Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Role of induced-strain and interlayer coupling in contact resistance of VS2–BGaX2 (X = S, Se) van der Waals heterostructures

Umair Khana, Basit Alia, Tahani A. Alrebdib, M. Bilala, M. Shafiqa, M. Idreesc and Bin Amin*a
aDepartment of Physics, Abbottabad University of Science & Technology, Abbottabad 22010, Pakistan. E-mail: binukhn@gmail.com; Tel: +92-333-943-665
bDepartment of Physics, College of Science, Princess Nourah Bint Abdulrahman University, P. O. Box 84428, Riyadh 11671, Saudi Arabia
cSchool of Chemistry and Chemical Engineering, Shandong University, Jinan, 250100, P. R. China

Received 15th April 2025 , Accepted 27th July 2025

First published on 28th July 2025


Abstract

Using Density Functional Theory (DFT) calculations, we explored the electronic band structure and contact type (Schottky and Ohmic) at the interface of VS2–BGaX2 (X = S, Se) metal–semiconductor (MS) van der Waals heterostructures (vdWHs). The thermal and dynamical stabilities of the investigated systems were systematically validated using energy–strain analysis, ab initio molecular dynamics (AIMD) simulations, as well as binding energy and phonon spectrum calculations. After analyzing the band structure, VS2–BGaX2 (X = S, Se) MS vdWHs metallic behavior with type-III band alignment is revealed. A p-type Schottky (Ohmic) contact in VS2–BGaS2 (VS2–BGaSe2) MS vdWHs with decreasing (increasing) tunneling probabilities (current) shows its potential uses in phototransistors, photodetectors and high-speed nanoelectronic devices. Additionally, the work function (ϕ), electrostatic potential and charge density difference are also investigated to gain detailed insights into the work function variations and charge transfer between layers during the fabrication of VS2–BGaX2 (X = S, Se) MS vdWHs. At equilibrium interlayer distance, strong interlayer coupling due to the vdW interactions is further confirmed via Bader charge analysis, showing that the electrons are transferred from BGaS2(VS2) to the VS2(BGaS2) layer in VS2–BGaS2 (VS2–BGaSe2) MS vdWHs. These calculations give a new strategy for experimentalists to design advanced high-speed nanoelectronic devices based on VS2–BGaX2 (X = S, Se) MS vdWHs.


1. Introduction

Practical applications of semiconductors require direct contact with the metal electrode in the form of a metal–semiconductor (MS) contact to achieve efficient carrier injection for high-performance nanoelectronic devices.1–3 In the MS contact, although the Schottky barrier (SB) allows carriers to pass between the metal and semiconductor,4 its height (SBH) poses critical problems to the charge carriers’ injection capability in devices.5 The formation of proficient nanodevices depends on the modulation of contact resistance (Schottky or Ohmic) at the MS interface.6

Heavy doping is used in conventional bulk materials to achieve low contact resistances, but this technique is not easy in silicon-based FETs (using bulk metal electrodes).7–10 The ultrathin nature of 2D materials also does not allow heavy doping in the contact area, as it would vary their structure and other properties. Fermi level pinning (FLP) at the interface of the 3D metal and 2D semiconductor also limits the drain current and charge injection efficiency.11–15 To find a metal with a suitable work function that also exhibits high conductivity and chemical and thermal stability to obtain a Schottky or Ohmic contact with semiconductors is also very difficult.16 Due to the low dimension scale of devices (less than 10 nm), tuning the SBH at the MS junctions via an external electric field also poses substantial challenges.17

Stacking of two-dimensional (2D) materials to form a van der Waals heterostructure (vdWH) has emerged as an effective strategy to enhance material properties and uncover novel physical phenomena. This approach is widely used in design and development of next-generation nanoelectronic devices with tailored functionalities, which often depend on the type of band alignment at the interface. Typically, the type of band alignment is categorized into three types on the basis of the relative positions of the conduction band (CB) and valence band (VB) edges of the constituent layers. In type-I (straddling gap) band alignment the contributions of the CB and VB originate from same layer, effectively confining holes and electrons within one material. While in type-II (staggered gap) alignment the CB and VB are contributed by distinct layers, facilitating efficient charge separation across the interface, and in type-III (broken gap) band alignment either the CB or VB of one layer crosses the Fermi level during the contact, resulting in a band overlap at the Fermi level which can facilitate tunneling behavior and is particularly useful in designing photodetectors, tunneling transistors, and other advanced electronic devices. In this context, understanding the band alignment at the interface of metal–semiconductor heterostructures is crucial, significantly influencing charge transfer and contact resistance.18,19

Stacking of metals with semiconductors in the form of 2D vdWHs also prevents the development of localized electronic states and allows the Fermi level freely to align with the semiconductor’s valence or conduction bands forming a p-/n-type Schottky or Ohmic contact at the interface.20,21 This reduces the FLP and potential scattering, thereby enhancing device performance.22,23 Tuning of the SBH by varying the interfacial distance in vdWHs has been previously reported in ref. 24 and 25, offering a promising approach for SBH modulation and optimizing charge injection and carrier transport. Although, in the 2D family, transition metal dichalcogenides (TMDCs) are widely used in MS contacts,26–28 members of the same family with metallic behaviour have also received significant attention for a variety of nanodevice applications.29–33 The work functions of TMDCs align well with several other semiconducting materials, making them suitable for contact electrodes in FETs, lowering the SBH and enhancing the device efficiency.6 The unique and complicated electronic structure of VS2 has inspired us to explore this layered compound as a promising functional material. The 2D electron–electron correlations among the V atoms are expected to induce more intricate planar electric transport properties.34,35 In the same family of 2D materials, Janus TMDCs (MoSSe)36,37 have opened a pathway for the designing of novel Janus-materials.38–40 Structural and mechanical stability, optical and ferro-piezoelectric properties of Janus BGaS2 and BGaSe2 have recently been explored via first-principles calculations.41 Both BGaS2 and BGaSe2 are semiconductors with indirect bandgaps of 2.13 eV and 1.63 eV, respectively.

