Layer-number-dependent performance of metal–semiconductor junctions (MSJs) with multilayered two-dimensional (2D) semiconductors has attracted increasing attention for their potential in ultrathin electronics and optoelectronics. However, the mechanism of the interaction and the resulting charge transfer/redistribution at the two kinds of interfaces in MSJ with multilayered 2D semiconductors, namely, the metal–semiconductor (M–S) and the semiconductor–semiconductor (S–S) interfaces, have not been well understood until now, although that is important for the overall Schottky barrier height and the energy-band-offset between different layers of the 2D semiconductors. Here, based on state-of-the-art density functional theory calculations, the mechanisms of bonding and asymmetric electron redistribution at the M–S and S–S interfaces of metal–bilayer MoS2 junctions are revealed. Multiple mechanisms collectively contribute to the electron redistribution at the two kinds of interfaces, and the dominant mechanism depends on both the dimensionality (2D vs 3D) and the work function of metal electrodes. For the M–S interface, the pushback effect and metal-induced gap states play a dominant role for MSJs with 3D metal, while the covalent-like quasi-bonding feature appears for MSJs with medium-work-function 2D metals, and charge transfer plays a main role for MSJs with 2D metals that have very large or small work functions. For the S–S interface, it inherits the electron-redistribution behavior of the M–S interface for MSJs with 2D metal, while opposite electron-redistribution appears in MSJs with 3D metal. These mechanisms provide general insights and new concepts to better understand and use MSJs with multilayered 2D semiconductors.

The potential applications of two-dimensional (2D) semiconducting transition-metal dichalcogenides (TMDs) in electronics1–5 and optoelectronics6–8 have stimulated intensive research of their fundamental properties as well as the metal–semiconductor junctions (MSJs) based on them. MoS2, as the representative TMDs, shows strong optical absorption9,10 and high on/off current ratio.11 However, in contacting MoS2 to metal electrodes, there is an unexpectedly high electrical contact resistance, which reduces the field-effect mobility of charge carriers across the interface of MSJ and degrades device performance.12 The non-ohmic behavior at metal–MoS2 interfaces indicates the existence of significant Schottky barrier height (SBH), which depends on the energy level alignment between isolated MoS2 and metal in the ideal Schottky–Mott limit.13 Realistically, there is a serious deviation from the Schottky–Mott limit especially for MoS2 contacts to 3D metal electrodes due to the strong Fermi-level pinning (FLP) caused by the interface dipole formed due to the asymmetric charge redistribution at the metal–MoS2 interface, even for a high-quality interface without defects/disorder.14 The charge density redistribution at the metal–MoS2 interface is collectively contributed by several factors, including (quasi-)chemical interaction, metal-induced gap states (MIGS), pushback effect that pushes electrons back to metal, and charge transfer induced by the tendency of MoS2 band edges aligning with the metal Fermi-level.15 

Recently, the study of metal–multilayered 2D semiconductor junctions and their thickness-dependent performance has attracted increasing attention.16–20 Layer-number engineering of 2D semiconductors offers a new controlling factor to tune device-performance. For example, the electric current and photocurrent varying as a function of the layer-number in devices based on multilayered 2D semiconductors were reported,16 and we proposed type II band alignment in 2D multilayered semiconductor homojunctions by supporting them on metals.19 It is worth noting that layer-number engineering plays a leading role only for few-layer 2D semiconductors, especially for less than four layers.21 Therefore, the mechanisms of layer-number-dependence of MSJs with multilayered MoS2, which contains the metal–MoS2 and MoS2–MoS2 interfaces, deserve to be well understood.

In the current work, by first-principles density functional theory (DFT) calculations, we probe the mechanisms of asymmetric electron redistribution at metal–MoS2 (M–S) and MoS2–MoS2 (S–S) interfaces in MSJ with bilayer H–MoS2. The electron density redistribution determines both the interface dipole at M–S interfaces and the band offset between 2D semiconductor layers, which collectively determine the Schottky barrier height of metal–bilayer MoS2 junctions. To give the dominant mechanisms for different types of MSJ, we perform a systematic study of monolayer (ML) and bilayer (BL) MoS2 contact to 3D and 2D metals with metal work functions adjustable in a wide range. At the M–S interfaces, the pushback effect and MIGS dominate in 3D metal–MoS2 junctions, while in 2D metal–MoS2 junctions, significant charge transfers across the interface for MSJs with 2D metals that have extremely large or small work functions, and the “covalent-like quasi-bonding” is formed for MSJs with medium-work-function 2D metals. At the S–S interfaces, the electron density redistribution is closely related to the M–S interface (i.e., change sign or not between the two kinds of interfaces) that depends on the dimensionality (3D vs 2D) of metals, and the resulting band offset generates a type-II band alignment between MoS2 layers19 with different mechanisms for 2D vs 3D metal contacts. General insights and new concepts are developed for MSJs with bilayer MoS2, and the importance of semiconductor–semiconductor interlayer interaction is highlighted.

First-principles calculations were carried out using density functional theory as implemented in the Vienna ab initio Simulation Package (VASP) with periodic boundary conditions.22–24 The projected augmented wave (PAW) method25 was adopted to describe the core electrons. The valence electrons were described by plane waves with a cut-off kinetic energy of 500 eV. All the calculations were carried out using the newly developed strongly constrained and appropriately normed (SCAN) meta-generalized gradient approximation (meta-GGA) density functional26 with rVV10 (the revised Vydrov–van Voorhis nonlocal correlation functional) for van der Waals (vdW) correction.27 The SCAN + rVV10 method have offered a good performance for layered-materials compared with experimental results28 and especially appropriate for metal–semiconductor contacts19 (see Table SI in the supplementary material for more details). The atomic positions were fully relaxed until the force on each atom was less than 0.01 eV Å−1. The energy convergence value between two consecutive self-consistent steps was set as 10−4 eV. The Brillouin zone (BZ) was sampled in the Monkhorst–Pack scheme with a K-point grid spacing of 0.01 Å−1. A vacuum region of 15 Å is added and a dipole correction is applied to avoid spurious interactions between periodic images of the slab.29 

