and PE-ALD (right) HfO2, and (d) the calculated binding energy of TDMAH to (c) pristine MoS2, (d) sulfur vacancies, and (e) hydrogen-passivated sulfur vacancies. (a) Reprinted with permission from Cheng et al., ACS Appl. Mater. Interfaces 6, 11834 (2014). Copyright 2014, Cheng et al. licensed under a Creative Commons Attribution License. (b) Reprinted with permission from Price et al., ACS Appl. Mater. Interfaces 9, 23072 (2017). Copyright 2017, American Chemical Society. (c) and (d) Reprinted with permission from Price et al., ACS Appl. Nano Mater. 2, 4085 (2019). Copyright 2019, American Chemical Society.

FIG. 7.

(a) AFM images and height profiles of ALD Al2O3 on MoS2 using H2O (30 cycles) (left), O3 (30 cycles) at 200 °C (middle), and (c) O3 (five cycles of seeding layer at 30 °C, followed by 45 cycles of ALD at 200 °C) (right). (b) AFM images of ALD Al2O3 on MoS2 using water (left) and O2 plasma (right). (c) Cross-sectional TEM images of thermal ALD (left) and PE-ALD (right) HfO2, and (d) the calculated binding energy of TDMAH to (c) pristine MoS2, (d) sulfur vacancies, and (e) hydrogen-passivated sulfur vacancies. (a) Reprinted with permission from Cheng et al., ACS Appl. Mater. Interfaces 6, 11834 (2014). Copyright 2014, Cheng et al. licensed under a Creative Commons Attribution License. (b) Reprinted with permission from Price et al., ACS Appl. Mater. Interfaces 9, 23072 (2017). Copyright 2017, American Chemical Society. (c) and (d) Reprinted with permission from Price et al., ACS Appl. Nano Mater. 2, 4085 (2019). Copyright 2019, American Chemical Society.

Close modal

4. Insertion of a buffer layer as a nucleation promoter

Even though a uniform thin film could be deposited using the plasma or UV and O3 treatment, slight damage during the process is inevitable, resulting in a deterioration of device properties because of its high reactivity with TMDs. To avoid this, in 2014, Zou et al. reported a new way to deposit a uniform thin film on 2D TMDs by insertion of ultrathin buffer layer, which can promote nucleation and growth of ALD-grown film between TMD and dielectric [simple schematic is shown in Fig. 6(a)].100 For the ALD film growth promotion, three different metals (Mg, Al, Y) were thermally evaporated on the MoS2 substrate with a nominal thickness of 1 nm and then oxidized for several hours in the drying oven. Nucleation with a buffer layer, shown in Fig. 6(b), showed quite different growth nature depending on what the buffer layer material was. When HfO2 was deposited on the MgO buffer layer, numerous granularlike nucleation of HfO2 was observed with voids and pinholes and this might be due to the gap between large MgO nanoparticles, while smooth film growth on Al2O3 or Y2O3 buffer layers. In contrast to MoS2, buffer metal oxide has numerous reaction sites on top of the surface, reaction with subsequent chemical species is favorable, making a uniform film on top of the buffer layer. In terms of melting point, Y has the highest melting point among three different metals, which means less probability of coalescence during the ALD process. Even the melting point of a thin film is lower than that of the bulk film, the thin Y2O3 film could resist any nucleation or coalescence during the ALD process. In addition, since Y is thought to have a good wettability to 2D materials,101 uniform and compact formation of buffer and MoS2 could be possible. From Raman analysis [Fig. 6(c)], no notable change was observed after buffer layer and dielectric deposition, indicating a buffer layer deposition without damaging the original MoS2 structure. Utilizing the MoS2/Y2O3/HfO2 stack, they could obtain a uniform HfO2 thin film by further scaling down HfO2 to 9 nm, and the fabricated device exhibits excellent electrical properties, showing the excellent interface quality and scalability.