In this work, using first principles calculations, vdWHs in the form of MS contacts of VS2 with Janus BGaS2 and BGaSe2 layers are modelled, and their contact properties are explored. The experimentally accessible lattice mismatch between the constituent layers ensures that the VS2–BGaX2 (X = S, Se) vdW interfaces remain energetically viable while preserving their intrinsic properties. These findings demonstrate that the VS2–BGaX2 (X = S, Se) vdW interfaces exhibit an adjustable SBH, and the contact type can be switched from p- to n-type and from Schottky to Ohmic. Therefore, due to their tunable barrier height, these vdW interfaces are highly suitable for enhancing the carrier injection capability with promising applications in light-emitting diodes (LEDs) photodetectors and solar cells.

2. Computational details

Density Functional Theory (DFT) calculations were carried out using the PWSCF code42 with employing the Perdew–Burke–Ernzerhof (PBE) exchange–correlation functional.43 A plane-wave cutoff energy of 700 eV was adopted to ensure accurate total energy and forces evaluations. For the geometry optimization a Monkhorst–Pack K-point grid of 8 × 8 × 1 was used while a denser grid of 16 × 16 × 1 was applied for the self-consistent field. The convergence thresholds were set to 10−3 eV Å−1 and 10−4 eV, for force and energy respectively, ensuring reliable optimization of lattice parameters and atomic position. To eliminate spurious interactions between periodic images, a vacuum space of 25 Å was introduced along the out-of-plane (z-direction). All binding energy and structural relaxation calculations were carried out using the PBE functional combined with Grimme’s DFT-D2 dispersion corrections to accurately capture long-range van der Waals interactions.44

Using ab initio molecular dynamics (AIMD) simulations,45 the thermal stability was investigated via the Nosé thermostat algorithm at a temperature of 300 K, and the dynamical stability through phonon spectra using density functional perturbation theory (DFPT).46 The mechanical stability of these systems is confirmed from the elastic constants calculated via the energy–strain approach.47 The interlayer distance and binding energies for all stacking configurations of VS2–BGaX2 (X = S, Se) are calculated via:48 Eb = E(vdWH)E(semiconducting-layer)E(metallic-layer).

Using the Schottky–Mott rule,49 the heights of the n-type (ΦBn = ECBMEF) and p-type (ΦBp = EFEVBM) Schottky contacts are calculated (ECBM and EVBM are the energies of the band edges of the semiconducting material, EF is the Fermi level of the vdWH). To confirm the charge transfer and type of contact, band bending is also calculated50 via: Δϕ = ϕVS2ϕBGaX2, where ϕVS2 and ϕBGaX2 denote the work function of the corresponding metallic (VS2) and semiconducting (BGaX2) layers, respectively in the MS vdWHs. The tunneling barrier probability via:49 image file: d5na00356c-t1.tif (m is the mass of a free electron, and ℏ is the reduced Planck constant) is expressed in terms of the comprehensive factor: C = (WTB)2ΦTB.

3. Results and discussion

BGaX2 (X = S, Se) in hexagonal close-packing has a four-layer atomic structure, composed of BX and GaX layers. Each B atom is covalently bonded to two X atoms and one Ga atom (X–B–Ga–X), leading to a trigonal prismatic geometry (space group no. 187 (P[6 with combining macron]m2)). Substituting B with a Ga atom disrupts the out-of-plane symmetry from D3h to C3v, which significantly affects the electronic structure and optical properties of these materials.41,51 BGaS2 (Fig. 1(a)) and BGaSe2 (Fig. 1(b)) are indirect bandgap (ΓM) semiconductors, while VS2 is metal (Fig. 1(c)). In the case of the BGaS2 (BGaSe2) layers, the valence band near the Fermi level is mainly due to B-p and Ga-p orbitals, while the conduction band is due to B-p and S(Se)-p orbitals confirming the semiconducting behavior, as shown in Fig. 1(d) and (e). In the VS2 layer (see Fig. 1(f)), V-d and S-p orbitals cross the Fermi level, confirming the metallic behaviour. The lattice constants, bond lengths, bond angles along with the band gap energies of BGaS2 and BGaSe2 in Table 1 are in good agreement with previous work.41
image file: d5na00356c-f1.tif
Fig. 1 Electronic band structure (first row) and density of states (second row) of: (a and d) BGaS2, (b and e) BGaSe2, and (c and f) VS2 (see text for details).
Table 1 Lattice constant (a in Å), bond length (B–Ga, B–X, Ga–X (X = S, Se) in Å), layer thickness (S–Se in Å) between chalcogen atoms, bond angles, the energy band-gap (Eg in eV) and the elastic constants (C11, C12, C66) of BGaS2, BGaSe2, VS2–BGaS2 and VS2–BGaSe2 MS vdWHs
Material a B–Ga B–X Ga–X S–Se

image file: d5na00356c-t4.tif

image file: d5na00356c-t5.tif

Eg C11 C12 C66
BGaS2 3.29 2.09 2.06 2.26 4.11 105.97 112.77 1.33 58.67 14.43 22.12
BGaSe2 3.46 2.07 2.19 2.39 4.38 104.51 114.06 0.91 51.22 10.25 20.48
VS2–BGaS2 3.23 2.08 2.03 2.25 4.15 105.51 113.18 1.327 88.36 28.67 29.84
VS2–BGaSe2 3.32 2.07 2.12 2.38 4.39 102.94 115.40 0.898 74.23 48.52 12.85


The carrier effective mass (m) is a crucial parameter in semiconductor physics, as it influences charge carrier mobility, electrical conductivity, and overall device performance. It depends on the curvature of the conduction and valence bands. A smaller effective mass leads to higher carrier mobility via: image file: d5na00356c-t2.tif,52 making materials more suitable for high-speed electronic and optoelectronic applications. The carrier effective mass in BGaX2 (X = S, Se) layers is investigated via image file: d5na00356c-t3.tif by parabolic fitting of the CBM and VBM.53,54 The calculated effective mass of electrons (holes) in BGaS2 is 0.0722 (0.205) and in BGaSe2 is 0.091 (0.187).

