Brought to you by:
Topical Review

Light–matter interaction in transition metal dichalcogenides and their heterostructures

, , and

Published 23 March 2017 © 2017 IOP Publishing Ltd
, , Citation Ursula Wurstbauer et al 2017 J. Phys. D: Appl. Phys. 50 173001 DOI 10.1088/1361-6463/aa5f81

0022-3727/50/17/173001

Abstract

The investigation of two-dimensional (2D) van der Waals materials is a vibrant, fast-moving and still growing interdisciplinary area of research. These materials are truly 2D crystals with strong covalent in-plane bonds and weak van der Waals interaction between the layers, and have a variety of different electronic, optical and mechanical properties. Transition metal dichalcogenides are a very prominent class of 2D materials, particularly the semiconducting subclass. Their properties include bandgaps in the near-infrared to the visible range, decent charge carrier mobility together with high (photo-) catalytic and mechanical stability, and exotic many-body phenomena. These characteristics make the materials highly attractive for both fundamental research as well as innovative device applications. Furthermore, the materials exhibit a strong light–matter interaction, providing a high sunlight absorbance of up to 15% in the monolayer limit, strong scattering cross section in Raman experiments, and access to excitonic phenomena in van der Waals heterostructures. This review focuses on the light–matter interaction in MoS2, WS2, MoSe2 and WSe2, which is dictated by the materials' complex dielectric functions, and on the multiplicity of studying the first-order phonon modes by Raman spectroscopy to gain access to several material properties such as doping, strain, defects and temperature. 2D materials provide an interesting platform for stacking them into van der Waals heterostructures without the limitation of lattice mismatch, resulting in novel devices for applications but also to enable the study of exotic many-body interaction phenomena such as interlayer excitons. Future perspectives of semiconducting transition metal dichalcogenides and their heterostructures for applications in optoelectronic devices will be examined, and routes to study emergent fundamental problems and many-body quantum phenomena under excitations with photons will be discussed.

Export citation and abstract BibTeX RIS

1. General introduction to two-dimensional materials and, in particular, semiconducting transition metal dichalcogenides (SC-TMDs)

In two-dimensional (2D) charge carrier systems electrons can freely move within the 2D plane; however, their motion is restricted in the third dimension by quantum mechanics. Besides the constant density of states in (quasi-) 2D systems, high charge carrier mobility and reduced screening are both of technological and fundamental physical relevance. The reduced phase space also causes enhanced quantum effects and is therefore of importance for many-body phenomena such as Coulomb-driven formation of excitons—bosonic quasiparticles of a bound electron–hole pair. A prominent example of such a systems is a highly mobile 2D electron system confined in a silicon inversion layer or a GaAs quantum well, where increased electron correlations account for the observation of the fractional quantum Hall effect, electron solids and other correlated phases at low temperatures and high magnetic fields [1]. Dimensionality-related phenomena are even more striking in real materials that have 2D character not only in momentum space, but also in real space. The 3D parental versions of 2D materials are van der Waals (a.k.a. layered) materials characterized by strong covalent or ionic in-plane bonds and weak van der Waals coupling between individual layers. The interaction between the planes is low and has been calculated for graphene to scale with distance d as 1/dα with α  ⩾  4 [2]. As a result, single or few layers can be peeled off from bulk crystals with an adhesive tape as demonstrated for the first time for graphene in 2004 [3] and for other 2D materials [4].

Graphene's peculiar linear band structure, resulting in relativistic Dirac fermions with record room temperature mobility in combination with a series of further superior properties such as e.g. record-breaking mechanical strength, stiffness, flexibility, heat conductance, broadband light absorption and chemical inertness [5], together with its ease of access using only a piece of graphite and an adhesive tape, has engendered the avalanche-like increase in research efforts of laboratories all over the globe. Despite its superior properties, the lack of a band gap limits graphene's suitability in the areas of digital electronics which require high on–off ratios, for optoelectronic and light-emitting devices, and for solar energy conversion applications [68]. Inspired by graphene, many other 2D materials have been explored and the list of theoretically predicted and experimentally realized atomistic thin crystals is steadily growing, attaining several hundred different derivatives that can be classified as 'families' of 2D materials with different physical properties even varying with the number of layers [911]. The main families are the graphene family with graphene and its chemical modifications along with hexagonal boron nitride (hBN), the layered chalcogenides and the layered oxides [12]. 2D materials span the electronic classes from metals, half-metals, insulators, semiconductors, superconductors, to magnetic and ferroelectric materials and even more exotic ones such as topological insulators [914]. Consequently, for almost any application a conceptually suitable 2D material is likely to exist.

A major advance towards the realization of future 2D materials-based flexible nano- and optoelectronic devices was the observation by Mak et al [15] in 2010. A single layer of a MoS2 crystal turns into a direct gap semiconductor emitting in the visible range [15, 16], followed by the realization of a single-layer MoS2 field effect transistor (FET) operating at room temperature with device characteristics comparable to silicon-based FETs [17, 18]. MoS2 belongs to the class of transition metal dichalcogenides (TMDs) consisting of a triple of a transition metal M surrounded by two chalcogenide atoms X with the formula MX2. Particularly, the four stable semiconducting TMDs (SC-TMDs) MoS2, MoSe2, WS2 and WSe2 excel due to their inertness and high stability in ambient conditions similar to graphene or hBN, an insulating material with a large band gap of about ~6 eV [19]. SC-TMDs possess outstanding electronic [7, 8, 17, 20, 21], excitonic [2228], mechanical [29] and fascinating spin- and valley [3033] properties. They also feature various exotic properties including single-photon emission e.g. in WSe2 at low temperatures [3438], photoluminescence efficiency with near-unity quantum yield achieved by chemical treatment of sulfur-based SC-TMDs [39, 40], and layer-dependent superconductivity in highly doped MoS2 [4145], to name just a view. In addition, more technologically relevant characteristics are their outstanding FET performance [17], high sunlight absorbance of up to 15% in the visible range [4648], ultrahigh photosensitivity [4951], catalytic activity [52, 53] and photocatalytic stability in aqueous electrolytes [54], making the material not only interesting as a candidate to replace silicon in electronics but also as photocatalyst, e.g. for solar hydrogen evolution [55] or carbon dioxide reduction [56], and also for cheap optical water disinfection [57].

In this review, we focus on the optical properties of the stable SC-TMDs and their van der Waals heterostructures with special emphasis on how these materials interact with light. We begin with a recap of the crystal structure, the striking changes in the electronic band structure by changing the number of layers, and the most prominent fabrication methods. The fundamental light–matter interaction is determined by the materials' complex dielectric function. We discuss this function—a tensor entity—of the SC-TMDs in relation to its dependence on the number of layers, and highlight differences between the four most prominent examples. Similar to graphene, Raman spectroscopy turns out to be a highly versatile, non-destructive and fast tool for studying 2D materials. The Raman active phonon modes of the TMDs are introduced and the phonon fingerprint for number of layers, strain, temperature, doping and defects is summarized. Furthermore, we demonstrate, using MoS2 as an example, the sensitivity of Raman experiments in studying the impact of the environment on the properties of 2D crystals. The combination of different SC-TMD monolayers in vertically stacked van der Waals bilayer heterostructures with type-II band alignment offers the potential for device applications such as light-emitting diodes [58, 59] and solar cells [60, 61], but also to study correlation phenomena in dense systems of bosonic quasiparticles—interlayer excitons, where the photoexcited electron resides in one layer spatially separated from the photoexcited hole in the other layer. The reduced overlap of the electron and hole wave functions results in an increased lifetime of the excitons and hence, in a high-density exciton ensemble, that is predicted to exhibit macroscopic quantum phenomena such as a superfluid phase at high temperatures of several tens of kelvin [62].

We review recent advances in the direction of van der Waals heterostructures with SC-TMDs and provide typical Raman and photoluminescence measurements on a MoS2/WSe2 van der Waals bilayer. The review will close with a short summary and some perspectives given by the fact that the optoelectronic properties and light–matter interaction in SC-TMDs are highly tunable by interfacial engineering [6368], chemical treatment [39, 40], strain [6971], defect engineering [72], doping [4145, 7376], in-field effect devices [77] and in combination with suitable heterostructures [12, 14]. This directly leads to an outlook on the great potential of SC-TMDs in several areas bridging fundamental science and future device applications due to their rich, manifold properties.