Son et al. also reported a uniform thin film deposition on MoS2 flake using buffer layer insertion prior to ALD deposition.62 For the buffer layer, which promoted nucleation of Al2O3, 1-nm-thick Al was evaporated on the MoS2 flake and naturally oxidized. Since the surface energy of Al is much higher (1.27 × 103mJ/m2)102 than that of MoS2 (46.5 mJ/m2),103 granular Al was formed on top of the MoS2, not a film. However, through the oxidation process, the thickness of Al was increased around 4 nm due to the metal-to-oxide transformation,104 resulting in a strain on MoS2 that could modulate the electrical properties.105 The AFM images of 3-nm-thick Al2O3 deposition on MoS2 with and without the buffer layer are shown in Fig. 6(d). Without this buffer layer, surface roughness and coverage of ALD Al2O3 on MoS2 without buffer layer were 1.969 nm and 61.89%, respectively, while those with buffer layer were 0.649 nm and 99.29%, respectively. These results confirm that the insertion of an ultrathin buffer layer between 2D TMD and dielectric not only could promote the film growth on TMD but leave the pristine nature of TMD intact. However, by inserting an additional layer, the increase of total thickness which might be detrimental for nanoscale device fabrication is also inevitable.

5. Ozone ALD (O3-ALD) or direct plasma-enhanced ALD (PE-ALD) on 2D TMDs

As discussed in Sec. IV B 3, UV-O3 treatment showed optimistic results for the generation of reaction sites on MoS2 without damaging the host material. Nevertheless, some papers have reported that long exposure of UV on TMDs may induce photodegradation, which could deteriorate the electrical properties of TMDs.106,107 Thus, researchers tried to find another way to make full of reaction sites on 2D TMDs without using UV light.

In 2014, Cheng et al. reported an improvement of ALD nucleation on MoS2 using O3 as a reactant. O3 has been known as a strong oxidant that is frequently used for the ALD process.108–111 Moreover, because of the high reactivity, various advantages including low impurities and high density would be anticipated when it was used as the reactant of ALD. However, since it was revealed that O3 itself could not form surface S–O functional groups, reaction sites for subsequent precursors, on MoS2,17 they used a two-step deposition process for ALD Al2O3. At 200 °C, ALD Al2O3 using TMA and H2O with 3 nm of nominal thickness showed poor uniformity on the MoS2 flake, similar to the previous reports that mentioned before. However, the uniformity of ALD Al2O3 significantly improved when O3 was adopted for reactant instead of H2O, but still a lot of pinholes existed in the film. To improve the surface morphology, five cycles of ALD Al2O3 was conducted at 30 °C to form a seeding layer prior to ALD Al2O3 at 200 °C, and it showed fairly smooth RMS roughness (0.23 nm) compared to that of pristine MoS2 (0.18 nm). Since reactive O* that were generated from O3 dissociation could provide the following precursors to reaction sites by forming weak S–O bond, artificially generated surface reaction species could be also desorbed or diffused on the surface as the temperature increases, leading a high roughness. Therefore, by employing a low temperature process in advance, undesirable effects such as thermal desorption of reaction sites would be avoided. From Raman spectra of MoS2 after O3 exposure as a function of process temperature, there was nearly no changes before and after O3 treatment since MoS2 is thermally stable against the oxygen-rich atmosphere up to 370 °C112 due to its high dissociative binding energy.98 Therefore, using O3 as a reactant could supply surface reaction sites on the 2D TMD surface without damaging or braking its original structures, which has great potential for future TMD-based electronics.

As mentioned in Sec. IV B 4, plasma treatment using O2 or H2O can make surface hydroxyl species, which enables reaction with the following chemical species for ALD. However, it may also damage the pristine material, resulting in a device performance deterioration.87 In particular, since the total thickness of 2D flakes that are used for FET fabrication is quite thin (<10 nm), partially oxidized or damaged 2D flake result in quite different characteristics with those of undamaged materials.