The lattice mismatch of VS2 and BGaS2 monolayers in the modelling of the VS2–BGaS2 vdWH is 3.5%. and in the case of the VS2–BGaSe2 vdWH it is 8.3%, which realizes experimental fabrication. Therefore, to understand the formation of the contact between the 2D metallic VS2 and semiconducting BGaX2 (X = S, Se) layers, we modelled six possible stackings ((a)–(f)) of VS2–BGaX2 vdWHs in a 1 × 1 unit cell of each layer conforming to the minor lattice mismatch (<10%),55 which does not substantially impact the fundamental properties of the constituent layers, see Fig. 2. In the case of (a), an X(B and Ga) atom of BGaX2 is placed on top of an S(V) atom of the VS2 layer. In (b), the X(B and Ga) atom of BGaX2 is settled on top of the V(S) atom of the VS2 layer. In (c), the B and Ga(X) atoms of BGaX2 are positioned on top of an S(hollow-site) of the VS2 layer. Similarly in (d), the B and Ga(X) atoms of BGaX2 are placed on top of a V(hollow-site) of the VS2 layer. In (e), the BGaX2 X(B and Ga) atoms are placed on top of a V(hollow-site) of the VS2 layer. In (f), the X(B and Ga) atoms of BGaS2 are placed on top of an S(hollow-site) of the hexagonal lattice site of the VS2 layer.


image file: d5na00356c-f2.tif
Fig. 2 Geometrical structures of the VS2–BGaX2 (X = S, Se) van der Waals heterostructure in various stacking patterns (see text for details).

The calculated binding energy and the interlayer distance in Table 2 show the experimental fabrication of all possible stacking configurations. The most stable stacking configuration is the one with the most negative binding energy and the shortest interlayer distance, see Table 2. Among these six stacking configurations, the stacking (b) is found to be the most favorable for these MS vdWHs. The calculated binding energies (−0.24 to −0.33 eV) fall well within the range typical for vdWHs, confirming the non-covalent bonding between the interlayers, hence supporting the physical nature of the heterostructure interface without significant chemical reconstruction. During the modeling of VS2–BGaX2 MS vdWHs, strain is induced in the corresponding layers due to their lattice mismatch that significantly impacts the stability of the various stacking configurations of the MS vdWHs. Therefore, we re-optimized the lattice constant of the VS2–BGaX2 MS vdWHs, and the results are presented in Table 1.

Table 2 Binding energy (Eb in eV) and inter-layer distance (d in Å) of VS2–BGaS2 and VS2–BGaSe2 MS vdWHs
Material (a) (b) (c) (d) (e) (f)
VS2–BGaS2 Eb (eV) −0.331 −0.333 −0.329 −0.331 −0.332 −0.328
d (Å) 3.320 3.170 3.240 3.301 3.302 3.250
VS2–BGaSe2 Eb (eV) −0.240 −0.242 −0.238 −0.239 −0.241 −0.240
d (Å) 3.320 3.200 3.450 3.430 3.250 3.290


Using AIMD simulations, the geometrical structure is found to remain unchanged (no structural distortions are observed) after 4000 step simulations at room temperature (300 K) with nearly constant fluctuation energy, see Fig. 3. This confirms the thermal stability of VS2–BGaX2 MS vdWHs.56


image file: d5na00356c-f3.tif
Fig. 3 Geometrical structures before heating (first row), with fluctuating energy (second row) and after heating (third row) of (a) VS2–BGaS2 and (b) VS2–BGaSe2 (see text for details).

In agreement with ref. 34, 35 and 41, the phonon dispersion curves of the 2D BGaS2, BGaSe2, and VS2 layers, along with their vdWHs (VS2–BGaS2 and VS2–BGaSe2) are calculated in the high-symmetry directions (ΓKMΓ) of the BZ. The absence of imaginary frequencies confirm the dynamical stability of these systems. The BGaX2 (X = S, Se) layer contains four atoms per unit cell exhibiting a total number of 12 modes, see Fig. 4(a) and (b). The three lower frequency modes correspond to the in-plane longitudinal (LA), transverse (TA) and out-of-plane acoustic (ZA) modes. The remaining nine modes indicate the optical branches are well-separated, having a higher frequency of ∼800 cm−1. VS2 in Fig. 4(e) and (f) consists of three atoms per unit cell having a total number of nine modes. The highest optical phonon frequency is ∼300 cm−1, significantly lower than BGaX2 compounds. This suggests weaker bonding and stronger electron–phonon coupling, consistent with the VS2 metallic nature (metals often have strong interactions between electrons and phonons). In the case of VS2–BGaX2 MS vdWHs, there are a total of 21 modes of frequency, due to the seven atoms per unit cell see Fig. 3(c) and (d). Due to the weak vdW interactions, the 3 lower frequency modes are acoustical (LA, TA and ZA) and the remaining 18 modes are optical. The general shapes of the phonon dispersion for VS2–BGaX2 MS vdWHs resemble a simple superposition of the phonon dispersions of the corresponding layers (VS2, BGaX2). Due to the metallic nature of the VS2 monolayers and interlayer coupling with BGaX2, the frequency ranges become lower in vdWHs up to ∼500 cm−1. These investigations confirm that the structure of the VS2–BGaX2 MS vdWHs do not impulsively collapse, making these materials well suited for further tuning, thereby opening a pathway for feature field effect transistors, optoelectronics and nanoelectronic devices.57,58


image file: d5na00356c-f4.tif
Fig. 4 Phonon spectra of (a) BGaS2, (b) BGaSe2, (c) VS2–BGaS2, (d) VS2–BGaSe2 and (e and f) VS2 (see text for details).