The optimized lattice constant of monolayer (ML) MoS2 is 3.16 Å, which agrees with theoretical and experimental results in the literature.27,30,31 The geometry structures of MSJs with Mo2C(OH)2–MoS2 are shown as an example in Fig. S1 and the supercells used to model the metal–MoS2 junctions are listed in Table SII in the supplementary material. Using these supercells, the in-plane lattices mismatch between metal and MoS2 is in the range of 2.5%–4.6%. Since the property of MoS2 is sensitive to strain, we strain the metal lattice to adapt the MoS2 lattice, as done in this way in the literature.32–34 The electron affinity energy (EAE) and ionization energy (IE) of isolated ML and bilayer (BL) MoS2 are shown in Fig. 1. The EAE is about 4.3 eV and almost unchanged from ML to BL, while the IE decreases greatly from 6.20 eV to 5.88 eV. When contacting MoS2 with frequently used 3D metal electrodes, a sizable n-type Schottky barrier (SB) is often formed due to the strong FLP at the M–S interfaces.32,34,35 With modulated metal work functions and dangling-bond-free surfaces, the 2D metallic MXenes contact to MoS2 can get vanishing n-type and p-type SBH.36,37 High-quality MXenes Mo2CT2 (T = O, OH, or F is the surface terminating group) acting as 2D metal electrodes have been reported both in experimental38–40 and theoretical works.41 The work functions of 3D and 2D metals are shown in the middle panels of Fig. 1, with the geometric structures of 2D Mo2CT2 shown in the right. The work functions of MXenes are mainly determined by the different functional groups, which produced surface dipoles with different magnitudes and signs:41 with O-termination, it has the largest work function of 7.79 eV and with OH-termination, the smallest work function of 2.12 eV is given, whereas F-termination exhibits a medium work function of 4.81 eV, which locates within the bandgap of MoS2 (Fig. 1). The work function of 3D metals (Ag and Pt) lies within the bandgap of monolayer MoS2, either close to the EAE or to IE (Fig. 1).

FIG. 1.

Band alignment between isolated semiconductors and metals. Energy band diagram of monolayer (ML) and bilayer (BL) MoS2 and work functions (WM) of several 3D and 2D metals; the geometric structures of 2D MXenes Mo2CT2 (T = OH, F, O) are shown on the right. EAE and IE represent electron affinity and ionization energies.

FIG. 1.

Band alignment between isolated semiconductors and metals. Energy band diagram of monolayer (ML) and bilayer (BL) MoS2 and work functions (WM) of several 3D and 2D metals; the geometric structures of 2D MXenes Mo2CT2 (T = OH, F, O) are shown on the right. EAE and IE represent electron affinity and ionization energies.

Close modal

In the MSJs with ML MoS2, the equilibrium distances (D) between metal and MoS2 (defined as the separation between the metal surface layer and the S atom layer of MoS2 close to the metal) are listed in Table I, which fall in the range of typical vdW interaction for 2D metal Mo2CF2 and Mo2CO2 and stronger than typical vdW interaction for 2D metal Mo2C(OH)2 and 3D metal Ag and Pt (note that the differences in atomic radius of metal surface atoms should also be considered). The typical vdW interaction character (or stronger than typical vdW interaction) is further confirmed by binding energies in Table I. Here, the binding energy (Eb) between metals and MoS2 is defined as Eb=EtotEMEMoS2/A, where Etot, EM, and EMoS2 are the total energies of metal–MoS2, isolated metal, and MoS2, respectively, and A is the contact area in the supercell. The calculated Eb values are in the range of −0.26 J/m2 to −0.37 J/m2 for 2D metals Mo2CF2 and Mo2CO2, which suggests that MoS2 is physisorbed on these 2D metals, while the Eb are −0.62 J/m2 and −0.96 J/m2 for Ag and Pt, which are stronger than pure vdW interactions. The stronger Eb for Mo2C(OH)2–ML MoS2 (−0.82 J/m2 or corresponds to −0.44 eV) is due to the synergistic effect of vdW interaction (typical vdW interstratification is −0.3 J/m2 or corresponds to −0.16 eV) and hydrogen bonding (here O–H⋯S, the typical strength is −0.26 eV).42,43 The “bond length” of hydrogen bonding O–H⋯S in Mo2C(OH)2–MoS2 is 2.67 Å, which is the competition between typical hydrogen bonding (the typical bond length of O–H⋯S is about 2.33 Å42) and van der Waals interaction (the sum of vdW radii of H and S is 2.95 Å).43 Although the binding energies in MSJ with 3D metal or 2D metallic Mo2C(OH)2 are stronger than typical van der Waals interaction, they are far from chemical bonding, and all the systems we studied belong to weak interactions (or mainly physisorption).

TABLE I.

Equilibrium distance (D) between metal and MoS2 and binding energy (Eb) of 2D and 3D metal–ML MoS2 junctions.

MetalsD (Å)Eb (J/m2)
2D Mo2CO2 2.82 -0.37 
2D Mo2CF2 3.12 -0.26 
2D Mo2C(OH)2 1.98 -0.82 
3D Pt(111) 2.51 -0.96 
3D Ag(111) 2.79 -0.62 
MetalsD (Å)Eb (J/m2)
2D Mo2CO2 2.82 -0.37 
2D Mo2CF2 3.12 -0.26 
2D Mo2C(OH)2 1.98 -0.82 
3D Pt(111) 2.51 -0.96 
3D Ag(111) 2.79 -0.62 

It is noticeable here that the strong FLP at the interfaces of 3D metals (e.g., Ag and Pt) and MoS2 becomes very weak at the interfaces of 2D metallic MXene–MoS2. The surfaces of both 2D MoS2 and MXene are dangling-bond-free, and hence, the weak MXene–MoS2 interactions suppress the formation of MIGS (Figs. S3 and S4 in the supplementary material and Fig. 5 below). The vdW interaction at 2D MXene–MoS2 interfaces makes the electron redistribution between them simple relative to that at the 3D metal–MoS2 interfaces. For 3D metal contacts, the interaction at the M–S interface is much more complicated, and hence, multiple mechanisms work together to determine the asymmetric electron density redistribution at the interface, which will be detailed below.