Recently, some researchers have reported a new way of uniform dielectric deposition on 2D TMDs using PE-ALD. Unlike thermal ALD, PE-ALD uses highly reactive chemical species that are generated during plasma reaction (e.g., O*, H*, etc.).113–115 Because of high reactivity of the radicals, the use of PE-ALD not only enables lower process temperature but also can make a high density film, compared with thermal ALD.116,117 In this research, they tried to minimize the damage of 2D TMDs from highly reactive plasma species, such as the use of remote plasma or relatively mild reactant source. In 2017, Price et al. reported uniform Al2O3 and HfO2 deposition on exfoliated multilayer MoS2 (6–8 nm) using remote PE-ALD.118 They used TMA and TEMAHf as precursors, and H2O and O2 plasma as reactants for ALD Al2O3 and HfO2, respectively. Both Al2O3 and HfO2 films grown by PE-ALD showed a smooth surface on several layers of MoS2, regardless of growth temperature, while thermal ALD showed numerous nucleation on top of the MoS2 flake (Fig. 7 shows AFM images of thermal and PE-ALD of Al2O3 on MoS2). Using PE-ALD, they were able to obtain uniform thin film deposition down to 3.5 nm with subnanometer scale roughness. Although the uniform growth using thermal ALD was determined by the coverage of physisorbed molecules, since PE-ALD induce dangling bonds on the surface directly, physisorption is not the factor that affects the initial growth characteristics anymore. They could acquire ultrathin dielectric thin film (∼3.5 nm) on MoS2 with subnanometer scale roughness. As they fabricated and characterized the MoS2 FET with these dielectrics, the change of electrical properties was quite different. When PE-ALD Al2O3 was deposited on MoS2 FET, PE-ALD HfO2 process did not degrade the electrical properties of MoS2 FET, rather improved the performance, contrary to ALD HfO2 or PE-ALD Al2O3. From XPS analysis, they found that there was nearly no oxidation on MoS2 flake after the PE-ALD process since MoO3-related peak was not observed on Mo3d and S2s spectra, respectively. However, it is hard to say that there was no oxidation since the oxidation of the atop MoS2 layer could not be detected because of the XPS detection limit. Instead, there was a slight negative peak shift of Mo3d and S2s after the PE-ALD process, indicating p-type doping.119,120 PE-ALD Al2O3 showed a larger shift (∼1.2 eV), whereas, the PE-ALD HfO2 showed a much smaller shift (∼0.4 eV), which indicates that PE-ALD HfO2 affected less on MoS2 than PE-ALD Al2O3.

Even this PE-ALD does not deteriorate the electrical characteristics of MoS2-based FET, rather enhanced the performance, it could be possible since the number of MoS2 was around 6–8 nm, which correspond 9–12 layers of MoS2. In the case of thick flake, the effect of oxidation or damage could be negligible since only the uppermost TMD layer was affected by plasma species. Thus, they also investigated the effect of PE-ALD on relatively thin, mono-, bi-, and trilayer of MoS2.121 They used CVD-grown MoS2, which has numerous inherent defects on the surface compared to mechanically exfoliated. Similar to the previous reports with exfoliated MoS2, electrical property degradation of monolayer MoS2 FET, such as the Vth shift, Ion/Ioff ratio decrease, was observed. However, in the case of FET using bi- and trilayer MoS2, the deterioration of electrical properties was quite decreased, and this was clarified by TEM analysis that the uppermost layer of MoS2 was damaged after the PE-ALD process. As shown in Fig. 7(c), the TEM image of trilayer MoS2 after thermal ALD shows clear and distinct three layers between SiO2 and HfO2, which indicates no damage after the thermal ALD process. In contrast to thermal ALD, however, bleary three layers of MoS2 were observed which possibly indicates plasma damage. The high-resolution TEM image [right inset image of Fig. 7(c)] clearly showed two MoS2 layers on the bottom side and one blurry layer on the uppermost MoS2 flake, verifying that the uppermost layer of MoS2 was damaged during the PE-ALD process, and this result corresponds to the previous results of MoS2 FET. From their previous reports, PE-ALD HfO2 showed less deterioration of FET characteristics when it applied to MoS2 FET than PE-ALD Al2O3. Thus, they also calculated adsorption energy of Hf precursor of MoS2 depending on different surface configuration, pristine, sulfur vacancies, and H-passivated vacancies in MoS2 using the first-principle calculation. From there calculation, the adsorption energy of the precursor was −1.06 eV for pristine, −1.33 for sulfur vacancies, and −2.52 eV for H-passivated vacancies in MoS2. This result indicates that H-terminated vacancies in MoS2 that were generated in the PE-ALD process are most stable among them, indicating PE-ALD could promote nucleation and growth film using ALD by forming these species on the topmost 2D TMD layer.

In this review, we discussed the extraordinary surface nature of 2D TMDs and the way to obtain uniform films on the TMD surface under 10 nm. Because of its fascinating and versatile characteristics of 2D TMDs, these can be applied not only to future electronics that need characteristics beyond the limits of conventional materials but also to diverse fields that cannot be acquired because of physical limits of conventional materials. For these applications, not only the development of 2D TMD synthesis for large-scale fabrication but also thin film deposition with excellent uniformity are both essential. Given the above, an ultrathin film is able to be obtained on 2D TMDs with negligible effect from the process, and the adequate process and chemical species can be selected depending on the host materials. These results provide a perspective of the ALD growth mechanism on 2D materials and solution for the deposition of ultrathin layers with high quality. Since the various methods have their pros and cons, we cannot suggest which method is the most promising for ALD on 2D TMDs. However, we predict that a new novel method, which can deposit a high quality and ultrathin film without the degradation of 2D TMDs properties, would be developed by comprehensive complementation of previous methods, which enables a breakthrough for the demonstration of various electrical or optical devices based on 2D materials.