The lattice mismatch introduced induced strain in BGaS2, BGaSe2 and VS2 layers may significantly alter the bond length. Therefore, elastic constants (Cij)59,60 of the vdWHs and corresponding layers are calculated. As shown in Table 1, C11 > 0, C12 > 0, C11 > ∣C12∣, satisfying the Born criteria61,62 for both vdWHs and their corresponding layers, which indicates high mechanical stability. The higher values of C11 in Table 1 (BGaS2 > BGaSe2 > VS2) suggest that BGaS2 is stiffer than BGaSe2 and VS2, hence confirming that BGaS2 is more resistant to deformation. In the case of VS2–BGaX2 (X = S, Se) MS vdWHs, additional stress fields are generated at the interface between the corresponding layers. These stress fields can introduce interfacial dipoles that significantly alter the elastic properties such as stiffness and shear resistance, making these material more anisotropic compared to the individual layers. For the vdWHs, the higher value of C11 compared to the corresponding layers indicates that the interaction between layers leads to enhanced in-plane stiffness, especially if the interlayer coupling is strong. The C66 parameter, derived from C11 and C12, is used to understand shear deformation within the basal plane, helping to assess the mechanical stability of the vdWHs. This parameter also provides insight into mechanisms of energy dissipation, such as interlayer sliding, which may occur under mechanical stress. These calculations suggest that the vdWHs are mechanically stronger due to interlayer vdW interactions and strain effects.

The calculated electronic band structures of VS2–BGaS2 (Fig. 5(a)) and VS2–BGaSe2 (Fig. 5(b)) vdWHs display energy dispersion along ΓKMΓ, where the blue (orange) lines represent the contribution of VS2 (BGaX2 (X = S, Se)) layers. Fig. 5(a) and (b) demonstrate that VS2–BGaX2 are metals with type-III band alignment.63 Due to the weak vdW forces at the interface, the band structures of VS2–BGaX2 vdWHs closely resemble the superposition of the band structures of their corresponding layers. Stacking in the form of VS2–BGaX2 MS vdWHs induces strain due to lattice mismatch between individual layers, leading to tuning of overall properties of the materials (the VBM of the BGaS2 layer shifts near to the Fermi level, while that of the BGaSe2 layer crosses the Fermi level). Additionally, the strong differences in electronegativity between the intrafacial atoms of the VS2 and BGaS2 (BGaSe2) layers encourage electron transfer across the interface, affecting the Schottky (Ohmic) contact resistance and also contributing to alteration of the type of contact (p-type and n-type) depending on the relative position of the metal’s Fermi level with respect to the semiconductor’s band edges (CBM and VBM) of MS vdWHs. Schottky contact occurs when there is a significant energy barrier for electron and hole transfer across the metal and semiconductors, while in the case of Ohmic contact the barrier for electron and hole injection is negligible or absent, allowing holes to move freely across the interface, which leads to low contact resistance and efficient carrier transport. In the case of n-type SBH, the conduction band of the semiconductors aligns closer to the Fermi level of the metal, facilitating electron flow from the metal to semiconductor. While in the case of p-type (SBH) contact, the valence band of the semiconducting layer is positioned closer to the metal Fermi level with holes as the majority carriers at the interface and allows hole transport from semiconductor to metal. Moreover, interfacial charge transfer creates a dipole layer at the interface, showing potential applications in FETs. The higher (lower) bandgap of BGaS2 (BGaSe2) implies that the VS2–BGaS2 (VS2–BGaS2) vdWHs exhibit Schottky (Ohmic) contact behavior. However, for future studies, the strong hybridization suggests that the barrier could be reduced, potentially leading to a transition toward Ohmic behavior under certain doping conditions or strain engineering. We have also calculated the band structure of other stacking configurations, see Fig. 3S in the supplementary information (SI).


image file: d5na00356c-f5.tif
Fig. 5 Electronic band structure (first row) and density of states (second row) of (a and c) VS2–BGaS2, (b and d) VS2–BGaSe2 vdWHs (see text for details).

The partial density of states (PDOS) of VS2–BGaX2 in Fig. 5 shows that the primary contribution to the Fermi level originates from the V-d and S-p orbitals of the VS2 layer (crossing the Fermi level). In the case of VS2–BGaS2, the S-p and Ga-p orbitals mainly contribute in the VBM, while the B-p orbitals dominate in the CBM, without crossing the Fermi level. In contrast, for VS2–BGaSe2, the Se-p orbitals cross the Fermi level, whereas the B-p and Ga-p orbitals do not significantly contribute to the band edges of VS2–BGaSe2, see Fig. 5(c) and (d) respectively. Both the electronic band structure and PDOS reveal that changing the chalcogen atom from S (BGaS2) to Se (BGaSe2), alters the band alignment, affecting the degree of band hybridization. In the case of VS2–BGaSe2 vdWH, Se leads to stronger hybridization with VS2 compared to S in VS2–BGaS2, potentially affecting interfacial charge transfer and contact resistance. These findings suggest that transition metal dichalcogenides (TMDs) like VS2 can serve as an excellent contact material for semiconductor-based heterostructures, enabling efficient charge transport in nanoelectronic and optoelectronic devices.

The electrostatic potential (ΔV) plays a crucial role in understanding the charge distribution, band alignment, and interfacial interactions in vdWHs. Fig. 6 illustrates the variation in potential energy along the Z-direction for (a) BGaS2, (b) BGaSe2, (c) VS2–BGaS2 and (d) VS2–BGaSe2 vdWHs, while the green line is for the VS2 layer. Although, the ΔV of the VS2 layer exhibits a deeper potential then BGaX2 (X = S, Se) due to its strong electron localization and high carrier density. The stacking of VS2–BGaX2 (X = S, Se) MS vdWHs, due to strong electronegativity differences and charge redistribution at the interface, significantly modifies ΔV. Fig. 6(c), shows that the VS2 layer exhibits a deeper potential then BGaS2, confirming transportation of charge from BGaS2 to the VS2 layer in the VS2–BGaS2 MS vdWH. In contrast, for the VS2–BGaSe2 MS vdWH (Fig. 6(d)), charges are transferred from VS2 to the BaGaSe2 layer. This charge transport indicates a strong interlayer coupling arising from the vdW interfacial forces of VS2 and BGaX2 layers in the VS2–BGaX2 (X = S, Se) MS vdWHs.


image file: d5na00356c-f6.tif
Fig. 6 Electrostatic potential of (a) BGaS2, (b) BGaSe2, (c) VS2–BGaS2 and (d) VS2–BGaSe2 vdWHs. Where the green line is for the VS2 monolayer (see text for details).