In the Schottky–Mott limit, the Schottky barrier height (SBH) ΦB is obtained by band alignment of the non-interacting subsystems (as in Fig. 1),

(1)

where ΦBe and ΦBh are the SBH for electrons (n-type) and holes (p-type), WM is the work function of the metal, EAE and IE are the electron affinity and ionization energies of the semiconductor, respectively (IEEAE = bandgap). In the Schottky–Mott limit, ΦB linearly depends on the metal work function WM and changes in the same amounts as WM does. In many cases, however, even for high-quality interface without defects or disorder in weakly interacting systems, the Schottky–Mott limit may not be obeyed and the hole and electron injection barriers are different from those in Eq. (1). The origin of the differences for MSJs with 2D semiconductors can be attributed mainly to the existence of an interface dipole, ΔV,32,33 which changes the SBHs to

(2)

The ΔV is caused by the electron density rearrangement at the M–S interface of MSJs. The electron density difference, Δρ(x, y, z), is an effective tool to analyze the bonding at metal–MoS2 interfaces, which is defined as the difference of the electron density distribution of the composite full system and the isolated subsystems,

(3)

where ρtot, ρM, and ρMoS2 are the electron density of the metal–MoS2 junction, the metal, and the free-standing MoS2, respectively. The Δρ mainly localized around the metal–MoS2 interface, which includes the contributions from the energy level shift induced by charge transfer, the pushback effect, and the energy level broadening induced by metal interaction, which collectively changes the SBH at the metal–semiconductor interface through interface dipole, ΔV. The ΔV (which can be also described as electrostatic potential change) satisfy the Poisson equation below as a function of Δρ.44–46 By solving the Poisson equation, a potential step across the interface is achieved,

(4)

in which z is the distance from the middle of the M–S interface and A is the contact area. In density functional theory (DFT) calculation with a slab model, ΔV can be alternately obtained as the difference between the asymptotic values of the potential difference in the vacuum with an expression

(5)

where WM/WM/S are the work function of the clean metal surface/the metal surface covered by MoS2.

The plane-averaged electron density differences along the z-direction perpendicular to the M–S interfaces, Δρ(z), are shown in Fig. 2. The metal–ML MoS2 junctions are used for the analysis of the M–S interface of metal–bilayer MoS2 junctions, since we find that in metal–BL MoS2 junctions, the behavior at the M–S interface largely unchanged relative to that in metal–ML MoS2 (also refer to Fig. 7 below).

FIG. 2.

Electron density redistribution Δρ(z) and Δρ(x, y, z) at the metal–ML MoS2 interface. Δρ(z) is the plane-averaged electron density difference along the z-direction perpendicular to the M–S interfaces. Red (blue) region represents electron accumulation (depletion). The region of the M–S interface is indicated by two dotted black lines representing the surface-atom-layers of metal and MoS2. The insets show the corresponding electron density differences Δρ(x, y, z) in 3D (side view); to ensure that each picture shows clearly the 3D electron density difference, different isovalues are used, which are 10−4, 3 × 10−3, 10−3, and 7×10−4 |e|/bohr3 in [(a)–(d)], respectively.

FIG. 2.

Electron density redistribution Δρ(z) and Δρ(x, y, z) at the metal–ML MoS2 interface. Δρ(z) is the plane-averaged electron density difference along the z-direction perpendicular to the M–S interfaces. Red (blue) region represents electron accumulation (depletion). The region of the M–S interface is indicated by two dotted black lines representing the surface-atom-layers of metal and MoS2. The insets show the corresponding electron density differences Δρ(x, y, z) in 3D (side view); to ensure that each picture shows clearly the 3D electron density difference, different isovalues are used, which are 10−4, 3 × 10−3, 10−3, and 7×10−4 |e|/bohr3 in [(a)–(d)], respectively.

Close modal

For the 3D metal Pt–ML MoS2 junction [Fig. 2(d)], the electron density distribution at the interface is very complicated. The pushback effect,32,47 which is observed in the physisorption of molecules or 2D layers on metals, plays a main role at the Pt–ML MoS2 interface. The reason for the pushback effect is the antisymmetrization of the product of the 3D metal and 2D semiconductor wave functions.48 When 2D monolayer MoS2 is physisorbed onto the metal surface, the two wave functions overlap (but no rehybridization as in chemical bonding). In order to reduce the Pauli repulsion at the interface overlap region, a rearrangement of the electron density occurs. Since the wave function of transition metal Pt is more extended and deformable than that of MoS2, the electrons are pushed back into Pt, leading to the phenomenon that electrons accumulate at the Pt side and deplete at the MoS2 side of the M–S interface. As a result, the metal work function is effectively lowered with a negative ΔV for the Pt(111) surface absorbed with MoS2, as illustrated in Fig. 3(d) [also see Eq. (5)]. The pushback effect also occurs at the Ag–MoS2 interface (Figs. S5 and S6), and the other factors of electron redistribution at the 3D metal–MoS2 interface are detailed in Fig. S5 and the related text.

FIG. 3.

Schematic illustration of band alignment at the interface of Mo2C(OH)2–ML MoS2 (a), Mo2CO2–ML MoS2 (b), Mo2CF2–ML MoS2 (c), and Pt(111)–ML MoS2 (d). EF represents the Fermi level of metal.

FIG. 3.

Schematic illustration of band alignment at the interface of Mo2C(OH)2–ML MoS2 (a), Mo2CO2–ML MoS2 (b), Mo2CF2–ML MoS2 (c), and Pt(111)–ML MoS2 (d). EF represents the Fermi level of metal.

