Open Access Article
Matjaž
Malok
*ab,
Janez
Jelenc
a,
Anja
Pogačnik Krajnc
ab and
Maja
Remškar
a
aSolid State Physics Department, Jozef Stefan Institute, Ljubljana, Slovenia. E-mail: matjaz.malok@ijs.si
bFaculty of Mathematics and Physics, University of Ljubljana, Ljubljana, Slovenia
First published on 4th November 2025
Molybdenum disulfide (MoS2) has emerged as a promising material for next-generation electronics and optoelectronic devices. MoS2 nanotubes (NTs) and their collapsed ribbon-like shapes (collapsed NTs) synthesized via chemical vapour transport (CVT) under chemical equilibrium typically exhibit low structural defect densities. However, defects and surface damage can arise during device fabrication or operation, leading to a significant degradation in performance, stability, and operational lifetime. These imperfections also induce hysteresis, which adversely affects the device switching behaviour. While the influence of charge trapping at the MoS2/substrate interfaces, on the MoS2 surface, and at intrinsic defects, such as sulfur vacancies and dangling bonds, on device performance has been extensively studied, MoS2 NTs, with their unique curved morphology, introduce additional charge-trapping mechanisms not observed in planar MoS2 structures. In this work, a combination of scanning tunnelling microscopy (STM), Kelvin probe force microscopy (KPFM), and conductive atomic force microscopy (c-AFM) was employed to examine how structural irregularities, including terminated layers, surface-grown flakes or NTs, and highly strained areas, affect charge injection, redistribution, and the resulting effects on electrical characteristics in collapsed NTs. The results reveal that structural defects act as charge traps, scattering centres, and transport barriers, giving rise to a reduced carrier mobility, localized charge accumulation, and spatially inhomogeneous charge distribution. These findings underscore the crucial role of structural and electrical characterization with nanoscale resolution in the design of defect-tolerant, high-performance devices based on transition metal dichalcogenides (TMDs).
MoS2 exhibits diverse morphologies, including hollow and collapsed nanotubes (NTs).3 Among various synthesis methods,4 chemical vapour transport (CVT) under near-thermodynamic equilibrium condition produces high-quality crystalline structures.5 Collapsed NTs form during growth, adopting a ribbon-like morphology with rounded edges and a chiral 2Hb stacking sequence.6 The degree of structural collapse varies from slightly deformed tubes, which retain a small interlayer gap6 to fully collapsed NTs resembling a single-crystal structure.7 At edges of the collapsed NTs, the molecular layers are partially split, but despite a significant bending, they remain intact7–9 and form rounded edge shapes.
Electronic structure calculations based on density-functional tight-binding (DFTB) theory indicates that MoS2 NTs remain semiconducting regardless of diameter.10 The zigzag NTs typically exhibit a small direct bandgap, whereas the armchair NTs may display either direct or indirect gaps depending on their chirality. The bandgap is dependent on geometry and is generally smaller than that of bulk MoS2. A tensile strain further modifies the electronic structure, as the bandgap closes completely under 11% elongation, inducing a semiconductor-to-metal transition.11
MoS2 demonstrates highly anisotropic conductivity due to its layered structure.12 The in-plane conductivity exceeds the out-of-plane conductivity by more than three orders of magnitude.13,14 Conductivity increases with decreasing flake thickness and varies with crystal orientation. In the zigzag direction, MoS2 exhibits higher conductivity than in the armchair direction due to a wider band gap along the armchair direction and differences in depletion regions.15 Interlayer rotation affects the out-of-plane transport, particularly for electrons, whose mobility decreases exponentially with increasing twist angle.16 Transport measurements at cryogenic temperatures indicate a quantized conductance in MoS2 NTs, suggesting transport through discrete quantum levels.17
Current flow in 2D devices is influenced by both vertical carrier distribution and resistance. While bottom-gate fields concentrate carriers near the substrate, surface n-doping (up to 10
000 times stronger than in the bulk) favours conduction near the surface.14 This effect, combined with the high out-of-plane resistance and higher mobility in upper layers, results in a dominant current flow along the top layers.12,18 Furthermore, the injected carriers can be segregated in thick MoS2 samples due to layer-dependent Fermi-level pinning. In air-exposed MoS2, electrons dominate the top layers, resulting in the n-doping at the surface, whereas holes reside in the inner layers, which remain weakly n-doped or intrinsic, minimizing the impact of the built-in potential barriers.19
CVT-grown MoS2 NTs have demonstrated potential as electric field emitters20 and as field-effect transistors (FETs).9 These NTs can confine electromagnetic fields, supporting whispering gallery modes (WGM),21 which evolve with changes in the NT's cross-section.22 Additionally, they exhibit intrinsic piezoelectricity, enabling direct coupling between mechanical deformation and electronic response.23
Despite this potential, the device performance is challenged not only by inherent contact limitations,24,25 but also by pervasive charge trapping. This phenomenon is a dominant factor in device degradation, significantly impairing performance, stability, reliability, and lifespan, while it also induces a performance-altering hysteresis.26,27
In this study, scanning tunnelling microscopy (STM), Kelvin probe force microscopy (KPFM), and conductive atomic force microscopy (c-AFM) are employed to investigate how surface defects, such as terminated layers, surface-grown structures, and highly strained edges, influence a charge spreading across collapsed NTs, and to understand the subsequent implications for the MoS2-based electronic device performance.