In accordance with ref. 64 (for GaSe/graphene) and ref. 65 (for MoS2/graphene), charge transport induces a region of charge separation, which create an interfacial electric field within the heterostructure interface.66 This interfacial electric field increases the number of carriers, strengthens mobility, and tunes the work function (ϕ). The calculated ϕ of VS2 is 4.711 eV, BGaS2 is 4.099 eV and BGaSe2 is 3.867 eV. Although, due to the vdW interaction and potential alignment, the Fermi levels of the individual layers equilibrate upon stacking, resulting in band bending at the interfaces, having profound implications for charge carrier transport (discussed in detail below). The work function of these heterostructures becomes optimized and inferior to that of their respective monolayers. The ϕ values of VS2–BGaX2 (X = S, Se) vdWHs are 4.399 eV and 4.293 eV, respectively, which are lower than that of their corresponding layers facilitating efficient charge transfer. These materials show potential for device applications due to provision of an optimal threshold voltage that enhances carrier injection control, and allows for further tuning. The ΔV across the interface of VS2–BGaX2 (X = S, Se) vdWHs ranges from 0.143–0.244 eV. The results, along with the tunable work function and improved carrier mobility, indicate that these vdWHs are promising candidates for electronic and optoelectronic applications including transistors, sensors and energy-harvesting devices.58,67

The charge transportation is further validated through charge density difference: △ρ = ρ(VS2–BGaX2)ρ(VS2)ρ(BGaX2) and Bader charge analysis, as shown in Fig. 7. In the case of VS2–BGaS2 MS vdWHs, the charge depletion is absorbed around the BGaS2, while charge accumulation around the VS2 layer highlights that BGaS2 loses and VS2 gains electrons, which creates a hole-rich environment in BGaS2 and an electron-rich region around VS2 layers.68 Although , in the case of VS2–BGaSe2 MS vdWHs, a hole (electron) rich region was established around the VS2 (BGaSe2) layer. The Bader charge analysis was used to quantitatively evaluate the charge transfer and shows that, in the case of (a) VS2–BGaS2, 0.003 (0.003) h (e) are transferred between VS2 (BGaS2), while in the case of (b) VS2–BGaSe2, 0.0104 (0.0105) h (e) are transferred between VS2 (BGaSe2). The phenomenon of charge transfer demonstrates a robust interlayer coupling and vdW interactions at the interface of VS2 and BGaX2 layers within VS2–BGaX2 MS vdWHs. This charge transport facilitates the formation of a built-in-electric field at the interface, creating a charge separation region and generating an interfacial electric field66 that enhances the carrier mobility and increases the concentration of carriers (holes and electrons). Similar experimental results have been confirmed in graphene/GaSe64 and graphene/MoS2.65


image file: d5na00356c-f7.tif
Fig. 7 Charge density differences of (a) VS2–BGaS2 and (b) VS2–BGaSe2 (see text for details).

The contact resistance can be tuned through modulation of the Schottky barrier height (SBH) using the electronic structure to regulate the device efficiency.53,54 Therefore, it is crucial to understand the barrier height and type at the interface of VS2–BGaX2 (X = S, Se) MS vdWHs. Fig. 8 shows that a Schottky contact is demonstrated at the interface of the VS2–BGaS2 MS vdWH,69 where ΦBp (0.0564) is found to be lower than ΦBn (1.0594) (see Fig. 8 and computational details), highlighting the formation of a p-type Schottky contact that favors holes conduction over electrons, and plays a significant role in enhancing the performance of phototransistors and photodiodes.70–72 In the case of VS2–BGaSe2 MS contacts, the induced strain and strong electronegativity differences between the VS2 and BGaSe2 layers cause a shift in the VBM of the semiconducting BGaSe2 layer toward the Fermi level, hence, affecting the interlayer coupling and band alignment at the interface. These changes provide adequate control over the contact’s properties and reduce the contact resistance, ultimately leading to the formation of an Ohmic contact. WSe2/Ti2CF2 and WSe2/Ti2C(OH)2 in ref. 73 have already been shown to be Ohmic contacts. The calculated band bending (Δϕ) for VS2–BGaS2 vdWHs is −0.223 eV, and for VS2–BGaSe2 vdWHs is −0.353 eV further confirming a p-type Schottky contact at the interface of VS2–BGaX2 (X = S, Se) vdWHs.


image file: d5na00356c-f8.tif
Fig. 8 Calculated contact resistance in VS2–BGaS2 (Schottky) and in VS2–BGaSe2 (Ohmic) MS vdWHs.

Unlike conventional TMDC systems, the asymmetric VS2–BGaX2 (X = S, Se) heterostructures disrupt vertical symmetry due to the differences in the atomic size and electronegativity of X = S, Se, leading to the formation of inequivalent bond lengths across the interface. Therefore, more electrons are transferred from the V atoms to S than Se. This charge localization creates an intrinsic dipole (ID) in the out-of-plane directions, which further alter the transportation of charge at the interface of the VS2 and BGaX2 layers. The ID is also enhanced due to the weak vdW forces, which promote charge density redistribution at the interface, (as already discussed; see Fig. 6).

Fermi level pinning (FLP), effected by ID at the MS interface, is quantitatively analyzed using the slope parameter S:74 image file: d5na00356c-t6.tif, where ΦBp is the Schottky barrier height (SBH) for a p-type semiconductor and ϕm is the metal work function. Generally, in the Schottky–Mott limit, S approaches 1 reflecting no Fermi level pinning (insignificant charge transfer between semiconductor and metal), while S approaches 0 for the strongly pinned Fermi level, called the Bardeen limit, where the SBH becomes unaffected by the metal work function. These findings suggest that the FLP profoundly influences the SBH and charge injection, which enhances the overall performance of field-effect transistors (FETs). The calculated S parameter for VS2–BGaS2 is 0.0114, and for VS2–BGaSe2 vdWHs is 0.056, indicating very low pinning at the interface.75