Close modal

The charge transfer is induced by the alignment of MoS2 band edges and 2D metal Fermi levels, which takes place at the physisorbed interface of Mo2C(OH)2–MoS2 and Mo2CO2–MoS2. The Fermi level of Mo2C(OH)2 is higher than the EAE of MoS2, while that of Mo2CO2 is lower than IE of MoS2 (Fig. 1). In the case of Mo2C(OH)2–MoS2 [Fig. 2(a)], the electron is transferred from the 2D metal to MoS2, and MoS2 becomes negatively charged, which is further confirmed by Bader charge analysis,49 which shows that MoS2 obtain 0.11 electrons. The charge transfer creates a positive ΔV at the M–S interface that increases the vacuum level, as shown in Fig. 3(a). The electrons continue transfer until the conduction-band minimum (CBM) level of MoS2 aligns close to the Fermi level of the 2D metal. Hence, an n-type vanishing SBH is obtained [Figs. 3(a) and S3(a)], and then, MoS2 is degenerately doped [Fig. 5(a)]. In Fig. 2(b), on the contrary to Fig. 2(a), electrons transfer from MoS2 to the 2D metal Mo2CO2. Bader charge analysis shows that MoS2 lose 0.34 electrons due to the Mo2CO2 Fermi level that is below the IE of MoS2. This charge transfer creates a negative ΔV at the interface and reduces the vacuum level. Therefore, a p-type vanishing SBH is obtained [Figs. 3(b) and S3(b)], and then, MoS2 is degenerately hole doped [Fig. 5(b)].

For the Mo2CF2–MoS2 interface, since the Fermi level of Mo2CF2 is situated in-between EAE and IE of MoS2, there is nearly no net charge transfer across the interface and hence no interface dipole formation, and the vacuum levels of the Mo2CF2 and MoS2 remain aligned [Fig. 3(c)] as that in the isolated subsystems (Fig. 1). Interestingly, we found that at the Mo2CF2–MoS2 interface, electrons accumulate at the middle of the interface region and deplete at both the fluorine and the sulfur surface-layers [Fig. 2(c)]. This phenomenon between vdW 2D systems is called the formation of “covalent-like quasi-bonding,” which is similar to the literature reported electronic interlayer-hybridization in bilayer PtS2 and in few-layered black phosphorus.50 The electron aggregation region is a bit close to fluorine than sulfur since fluorine is more electronegative than sulfur. The polarized “covalent-like quasi-bonding” in our system occurs under the following conditions: the interaction between the two layers of different 2D materials is stronger than pure vdW interaction but much weaker than chemical interaction, and the electronegativity of the atomic layers at the left and right sides of the interface is similar to each other. The covalent-like quasi-bonding at the Mo2CF2–MoS2 interface has no net charge transfer and hence will not create interface dipole. Hence, the SBH in Mo2CF2–MoS2 junction follows Eq. (1) (the Schottky–Mott limit). It is interesting to note that here, we have generalized the concept of “covalent-like quasi-bonding” between the same 2D semiconductor50 and polarized “covalent-like quasi-bonding” between two different 2D materials, and the 2D material is not limited to semiconductors but can also to metals in our case.

A comparison of the mechanisms at the M–S interfaces with 3D metal vs 2D metal: the pushback effect dominates in 3D metal–MoS2 junctions while is not apparent in 2D metal–MoS2 junctions, since 2D metal Mo2CT2 have a similar electron-accepting capability to that of 2D MoS2, namely, the electrons spilled out of the surface are symmetrically pushed back to 2D metal and 2D MoS2 for the 2D metal–MoS2 junctions.

Based on all the above analyses, one can analyze the mechanism of electron density redistribution Δρ at the M–S interface and can qualitatively understand the resulted interface dipole ΔV. However, one may argue that ΔV is directly related to zΔρ(z) but not simply Δρ. Hence, in the following, we analyze zΔρ(z) and show that the conclusion simply from Δρ holds. The integration of zΔρ(z) can be derived from Eq. (4),

(6)

Figure 4 shows the zΔρ(z) plots around the M–S interfaces. In Fig. 4(a) for Mo2C(OH)2–MoS2, the positive zΔρ(z) region is much more than the negative region, which gives a positive ΔV and is consistent with the analysis of Δρ(z) only in Fig. 2. For Mo2CO2–MoS2, the negative zΔρ(z) region is much larger than the positive one, which gives a negative ΔV. For Mo2CF2–MoS2 [Fig. 4(c)], the positive and negative regions are comparable, which give rise to a zero ΔV. In Pt(111)–MoS2, the distribution of zΔρ(z) is complicated, and therefore, a semi-quantitative analysis from the integral of ∫zΔρ(z)dz is needed and has been carried out. The values of the integral for the junctions in Figs. 4(a)–4(d) are 0.14, −0.08, 0, and −0.05, respectively. The −0.05 for Pt(111)–MoS2 means the formation of negative ΔV. In summary, the analysis of Δρ(z) in Figs. 2 and zΔρ(z) in Fig. 4 gives the same conclusion for ΔV.

FIG. 4.

The zΔρ(z) plots around the M–S interface in Mo2C(OH)2–ML MoS2 (a), Mo2CO2–ML MoS2 (b), Mo2CF2–ML MoS2 (c), and Pt(111)–ML MoS2 (d).

FIG. 4.

The zΔρ(z) plots around the M–S interface in Mo2C(OH)2–ML MoS2 (a), Mo2CO2–ML MoS2 (b), Mo2CF2–ML MoS2 (c), and Pt(111)–ML MoS2 (d).