2. Crystal structure and electronic bands of semiconducting TMDs

The TMD family forms a vast group of more than 60 different compounds that have been studied in the bulk and multilayer forms for decades. The research is motivated by their variety of fascinating properties including metallic, semiconducting, insulating, superconducting and magnetic behavior as well as their thermal, mechanical, catalytic and optical properties including exciton screening [78]. Perhaps the longest known compound is MoS2, a.k.a. molybdenite, similar to graphite and a naturally occurring mineral that can be mined. Due to its layered structure, MoS2 is used as a dry lubricant [79], but is also of interest for catalysis [52, 80] and photovoltaics [81]. The observation of the exciting change of the physical properties with the number of layers [15, 16, 35, 82] was the advent of the recent, intense, and worldwide research on TMDs.

2.1. Crystal structure

TMDs are layered materials with a triple X-M-X layer as the smallest unit, where the transition metal M is surrounded by two chalcogen atoms X with strong covalent in-plane bonds, whereas the individual layers hold together by weak van der Waals interaction. The metal atoms are condensed either in octahedral or trigonal prismatic coordination, resulting in semiconducting or metallic behavior [20]. An individual layer can be seen as hexagonally arranged MX2 molecules [83]. Different stacking sequences in multilayer structures result in a variety of polytypes with diverse properties. Here we focus on the semiconducting polytype of MoS2, WS2, MoS2 and WSe2 that condenses in a 2H-configuration in an ABAB stacking sequence with the second layer superimposed on the first one but rotated by 180° around the c-axis as depicted in figure 1(a) [20]. The unit cell of a single layer consists of one MX2 molecule and hence three atoms. The lattice symmetry of a single layer is D3h [84] without an inversion center as visualized in figure 1(a). The unit cell of bulk materials consists of two MX2 molecules and hence six atoms (see figure 1(a)) and exhibits D6h symmetry with an inversion center. Generally, crystals with an odd number of layers have D3h symmetry and those with an even number of layers have D3d symmetry, but are often treated as a bulk material that has D6h symmetry [84].

Figure 1.

Figure 1. (a) Schematic top and side view of a monolayer transition metal dichalcogenide (MX2) crystal in the 2H phase. The smallest unit is a triple layer, and the atoms are arranged hexagonally in-plane. Stacking order in the 2H phase: a monolayer belongs to the D3h point group, a bilayer to the D3d point group and bulk material to the $D_{\text{6h}}^{4}$ point group. (b) Sketch of the hexagonal Brillouin zone of a monolayer MoS2 with the most important symmetry points K, M, Γ. (c) Calculated band structure of bulk, bi- and monolayer MoS2 without spin–orbit coupling effects. The lowest-energy transitions at the K point (direct) and close to the Γ-point (indirect) are indicated. By lowering the number of layers from bi- to monolayers, the direct transition becomes the lowest-energy transition in the band structure, causing a transition from an indirect to a direct semiconductor in the monolayer limit. (Adapted with permission from [16]. Copyright 2010, American Chemical Society.) (d) False-color representation of an optical micrograph of a MoS2 flake on a Si/SiO2 substrate with mono- (1L), bi- (2L), trilayer (3L) and bulk regions.

Standard image High-resolution image

In 1923 Dickinson and Pauling studied the crystal structure of MoS2 by x-ray diffraction and determined the lattice constants of bulk material to be a  =  3.15 Å and c  =  12.30 Å [85] in the 2D plane and perpendicular to it, respectively. The distance between Mo and S atoms constitutes dMo-S  =  2.42 Å and between two sulfur atoms dS–S  =  3.17 Å [86]. The interlayer distance is given by d  =  6.5 Å and is commonly regarded as the thickness of a single layer [20].

2.2. Electronic structure

The electronic band structures of the four SC-TMDs in the focus of this review exhibit very similar signatures with a remarkable dependence on the number of layers due to quantum confinement [15]. The Bloch states defining the single-particle band structure are defined by the electron configuration of the atoms together with the bonding structure. In figure 1(c) the band structure determined from density functional theory (DFT) calculations is plotted for bulk, bilayer and monolayer MoS2 [16]. Bulk and bilayer MoS2 exhibit an indirect band gap close to the Γ-point that is increased by decreasing the number of layers. Nearby is a direct transition located both at the K- and the K'-points that remain almost unaffected by changing the number of layers. The transitions at the K- and K'-points are the lowest-energy transitions for a monolayer MoS2, because the valence band maximum at the indirect transition is further lowered and simultaneously the conduction band minimum increased so that the indirect transition holds a larger gap compared to the direct transition in the monolayer limit. The transition from an indirect to a direct semiconductor by reducing the number of layers from a bi-layer to a monolayer manifests in a significant increase in the photoluminescence quantum efficiency for monolayers compared to bi- and multilayers, ranging from just one order of magnitude up to several orders [15, 16] depending on sample and experimental parameters such as temperature, doping density, defects, strain and environment.

The origin of this transition is explained by changes in the hybridization between chalcogen pz-orbitals and transition metal d-orbitals together with quantum confinement phenomena. The valence band and conduction band states near the Γ-point are predominantly formed by a linear combination of transition metal d-orbitals and antibonding pz-orbitals from the chalcogen atoms. As a consequence, these states are significantly affected by interlayer interaction as well as quantum confinement [87]. In turn, the energy of the conduction band increases near the Γ-point, whereas the energy of the valence band decreases with lowering the number of layers [16]. In contrast, both the conduction and valence band at the K- and K'-points are dominated by localized transition metal d-orbitals that lie in the middle of each X-M-X triple unit. Therefore, the energies of the electronic bands at the K- and K'-points are almost unaffected by interlayer coupling effects, and therefore mostly independent of the number of layers. The excitonic transitions at the K- and K'- points are much brighter in optical emission experiments compared to the indirect transitions. This qualitative description holds for all SC-TMDs [20, 87]. The bandgaps for monolayers MoS2, WS2, MoSe2 and WSe2 constitute 1.8–1.9 eV, 1.8–2.1 eV, 1.5–1.6 eV and 1.6–1.7 eV, respectively, determined from DFT band structure calculations [13]. Whereas DFT calculations might underestimate the single-particle band gap from TMDs, the numbers agree reasonably well with the lowest energy transitions in optical absorption or photoluminescence measurements with the transition energy significantly lowered by excitonic effects compared to the single-particle band gap. The single-particle electronic bands, as well as the excitonic effects, are highly sensitive to environmental and doping-induced screening and renormalization effects [24, 42, 73, 76, 8890] and they can be also altered by strain [70, 71, 91, 92] which enables us to engineer the optoelectronic properties.

Generally, the d-orbitals of the transition metal cause strong spin–orbit coupling (SOC). As a direct consequence, the valence band is spin-split due to SOC by ~150 meV for the MoX2 and by ~450 meV for the WX2 compounds [13, 47, 9396]. The two spin-split optical transitions are called A- and B-excitonic transitions in optical emission and absorption spectra. The band structure for the monolayer in the close vicinity of the K- and K'- valley can be modelled by an effective 2D massive Dirac Hamiltonian consisting of two additional terms taking into account the finite band gap due to broken symmetry and SOC due to the transition metal d-orbitals [94]. The two degenerate high-symmetry points in the Brillouin zone at the K- and K'- valley and the lack of an inversion center for monolayers locks the spin and valley pseudospin degree of freedom [30, 32, 33, 93, 94]. The valley pseudospin can be controlled by excitation with circularly polarized light, resulting in interesting valley and spintronic studies [33, 93] such as valley polarization [30, 31] and the observation of the valley Hall effect [32]. Furthermore, the valley pseudospin can be controlled by applied magnetic fields [97, 98] and the degeneracy can be broken either by the magnetic field [99, 100] or the optical Stark effect [101, 102].

2.3. Fabrication

For fundamental studies, mono- and few-layer TMD flakes are mostly prepared by micromechanical cleavage from bulk crystals using an adhesive tape [3, 4]. The flakes are then transferred to an appropriate flat substrate material, e.g. Si/SiO2, sapphire or glass. The thin crystals can then be identified by optical microscopy, utilizing light interference contrast, as done for graphene [103107]. An example of an optical micrograph of a MoS2 flake deposited on a Si/SiO2 substrate with 300 nm SiO2 is shown in figure 1(d). The contrast is highlighted by a false-color representation demonstrating that the mono-, bi-, tri- and few-layer regions are clearly distinguishable. This exfoliation method is applicable for almost all 2D materials existing as bulk crystals and results in high-quality and ultra-pure samples perfectly suited for fundamental studies. Exfoliated or grown flakes can be precisely positioned on a substrate using several transfer methods for van der Waals assembly [12, 14] e.g. by colamination [19, 108, 109], PDMS stamping [110] or by a pick-and-place method [111, 112]. The various stacking methods also allow the preparation of vertically stacked van der Waals heterostructures with precise control over the individual layers as well as the rotational alignment [113]. However, the micromechanical cleavage method is not scalable; the yield on large single-layer flakes is small and lacks control over size and thickness of the flakes.