Induced strain in pristine monolayers while fabricating VS2–BGaX2 MS vdWHs and strong differences in electronegativity also alter the orbitals overlapping, creating a tunneling barrier (TB) at the interface. Lowering the TB can improve carrier mobility, subsequently increasing the current at the contact interface. The TB involves both the tunnel barrier height (ΦTB is the potential difference between the vdW gap (Φgap) and the potential energy of BGaX2 (ΦBGaX2)) and the tunnel width (WTB indicate the width of the square potential barrier at the interface).73 Tuning the Schottky barrier height and minimizing the tunneling barrier width are effective strategies to enhancing charge injection and reducing contact resistance at the MS interface. Therefore, the tunneling barrier probability,49 PTB, can be expressed as a comprehensive factor C (see computational details). A smaller C value is beneficial, as it indicates a higher tunneling barrier probability. In the case of VS2–BGaS2 and VS2–BGaSe2 the tunnel barrier height, ΦTB (tunnel width, WTB) is 3.009 eV (3.015 Å) and 2.341 eV (2.73 Å) respectively. For these vdWHs the tunneling barrier probability PTB and the comprehensive factor (C) are 23.83% (27.361) and 60.82% (17.448). These calculations suggest that VS2–BGaSe2 exhibits greater charge transfer than in VS2–BGaS2.

4. Conclusion

In this study, we systematically explored the role of interface-induced strain and interlayer coupling in modulating the contact resistance of VS2–BGaX2 (X = S, Se) vdWHs using DFT calculations. Our findings reveal that these materials exhibit mechanical, thermal and dynamical stability confirmed via an energy–strain approach, AIMD simulations, binding energy analysis, and phonon spectrum calculations. The electronic band structure demonstrates that all VS2–BGaX2 (X = S, Se) MS vdWHs are metal with type-III band alignment, providing a deeper understanding of the interface mechanisms, tunneling characteristics and electronic transport properties across these heterostructures. The work functions of VS2–BGaS2 and VS2–BGaSe2 MS vdWHs are 4.399 eV and 4.293 eV, respectively, which are lower than their respective monolayers, modulating charge transfer efficiency. Charge depletion (accumulation) is observed in BGaS2 (VS2) and VS2 (BGaSe2) layers confirming the loss (gain) of electrons that creates hole (electron) rich environments. This charge transportation highlights strong interlayer coupling driven by the vdW interactions at the interface of the VS2–BGaX2 (X = S, Se) MS contact. Importantly, a p-type Schottky contact is identified at the VS2–BGaS2 interface and an Ohmic contact at the VS2–BGaSe2 interface with lowering (increasing) of the tunneling probabilities (current), confirming its potential in high-speed nanoelectronic devices. Our findings show that the differences in the contact resistance behavior arise from the impact of induced strain and electronic coupling at the interfaces, enabling tunable functionalities for minimizing power losses and ensuring efficient current injection in transistors and other semiconductor devices due to their built-in potential for charge carriers selectivity. Overall, our results offer valuable insights into the mechanism of contact resistance in 2D MS vdWHs, and serve as a strategic framework for experimentalists to design high-performance solar cells, photodetectors, phototransistors, sensors, and high-speed rectifying devices for next-generation technologies.

Conflicts of interest

There are no conflicts to declare.

Data availability

The data that support the findings of this study are available on request.

The data that support the findings of this study in the manuscript and supplementary informations are available on request. See DOI: https://doi.org/10.1039/d5na00356c.

Acknowledgements

The authors extend their sincere appreciation to Princess Nourah bint Abdulrahman University Researchers Supporting Project number (PNURSP2025R71), Princess Nourah bint Abdulrahman University, Riyadh, Saudi Arabia.