Fig. 2c shows the initial state before charge injection. First, electrons were injected into the rhomboidal region. As shown in Fig. 2d, the CPD decreased from the initial 420 mV to 190 mV, but this change was confined to the rhomboidal region. Charge spreading was limited by the terminated MoS2 edges on the left, bottom, and top, as well as by the NT present on the surface. In 24 h (Fig. 2e), the CPD of the rhomboidal region increased to 350 mV, while the CPD of the surrounding NT area decreased to 380 mV. Then, holes were injected into the rhomboidal region. As shown in Fig. 2f, the CPD of the rhomboidal flake increased from 370 mV to 800 mV. In the next 24 h (Fig. 2g), the CPD of both the rhomboidal region and surrounding areas returned to their pre-injection value (around 420 mV). These results demonstrate that the terminated layers, as well as surface-grown structures such as collapsed NTs, effectively restrict a rapid charge spreading across the surface.
To verify that the NT situated on the collapsed NT surface restricted charge spreading, the charge injection experiment was repeated after the NT was inadvertently removed (SI Fig. S1) from the rhomboidal region by the AFM tip. Although the NT was largely removed, some remnants remained on the surface, affecting subsequent scanning. Fig. 3 presents CPD profiles of the same area recorded before (bright lines) and after (dark lines) injection of electrons at −8 V bias, with green lines corresponding to the case when the NT was still present, and red lines after its removal (SI Fig. S2). When the NT was in place, the injected charge did not propagate beyond it, indicating that the NT acted as a physical barrier to charge distribution. After the NT was removed, the charge spread into the area previously occupied by the NT (Fig. 3, red). However, the CPD change right from the former NT location was less pronounced than on the left side. This reduced CPD change is likely due to residual fragments of the NT, which continued to partially obstruct charge transport across the surface.
In contrast to the previously described rhomboidal flake, which remained connected to the bulk at its bottom and at the upper side, Fig. 4a shows a triangular and folded flake (2 nm thick), which was completely separated from the side surface of a collapsed NT on which it was laying. The KPFM image (Fig. 4a) reveals that the work function (WF) at the edges of the flake was lower than that of its central surface. Both edges of the fold, upper (A) and lower (B) one, are also resolved as dark lines corresponding to decreased WF. Fig. 4b presents the extracted I–V spectra from the flake's surface, its edges, and the underlying collapsed MoS2 NT surface. The I–V spectra indicate that the edges of the flake exhibited a higher conductivity than both the flake surface and the collapsed NT surface. While the surfaces displayed n-type behaviour, the edges exhibited p-type behaviour.
In 24 h after injection of electrons, the CPD of the folded region (Fig. 5c) exhibited only partially recovered towards its original level, in contrast to the nearly complete recovery observed after injection of holes (Fig. 5e). Notably, no significant change in CPD was detected on the adjacent regions until the following day. These observations suggest that the highly curved longitudinal edges of the collapsed NT, enclosing the folded segment, acted as an effective potential barrier that confined charge during and immediately after injection by suppressing the lateral charge diffusion.