Close modal

The metal-induced gap states (MIGS)21,33 also play a significant role in 3D metal–MoS2 junctions in forming ΔV51 in addition to the above discussed pushback effect in 3D metal–MoS2 junctions. The partial density of states (PDOS) of Mo2CT2–MoS2 and Pt(111)–MoS2 are displayed in Fig. 5 to explore whether the MIGS in MoS2 can be induced by 2D and 3D metals. In Fig. 5(d) for Pt(111)–ML MoS2, a distinct overlap between the Pt 5d, Mo 4d, and S 3p orbitals can be seen around the Fermi energy, as indicated by the arrow, suggesting orbital hybridization upon interface formation,34 and MoS2 becomes partially metallized in contacts to Pt(111). Note that in the Mo2CF2–MoS2 junction [Fig. 5(c)], the MoS2 layer remains semiconducting with a bandgap. In Figs. 5(a) and 5(b), the bandgaps exist but the Fermi level locates above the conduction band edge of MoS2 for Mo2C(OH)2–MoS2 (locates below the valence band edge of MoS2 for Mo2CO2–MoS2). Hence, MoS2 is degenerately electron doped or degenerately hole doped in these two junctions.52 

FIG. 5.

Partial DOS of Mo2C(OH)2–ML MoS2 (a), Mo2CO2–ML MoS2 (b), Mo2CF2–ML MoS2 (c), and Pt(111)–ML MoS2 (d). The arrow in panel (d) at Fermi level indicates the metal-induced gap states in MoS2.

FIG. 5.

Partial DOS of Mo2C(OH)2–ML MoS2 (a), Mo2CO2–ML MoS2 (b), Mo2CF2–ML MoS2 (c), and Pt(111)–ML MoS2 (d). The arrow in panel (d) at Fermi level indicates the metal-induced gap states in MoS2.

Close modal

The interface dipole ΔV modifies the SBH [by Eq. (2)] and deviates from the Schottky–Mott limit [Eq. (1)]. For 2D metal contacts with very large/small work functions, the interface dipole ΔV is significant [Figs. 3(a) and 3(b)]; if the Fermi level of 2D metal is in-between EAE and IE, the ΔV is negligible [Fig. 3(c)], and the SBH can be well described by the Schottky–Mott limit [Eq. (1)]. However, for 3D metal contacts, even their work functions are located in-between EAE and IE of ML MoS2 (such as Ag, Au, Cu, and Pt),32 and a significant ΔV appear due to the pushback effect at the 3D metal–MoS2 interface.

The SBHs from the projected band structure (or the PDOS)32,35,53,54 calculated by DFT, for the MoS2–metal junctions, can be defined as

(7)

in which EF is the Fermi level of the junction, χMoS2 is the EAE of MoS2 in the MoS2–metal junction [which is affected by the metal, but not the EAE of isolated MoS2 as used in Eqs. (1) and (2) and in Fig. 1], and Eg is the bandgap of MoS2 in the junction. The projected band structures of the ML MoS2 supported on 2D metals Mo2CT2 and 3D metal Pt are shown in Fig. S3, and the SBHs from the DFT projected band structures are shown in Fig. 6, along with the SBH from Eqs. (1) and (2). The SBHs from DFT projected band structures are close to the ones from Eq. (2), which includes the effects of ΔV beyond the Schottky–Mott limit [Eq. (1)]. ΔV has negligible effects on SBH in the Mo2CF2–MoS2 junction, since there is no MIGS and no charge transfer at its M–S interface. For all the other studied junctions, the interface dipole ΔV has a great effect on SBH, which largely makes up the difference between the result from DFT and from Eq. (1). It is worth noting that for Ag, the ΔV is small since the interaction of Ag with MoS2 is weaker than that of Pt (Table I) due to the fully filled d bands of Ag.

FIG. 6.

Comparison of SBHs in metal–ML MoS2 junctions from DFT projected band structures from Eq. (2) (corrected by ΔV) and from Eq. (1) (the Schottky–Mott limit). All are electron SBHs except for Mo2CO2 where the hole SBH is used.

FIG. 6.

Comparison of SBHs in metal–ML MoS2 junctions from DFT projected band structures from Eq. (2) (corrected by ΔV) and from Eq. (1) (the Schottky–Mott limit). All are electron SBHs except for Mo2CO2 where the hole SBH is used.

Close modal

Although ΔV can improve the description of SBH by Eq. (2) (close to the SBH from DFT), there is some deviation between the SBH from Eq. (2) and that from DFT, which may be resulted from other factors beyond Eq. (2), such as the magnitude of bandgap variation due to interaction with metal surface and the hydrogen bonding in Mo2C(OH)2–MoS2. The bandgap of 1.88 eV from the isolated ML MoS2 is used in Eqs. (1) and (2), while the bandgaps vary from 1.55 eV to 1.98 eV for ML MoS2 contacting to 3D and 2D metals. Simultaneously, the work functions of metal also change under strain (see Table SIII in the supplementary material).

In the above, the various mechanisms of electron density redistribution at M–S interfaces have been discussed in metal–ML MoS2 junctions. Comparing the electron density redistributions at the M–S interfaces in metal–BL MoS2 junctions (Fig. 7) with that in metal–ML MoS2 junctions (Fig. 2), one finds that the electron density redistribution is almost identical at the M–S interfaces in both junctions with ML and BL MoS2. Then, in the following, we focus on the electron density redistributions at the S–S interfaces in Fig. 7, which give rise to band offsets between the two semiconducting layers of MoS2 and hence may change the overall SBH in metal–BL MoS2 junctions.

FIG. 7.

Electron density redistribution at the S–S interfaces in metal–BL MoS2 junctions. [(a)–(d)] Plane-averaged electron density difference along the vertical z-direction of the metal–BL MoS2 junctions. Red (blue) regions represent electron accumulation (depletion) regions. The interface region between metal and MoS2 is within the two black dotted lines, while the interface region between the two MoS2 layers is within the two red dotted lines. The insets in [(a)–(d)] illustrate the various band alignments between the two MoS2 layers, where red (green) lines indicate band edges of the first (second) MoS2 layer.

FIG. 7.