This work was supported by the Materials and Components Technology Development Program of MOTIE/KEIT (No. 10080642, Development on precursors for carbon/halogen-free thin film and their delivery system for high-k/metal gate application and No. 10080527, Development of commercialization technology of high sensitive gas sensor based on chalcogenide 2D nanomaterial).

1.
W.
Choi
,
N.
Choudhary
,
G. H.
Han
,
J.
Park
,
D.
Akinwande
, and
Y. H.
Lee
,
Mater. Today
20
,
116
(
2017
).
2.
K. F.
Mak
and
J.
Shan
,
Nat. Photonics
10
,
216
(
2016
).
3.
J. R.
Schaibley
,
H.
Yu
,
G.
Clark
,
P.
Rivera
,
J. S.
Ross
,
K. L.
Seyler
,
W.
Yao
, and
X.
Xu
,
Nat. Rev. Mater.
1
,
16055
(
2016
).
4.
D.
Akinwande
 et al,
Extrem. Mech. Lett.
13
,
42
(
2017
).
5.
A.
Geim
and
K.
Novoselov
,
Nat. Mater.
6
,
183
(
2007
).
6.
M. J.
Allen
,
V. C.
Tung
, and
R. B.
Kaner
,
Chem. Rev.
110
,
132
(
2010
).
7.
Y.
Zhang
,
T. T.
Tang
,
C.
Girit
,
Z.
Hao
,
M. C.
Martin
,
A.
Zettl
,
M. F.
Crommie
,
Y. R.
Shen
, and
F.
Wang
,
Nature
459
,
820
(
2009
).
8.
R.
Balog
 et al,
Nat. Mater.
9
,
315
(
2010
).
9.
A.
Nourbakhsh
 et al,
Nanotechnology
21
,
43
(
2010
).
10.
M.
Chhowalla
,
H. S.
Shin
,
G.
Eda
,
L.-J.
Li
,
K. P.
Loh
, and
H.
Zhang
,
Nat. Chem.
5
,
263
(
2013
).
11.
B.
Radisavljevic
,
A.
Radenovic
,
J.
Brivio
,
V.
Giacometti
, and
A.
Kis
,
Nat. Nanotechnol.
6
,
147
(
2011
).
12.
K. F.
Mak
,
C.
Lee
,
J.
Hone
,
J.
Shan
, and
T. F.
Heinz
,
Phys. Rev. Lett.
105
,
136805
(
2010
).
13.
A.
Splendiani
,
L.
Sun
,
Y.
Zhang
,
T.
Li
,
J.
Kim
,
C. Y.
Chim
,
G.
Galli
, and
F.
Wang
,
Nano Lett.
10
,
1271
(
2010
).
14.
A.
Castellanos-Gomez
,
M.
Poot
,
G. A.
Steele
,
H. S. J.
Van Der Zant
,
N.
Agraït
, and
G.
Rubio-Bollinger
,
Adv. Mater.
24
,
772
(
2012
).
15.
A.
Castellanos-Gomez
,
R.
Van Leeuwen
,
M.
Buscema
,
H. S. J.
Van Der Zant
,
G. A.
Steele
, and
W. J.
Venstra
,
Adv. Mater.
25
,
6719
(
2013
).
16.
S.
Bertolazzi
,
J.
Brivio
, and
A.
Kis
,
ACS Nano
5
,
9703
(
2011
).
17.
S.
McDonnell
 et al,
ACS Nano
7
,
10354
(
2013
).
18.
H.
Liu
,
M.
Si
,
S.
Najmaei
,
A. T.
Neal
,
Y.
Du
,
P. M.
Ajayan
,
J.
Lou
, and
P. D.
Ye
,
Nano Lett.
13
,
2640
(
2013
).
19.
X.
Huang
,
C.
Tan
,
Z.
Yin
, and
H.
Zhang
,
Adv. Mater.
26
,
2185
(
2014
).
20.
F. H.
Koppens
,
T.
Mueller
,
P.
Avouris
,
a C.
Ferrari
,
M. S.
Vitiello
, and
M.
Polini
,
Nat. Nanotechnol.
9
,
780
(
2014
).
21.
S. M.
George
,