Regarding scalable fabrication methods, there is great development in both bottom-up and top-down approaches to produce large amounts of high-quality 2D materials for different applications [20, 114]. Two typical top-down approaches use bulk materials or powder for liquid exfoliation in solvents [115] or for chemical delamination by chemical intercalation [116] or galvanostatic discharging [117] to produce solutions with a large quantity of TMD nanosheets. Thin films can be prepared from such solutions, e.g. by inkjet printing, spray or dip-coating and filtration [114]. These approaches are scalable and the achieved nano-meshes are suitable for some applications. However, these methods do not produce large ultra-pure single crystalline materials.

The synthesis of large-area wafer scale and uniform layers of single-layer TMDs is approached by bottom-up methods such as chemical vapor deposition (CVD) [118121], metal organic chemical vapor deposition (MOCVD) [122], physical vapor phase growth (PVG) [123, 124] or dip-coating of a precursor and thermolysis crystallization by annealing in sulphur gas [125]. Great progress has been made, and it is possible to grow single-layer, high-quality flakes up to 300 µm in size covering a cm-sized substrate surface [126]. It is still challenging to avoid large gradients across the substrate and to grow homogenously connected single-layer films with high crystal quality [20]. Nevertheless, lateral and vertical van der Waals heterostructures with SC-TMDs that are optically and electrically active have been realized by means of CVD [127130], indicating the realistic potential to overcome the scalability issue by the fabrication of single-layer SC-TMDs and their heterostructures for future applications.

3. Complex dielectric function and absorbance of TMDs

Optical properties including absorption efficiency, optical transitions and excitonic phenomena are of particular interest for fundamental studies, and key to realizing optoelectronic applications. SC-TMDs exhibit strong light–matter interaction, particularly in the visible range [27]. The light–matter interaction is basically described by the complex dielectric function of a material that helps to link experimental observations and theory to interband excitations. The dielectric function is a tensor entity and therefore the light–matter interaction for a light field parallel and perpendicular to a 2D crystal is highly anisotropic [48]. The optical properties of SC-TMDs strongly depend on the number of layers [15, 16]. For this reason, detailed knowledge of the dependence of the complex dielectric function on the number of layers is highly desirable. Typical methods to extract the dielectric function are ellipsometry measurements under a certain angle of incidence to probe in-plane and out-of-plane components of the light–matter interaction with macroscopically large light spots [131133]. The fabrication of flakes with precise control over the number of layers is thus far only achievable by micron-sized mechanically exfoliated flakes. Imaging ellipsometry (IE) offers a high lateral resolution while maintaining precise control over the angel of incidence. IE has been proven to allow access to the dielectric function of 2D materials such as graphene [134], graphene oxide [135] and mono- and multilayer MoS2 [48]. In particular, for monolayer flakes with a thickness of less than 1 nm, it can be assumed that interaction with a light field perpendicular to the 2D sheet can be neglected. Focused reflectivity measurements with the light field parallel to the 2D plane are suitable to extract the light–matter interaction of SC-TMDs utilizing a Kramers–Kroning constrained variational analysis [47].

3.1. Layer-dependent anisotropic dielectric function and absorbance of MoS2

In a recent study, we investigated the complex dielectric function of mechanically exfoliated mono-, tri- and few-layer MoS2 on a transparent sapphire substrate by imaging ellipsometry in the spectral range from 1.7 eV–3.1 eV. The used angle of incidence of 50° allowed us to access the optical response of MoS2 parallel (x, y) and perpendicular (z) to the plane of the flake. A lateral resolution of better than 2 µm enabled us to probe ultra-high-quality crystalline regions. The complex dielectric functions for different numbers of layers were extracted from the measured ellipsometric angles Δ and ψ as input to a multilayer model using Lorentz as well as Tauc–Lorentz fit approaches [48]. In the model, the thickness of the MoS2 monolayer was kept at the theoretical value of d  =  6.15 Å [78, 85]. The resulting anisotropic complex dielectric functions εx  =  εy  ≠  εz with real part ε1(E) and imaginary part ε2(E) are plotted for mono-, tri- and few-layer MoS2 in figure 2 [48]. The in-plane component of the dielectric function εx(E)  =  εy(E) is best described by five Lorentz profiles and the out-of-plane component εz(E) by one Tauc-Lorentz profile. As visualized in figures 2(a) and (b) the in-plane component exhibits three well pronounced critical points and two rather sharp peaks centered at 1.9 eV and 2.05 eV assigned to the A and B spin-split excitonic transition at the K- and K'-points in the Brillouin zone. The prominent but rather broad resonance at around 2.9 eV is called the C excitonic transition and indicates a high joint density of states [23, 78]. This energy range comprises a combination of several individual bright and dark excitonic transitions between 2.2 eV and 3 eV that are located near the Γ-point [23] in addition to a band nesting region—a region with nearly parallel conduction and valence bands between the M- and Γ- points with a high joint density of states. The individual transitions are significantly broadened by the interaction of electrons with optical phonons [23, 136].

Figure 2.

Figure 2. Complex dielectric function and absorbance of MoS2 from spectroscopic ellipsometry measurements. (a) Real part of the in-plane component of the dielectric tensor ε1(x, y) from an anisotropic model for mono-, tri-, four- and few-layer MoS2. (b) Imaginary part of the in-plane component of the dielectric tensor ε2(x, y) as in (a) and (c) Real part ε1(z) and (d) imaginary part ε2(z) of the dielectric tensor in the out-of plane direction from the anisotropic model to the ellipsometry measurements. (Adapted with permission from [48]. Copyright 2016, IOP Publishing.)

Standard image High-resolution image

The out-of-plane component plotted in figures 2(c) and (d) exhibits one critical point that is located for the monolayer MoS2 at 1.9 eV and hence slightly above the A excitonic transition at the direct band gap. The energy of the critical point is reduced in energy for increasing number of layers to 1.7 eV for a trilayer and to 1.67 eV for few-layer flake. Consequently, this optical transition is below the direct optical transition at the K- and K'- points. Therefore, it is most likely related to the states in the Brillouin zone that form the indirect band gap in agreement with the trend that the band gap is reduced for increasing number of layers. The magnitude of the out-of-plane contribution of the dielectric functions is significantly reduced in amplitude compared to the in-plane component and seems to be inversely proportional to the number of layers. This behavior can be intuitively understood such that the interaction of a light field perpendicular to the 2D film is much stronger for a thicker film compared to an atomically thin monolayer. Independent of the number of layers, the dielectric functions of the out-of-plane component approach constant values of ε1(z)  ≈  1 and ε2(z)  ≈  0 above approximately 2 eV [48]. An alternative explanation for the weak out-of plane component of the dielectric function compared to the in-plane component might be also found in the fact that in monolayer TMDs the conduction and valence bands are symmetric with respect to the horizontal mirror plane of the crystal lattice [93, 137]. As a direct consequence dipolar transitions due to out-of plane electric (light) fields are only allowed by breaking the symmetry, e.g. by placing a monolayer TMD on a substrate or by covering the layer with a medium with a different dielectric function. Similar symmetry arguments might be developed for multilayers.

The absorbance α of the isolated MoS2 mono-, tri-, and few-layer films can be directly calculated from the dielectric function, taking the relation between the extinction coefficient κ  =  2ε1ε2 and the absorbance α  =  4πκ/λ at a light wavelength λ. The in-plane and out-of-plane absorbance for mono- and trilayer MoS2 is plotted in figure 3. A monolayer MoS2 with a thickness of less than 1 nm absorbs about 4% of the parallel incoming light field at the energy of the A- and B-exciton and almost 15% at the C-exciton [15, 4648]. In the whole visible range, the absorbance is reduced by increasing the number of layers. In particular, the excitonic resonance at the A- and B-transitions are broadened and reduced in intensity, indicating an increased dielectric screening [138]. While the energy positions of the A- and B-excitons are less affected by the number of layers, the energy position of the C-resonance shifts to lower energies by increasing the number of layers, demonstrating that the electronic band structure and many-body interactions, such as the Coulomb interaction responsible for the exciton formation, are located at different k-values in the Brillouin zone. These effects are unequally affected by changing the number of layers depending on the atomic orbitals dominating the Bloch states.