Electron density redistribution at the S–S interfaces in metal–BL MoS2 junctions. [(a)–(d)] Plane-averaged electron density difference along the vertical z-direction of the metal–BL MoS2 junctions. Red (blue) regions represent electron accumulation (depletion) regions. The interface region between metal and MoS2 is within the two black dotted lines, while the interface region between the two MoS2 layers is within the two red dotted lines. The insets in [(a)–(d)] illustrate the various band alignments between the two MoS2 layers, where red (green) lines indicate band edges of the first (second) MoS2 layer.

Close modal

For the Mo2C(OH)2–BL MoS2 and Mo2CO2–BL MoS2 junctions as shown in Figs. 7(a) and 7(b), acting as a continuation of the M–S interface, the electron density redistribution at the S–S interfaces in these two junctions is the same as that in their M–S interfaces, although the magnitude decayed considerably. For Mo2C(OH)2–BL MoS2 [Figs. 7(a) and 8(a)], at the S–S interface, electrons accumulate at the second layer MoS2 side, and therefore, the energy band of the second layer moves up, leading to a positive band offset (the band offset ΔCBM = CBM2CBM1 as shown in the inset of Fig. 7(a). On the contrary, for Mo2CO2–BL MoS2 shown in Figs. 7(b) and 8(b), electrons deplete at the second layer MoS2 side and a negative band offset is formed. A perfect type II band alignment is demonstrated in these BL MoS2 homojunctions supported on Mo2C(OH)2 and Mo2CO2 with the CBM and VBM of the bilayer sharply located at the two different layers of MoS2.19 In the Mo2CF2–BL MoS2 junction, the net electron transfer across the M–S interface is negligible, so the electron density redistribution at the S–S interface is also negligible, and hence, the degeneracy in energy of band edges in the two MoS2 layers is kept [Figs. 8(c) and S4(c) in the supplementary material].

FIG. 8.

Schematic illustration of band alignment between the two MoS2 layers in junctions of Mo2C(OH)2–BL MoS2 (a), Mo2CO2–BL MoS2 (b), Mo2CF2–BL MoS2 (c), and Pt(111)–BL MoS2 (d).

FIG. 8.

Schematic illustration of band alignment between the two MoS2 layers in junctions of Mo2C(OH)2–BL MoS2 (a), Mo2CO2–BL MoS2 (b), Mo2CF2–BL MoS2 (c), and Pt(111)–BL MoS2 (d).

Close modal

For the S–S interface in the 3D metal Pt(111)–BL MoS2 junction, a different mechanism appears relative to 2D metal–BL MoS2 junctions. Figures 7(d) and 8(d) show an opposite direction of dipole at the S–S interface relative to that at the M–S interface. As discussed above [Fig. 5(d)], MIGS appears in the first MoS2 layer in 3D metal–MoS2 junction, which means partial metallization of the first-layer of MoS2. The screening effect occurs in the first-layer of MoS2 and it tends to screen the dipole at the metal–MoS2 interface, which is indicated by the arrow in Fig. 7(d). Originated from the screening effect inside the first MoS2 layer, at the S–S interface, the electrons deplete at the first layer side and accumulate at the second layer side [Figs. 7(d) and 8(d)], which creates an positive band offset and acts as the source of the de-pinning effect between MoS2 layers.21 As a result, the band edges of the second layer MoS2 is moved up in Pt(111)–BL MoS2, which leads to the SB-type transition from n-type in Pt(111)–ML MoS2 [Fig. 3(d)] to p-type in Pt(111)–BL MoS2 [Fig. 8(d)].

The electron density redistribution at interfaces in Mo2C(OH)2-trilayer MoS2 is also considered in Fig. S7. The electron density distribution at the Ssecond-layer–Sthird-layer interface is the same as that in the Sfirst-layer–Ssecond-layer interface but with an even smaller magnitude. Therefore, the electron density redistribution of the third layer and thicker-layers can be inferred to be similar with that in the second layer, and the influence on the MSJ tends to be negligible due to the decay of charge redistribution.

A brief summary for this subsection: the electron density redistributions at the S–S interface inherit the behavior at the M–S interface in 2D metal–BL MoS2 junctions, while an opposite electron redistribution occurs at the S–S interface relative to that at the M–S interface in 3D metal–BL MoS2 junctions due to the screening effect of the partially metallized first-layer of MoS2 and can result in a contact-type-transition from n- to p-type for the junctions with Pt(111).

We have revealed the mechanisms of the electron density redistribution at the M–S and S–S interfaces of bilayer MoS2 contact to metal electrodes via physisorption. At the M–S interfaces, the pushback effect plays a major role in contacting to 3D metals; charge transfer or polarized covalent-like quasi-bonding appears in 2D metal–MoS2 junctions as a result of three different types of band alignments from their isolated subsystems. At the S–S interfaces, the electron density redistribution is the same as that at the M–S interfaces for junctions with 2D metals, while an opposite electron density redistribution at the S–S interface from that at the M–S interface is found in junctions with 3D metal. The electron density redistribution at M–S interfaces causes the interface dipole ΔV and ultimately affected the SBH at the M–S interfaces, while the electron density distribution at the S–S interface gives rise to band offset between the semiconducting layers of MoS2, which may result in a contact-type-transition. Our results offered a comprehensive understanding of the mechanisms of the electron density redistribution at the interfaces between bilayer MoS2 and 3D/2D metals with a wide-range of work functions, which paves the way to understand and manipulate the performance of metal-2D multilayered semiconductor junctions for electronics and optoelectronics.

See the supplementary material for the reason of using SCAN + rVV10 functional in our manuscript, a list of supercells and lattice mismatch for metal–MoS2 junctions, and metal work function with strain, figures of the interface potential step and projected band structure of monolayer and bilayer MoS2 supported on metals, the generation mechanism of the interface dipole at 3D metal–MoS2 interfaces and charge distribution at Ag–MoS2 interfaces, electron density redistribution at the MoS2–MoS2 interfaces in Mo2C(OH)2–trilayer MoS2 junctions, and the effect of 3R stacking MoS2.