Chem. Rev.
110
,
111
(
2010
).
22.
H.
Kim
,
H.-B.-R. B. R.
Lee
, and
W.-J. J.
Maeng
,
Thin Solid Films
517
,
2563
(
2009
).
23.
C.
Zhu
,
X.
Xia
,
J.
Liu
,
Z.
Fan
,
D.
Chao
,
H.
Zhang
, and
H. J.
Fan
,
Nano Energy
4
,
105
(
2014
).
24.
M. Z.
Ansari
,
N.
Parveen
,
D. K.
Nandi
,
R.
Ramesh
,
S. A.
Ansari
,
T.
Cheon
, and
S. H.
Kim
,
Sci. Rep.
9
,
10225
(
2019
).
25.
V.
Cremers
,
R. L.
Puurunen
, and
J.
Dendooven
,
Appl. Phys. Rev.
6
,
021302
(
2019
).
26.
D.
Jena
,
Proc. IEEE
101
,
1585
(
2013
).
27.
H. J.
Chuang
 et al,
Nano Lett.
14
,
3594
(
2014
).
28.
H.
Liu
,
K.
Xu
,
X.
Zhang
, and
P. D.
Ye
,
Appl. Phys. Lett.
100
,
152115
(
2012
).
29.
G. Z.
Magda
,
J.
Petõ
,
G.
Dobrik
,
C.
Hwang
,
L. P.
Biró
, and
L.
Tapasztó
,
Sci. Rep.
5
,
3
(
2015
).
30.
V.
Nicolosi
,
M.
Chhowalla
,
M. G.
Kanatzidis
,
M. S.
Strano
, and
J. N.
Coleman
,
Science (80-.)
340
,
1226419
(
2013
).
31.
J. N.
Coleman
 et al,
Science (80-.)
331
,
568
(
2011
).
32.
K. I.
Bolotin
,
Graphene: Properties, Preparation, Characterisation and Devices
(
Woodhead Publishing
,
Sawston, Cambridge
,
2014
), p.
199
.
33.
R.
Addou
,
L.
Colombo
, and
R. M.
Wallace
,
ACS Appl. Mater. Interfaces
7
,
11921
(
2015
).
34.
T.
Verhagen
,
V. L. P.
Guerra
,
G.
Haider
,
M.
Kalbac
, and
J.
Vejpravova
,
Nanoscale
12
,
3019
(
2020
).
35.
J.
Hong
 et al,
Nat. Commun.
6
,
6293
(
2015
).
36.
H.
Li
 et al,
Nat. Mater.
15
,
48
(
2016
).
37.
J.
Feng
 et al,
Nature
536
,
197
(
2016
).
38.
S.
Xiao
,
P.
Xiao
,
X.
Zhang
,
D.
Yan
,
X.
Gu
,
F.
Qin
,
Z.
Ni
,
Z. J.
Han
, and
K. K.
Ostrikov
,
Sci. Rep.
6
,
19945
(
2016
).
39.
J.
Deng
 et al,
Energy Environ. Sci.
8
,
1594
(
2015
).
40.
X.
Wang
,
S. M.
Tabakman
, and
H.
Dai
,
J. Am. Chem. Soc.
130
,
8152
(
2008
).
41.
L.
Cheng
,
X.
Qin
,
A. T.
Lucero
,
A.
Azcatl
,
J.
Huang
,
R. M.
Wallace
,
K.
Cho
, and
J.
Kim
,
ACS Appl. Mater. Interfaces
6
,
11834
(
2014
).
42.
W.
Zhou
 et al,
Nano Lett.
13
,
2615
(
2013
).
43.
K.
Kim
 et al,
Nat. Commun.
5
,
4781
(
2014
).
44.
C.
González
,
B.
Biel
, and
Y. J.
Dappe
,
Phys. Chem. Chem. Phys.
19
,
9485
(
2017
).
45.
Y.
Hajati
,
T.
Blom
,
S. H. M.
Jafri
,
S.
Haldar
,
S.
Bhandary
,
M. Z.
Shoushtari
,
O.
Eriksson
,
B.
Sanyal
, and
K.
Leifer
,
Nanotechnology
23
,
505501
(
2012
).
46.
R. L.
Puurunen
,
J. Appl. Phys.
97
,
121301
(
2005
).
47.
D. B.
Farmer
and
R. G.
Gordon
,
Nano Lett.
6
,
699
(
2006
).
48.
R.
Rammula