References

  1. Y. Wang, S. Liu, Q. Li, R. Quhe, C. Yang, Y. Guo, X. Zhang, Y. Pan, J. Li and H. Zhang, Rep. Prog. Phys., 2021, 84, 056501 CrossRef CAS PubMed.
  2. A. Allain, J. Kang, K. Banerjee and A. Kis, Nat. Mater., 2015, 14, 1195 CrossRef CAS PubMed.
  3. S. B. Mitta, M. S. Choi, A. Nipane, F. Ali, C. Kim, J. T. Teherani, J. Hone and W. J. Yoo, 2D Mater., 2021, 8, 012002 CrossRef CAS.
  4. R. T. Tung, Appl. Phys. Rev., 2014, 1, 54 Search PubMed.
  5. T. Shen, J. C. Ren, X. Liu, S. Li and W. Liu, J. Am. Chem. Soc., 2019, 141, 3110 CrossRef CAS PubMed.
  6. X. Ding, Y. Zhao, H. Xiao and L. Qiao, Appl. Phys. Lett., 2021, 118, 091601 CrossRef CAS.
  7. F. A. Zwanenburg, A. S. Dzurak, A. Morello, M. Y. Simmons, L. C. L. Hollenberg, G. Klimeck, S. Rogge, S. N. Coppersmith and M. A. Eriksson, Rev. Mod. Phys., 2013, 85, 961 CrossRef CAS.
  8. S. Najmaei, J. Yuan, J. Zhang, P. Ajayan and J. Lou, Acc. Chem. Res., 2015, 48, 31 CrossRef CAS PubMed.
  9. W. Ahmad, Y. Gong, G. Abbas, K. Khan, M. Khan, G. Ali, A. Shuja, A. K. Tareen, Q. Khan and D. Li, Nanoscale, 2021, 13, 5162 RSC.
  10. C. Sheng, X. Dong, Y. Zhu, X. Wang, X. Chen, Y. Xia, Z. Xu, P. Zhou, J. Wan and W. Bao, Adv. Funct. Mater., 2023, 33(50), 2304778 CrossRef CAS.
  11. I. Popov, G. Seifert and D. Tomanek, Phys. Rev. Lett., 2012, 108, 156802 CrossRef PubMed.
  12. C. Gong, L. Colombo, R. M. Wallace and K. Cho, Nano Lett., 2014, 14, 1714 CrossRef CAS PubMed.
  13. J. H. Kang, W. Liu, D. Sarkar, D. Jena and K. Banerjee, Phys. Rev. X, 2014, 4, 031005 CAS.
  14. R. H. Quhe, X. Y. Peng, Y. Y. Pan, M. Ye, Y. Y. Wang, H. Zhang, S. Y. Feng, Q. X. Zhang, J. J. Shi, J. B. Yang, D. P. Yu, M. Lei and J. Lu, ACS Appl. Mater. Interfaces, 2017, 9, 3959 CrossRef CAS PubMed.
  15. C. He, M. Cheng, T. Li and W. Zhang, ACS Appl. Nano Mater., 2016, 2, 2767 CrossRef.
  16. F. Daneshvar, H. Chen, K. Noh and H. J. Sue, Nanoscale Adv., 2021, 3(4), 942 RSC.
  17. W. Zhang, G. Hao, R. Zhang, J. Xu, X. Ye and H. Li, J. Phys. Chem. Solids, 2021, 157, 110189 CrossRef CAS.
  18. B. T. Beshir, K. O. Obodo and G. A. Asres, RSC Adv., 2022, 12(22), 13749–13755 RSC.
  19. U. Khan, B. Ali, B. Yuxiang, M. Idrees and B. Amin, J. Phys. Chem. Solids, 2025, 112832 CrossRef CAS.
  20. H. Hasegawa and T. Sawada, Thin Solid Films, 1983, 103, 119 CrossRef CAS.
  21. J. Bardeen, Phys. Rev., 1947, 71, 717 CrossRef.
  22. Y. Liu, P. Stradins and S. H. Wei, Sci. Adv., 2016, 2(4), e1600069 CrossRef PubMed.
  23. Y. C. Lin, J. Li, S. C. de la Barrera, S. M. Eichfeld, Y. Nie, R. Addou, P. C. Mende, R. M. Wallace, K. Cho, R. M. Feenstra and J. A. Robinson, Nanoscale, 2016, 8, 8947 RSC.
  24. L. Peng, Y. Cui, L. Sun, J. Du, S. Wang, S. Zhang and Y. Huang, Nanoscale Horiz., 2019, 4, 480 RSC.
  25. M. L. Sun, J. P. Chou, Q. Ren, Y. Zhao, J. Yu and W. C. Tang, Appl. Phys. Lett., 2017, 110, 173105 CrossRef.
  26. C. Kim, I. Moon, D. Lee, M. S. Choi, F. Ahmed, S. Nam, Y. Cho, H.-J. Shin, S. Park and W. J. Yoo, ACS Nano, 2017, 11, 1588 CrossRef CAS PubMed.
  27. S. Das, H.-Y. Chen, A. V. Penumatcha and J. Appenzeller, Nano Lett., 2012, 13, 100 CrossRef PubMed.
  28. Y. Guo, D. Liu and J. Robertson, ACS Appl. Mater. Interfaces, 2015, 7, 25709 CrossRef CAS PubMed.
  29. J.-A. Yan, M. A. Dela Cruz, B. Cook and K. N. Varga, Sci. Rep., 2015, 5, 16646 CrossRef CAS PubMed.
  30. N. Guo, X. Fan, Z. Chen, Z. Luo, Y. Hu, Y. An, D. Yang and S. Ma, Comput. Mater. Sci., 2020, 176, 109540 CrossRef CAS.
  31. T. Jiang, T. Hu, G.-D. Zhao, Y. Li, S. Xu, C. Liu, Y. Cui and W. Ren, Phys. Rev. B, 2021, 104, 075147 CrossRef CAS.
  32. S. Pratap, N. Joshi, R. Trivedi, C. S. Rout and B. Chakraborty, Micro Nanostruct., 2023, 181, 207627 CrossRef.
  33. N. Zhao, S. Tyagi and U. Schwingenschlögl, NPG Asia Mater., 2023, 15, 1343 Search PubMed.
  34. H. L. Zhuang and R. G. Hennig, Phys. Rev. B, 2016, 93(5), 054429 CrossRef.
  35. H. Zhang, L. M. Liu and W. M. Lau, J. Mater. Chem. A, 2013, 1(36), 10821 RSC.
  36. A.-Y. Lu, H. Zhu, J. Xiao, C.-P. Chuu, Y. Han, M.-H. Chiu, C.-C. Cheng, C.-W. Yang, K.-H. Wei, Y. Yang, Y. Wang, D. Sokaras, D. Nordlund, P. Yang, D. A. Muller, M.-Y. Chou, X. Zhang and L.-J. Li, Nat. Nanotechnol., 2017, 12, 744 CrossRef CAS PubMed.
  37. J. Zhang, S. Jia, I. Kholmanov, L. Dong, D. Er, W. Chen, H. Guo, Z. Jin, V. B. Shenoy, L. Shi and J. Lou, ACS Nano, 2017, 11, 8192 CrossRef CAS PubMed.
  38. N. C. Frey, A. Bandyopadhyay, H. Kumar, B. Anasori, Y. Gogotsi and V. B. Shenoy, ACS Nano, 2019, 13, 2831 CrossRef CAS PubMed.
  39. L. L. Li, C. Bacaksiz, M. Nakhaee, R. Pentcheva, F. M. Peeters and M. Yagmurcukardes, Phys. Rev. B, 2020, 101, 134102 CrossRef CAS.