Q.W., Y.S., and X.S. contributed equally to this work.

The data that support the findings of this study are available within the article and its supplementary material.

This work was financially supported by the Shenzhen Fundamental Research Foundation (Grant No. JCYJ20170817105007999), the Natural Science Foundation of Guangdong Province of China (Grant No. 2017A030310661), and the Guangdong Provincial Key Laboratory for Computational Science and Material Design (Grant No. 2019B030301001). The computational resources were provided by the Center for Computational Science and Engineering of Southern University of Science and Technology.

1.
S.
Kang
,
D.
Lee
,
J.
Kim
,
A.
Capasso
,
H. S.
Kang
,
J.-W.
Park
,
C.-H.
Lee
, and
G.-H.
Lee
,
2D Mater.
7
,
022003
(
2020
).
2.
S.
Aftab
and
J.
Eom
,
2D Mater.
6
,
035005
(
2019
).
3.
W.
Zhu
,
T.
Low
,
H.
Wang
,
P.
Ye
, and
X.
Duan
,
2D Mater.
6
,
032004
(
2019
).
4.
G.-H.
Lee
,
Y.-J.
Yu
,
X.
Cui
,
N.
Petrone
,
C.-H.
Lee
,
M. S.
Choi
,
D.-Y.
Lee
,
C.
Lee
,
W. J.
Yoo
,
K.
Watanabe
,
T.
Taniguchi
,
C.
Nuckolls
,
P.
Kim
, and
J.
Hone
,
ACS Nano
7
(
9
),
7931
7936
(
2013
).
5.
Y.
Yoo
,
Z. P.
Degregorio
, and
J. E.
Johns
,
J. Am. Chem. Soc.
137
(
45
),
14281
14287
(
2015
).
6.
L.
Britnell
,
R. M.
Ribeiro
,
A.
Eckmann
,
R.
Jalil
,
B. D.
Belle
,
A.
Mishchenko
,
Y. J.
Kim
,
R. V.
Gorbachev
,
T.
Georgiou
,
S. V.
Morozov
,
A. N.
Grigorenko
,
A. K.
Geim
,
C.
Casiraghi
,
A. H. C.
Neto
, and
K. S.
Novoselov
,
Science
340
(
6138
),
1311
1314
(
2013
).
7.
Y.
Pan
,
S.
Fölsch
,
Y.-C.
Lin
,
B.
Jariwala
,
J. A.
Robinson
,
Y.
Nie
,
K.
Cho
, and
R. M.
Feenstra
,
2D Mater.
6
,
021001
(
2019
).
8.
F.
Withers
,
O.
Del Pozo-Zamudio
,
A.
Mishchenko
,
A. P.
Rooney
,
A.
Gholinia
,
K.
Watanabe
,
T.
Taniguchi
,
S. J.
Haigh
,
A. K.
Geim
,
A. I.
Tartakovskii
, and
K. S.
Novoselov
,
Nat. Mater.
14
(
3
),
301
306
(
2015
).
9.
K. F.
Mak
,
C.
Lee
,
J.
Hone
,
J.
Shan
, and
T. F.
Heinz
,
Phys. Rev. Lett.
105
(
13
),
136805
(
2010
).
10.
Q. H.
Wang
,
K.
Kalantar-Zadeh
,
A.
Kis
,
J. N.
Coleman
, and
M. S.
Strano
,
Nat. Nanotechnol.
7
(
11
),
699
712
(
2012
).
11.
D.
Dumcenco
,
D.
Ovchinnikov
,
K.
Marinov
,
P.
Lazić
,
M.
Gibertini
,
N.
Marzari
,
O. L.
Sanchez
,
Y.-C.
Kung
,
D.
Krasnozhon
,
M.-W.
Chen
,
S.
Bertolazzi
,
P.
Gillet
,
A.
Fontcuberta i Morral
,
A.
Radenovic
, and
A.
Kis
,
ACS Nano
9
(
4
),
4611
4620
(
2015
).
12.
R. T.
Tung
,
Appl. Phys. Rev.
1
(
1
),
011304
(
2014
).
13.
W.
Shockley
,
Phys. Rev.
56
(
4
),
317
323
(
1939
).
14.
Y.
Liu
,
P.
Stradins
, and
S.-H.
Wei
,
Sci. Adv.
2
(
4
),
1600069
(
2016
).
15.
F.
Flores
,
J.
Ortega
, and
H.
Vázquez
,
Phys. Chem. Chem. Phys.
11
(
39
),
8658
8675
(
2009
).
16.
Y.
Liu
,
J.
Guo
,
E.
Zhu
,
L.
Liao
,
S.-J.
Lee
,
M.
Ding
,
I.
Shakir
,
V.
Gambin
,
Y.
Huang
, and
X.
Duan
,
Nature
557
(
7707
),
696
700
(
2018
).
17.
M.
Zhong
,
Q.
Xia
,
L.
Pan
,
Y.
Liu
,
Y.
Chen
,
H.-X.
Deng
,
J.
Li
, and
Z.
Wei
,
Adv. Funct. Mater.
28
(
43
),
1802581
(
2018
).
18.
G.-S.
Kim
,
S.-H.
Kim
,
J.
Park
,
K. H.
Han
,
J.
Kim
, and
H.-Y.
Yu
,
ACS Nano
12
(
6
),
6292
6300
(
2018
).
19.
Q.
Wang
,
K.
Dou
, and
X.
Shi
,
J. Mater. Chem. C
8
,
959
967
(
2020
).
20.
J.
Ji
and
J. H.
Choi
,
Adv. Mater. Interfaces
6
(
17
),
1900637
(
2019
).
21.
Q.
Wang
,
Y.
Shao
,
P.
Gong
, and