,
L.
Aarik
,
A.
Kasikov
,
J.
Kozlova
,
T.
Kahro
,
L.
Matisen
,
A.
Niilisk
,
H.
Alles
, and
J.
Aarik
,
IOP Conf. Ser. Mater. Sci. Eng.
49
,
012014
(
2013
).
49.
S.
Jandhyala
 et al,
ACS Nano
6
,
2722
(
2012
).
50.
Y.
Xuan
,
Y. Q.
Wu
,
T.
Shen
,
M.
Qi
,
M. A.
Capano
,
J. A.
Cooper
, and
P. D.
Ye
,
Appl. Phys. Lett.
92
,
013101
(
2008
).
51.
T.
Nam
,
Y. J.
Park
,
H.
Lee
,
I. K.
Oh
,
J. H.
Ahn
,
S. M.
Cho
,
H.
Kim
, and
H. B. R.
Lee
,
Carbon
116
,
553
(
2017
).
52.
P.
Lazic
,
Physics of Surface, Interface and Cluster Catalysis
(
IOP
,
Bristol, England
,
2016
), pp.
2.1
2.25
.
53.
E.
Lippert
,
Angew. Chemie
72
,
602
(
1960
).
54.
S.-J.
Jeong
 et al,
2D Mater.
3
,
035027
(
2016
).
55.
T.
Muneshwar
and
K.
Cadien
,
J. Appl. Phys.
124
,
095302
(
2018
).
56.
A.
Konar
,
T.
Fang
, and
D.
Jena
,
Phys. Rev. B
82
,
115452
(
2010
).
57.
A. K. M.
Newaz
,
Y. S.
Puzyrev
,
B.
Wang
,
S. T.
Pantelides
, and
K. I.
Bolotin
,
Nat. Commun.
3
,
734
(
2012
).
58.
F.
Chen
,
J.
Xia
,
D. K.
Ferry
, and
N.
Tao
,
Nano Lett.
9
,
2571
(
2009
).
59.
D.
Jena
and
A.
Konar
,
Phys. Rev. Lett.
98
,
136805
(
2007
).
60.
J.
Robertson
,
Eur. Phys. J. Appl. Phys.
28
,
265
(
2004
).
61.
N. P.
Kobayashi
,
C. L.
Donley
,
S.-Y.
Wang
, and
R. S.
Williams
,
J. Cryst. Growth
299
,
218
(
2007
).
62.
S.
Son
,
S.
Yu
,
M.
Choi
,
D.
Kim
, and
C.
Choi
,
Appl. Phys. Lett.
106
,
021601
(
2015
).
63.
M. D.
Groner
,
F. H.
Fabreguette
,
J. W.
Elam
, and
S. M.
George
,
Chem. Mater.
16
,
639
(
2004
).
64.
J.
Swerts
,
N.
Peys
,
L.
Nyns
,
A.
Delabie
,
A.
Franquet
,
J. W.
Maes
,
S.
Van Elshocht
, and
S.
De Gendt
,
J. Electrochem. Soc.
157
,
26
(
2010
).
65.
H.
Fang
,
S.
Chuang
,
T. C.
Chang
,
K.
Takei
,
T.
Takahashi
, and
A.
Javey
,
Nano Lett.
12
,
3788
(
2012
).
66.
H.
Qiu
,
L.
Pan
,
Z.
Yao
,
J.
Li
,
Y.
Shi
, and
X.
Wang
,
Appl. Phys. Lett.
100
,
123104
(
2012
).
67.
Y.-C.
Lin
,
C.-C.
Lu
,
C.-H.
Yeh
,
C.
Jin
,
K.
Suenaga
, and
P.-W.
Chiu
,
Nano Lett.
12
,
414
(
2012
).
68.
J. H.
Park
 et al,
ACS Nano
10
,
6888
(
2016
).
69.
J. H.
Park
 et al,
Nano Lett.
15
,
6626
(
2015
).
70.
J.
Shul
and
S. L. P.
Pearton
,
Plasma Phys. Control. Fusion
43
,
372
(
2001
).
71.
W.
Steckelmacher
,
Vacuum
42
,
779
(
2002
).
72.
N.
Kumar
,
A.
Yanguas-Gil
,
S. R.
Daly
,
G. S.
Girolami
, and
J. R.
Abelson
,
Appl. Phys. Lett.
95
,
144107
(
2009
).
73.
Y. C.
Hsueh
,
C. C.
Wang
,
C.
Liu
,
C. C.
Kei
, and
T. P.