Figure 3.

Figure 3. Absorbance α for mono- and trilayer MoS2 extracted from the optical constants, i.e. from the extinction coefficient in the visible range displaying high absorption efficiencies at A, B and C exciton transitions. The solid spheres are the results deduced from the in-plane component (xy-plane) from the anisotropic model (black monolayer and red trilayer). The grey and green spheres represent the absorbance perpendicular (z-direction) to the 2D crystal determined from the anisotropic model (monolayer green, trilayer grey). (Adapted with permission from [48]. Copyright 2016, IOP Publishing.)

Standard image High-resolution image

The absorbance of a light field perpendicular to the 2D film is 0.2% at the excitonic resonance at 1.9 eV and zero elsewhere. Only 1% of an out-of-plane light field is absorbed by the excitonic peak for trilayer and about 4% for few-layer MoS2.

The described characteristics of the in-plane dielectric functions εx(E)  =  εy(E) are in good agreement with the absorbance spectra for exfoliated and CVD-grown ultrathin MoS2 layers determined either by spectroscopic ellipsometry with isotropic or anisotropic fit approaches or extracted from reflectance measurements [47, 48, 131, 132, 139, 140]. The universality of the main characteristics of the dielectric functions underlines the robustness of the strong light–matter interaction in atomically thin MoS2 layers. However, we would like to emphasize that the strength as well as the specific spectral position of the A-, B- and C-excitonic transitions are unique for each sample and highly affected by the individual situation such as environment, substrate, doping or strain just to name some major sources modifying the light–matter interaction. As described above, the different critical points in the electronic band structure with a high joint density of states are formed by different atomic orbitals of the transition metal and sulfur atoms, respectively. The A- and B-excitons can consequently be affected differently by perturbations and changes compared to the higher-energy transitions at the C-excitonic resonance or to the spectral region between the B- and C-exciton transitions.

3.2. Comparison of the in-plane dielectric function and absorbance for MS2 and MSe2 (M  =  Mo, W)

All four SC-TMDs are characterized by strong light–matter interaction, resulting in similar values and overall characteristics in absorbance spectra as well as dielectric functions in the visible range as demonstrated in figure 4 for MoSe2, WSe2, MoS2 and WS2 monolayers [47]. The 2D crystals were prepared from bulk crystals by micromechanical exfoliation and placed on fused silica substrates. The optical parameters were extracted from reflection measurements with the incident focused light beam perpendicular to the 2D plane. The light field in such a configuration is parallel to the sample plane and therefore only the in-plane light–matter interaction can be probed. The in-plane components of the dielectric functions εx(E)  =  εy(E) are derived from the reflection spectra by a Kramers–Kroning constrained analysis [47].

Figure 4.

Figure 4. (a) In-plane components of the real part ε1 (left) and imaginary part ε2 (right) of the dielectric functions determined from reflection measurements for micro-mechanically exfoliated MoSe2, WSe2, MoS2 and WS2 monolayers on fused silica [47]. (b) Absorbance of micro-mechanically exfoliated MoSe2, WSe2, MoS2 and WS2 monolayers that are free standing (left) and supported on fused silica (right) [47]. (All panels reprinted with permission from [47]. Copyright 2014 by the American Physical Society.)

Standard image High-resolution image

As already introduced above, the absorbance can be directly extracted from the dielectric functions. The absorbance spectra for quasi free-standing and supported monolayers are displayed in figure 4(b), respectively. The authors of [47] have calculated the absorbance spectra for the supported monolayers starting from the identical dielectric function (shown in figure 4(a)), but taking into account a local field correction factor for the light intensity above the substrate that depends on the refractive index of the substrate material [47]. The absorbance of the free-standing material constitutes in the visible range up to 15% and is reduced by about 1/3 for the monolayers on fused silica [47]. The high absorbance values underline the strong light–matter interaction of SC-TMDs even in the monolayer limit with a thickness of less than 1 nm, as has been reported by several theoretical and experimental studies [15, 23, 27, 28, 4649, 131, 133, 139, 141]. The sizeable reduction of the absorbance for monolayers on a substrate points to the fact that the optical response can be significantly altered just by modification of the substrate or environment by dielectric engineering and screening effects [39, 63, 64, 66, 74, 90, 138, 142144]. The strong influence of the environment on the light–matter interaction of atomically thin semiconducting membranes motivates the great potential not only for sensing applications, but also for novel device architectures with precisely tailorable optical properties.

The critical points in the optical response found in the complex dielectric functions and absorbance spectra are similar for MoSe2, WSe2, MoS2 and WS2 monolayers. The lowest-energy transitions, labeled as A- and B-excitons, belong to the direct optical transitions at the K- and K'- points of the band structure. The energy of the A-exciton is the single-particle band gap minus the exciton binding energy which is of the order of several hundred meV [23, 26]. Since both the single-particle band structure and the Coulomb interaction determining the exciton binding energy are reduced by screening effects such as doping [24, 73, 75, 76], the A-excitonic transition energy is only minimally affected by screening. However, the oscillator strength in both absorption and photoluminescence is significantly altered by screening and doping [64, 73, 88, 90, 142]. The B-exciton belongs to the K-, K'- points in momentum space and is the transition from the lower-energy spin-split valence band state to the conduction band [93]. Since the Coulomb interaction is comparable for A-and B-excitons, the energy splitting between A- and B-excitons in the absorbance is a direct measure for the spin-splitting in the valence band due to spin–orbit coupling. The spin-splitting of the valence bands at the K- and K'-points constitutes between 150 meV and 200 meV and between 400 and 450 meV for the MoX2 and WX2 (X  =  S, Se) monolayers, respectively. The increase in the spin-splitting of the tungsten-based compounds is due to the larger atomic number of tungsten and, hence, an increase in the spin–orbit coupling strength of the tungsten d-orbitals. Similar to the discussion of the broad C-resonances for MoS2 monolayers above, the broad C-excitonic signatures for the other compounds belong to higher-lying interband transitions with a high joint density of states at different areas in momentum space. The same holds for the so-called D-resonance observable for WSe2 in the plotted spectral range [47].

MoSe2, WSe2, MoS2 and WS2 monolayers exhibit superior light–matter interaction properties and exciton-dominated optical response in the visible range with maxima in the peak absorbance for freestanding materials in the order of 15% for less than 1 nm thick crystals. As expected for 2D materials, the optical responses described by the complex dielectric functions are highly anisotropic for the interaction of the light field parallel or orthogonal to the plane of the 2D materials. The light–matter interaction is reduced for increasing number of layers and the excitonic signatures are broadened presumably due to a decreased Coulomb interaction and hence decreased excitonic phenomena. However, further decay and relaxation channels of the carriers by interlayer scattering exist. In particular, the C-excitonic transition is redshifted with an increasing number of layers. The same observation holds for the only critical point in the out-of-plane component of the dielectric function εz(E). It is found that this resonance is redshifted with increasing number of layers. The amplitude, however, is enlarged with the number of layers, as intuitively expected since the thickness in the z direction increases.

4. First-order phonon modes as a unique fingerprint for material properties

Raman spectroscopy is a very powerful tool for studying many structural but also electronic properties of 2D materials due to the sensitivity of the fundamental Raman active lattice vibrations to perturbations of the lattice, but also to changes in the electronic properties caused by electron–phonon interaction and phonon renormalization effects as well as, more indirectly, by resonance effects if the incoming or scattered light meets a fundamental optical interband transition of the electronic band structure. Raman scattering is a versatile, non-destructive, fast characterization tool without the need for extensive device fabrication. Its power has significantly supported the success of graphene research [145, 146]. The strength of Raman investigations continues for research on MoS2 and other TMDs. However, there is one fundamental difference between graphene and SC-TMDs: due to the gapless linear band structure of graphene, Raman spectroscopy is resonant for all excitation energies enabling, e.g., the activation of the double resonant 2D mode in graphene [145, 146]. SC-TMDs hold a finite band gap and therefore resonant Raman spectroscopy is only possible if the incoming or scattered light meets a fundamental optically active interband transition [145, 147]. In this respect, resonant Raman spectroscopy with varying photon energies provides a complementary tool to emission and absorption measurements to access the electronic structure in SC-TMDs as well as to study fundamental aspects of the electron–phonon interaction [145, 147, 148]. Non-resonant Raman spectroscopy, meaning that neither the exciting (incoming resonance) nor the scattered photon energy (outgoing resonance) meets an optically active interband transition of the electronic bands, provides access to the number of layers [82, 149153], a phase transition between 2H, 3R, 1T, 1T' MX2 crystal phases [148, 154, 155], strain [70, 156, 157], disorder and defects [158], doping [74, 159] and temperature as well as thermal conductivity [106, 160163]. Amazingly, this information can already be extracted by investigating only two Raman active zone center phonon modes in the higher-frequency range (ωph  >  200 cm−1). These are an in-plane mode with $E_{\text{2g}}^{1}$ (E') symmetry and a homopolar out-of-plane mode with A1g (A'1) symmetry for bulk (monolayer) crystals. The in-plane mode is the degenerate LO/TO phonon of the material. Overall, 2H-TMDs possess four Raman active phonon modes with the displacement of the atoms sketched in figure 5(a). Besides the two high-frequency modes, there are two lower-frequency modes (ωph  >  45 cm−1), the layer breathing mode and the interlayer shear mode with E1g (E'') and $E_{2\text{g}}^{2}$ symmetry, respectively. Obviously, for monolayers only the breathing mode with E'' symmetry exists.