X.
Shi
,
J. Mater. Chem. C
8
,
3113
3119
(
2020
).
22.
G.
Kresse
and
J.
Hafner
,
Phys. Rev. B
47
(
1
),
558
561
(
1993
).
23.
P. E.
Blöchl
,
Phys. Rev. B
50
(
24
),
17953
17979
(
1994
).
24.
G.
Kresse
and
J.
Furthmüller
,
Phys. Rev. B
54
(
16
),
11169
11186
(
1996
).
25.
G.
Kresse
and
D.
Joubert
,
Phys. Rev. B
59
(
3
),
1758
1775
(
1999
).
26.
J.
Sun
,
A.
Ruzsinszky
, and
J. P.
Perdew
,
Phys. Rev. Lett.
115
(
3
),
036402
(
2015
).
27.
H.
Peng
,
Z.-H.
Yang
,
J. P.
Perdew
, and
J.
Sun
,
Phys. Rev. X
6
(
4
),
041005
(
2016
).
28.
H.
Peng
and
J. P.
Perdew
,
Phys. Rev. B
95
(
8
),
081105
(
2017
).
29.
J.
Neugebauer
and
M.
Scheffler
,
Phys. Rev. B
46
(
24
),
16067
16080
(
1992
).
30.
T.
Böker
,
R.
Severin
,
A.
Müller
,
C.
Janowitz
,
R.
Manzke
,
D.
Voß
,
P.
Krüger
,
A.
Mazur
, and
J.
Pollmann
,
Phys. Rev. B
64
(
23
),
235305
(
2001
).
31.
A. A.
Al-Hilli
and
B. L.
Evans
,
J. Cryst. Growth
15
(
2
),
93
101
(
1972
).
32.
M.
Farmanbar
and
G.
Brocks
,
Phys. Rev. B
93
(
8
),
085304
(
2016
).
33.
M.
Farmanbar
and
G.
Brocks
,
Phys. Rev. B
91
(
16
),
161304
(
2015
).
34.
C.
Gong
,
L.
Colombo
,
R. M.
Wallace
, and
K.
Cho
,
Nano Lett.
14
(
4
),
1714
1720
(
2014
).
35.
D. S.
Schulman
,
A. J.
Arnold
, and
S.
Das
,
Chem. Soc. Rev.
47
(
9
),
3037
3058
(
2018
).
36.
J.
You
,
C.
Si
,
J.
Zhou
, and
Z.
Sun
,
J. Phys. Chem. C
123
(
6
),
3719
3726
(
2019
).
37.
L.-Y.
Gan
,
Y.-J.
Zhao
,
D.
Huang
, and
U.
Schwingenschlögl
,
Phys. Rev. B
87
(
24
),
245307
(
2013
).
38.
C.
Xu
,
L.
Wang
,
Z.
Liu
,
L.
Chen
,
J.
Guo
,
N.
Kang
,
X.-L.
Ma
,
H.-M.
Cheng
, and
W.
Ren
,
Nat. Mater.
14
(
11
),
1135
1141
(
2015
).
39.
R.
Meshkian
,
L.-Å.
Näslund
,
J.
Halim
,
J.
Lu
,
M. W.
Barsoum
, and
J.
Rosen
,
Scr. Mater.
108
,
147
150
(
2015
).
40.
M.
Naguib
,
V. N.
Mochalin
,
M. W.
Barsoum
, and
Y.
Gogotsi
,
Adv. Mater.
26
(
7
),
992
1005
(
2014
).
41.
Y.
Liu
,
H.
Xiao
, and
W. A.
Goddard
 III
,
J. Am. Chem. Soc.
138
(
49
),
15853
15856
(
2016
).
42.
H. S.
Biswal
and
S.
Wategaonkar
,
J. Phys. Chem. A
114
(
19
),
5947
5957
(
2010
).
43.
J.
Perlstein
,
J. Am. Chem. Soc.
123
(
1
),
191
192
(
2001
).
44.
P. A.
Khomyakov
,
G.
Giovannetti
,
P. C.
Rusu
,
G.
Brocks
,
J.
van den Brink
, and
P. J.
Kelly
,
Phys. Rev. B
79
(
19
),
195425
(
2009
).
45.
L. H.
Li
,
T.
Tian
,
Q.
Cai
,
C.-J.
Shih
, and
E. J. G.
Santos
,
Nat. Commun.
9
,
1271
(
2018
).
46.
M.
Oehzelt
,
N.
Koch
, and
G.
Heimel
,
Nat. Commun.
5
,
4174
(
2014
).
47.
R.
Otero
,
A. L.
Vázquez de Parga
, and
J. M.
Gallego
,
Surf. Sci. Rep.
72
(
3
),
105
145
(
2017
).
48.
M.
Bokdam
,
G.
Brocks
, and
P. J.
Kelly
,
Phys. Rev. B
90
(
20
),
201411
(
2014
).
49.
W.
Tang
,
E.
Sanville
, and
G.
Henkelman
,
J. Phys.: Condens. Matter
21
(
8
),
084204
(
2009
).
50.
Y.
Zhao
,
J.
Qiao
,
P.
Yu
,
Z.
Hu
,
Z.
Lin
,
S. P.
Lau
,
Z.
Liu
,
W.
Ji
, and
Y.
Chai
,
Adv. Mater.
28
(
12
),
2399
2407
(
2016
).
51.
M.
Stengel
,
P.
Aguado-Puente
,
N. A.
Spaldin
, and
J.
Junquera
,
Phys. Rev. B
83
(
23
),
235112
(
2011
).
52.
Z.
Gao
,
Z.
Zhou
, and
D.
Tománek
,
ACS Nano
13
(
5
),
5103
5111
(
2019
).
53.
W.
Chen
,
E. J. G.
Santos
,
W.
Zhu
,
E.
Kaxiras
, and
Z.
Zhang
,
Nano Lett.
13
(
2
),
509
514
(
2013
).
54.
J.
Kang
,
W.
Liu
,
D.
Sarkar
,
D.
Jena
, and
K.
Banerjee
,
Phys. Rev. X
4
(
3
),
031005
(
2014
).

Supplementary Material