Perng
,
Nanotechnology
23
,
405603
(
2012
).
74.
R. H. J.
Vervuurt
,
B.
Karasulu
,
M. A.
Verheijen
,
W. M. M.
Kessels
, and
A. A.
Bol
,
Chem. Mater.
29
,
2090
(
2017
).
75.
H.
Nan
,
R.
Zhou
,
X.
Gu
,
S.
Xiao
, and
K. K.
Ostrikov
,
Nanoscale
11
,
19202
(
2019
).
76.
J.
Yang
,
S.
Kim
,
W.
Choi
,
S. H.
Park
,
Y.
Jung
,
M. H.
Cho
, and
H.
Kim
,
ACS Appl. Mater. Interfaces
5
,
4739
(
2013
).
77.
G.
Lucovsky
,
J. Vac. Sci. Technol. B Microelectron. Nanom. Struct.
7
,
861
(
1989
).
78.
W.
Gulbinski
,
D.
Pailharey
,
T.
Suszko
, and
Y.
Mathey
,
Surf. Sci.
475
,
149
(
2001
).
79.
S.
Hong
,
H.
Im
,
Y. K.
Hong
,
N.
Liu
,
S.
Kim
, and
J. H.
Park
,
Adv. Electron. Mater.
4
,
1800308
(
2018
).
80.
S.
Kim
 et al,
Adv. Mater.
30
,
1705542
(
2018
).
81.
J.
Gusakova
,
X.
Wang
,
L. L.
Shiau
,
A.
Krivosheeva
,
V.
Shaposhnikov
,
V.
Borisenko
,
V.
Gusakov
, and
B. K.
Tay
,
Phys. Status Solidi
214
,
1700218
(
2017
).
82.
A. T.
Martín-Luengo
,
H.
Köstenbauer
,
J.
Winkler
, and
A.
Bonanni
,
AIP Adv.
7
,
015034
(
2017
).
83.
B.
Huang
,
M.
Zheng
,
Y.
Zhao
,
J.
Wu
, and
J. T. L.
Thong
,
ACS Appl. Mater. Interfaces
11
,
35438
(
2019
).
84.
Y.
Gorbanev
,
D.
O’Connell
, and
V.
Chechik
,
Chem. Eur. J.
22
,
3496
(
2016
).
85.
S. V. T.
Nguyen
,
J. E.
Foster
, and
A. D.
Gallimore
,
Rev. Sci. Instrum.
80
,
083503
(
2009
).
86.
N. A.
Alcantar
,
E. S.
Aydil
, and
J.
Israelachvili
,
Fundamental and Applied Aspects of Chemically Modified Surfaces
(
Woodhead Publishing
,
Sawston, Cambridge
,
1999
), pp.
212
222
.
87.
A.
Azcatl
 et al,
Appl. Phys. Lett.
104
,
111601
(
2014
).
88.
J. R.
Vig
,
J. Vac. Sci. Technol. A
3
,
1027
(
1985
).
89.
T.
Clark
,
J. D.
Ruiz
,
H.
Fan
,
C. J.
Brinker
,
B. I.
Swanson
, and
A. N.
Parikh
,
Chem. Mater.
12
,
3879
(
2000
).
90.
S.
Huh
,
J.
Park
,
Y. S.
Kim
,
K. S.
Kim
,
B. H.
Hong
, and
J.-M.
Nam
,
ACS Nano
5
,
9799
(
2011
).
91.
H.
Sun
 et al,
J. Mater. Chem. C
5
,
1880
(
2017
).
92.
N. M. D.
Brown
,
N.
Cui
, and
A.
McKinley
,
Appl. Surf. Sci.
134
,
11
(
1998
).
93.
A.
Angelica
 et al,
2D Mater.
2
,
14004
(
2015
).
94.
W.
Jaegermann
and
D.
Schmeisser
,
Surf. Sci.
165
,
143
(
1986
).
95.
C. L.
Hinkle
,
E. M.
Vogel
,
P. D.
Ye
, and
R. M.
Wallace
,
Curr. Opin. Solid State Mater. Sci.
15
,
188
(
2011
).
96.
A. J.
Bard
,
R.
Parsons
, and
J.
Jordan
,
Standard Potentials in Aqueous Solution
(
Routledge
,
New York
,
1985
).
97.
G. L.