Figure 5.

Figure 5. (a) Visualization of the displacement of the atoms for the Raman active phonon modes in TMDs. (b) Raman spectra for micro-mechanically exfoliated MoS2 mono-, bi-, tri-, four-, layer and bulk flakes. The solid lines are fits to the spectra with two Lorentzians (MoS2 on Si/SiO2, excited with λ  =  488 nm). (c) Energy difference between the two Raman active phonon modes A1g and $E_{\text{2g}}^{1}$ measured for MoS2 from (a), which is an explicit measure to determine the number of layers.

Standard image High-resolution image

In particular, the interlayer shear mode has proven to be a very sensitive measure of the interlayer coupling in artificially stacked bi- and multilayers [164] as well as of the dependence of the interlayer coupling strength from the twist angle between two monolayers [10, 165, 166]. Not only for monolithic devices but also for van der Waals hetero bi- and multilayers, the interlayer coupling strength can be evaluated by the energy of the interlayer shear and breathing modes [167]. Optical signatures for van der Waals heterostructures and the interlayer coupling strength will be discussed in detail in a separate section below.

In the course of this section, we focus on aspects of non-resonant Raman investigations of the higher-frequency intra-layer Raman active modes of SC-TMDs with special emphasis on the determination of the number of layers, on doping-induced phonon-renormalization, and the resulting options to study the change in doping density without the need for contacts and leads. Furthermore, we cover the impact of strain, defects and temperature on the phonon modes. The Raman tensor of the two modes of interest are for the A1g (A'1g) symmetries $\left(\begin{array}{c} a & 0 & 0 \\ 0 & a & 0 \\ 0 & 0 & b \end{array}\right)$ and for the $E_{\text{2g}}^{1}$ (E') symmetries $\left(\begin{array}{c} 0 & d & 0 \\ d & 0 & 0 \\ 0 & 0 & 0 \end{array}\right)$ , $\left(\begin{array}{c} d & 0 & 0 \\ 0 & -d & 0 \\ 0 & 0 & 0 \end{array}\right)$ .

As a result, the out-of-plane mode is for linearly and circularly polarized light observable in a co-polarization geometry and the in-plane mode is un-polarized for linearly polarized light and co-polarized for circularly polarized light [168]. For a more detailed description of the classical as well as quantum description of the Raman tensor of TMDs, the inelastic light scattering itself and consequences for the polarization dependence, we refer to a recent comprehensive review by Saito et al [148]. Moreover, we would like to point out that there are recent review articles with the primary focus on Raman spectra of TMD materials [148, 169, 170].

4.1. Counting the number of layers from Raman measurements

As introduced in 2010 by Lee et al [82] for MoS2, the frequencies of the first-order phonon modes with $E_{\text{2g}}^{1}$ and A1g symmetry change monotonously with the number of layers with opposite trends. Raman spectra taken on exfoliated MoS2 on Si/SiO2 substrate are plotted in figure 5(b). The in-plane mode $E_{\text{2g}}^{1}$ softens and the Raman signal is therefore redshifted with increasing number of layers, whereas the out-of-plane mode A1g is significantly stiffened and the Raman frequency is blueshifted. Therefore, the energy difference between the two modes ΔE  =  |ω(A1g)  −  ω($E_{\text{2g}}^{1}$ )| allows us to determine the number of layers with very high precision, as demonstrated in figure 5(c). In particular, mono-, bi- and trilayers can be unambiguously determined.

The stiffening of the out-of-plane A1g mode with increasing number of layers is well understood considering the classical model for coupled harmonic oscillators [82]. The unusual behavior of the in-plane $E_{\text{2g}}^{1}$ mode in softening with increasing number of layers is controversially discussed in the literature and was first explained by a modification of long-range Coulomb interlayer interactions induced by the additional layer, causing an enhanced dielectric screening between the effective charges in MoS2 [82, 153]. The unusual layer dependence of the $E_{\text{2g}}^{1}$ mode can be more likely explained by surface effects with a larger Mo-S force constant at the surface of thin film compared to the inner layers [170, 171].

These two interlayer phonon modes are useful to determine the number of layers for other SC-TMDs, too. In table 1, the relevant mode frequencies are summarized for monolayers and bulk materials of MoS2, WS2, MoSe2 and WSe2. The same trend as for MoS2 holds for the two interlayer phonon modes of WSe2, MoSe2 and WS2 [172]. Only for monolayers of WSe2, the E' and $A_{1}^{\prime}$ modes are degenerate with a frequency of ω  ≈  250 cm−1 [152] and cannot be separated in experiments. However, they are split for WSe2 bulk and the (degenerate) mode frequencies change with the number of layers so that an estimation of the number of layers is also possible for WSe2 [152]. For completeness we would like to mention that the low-frequency interlayer shear modes and layer breathing modes are very sensitive to the number of layers in TMD materials [149, 173175].

Table 1. Energies for the most prominent first-order Raman modes E'/$E_{\text{2g}}^{1}$ and $A_{1}^{'}$ /A1g belonging to the Raman active optical phonon branches of the phonon dispersion at the Γ-point for MoS2, WS2, MoSe2 and WSe2 monolayers and bulk materials, respectively. Notations are given with respect to the D3h/D6h point groups for monolayer/bulk materials condensed in 2H symmetry. All values are extracted from measurements on mechanically exfoliated flakes excited at room temperature with a laser wavelength of λ  =  514.5 nm.

  MoS2 [82] WS2 [150] MoSe2 [152] WSe2 [152]
E' (1L) 384.2 cm−1 355.2 cm−1 287.2 cm−1 ≈250 cm−1
$E_{\text{2g}}^{1}$ (bulk) 382.0 cm−1 355.3 cm−1 285.9 cm−1 248.0 cm−1
$A_{1}^{'}$ (1L) 403.0 cm−1 417.2 cm−1 240.5 cm−1 ≈250 cm−1
A1g (bulk) 407.1 cm−1 420.1 cm−1 242.5 cm−1 250.8 cm−1

4.2. Phonon renormalization due to doping

Figures 6(a) and (b) show the E' ($E_{\text{2g}}^{1}$ ) and $A_{1}^{\prime}$ (A1g) phonon mode frequencies for monolayer and bilayer MoS2 as a function of the top gate voltage using a solid electrolyte gate electrode [74]. The flakes have been placed on a highly p-doped Si back electrode covered with 300 nm SiO2 and contacted by two Ti/Au contacts. A solid electrolyte gate using polyethylene oxide (PEO) and cesium-perchlorate is utilized because its high capacity allows us to tune the charge carrier density over a wide range by applying a relatively small voltage [159]. With the top gate, the 2D electron density in the MoS2 flakes can be tuned by two orders of magnitude from approximately 1011 to 1013 cm−2 by applying a top gate voltage of VTG  =  ±1 V [74]. Raman measurements are carried out at room temperature with an excitation wavelength of λ  =  488 nm. For both, mono- and bilayer MoS2 the phonon frequencies of the in-plane modes are almost unaffected by changing the gate voltage and hence the carrier concentration, whereas the out-of-plane modes are redshifted by Δω(A')  ≈  3 cm−1 for the monolayer and Δω(A')  ≈  1.5 cm−1 for the bilayer by changing the charge carrier density by approximately two orders of magnitude. In both cases, the change in the mode frequencies seems to be directly proportional to the applied gate voltage and therefore to the 2D charge carrier density. A line-shape analysis using two Lorentz profiles to describe one Raman spectrum reveals that simultaneously the FWHM of the out-of-plane modes increases with increasing charge carrier densities. By increasing the charge carrier density from ~1011 to ~1013 cm−2, the FWHM increases from approximately 6.5 cm−1 to 10 cm−1 for the monolayer and from 8.8 cm−1 to 11.6 cm−1 for the bilayers (figures 6(c) and (d)). The in-plane modes are again unaffected by changing the charge carrier density and the FWHM for the mono- and bilayers is approximately 5.5 cm−1.