After injection of holes at +8 V, the I–V spectrum at negative voltages did not change, but at positive voltages it was suppressed with regard to the spectrum before the injection. This means that conductance of holes was slightly reduced. After injection of electron at −8 V, the conductance of electrons was slightly reduced, but conductance of holes was strongly enhanced. The I–V characteristics were virtually shifted towards right due to injection of holes and towards left due to injection of electrons.
MoS2, like many other layered materials, exhibits a strong electrical anisotropy, as its in-plane conductivity (σ∥) exceeds the out-of-plane conductivity (σ⊥) by more than three orders of magnitude.12,14 This anisotropy strongly favours in-plane current flow over the vertical transport across the vdW-stacked layers. Considering charge injection at the centre of a 3 µm wide and 17 nm thick NT (Fig. 1a), the ratio of perpendicular to lateral current can be estimated:
However, the experiments have shown that the injected charge accumulates predominantly on the NT surface rather than being transferred efficiently into the underlying silicon substrate. This behaviour can be attributed to the structural imperfections such as defects, dangling bonds, or adsorbates that act as charge traps;14 the presence of a Schottky barrier at the MoS2–Si interface that impedes charge transfer;29–31 and local structural defects, including terminated layers and highly strained regions, that can confine charge transport laterally. Collectively, these factors form potential barriers that govern both the charge distribution and current pathways within the collapsed NT.
Fig. 2 demonstrates how terminated layers of surface-grown flakes can hinder charge spreading. Edge states strongly influence carrier transport, as MoS2 zigzag edges exhibit vanishing band gaps with metallic states running parallel to the edges, whereas armchair edges retain semiconducting behavior,32–34 Passivation of the dangling bonds further modifies these electronic properties.35,36 As a result, edges can form natural p–n junctions, with p-doped edge states (originating from the dangling bonds) contrasting with the n-doped basal plane,37,38 which potentially restrict the charge transfer across the edges. Initially, the injected charge remains confined within the top MoS2 layer due to a weak interlayer coupling and electrostatic self-screening, which create a potential barrier for the interlayer transfer. However, as shown in Fig. 2, initially localized charge eventually redistributes across the entire collapsed NT surface. This redistribution may occur via charge hopping though defects or impurities,39 a slow quantum tunnelling across energy barriers, or diffusion driven by local electric fields.
Regarding the significant influence of edges on electronic transport, their properties were thoroughly characterized. Fig. 4a reveals that the CPD at edges is lower than in the surrounding basal plane, reflecting the modification of both the WF and the local density of states by highly charged edges.32,40,41 The I–V spectra presented in Fig. 4b indicate that the flake edges exhibit higher conductivity than both the flake surface and the collapsed NT surface. This behaviour, consistent with previous findings,32,33 arises because the terminated layers enable electron transport between the STM tip and the flake to occur predominantly in-plane relative to the MoS2 layers, whereas transport on the surfaces away from the edges is largely limited to out-of-plane conduction.
The presence of a MoS2 NT atop the collapsed NT further influenced charge transport by hindering the lateral carrier spreading (Fig. 3), analogous with the behaviour observed in carbon NTs, which can trap charges and act as confining barriers.42
STS studies (Fig. 7) reveal charge-injection-induced modifications of the sample's local electrostatic potential. Injection of holes corresponding to a net electron removal, causes a suppression of the electronic density of states relative to the tip. This realignment enhances tunnelling at negative tip biases by providing more empty sample states for electrons from the tip, while simultaneously suppressing tunnelling at positive biases, as the occupied sample states are pushed to lower energies, thus requiring a higher tip voltage for electron extraction.
![]() | ||
| Fig. 7 STS measurements: I–V characteristics of a collapsed MoS2 NT before charge injection (black), after injection of holes at +8 V (red), and after injection of electrons at −8 V (blue). | ||
Conversely, negative charge injection, corresponding to net electron addition, induces an upward shift of the density of states. This new alignment suppresses electron injection at negative tip biases due to lack of empty states, resulting in a leftward shift of the I–V curve in this regime. At positive biases, tunnelling is strongly enhanced because the occupied sample states move closer to the tip's Fermi level, steepening the I–V slope and contributing to the overall leftward shift of the I–V characteristics.
When materials with different WFs come into contact, charge redistribution occurs until Fermi level alignment is achieved, producing local electric fields, band bending at the interface, and potential formation of electron depletion and accumulation regions, which can affect conductivity.43 Additionally, the NT may introduce a local electrostatic gating effect doping the underlying collapsed NT and forming a conductivity bottleneck.44 The formation of moiré patterns between MoS2 nanoribbon and the MoS2 sheet can induce lattice strain on the nanoribbon surface45,46 modifying material electronic properties47–52 and introducing periodic potential barriers that scatter charge carriers.