Humphrey
,
J. Am. Chem. Soc.
75
,
2806
(
1953
).
98.
Y.-R.
Luo
,
Comprehensive Handbook of Chemical Bond Energies
(
CRC Press
,
Boca Raton, FL
,
2007
).
99.
Z.
Li
,
Z.
Fang
,
M. S.
Kelley
,
B. D.
Kay
,
R.
Rousseau
,
Z.
Dohnalek
, and
D. A.
Dixon
,
J. Phys. Chem. C
118
,
4869
(
2014
).
100.
X.
Zou
 et al,
Adv. Mater.
26
,
6255
(
2014
).
101.
Z.
Wang
 et al,
Nano Lett.
10
,
2024
(
2010
).
102.
H. L.
Skriver
and
N. M.
Rosengaard
,
Phys. Rev. B
46
,
7157
(
1992
).
103.
A. P. S.
Gaur
,
S.
Sahoo
,
M.
Ahmadi
,
S. P.
Dash
,
M. J. F.
Guinel
, and
R. S.
Katiyar
,
Nano Lett.
14
,
4314
(
2014
).
104.
J. A.
Robinson
 et al,
ACS Nano
4
,
2667
(
2010
).
105.
M. A.
Bissett
,
M.
Tsuji
, and
H.
Ago
,
Phys. Chem. Chem. Phys.
16
,
11124
(
2014
).
106.
W.
Choi
 et al,
Adv. Mater.
24
,
5832
(
2012
).
107.
S.
Ahn
,
G.
Kim
,
P. K.
Nayak
,
S. I.
Yoon
,
H.
Lim
,
H. J.
Shin
, and
H. S.
Shin
,
ACS Nano
10
,
8973
(
2016
).
108.
D. N.
Goldstein
,
J. A.
McCormick
, and
S. M.
George
,
J. Phys. Chem. C
112
,
19530
(
2008
).
109.
J.
Kwon
,
M.
Dai
,
M. D.
Halls
, and
Y. J.
Chabal
,
Chem. Mater.
20
,
3248
(
2008
).
110.
M.
Rose
,
J.
Niinistö
,
I.
Endler
,
J. W.
Bartha
,
P.
Kücher
, and
M.
Ritala
,
ACS Appl. Mater. Interfaces
2
,
347
(
2010
).
111.
S. D.
Elliott
,
G.
Scarel
,
C.
Wiemer
,
M.
Fanciulli
, and
G.
Pavia
,
Chem. Mater.
18
,
3764
(
2006
).
112.
B. C.
Windom
,
W. G.
Sawyer
, and
D. W.
Hahn
,
Tribol. Lett.
42
,
301
(
2011
).
113.
H.
Kim
,
Thin Solid Films
519
,
6639
(
2011
).
114.
S. M.
Rossnagel
,
A.
Sherman
, and
F.
Turner
,
J. Vac. Sci. Technol. B
18
,
2016
(
2000
).
115.
H.
Kim
and
S. M.
Rossnagel
,
J. Vac. Sci. Technol. A
20
,
802
(
2002
).
116.
S. E.
Potts
,
W.
Keuning
,
E.
Langereis
,
G.
Dingemans
,
M. C. M.
van de Sanden
, and
W. M. M.
Kessels
,
J. Electrochem. Soc.
157
,
P66
(
2010
).
117.
H.
Kim
,
I.
Chung
,
S.
Kim
,
S.
Shin
,
W.
Jung
,
R.
Hwang
,
C.
Jeong
, and
H.
Hwang
,
J. Vac. Sci. Technol. A
33
,
01A146
(
2015
).
118.
K. M.
Price
,
K. E.
Schauble
,
F. A.
McGuire
,
D. B.
Farmer
, and
A. D.
Franklin
,
ACS Appl. Mater. Interfaces
9
,
23072
(
2017
).
119.
A.
Azcatl
 et al,
Nano Lett.
16
,
5437
(
2016
).
120.
S.
McDonnell
,
R.
Addou
,
C.
Buie
,
R. M.
Wallace
, and
C. L.
Hinkle
,
ACS Nano
8
,
2880
(
2014
).
121.
K. M.
Price
,
S.
Najmaei
,
C. E.
Ekuma
,
R. A.
Burke
,
M.
Dubey
, and
A. D.
Franklin
,
ACS Appl. Nano Mater.
2
,
4085
(
2019
).