Figure 6.

Figure 6. Energy and full width at half maximum (FWHM) of the A1g and $E_{\text{2g}}^{1}$ modes as a function of the applied top gate voltage. (a) Phonon mode energies for monolayer MoS2 and (b) for bilayer MoS2. Solid lines are linear fits to the data points. (Panels (a) and (b) reprinted from [74], with the permission of AIP Publishing 2015.) (c) FWHM of the phonon modes extracted from Lorentzian fits to the Raman spectra for monolayer and (d) for bilayer MoS2. The Raman measurements are carried out at room temperature with an excitation wavelength λ  =  488 nm.

Standard image High-resolution image

Very similar qualitative and quantitative findings in Raman investigations on gated MoS2 monolayers have been reported in a comparable range of charge carrier densities by Chakraborty et al [159]. The authors calculated the charge carrier dependence of the electron–phonon coupling from first-principles DFT. They found an excellent agreement between theory and experiments for doped MoS2 monolayers: theoretically the strengthened electron–phonon coupling with increasing charge carrier density results in a significant softening and broadening of the out-of-plane phonon. The in-plane mode is hardly affected by doping [159]. The different reactions of the two optical phonon modes on doping result from their different symmetry. The A' (A1g) mode shares a symmetry with the lattice and consequently structural distortion in this mode does not break the symmetry of MoS2, resulting in a large electron–phonon coupling strength since all electronic states in the electron–phonon coupling matrix elements can have non-zero expectation value [159]. The conduction band states at the K- and K'- points in momentum space have the character of the Mo dz2 state. By doping with electrons, the antibonding states with the Mo dz2 orbital character are filled, which reduces the bond strength. As a result the out-of-plane phonon mode is softened. The occupation of the conduction band minimum states at the K- and K'-points induces an increase in the electron–phonon coupling [159]. Such a doping-induced phonon renormalization is absent for the in-plane mode with E' ($E_{\text{2g}}^{1}$ ) symmetry due to its orthogonality to the A' (A1g) symmetry. Therefore, the matrix element that couples electrons and phonons vanishes for the in-plane mode because of this orthogonality [159].

It is interesting to see that the general behavior of the two optical phonon modes is also valid for bilayers (compare figures 6(b) and (d)) and seemingly also for tri- and four-layer crystals [74] indicating that the conduction band minimum at the K- and K'-points matters for thin 2D crystals with an indirect bandgap. The symmetry argument that the phonon renormalization is only effective for the A1g modes is valid also for bi-, tri- and four-layer MoS2.

Overall, the sensitivity of the out-of-plane phonon mode with A1g symmetry empowers Raman experiments to identify the charge carrier density in atomically thin SC-TMDs without the need to fabricate Ohmic contacts and to apply magnetic fields to conduct Hall measurements—the standard probe to determine the charge carrier density in doped semiconductors. Since the explanation of the doping-induced phonon renormalization holds also for other SC-TMDs with identical lattice symmetry and direct band gaps at the K- and K'-points, the determination of the charge carrier density from Raman measurements is expected to be also suitable for all four TMDs in the focus of this review.

An example of this convenient method to determine the change in the charge carrier density of monolayer MoS2 is provided by the results shown in figure 7(a). Non-resonant Raman spectra (λlaser  =  488 nm at room temperature) taken on a micro-mechanically exfoliated MoS2 monolayer on a Si/SiO2 substrate are contrasted for measurements in vacuum (p  ⩽  5  ×  10−5 mbar), ambient conditions and with the flake immersed in DI-water (water measurements are performed using a water-dipping objective as described elsewhere [54]). The frequency and linewidth of the in-plane E'-mode are relatively constant for all environments. The frequency of the out-of-plane $A_{1}^{\prime}$ -mode is highest for measurements in water, but reduced by almost 3 cm−1 and broadened for spectra taken in vacuum. As shown by us in an earlier work [74], the $A_{1}^{\prime}$ -mode frequency and its broadening in ambient conditions strongly depend on the laser power and vary between the two limits for measurements in water and vacuum. The mode frequencies for lowest laser power approach the values measured in water, whereas those observed with higher laser power are close to the mode frequencies observed in vacuum. The experimental observations are consistent with a change of the charge carrier density. The free electron density for measurements in vacuum is up to two orders of magnitude higher compared to those with the MoS2 flakes completely immersed in DI water. We attribute this change to the effect of molecular doping. Due to their dipolar moments, some molecules, such as O2 and H2O, can act as molecular gates causing the transfer of a fraction of the charge from the 2D material to the molecules [64, 88], effectively depleting the 2D electron system in MoS2. By illuminating the flakes in vacuum, the surface of the 2D material is cleaned and physisorbed molecules are removed. In vacuum, we assume that the intrinsic charge carrier density is measured. Intrinsically, MoS2 is unintentionally n-type doped with electron densities exceeding 1013 cm−2 most likely because of a high density of sulfur vacancies [17, 25]. Immersed in water, the surface of a MoS2 flake is completely covered with H2O and O2 molecules, as sketched in figure 7(b), resulting in the most efficient molecular gating effect and a depletion of the MoS2 flake by up to two orders of magnitude. The laser power dependence is explained by the gradual removal of physisorbed molecules with a steady state between adsorption and removal depending on the laser power. Since this process is highly reversible, the charge carrier density of MoS2 in ambient conditions can be tuned just by the illumination intensity [74].

Figure 7.

Figure 7. (a) Environmental dependence on the Raman spectra for a micro-mechanically exfoliated monolayer of MoS2 on a Si/SiO2 substrate measured at room temperature under non-resonant excitation (λ  =  488 nm, P  =  100 µW, spot diameter  <1 µm). The measurements are performed in DI-water, ambient and vacuum (p  ⩽  5  ×  10−6 mbar). Filled spheres are the experimental data and solid lines are fits to the data using two Lorentz profiles. (b) Sketch of the MoS2 flake on a substrate covered with water molecules, causing a finite dipolar moment and hence acting as molecular gates.

Standard image High-resolution image

4.3. Impact of strain and defects on the phonon modes

In contrast to the doping that affects only the homopolar out-of-plane phonon with $A_{1}^{\prime}$ (A1g) symmetry, uniaxial strain predominantly affects the in-plane phonon with E' ($E_{\text{2g}}^{1}$ ) symmetry [70, 156]. In particular, uniaxial strain induces a reduction of the mode frequency per % tensile strain of Δω/ε  =  −2.1 cm−1 for the E' phonon mode and only Δω/ε  =  −0.4 cm−1 for the $A_{1}^{\prime}$ phonon mode [156], and another group reports no measurable shift of the $A_{1}^{\prime}$ phonon mode due to uniaxial tensile strain [70]. Bilayer MoS2 is affected in a similar way [70]. For the application of biaxial tensile strain to monolayer MoS2, only minor softening of the $A_{1}^{\prime}$ mode is reported; however the E'-mode is substantially softened and a mode splitting appears for tensile strain larger than 0.8% [157]. The splitting of the doubly degenerate E' mode into the LO and TO branch is assigned to symmetry breaking of the crystal caused by biaxial strain [157]. We note that the application of strain not only affects the zone center phonon modes but also the electronic band structure. The bandgap is lowered with increasing tensile strain and the strained material can even undergo the transition from a direct to an indirect semiconducting material [70, 157].