In the present configuration, with the MoS2 NT atop the collapsed NT, CPD images (Fig. 2c) show comparable CPD values for the NT and the underlying collapsed NT. Since both electron and hole conductivities were confined, WF mismatch and the associated gating effect are unlikely to be the primary cause of reduced charge spreading. Instead, a more plausible explanation is the presence of imperfections at the vdW interface, such as voids, wrinkles, or trapped adsorbates, which act as scattering centres and limit carrier mobility. The persistence of CPD anomalies even after the NT was removed supports this interpretation.
When charge was injected into the folded region bounded by highly strained edges, lateral spreading was significantly restricted (Fig. 5). Although strain in MoS2 is known to induce a semiconductor-to-metal transition by narrowing the bandgap and enhancing conductivity,11 our CPD measurements showed no lateral charge flow beyond the bounded regions. This suggests the presence of additional energy states, likely caused by microcracks or other structural disruptions at the strained edges, that effectively terminate conductive pathways. Moreover, the intrinsic piezoelectric response previously demonstrated in collapsed MoS2 NTs,7 which could be introduced by charge injection and formation of potential gradient, may induce permanent changes in molecular configuration via plastic deformations. Such rearrangements can suppress hopping transport within the conduction channel or even trigger a metal-to-insulator transition.53 These interpretations are supported by c-AFM measurements, which reveal increased conductivity at edges relative to flat regions, while also revealing localized zones of zero conductivity.7 Interestingly, CPD values within the folded region partially returned to the initial values over time, whereas adjacent areas remained unchanged. This behaviour suggests that out-of-plane charge redistribution takes place preferentially through MoS2 layers toward the substrate rather than laterally across structurally compromised edges.
The observed charge transport anomalies in collapsed MoS2 NTs have significant implications for device performance and reliability.53–56 It has been shown that charge trapping adversely affects electrical performance by degrading stability and reliability through reduced mobility, threshold voltage shift, and transient current decay,27 while also dictating the switching behaviour via significant hysteresis dependent on the gate sweep history and through the induction of memory-like effects from floating gates.26,57 Conversely, the same charge trapping mechanisms can be used for applications like analog synapses,58 non-volatile memory and neuromorphic computing,26,59,60 as well as multi-bit data storage.61
Charge trapping occurs at multiple physical sites, broadly categorized into interface, surface, and bulk defect-related sites. A substantial contribution originates from the interface between the MoS2 channel and the gate dielectric.27,57 For example, the SiO2 surface possesses a high density of dangling bonds that create interface traps, causing strong dielectric scattering.27 Persistent hysteresis in suspended devices confirms that trapping is not solely a substrate effect,62 but also the MoS2 surface itself represents a major trapping site.62 Ambient molecules such as H2O and O2 adsorb onto the surface, leading to water-assisted charge trapping.26 Furthermore, fabrication residues like carbon contamination, can also act as trap states.63 Intrinsic defects within the MoS2 lattice, such as sulphur and oxygen vacancies, create localized states that trap charge.26,63 Furthermore, extended structural defects, like pores, ripples, and nanobubbles formed during processing, contain dangling bonds or trapped ambient molecules, creating quasi-isolated regions that can store charge for extended periods.59,61,64
On the other hand, the geometry of MoS2 NTs with rounded edges enables new charge trapping mechanisms. Strained edges can act as effective barriers to lateral charge transport, confining charge within top layers and approximating the behaviour of a planar MoS2. Furthermore, terminated layers and other surface irregularities can impede or redirect current, resulting in a non-uniform charge distribution. Localized charge accumulation enhances electric field gradients, increases current density in confined regions and potentially accelerates device degradation. Such effects may form large electrostatic barriers that impair electrical transport,65 underscoring the importance for comprehensive nanoscale electrical and structural characterization studies prior to a device integration to ensure predictable performance and reliability.
All data supporting the findings of this study are provided in the main article and the supplementary information (SI). Supplementary information is available. See DOI: https://doi.org/10.1039/d5na00771b.
| This journal is © The Royal Society of Chemistry 2025 |