  40. A. Kandemir and H. Sahin, Phys. Rev. B, 2018, 97, 155410 CrossRef CAS.
  41. D. Bezzerga, E.-A. Haidar, C. Stampfl, A. Mir and M. Sahnoun, Nanoscale Adv., 2023, 5, 1425 RSC.
  42. P. Giannozzi, S. Baroni, N. Bonini, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli, G. L. Chiarotti, M. Cococcioni, I. Dabo, A. Dal Corso, S. de Gironcoli, S. Fabris, G. Fratesi, R. Gebauer, U. Gerstmann, C. Gougoussis, A. Kokalj, M. Lazzeri, L. Martin-Samos, N. Marzari, F. Mauri, R. Mazzarello, S. Paolini, A. Pasquarello, L. Paulatto, C. Sbraccia, S. Scandolo, G. Sclauzero, A. P. Seitsonen, A. Smogunov, P. Umari and R. M. Wentzcovitch, J. Phys.: Condens. Matter, 2009, 21, 395502 CrossRef PubMed.
  43. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865 CrossRef CAS PubMed.
  44. S. Grimme, Semiempirical GGA-type density functional constructed with a long-range dispersion correction, J. Comput. Chem., 2006, 27(15), 1787–1799 CrossRef CAS PubMed.
  45. R. Yuan, J. A. Napoli, C. Yan, O. Marsalek, T. E. Markland and M. D. Fayer, ACS Cent. Sci., 2019, 5, 1269 CrossRef CAS PubMed.
  46. S. Baroni, S. de Gironcoli and A. Dal Corso, Rev. Mod. Phys., 2001, 73, 515 CrossRef CAS.
  47. R. Golesorkhtabar, P. Pavone, J. Spitaler, P. Puschnig and C. Draxl, Comput. Phys. Commun., 2013, 184, 1861 CrossRef CAS.
  48. B. Amin, N. Singh and U. Schwingenschlögl, Phys. Rev. B:Condens. Matter Mater. Phys., 2015, 92, 075439 CrossRef.
  49. W. X. Zhang, Y. Yin and C. He, Phys. Chem. Chem. Phys., 2020, 22, 26231 RSC.
  50. B. Liu, L. Wu, Y. Q. Zhao, L. Z. Wang and M. Q. Caii, Phys. Chem. Chem. Phys., 2016, 18, 19918 RSC.
  51. P. Hess, Nanoscale Horiz., 2021, 6(11), 856 RSC.
  52. B. Amin, T. P. Kaloni and U. Schwingenschlögl, RSC Adv., 2014, 4, 34561 RSC.
  53. H. T. T. Nguyen, M. M. Obeid, A. Bafekry, M. Idrees, T. V. Vu, H. V. Phuc, N. N. Hieu, L. T. Hoa, B. Amin and C. V. Nguyen, Phys. Rev. B, 2020, 102, 075414 CrossRef CAS.
  54. C. V. Nguyen, M. Idrees, H. V. Phuc, N. N. Hieu, N. T. T. Binh, B. Amin and T. V. Vu, Phys. Rev. B, 2020, 101, 235419 CrossRef CAS.
  55. M. Liao, P. Nicolin, L. Du, J. Yuan and G. Zhang, Nat. Mater., 2022, 21, 47 CrossRef CAS PubMed.
  56. R. Yuan, J. A. Napoli, C. Yan, O. Marsalek, T. E. Markland and M. D. Fayer, ACS Cent. Sci., 2019, 5, 1269 CrossRef CAS PubMed.
  57. Y. Zhao, Q. Tan, H. Li, Z. Li, Y. Wang and L. Ma, Sci. Rep., 2024, 14(1), 10698 CrossRef CAS PubMed.
  58. R. Duflou, G. Pourtois, M. Houssa and A. Afzalian, npj 2D Mater. Appl., 2023, 7(1), 38 CrossRef CAS.
  59. R. C. Andrew, R. E. Mapasha, A. M. Ukpong and N. Chetty, Phys. Rev. B: Condens. Matter Mater. Phys., 2012, 85, 125428 CrossRef.
  60. M. Born, K. Huang and M. Lax, Am. J. Physiol., 1955, 23, 474 CrossRef.
  61. G. Barik and S. Pal, Phys. Chem. Chem. Phys., 2020, 22, 1701 Search PubMed.
  62. X. Zhu, H. Jiang, Y. Zhang, D. Wang, L. Fan, Y. Chen, X. Qu, L. Yang and Y. Liu, Molecules, 2023, 28, 5607 CrossRef CAS PubMed.
  63. C. Lei, Y. Ma, X. Xu, T. Zhang, B. Huang and Y. Dai, J. Phys. Chem. C, 2019, 123, 23089 CrossRef CAS.
  64. Z. Ben Aziza, H. Henck, D. Pierucci, M. G. Silly, E. Lhuillier, G. Patriarche, F. Sirotti, M. Eddrief and A. Ouerghi, ACS Nano, 2016, 10, 9679 Search PubMed.
  65. D. Pierucci, H. Henck, J. Avila, A. Balan, C. H. Naylor, G. Patriarche, Y. J. Dappe, M. G. Silly, F. Sirotti, A. C. Johnson, M. C. Asensio and A. Ouerghi, Nano Lett., 2016, 16, 4054 CrossRef CAS PubMed.
  66. X. Yue, J. Fan and Q. Xiang, Adv. Funct. Mater., 2022, 32, 2110258 Search PubMed.
  67. S. H. Mir, V. K. Yadav and J. K. Singh, ACS Omega, 2020, 5(24), 14203 CrossRef CAS PubMed.
  68. M. Kiguchi, M. Nakayama, K. Fujiwara, K. Ueno, T. Shimada and K. Saiki, Jpn. J. Appl. Phys., 2003, 42, L1408 CrossRef CAS.
  69. U. Khan, B. Ali, H. Ullah, M. Idrees, C. Nguyen and B. Amin, Micro Nanostruct., 2024, 187, 207765 CrossRef CAS.
  70. X. Qin, W. Hu and J. Yang, Phys. Chem. Chem. Phys., 2019, 21(42), 23611 RSC.
  71. C. Strobel, C. A. Chavarin, M. Knaut, M. Albert, A. Heinzig, L. Gummadi, C. Wenger and T. Mikolajick, Nanomaterials, 2024, 14(13), 1140 CrossRef CAS PubMed.
  72. P. Vabbina, N. Choudhary, A.-A. Chowdhury, R. Sinha, M. Karabiyik, S. Das, W. Choi and N. Pala, ACS Appl. Mater. Interfaces, 2015, 7(28), 15206 Search PubMed.
  73. Q. Li, J. Yang, R. Quhe, Q. Zhang, L. Xu, Y. Pan, M. Lei and J. Lu, Carbon, 2018, 135, 125 Search PubMed.
  74. A. M. Cowley and S. M. Sze, J. Appl. Phys., 1965, 36, 3212 CrossRef CAS.
  75. C. Kim, I. Moon, D. Lee, M. S. Choi, F. Ahmed, S. Nam, Y. Cho, H. J. Shin, S. Park and W. J. Yoo, ACS Nano, 2017, 11, 1588 CrossRef CAS PubMed.

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