The effect of disorder on the Raman spectrum of monolayer MoS2 is reproduced in figure 8 [158], where the defects are induced by a bombardment with Mn+ ions with varying dose, enabling a systematic study of the Raman response with respect to the interdefect distance LD. The observed changes in the Raman spectra contrasts the modifications in the spectra induced by strain or doping. With increasing defect density, the E'-phonon frequency is lowered, and the frequency of the $A_{1}^{\prime}$ mode is simultaneously increased [158]. Both modes are broadened with increasing defect density and the LA(M) mode, a longitudinal acoustic mode at the M point at the Brillouin zone boundary, is activated. A combination of doping caused by ion bombardment together with strain cannot explain the experimental signatures since the $A_{1}^{\prime}$ mode frequency not only increases with increasing ion bombardment but also significantly broadens, excluding a reduction of the electron density to be the main source for the modification of the out-of-plane phonon mode. Mignuzzi et al [158] adopted an already established phonon confinement model [176, 177] to explain the defect-induced changes of the first-order phonon modes. The authors find excellent agreement between experiment and numerical calculations of the phonon lineshape utilizing the phonon confinement model together with the phonon dispersion from DFT calculations [158]. Furthermore, the intensity ratios between the defect-activated phonon modes such as the LA(M) and the first-order E' and $A_{1}^{\prime}$ phonons increases monotonically with decreasing interdefect density LD [158] serving as a very suitable estimate for the defect density from Raman measurements in addition to the opposite frequency shifts of the E' and $A_{1}^{\prime}$ phonons.

Figure 8.

Figure 8. Raman spectra of pristine and ion-bombarded (Mn+-ions) monolayers of MoS2 on Si/SiO2 with varying interdefect distances LD. The asterisk refers to a Raman peak of the Si substrate. The spectra have been normalized to the intensity of the $A_{1}^{\prime}$ mode. The spectra are taken at room temperature in ambient conditions and have been excited using a laser with λlaser  =  532 nm. (Reprinted figure with permission from [158]. Copyright 2015 by the American Physical Society.)

Standard image High-resolution image

4.4. Temperature shift of the phonon modes

Temperature-induced changes in the lattice constant of a crystal have immediate consequences on the phonon mode frequencies. Typically, the lattice constant increases with increasing temperature. As a direct consequence the phonon modes are softened. Therefore, Raman measurements provide a fast, non-contact method to monitor local changes in the temperature with the spatial resolution of the laser spot that can be better than 1 µm. A rise in temperature can be directly induced by light irradiation, by electric currents in electronic devices and other heat sources. Raman measurements with a high spatial resolution can not only be utilized to measure the local temperature but also to investigate the thermal conductivity of a material. It has been shown that both optical phonon modes with E'/$E_{\text{2g}}^{1}$ and $A_{1}^{\prime}$ /A1g symmetry of SC-TMDs are lowered with increasing temperature [160163]. The first-order temperature coefficients χT for MoS2 are reported to be linear for both phonon modes, and to be very similar for supported and suspended monolayers. The temperature coefficients for suspended MoS2 mono-layers are χT (E')  =  −0.011 cm−1 K−1 and χT ($A_{1}^{\prime}$ )  =  −0.012 cm−1 K−1 for the E' and $A_{1}^{\prime}$ modes, respectively and, for sapphire-supported MoS2 monolayers, χT (E')  =  −0.017 cm−1 K−1 and χT ($A_{1}^{\prime}$ )  =  −0.012 cm−1 K−1 [160]. With the help of Raman measurements, the thermal conductivity κ of a monolayer MoS2 at room temperature was determined to be κ  =  34.5 W m−1 K−1 [160], which is significantly smaller than the thermal conductivity of graphene with κ  =  2000 W m−1 K−1 [178].

Overall, Raman investigations provide unique access to both number of layers and doping, as well as strain, defect densities and temperatures. Particularly, for monolayers and ultrathin SC-TMDs, the changes in the high-frequency, first-order Raman modes with $A_{1}^{\prime}$ (A1g) (out-of-plane) and E'($E_{\text{2g}}^{1}$ ) (in-plane) symmetries which are easily experimentally accessible, provide unique and unequivocal access to the above-mentioned parameters. A temperature increase shifts both modes to lower frequencies, whereas increasing defect densities shifts the in-plane mode to lower frequencies and the out-of-plane mode to higher frequencies, with both modes becoming broadened. A decrease in the charge carrier density also shifts the out-of-plane mode to higher frequencies; however the mode is then getting narrower and the in-plane mode is unaffected. This character of the two higher-frequency phonon modes enables us to clearly separate between temperature-, defect- and doping-related effects on the Raman response. Moreover, the effect of uni- and biaxial strain on the Raman modes can be clearly distinguished since strain mainly affects the in-plane mode, while the out-of-plane mode is mostly unchanged. With increasing tensile strain, the in-plane mode frequency is reduced, broadened, and eventually the degeneracy of the TO and LO phonon branches lifts, resulting in a splitting of the in-plane mode.

5. Optical properties of TMDC heterostructures

Engineering and controlling the optical and optoelectronic properties of semiconducting materials on demand is highly desirable for technological applications, e.g. for solar cells and light-emitting devices, but also to design host materials for studying emergent quantum phenomena, for instance in dense exciton ensembles. For the realization of optoelectronic devices, bipolar heterostructures are often required. Van der Waals heterostructures can be electrostatically doped by gate structures, because there are intrinsic n-type (e.g. MoS2) or p-type (e.g. WSe2) SC-TMDs available or the type of charge carriers can be induced by contacting the materials with metals with different work functions [179]. In such a way, atomically thin p–n diodes [58] or electrically tunable photovoltaic devices [61] can, for instance, be realized.

2D materials offer a major advantage compared to 3D solid-state-based heterostructures such as GaAs/AlGaAs; due to the absence of dangling bonds on the surface, 2D materials can be arbitrarily combined with atomistically precise interfaces and with an additional rotational degree of freedom [12, 113, 180]. Such 2D heterostructures can be prepared either by van der Waals assembly of micro-mechanically exfoliated or grown 2D crystals [14, 108] (sketched in figure 9(a)) or by a direct growth of lateral and vertical 2D heterostructures [128, 129]. The precise control of the interface is a challenge that needs to be overcome in order to build functional heterostructures with a sufficient interlayer coupling. Hereby, not only is the rotational control of importance, but also strategies need to be developed to avoid and remove contamination from the interfaces. The interface can either be polluted by thin water films or by polymer residues from the carrier materials used during the transfer process. In order to maintain a sufficient interlayer coupling, careful cleaning of the interface with suitable solvents, annealing procedures or performing the transfer process in inert atmosphere are possible routes. Interlayer coupling in vertical van der Waals heterostructures has been investigated by means of optical spectroscopy with reflectance [181], photoluminescence [180, 182184], electroluminescence [59], and Raman experiments [167], by measurements of the photovoltaic effect [58, 61] or by scanning tunnelling spectroscopy [185, 186].

Figure 9.

Figure 9. (a) Cartoon depicting a vertical van der Waal heterostructure consisting of monolayers of MoS2 and WSe2 with a photoexcited interlayer exciton (IX) with the electron localized in MoS2 and the hole localized in the WSe2 layer. (b) Scheme of the type-II band alignment of valence and conduction bands at the K- and K'-points of stacked MoS2 and WSe2 monolayers. The process of photoexcitation of charge carriers shows charge carrier relaxation across the heterostructure.

Standard image High-resolution image

Figure 9(b) schematically depicts a vertical van der Waals heterostructure built from MoS2 and WSe2 monolayers with type-II band alignment. The figure also sketches the process for the separation of photoexcited charge carriers across a heterojunction, the formation of interlayer excitons and radiative re-emission. Incoming photons are absorbed in both monolayers, and electrons are excited from the valence band to the conduction band. At the heterostructure-interface the electrons are transferred to the favorable lower potential conduction band in MoS2, while the holes are transferred to the favorable higher potential valence band in WSe2. The charge dynamics cause a separation of the e–h pairs into different layers. If the Coulomb interaction between the electrons and holes in the different layers is strong enough, bound bosonic quasiparticles called interlayer excitons (IXs) can form. The recombination of such an IX occurs under emission of a photon with characteristic energy that is different from the energy of an exciton stemming from the two individual layers. The band alignment of conduction band minima and valence band maxima versus the vacuum level at the K- and K'-points of the direct gap monolayers of MoS2, WSe2, MoSe2 and WS2 are summarized in figure 10. It is obvious that for any possible combination of two different monolayers, a type-II band heterojunction forms with the above-described mechanism for e–h separation, whereas the electrons and holes can be either electrically injected or photoexcited. By application of an in-plane electric field induced by the heterojunction, the charge carriers can either be forced away towards the contacts, as required for photovoltaic applications [61], or pushed towards the heterostructures' interface, as needed for light-emitting diodes.

Figure 10.

Figure 10. Work function and band alignment of valence and conduction band edges at the K- and K'-points for MoS2, WSe2, MoSe2 and WS2 monolayers (values taken from [162]). All possible combinations of vertically stacked van der Waals hetero-bilayers result in a type-II band alignement.

Standard image High-resolution image

Figure 11(a) shows an optical micrograph of a typical van der Waals heterostructure stacked from micromechanically exfoliated MoS2 and WSe2 layers, which were fabricated by the so-called dry transfer method using a PDMS stamp [110]. It has been proven by Raman investigations (figure 11(b)) that the several µm large area of the heterostructure is prepared from monolayers. Figure 11(c) depicts the emission spectra from a MoS2 monolayer, a WSe2 monolayer and the heterostructure region at room temperature. The weaker emission from MoS2 occurs at an energy of 1.81 eV. The emission of WSe2 at 1.64 eV exhibits a redshifted shoulder that is assigned to the emission of a charged exciton (trion) [25]. The emission from the heterostructure is further redshifted with a peak energy of 1.58 eV in agreement with earlier reports [180, 182]. We emphasize that the redshifted emission is constant over the whole heterostructure area on the sample. Compared to the emission from WSe2, the intensity from the heterostructure is reduced by a factor of six. In particular, the reduced emission energy is a strong hint that the coupling between the two TMD monolayers is strong enough to form a type-II heterostructure causing an effective e–h separation and the formation of interlayer excitons even at room temperature. Hereby, TMD heterostructures represent an intriguing possibility to study excitonic phenomena compared to conventional heterostructures, e.g. in GaAs with a much smaller exciton binding energy of only a few meV [187].

Figure 11.

Figure 11. (a) False-color representation of an optical micrograph of a vertically stacked MoS2/WSe2 heterostructure prepared by dry transfer of micro-mechanically exfoliated flakes. The region of the hetero-bilayer area consisting of a monolayer MoS2 and a monolayer WSe2 is marked by a blacked dashed line. The positions for optical measurements are indicated in black (hetero-bilayer), green (MoS2 monolayer) and red (WSe2 monolayer). (b) Raman spectra from a MoS2/WSe2 heterostructure (black), MoS2 monolayer (green) and WSe2 (red) monolayer displaying the Raman active high-frequency E' and $A_{1}^{'}$ phonon modes with suitable energies for the related monolayers (excited at room temperatures in vacuum with λLaser  =  488 nm). (c) Photoluminescence (PL) spectra from a MoS2/WSe2 heterostructure (black), MoS2 monolayer (green) and WSe2 (red). The red-shifted photoluminescence peak from the heterostructure is a strong hint for the formation of interlayer excitons. (Raman and PL spectra are taken at room temperatures in vacuum with λLaser  =  488 nm.)

Standard image High-resolution image

The IX excitons in vertical van der Waals heterostructures are long-lived with a lifetime of up to τ  =  1.8 ns (for MoSe2–WSe2 heterostructures at T  =  20 K) [183]. This lifetime is orders of magnitude longer than that of the direct excitons in the individual SC-TMDs monolayers that are in the range of tens of picoseconds [188, 189]. The increased lifetime can be explained by a reduced spatial overlap of the wave functions from electron and holes residing in different layers. Therefore, the interlayer excitons are of spatially indirect nature with reduced optical dipole moments similar to those created, e.g. in conventional GaAs-based double quantum well structures with a perpendicular applied electric field [190]. It has been shown by Rivera et al [183] that also for van der Waals heterostructures the IX emission energy can be controlled by the application of an electric field using top and bottom gate electrodes. Since in the device in [183], one electrode was in direct contact with the heterostructure and not separated by a dielectric layer, the impact of the electric field and doping cannot be separated [183]. To study the isolated influence of an electric field on the emission, a capacitor device with both gate electrodes electrically isolated from the hetrostructure is required similar to those used by Klein et al for their Stark effect spectroscopy studies of MoS2 [77].

Long-lived excitons are of great interest for the study of correlation and quantum phenomena in dense ensembles of bosonic quasiparticles [62, 191]. Such a multi-exciton system exhibits a rich phase diagram depending on temperature and quasiparticle density with different phases from a classical exciton gas to a degenerate bosonic exciton gas that can condense to a Bose–Einstein condensate (BEC) by further cooling [62, 191]. A BEC is a coherent state of bosons in the same ground state. A BEC of excitons in solid state systems is the subject of intense research thus far concentrated on low-dimensional heterostructures embedded in conventional 3D crystals studied below 4 K [191194]. It has been predicted by numerical investigations together with scaling arguments that hetero-bilayers prepared from SC-TMDs with an insulating interlayer prepared from atomically thin hBN sheets could result in a degenerate Bose gas with a superfluid phase with a record high temperature approaching 100 K [62].

This underlines that SC-TMD heterostructures are not only highly interesting and of significant potential from the application point of view but also for the study of correlation and many-body quantum phenomena in 2D systems.

6. Conclusion and perspectives

Atomically thin crystalline membranes are promising as a platform for novel technologies and also fundamental studies. These materials, with manifold electronic, optical, mechanical and thermal properties, are inherently fascinating due to the coincidence of dimensionality in real and momentum space—they are truly 2D membranes. Due to the absence of dangling bonds at the interface, 2D materials can be combined to vertical van der Waals heterostructures and integrated in 3D hybrid networks without limitation. The direct access to the 2D surface enables interfacial engineering, environmental sensing and a high degree of tunability. The integration of 2D materials is not restricted to solid state materials. The flexibility of most 2D materials enables a combination with bendable and foldable materials such as polymers, foils or even fabrics. The materials in the focus of this review are SC-TMDs with a direct band gap in the monolayer limit in the visible range. The materials feature high sunlight absorption of up to 15% for a single monolayer, high stability, catalytic activity, and a decent charge carrier mobility. They are stretchable and bendable. Most of the fundamental studies on SC-TMDs are still carried out on micro-mechanically exfoliated individual flakes. However, materials scientists are progressing quickly in developing routes for large-scale production. Investigations of the higher-frequency Raman active phonon modes in SC-TMDs already provide access to a variety of structural and electronic properties such as defects, strain, number of layers, doping and temperature. The impressive scattering cross-section for only a single crystal layer makes inelastic light-scattering experiments a versatile, non-destructive, contactless tool for advanced characterization without the need to prepare electric contacts. Also, the excitonic properties of the materials are superior. Stacked as van der Waals heterostructures, a heterojunction with type-II band alignment arises that is required for several applications including photovoltaics and effective photocatalysis, but the stacking is also suitable to generate p–n junctions for light-emitting diodes. Moreover, long-lived interlayer excitons can be photoexcited with the charge carriers separated in different layers. A reduced spatial overlap of the wave functions increases the IX lifetime and enables the investigation of correlated phenomena and many-body quantum effects in a dense exciton ensemble. A peculiar example is the condensation of the bosonic quasiparticle to a coherent phase that is predicted to occur at record high temperatures in SC-TMD heterostructures [183].

With all these 'superlatives', it remains thrilling to learn about the first real device made out of SC-TMD on the market that can be purchased, different from the usual application as dry lubricant. Such a SC-TMD device must either be much cheaper in production than existing products, the device characteristics must significantly exceed those of existing ones, or such a device must offer a completely new functionality. For these reasons, we surmise that it will not be an electronic device that first enters the market, because SC-TMD-based FETs have comparable key characteristics to silicon-based devices, although the silicon technology is already highly industrialized. On the basis of the above considerations, the first devices might be in the area of (photoactivated) bio-sensors, gas sensors, bio-medical applications, solar energy conversion or photocatalysis. Also, on the fundamental side, SC-TMDs harbour numerous exciting properties that are worth exploring further, such as doping-induced superconductivity in less than 1 nm thin crystals, many-body interactions in exciton ensembles, spin and valley properties and their control, interaction, e.g. in hybrid devices, defect engineering and the control of single-photon emitters, integration of SC-TMDs with strong light–matter interaction in photonic and plasmonic circuitry, just to mention some possible directions. There will be an exciting future in this highly active and interdisciplinary field by studying optically active SC-TMDs and other 2D materials.

Acknowledgments

We acknowledge financial support by Deutsche Forschungsgemeinschaft (DFG) via excellence cluster 'Nanosystems Initiative Munich' (NIM), through the TUM International Graduate School of Science and Engineering (IGSSE) and DFG projects WU 637/4-1, HO3324/9-1 and BaCaTeC.

Please wait… references are loading.
10.1088/1361-6463/aa5f81