Matjaž
Malok
*ab,
Janez
Jelenc
a 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 28th April 2025
Molybdenum disulfide (MoS2) is a promising material for future high-performance and ultra-low-power electronics. Growth from a vapor phase at chemical equilibrium enables the production of crystals possessing a relatively low density of structural defects. Besides thin MoS2 flakes, MoS2 nanotubes (NTs) and collapsed NTs in the shape of nanoribbons (NRs) are also synthesized in the same growth process. Here, we present the first study on the structural and electrical properties of the NRs. High resolution electron microscopy revealed a chiral structure of the NRs with no peculiarities at the inner interface where both walls are in contact. In contrast, resonant Raman spectroscopy revealed the presence of bands typical of a few layers thick MoS2, suggesting that some of the layers of the NR are partially split. Contact current imaging spectroscopy (CCIS) revealed longitudinal wrinkles on the NR surface, with elevated regions found to be more conductive than the depressed areas. The edges of the NR, where molecular layers are strongly curved but not broken, exhibit varying conductivity. While some parts exhibit zero conductivity, others show much higher conductivity than the central part of the NR, suggesting an electron confinement effect. Charge injections strongly altered the NR's work function and induced changes in the NR's topography. The surface wrinkling was intensified, and the NR tended to rotate around its longitudinal axis. This rotation is explained as the reverse piezoelectric effect.
CVT-grown MoS2 NTs, as well as those that have collapsed into NRs, have been investigated as channels in field-effect transistors, demonstrating n-type behaviour with ON/OFF current ratios exceeding 103,7 and have also been utilized as electron field emitters.8 Due to a low density of defects, the transport properties of MoS2 NTs studied at cryogenic temperatures indicate quantized single quantum-level transport.9 Furthermore, studies have shown that MoS2 NTs prepared by the CVT method can confine electromagnetic fields within their walls, leading to the appearance of whispering gallery modes (WGMs).10 If the NTs are flattened into a NR, the degree of WGM peak splitting is expected to vary depending on the aspect ratio of the NR's cross-section. Consequently, they can be employed as optical resonators for self-radiating light, with the ability to tune the frequency at which the radiation is amplified.11
Despite the recent interest in MoS2 monolayers and NTs, the impact of their collapse on electrical properties remains largely unexplored. Conductive atomic force microscopy (CAFM) has been used to investigate the electrical properties and Schottky barrier height at MoS2 interfaces with metals,12–15 with Si16 and GaN,17 as well as in the epitaxial MoS2–WSe2–graphene heterostructures.18 Additionally, CAFM has been employed to probe grain boundaries in multilayer MoS219 and to study the current injection mechanism at the surface of MoS2 thin films.20 Charge transport properties investigated by charge injection and Kelvin probe force microscopy (KPFM) have been reported for flat MoS2, including few-layered MoS2 films and flakes,21,22 and for a single MoS2 NT.23 On the other hand, extensive research has focused on carbon NTs. The effect of collapse on the atomic and electronic structure of carbon NTs has been characterized using tapping mode AFM24 and scanning tunneling microscopy (STM).25,26 By recording tunneling spectra at various distinct locations, researchers observed that deformation induces an electronic band gap in an otherwise metallic NT due to altered interlayer interactions.26 Furthermore, carbon NT FET has been characterized using CAFM.27
Recognizing the significant implications of the collapse of the MoS2 NTs for electronic applications, we investigate the electronic properties of the MoS2 NRs. These structures, featuring mutually rotated walls and edges, where the S–Mo–S molecular layers remain intact but are highly curved, represent an intriguing anisotropic system. Employing STM and CAFM, we examined their electronic characteristics. Additionally, charge transport properties were studied through charge injection experiments utilizing a conductive AFM probe in STM mode for injection and KPFM for observation of work function (WF) modulation.
Structural analysis of NRs was performed using HRTEM and transmission electron diffraction (TED). Narrow NRs allow observation of the interface between both walls without making a cross-section. The NR shown in Fig. 2a is twisted along its axis. A high-resolution TEM image of the narrowest area of the NR (24 nm wide), where MoS2 layers are parallel to the electron beam, reveals that the interface between both walls, composed of 18 molecular layers, is without peculiarities, making the NR appear as a single-crystal structure (Fig. 2b). The corresponding TED pattern (Fig. 2c) corresponds to a superposition of electrons scattered in two zone-axes: [010] and [110]. The diffraction peaks (hkl) in the [010] zone-axis satisfy the rule l = ±(2n + 1) that corresponds to the 2Hb polytype stacking. In the projection of the side (10l) reflections onto the [00l] axis, the spots appear exactly between (00l) spots.29 The widest part of the NR, where MoS2 layers in the central part are perpendicular to the electron beam, measures 65.5 nm in width. The TEM image (Fig. 2d) reveals walls with a thickness of 12 nm ± 1 nm and moiré patterns in the central part of the NR, formed by a superposition of two mutually rotated walls. Two chiral angles are identified in the electron diffraction pattern (Fig. 2e): 9.5° (α) and 2° (β). Additionally, some wrinkles are visible in the central part of the NR.
![]() | ||
Fig. 2 TEM images with corresponding electron diffraction patterns of a helically twisted MoS2 NR (a), taken at the narrowest part (b and c) and at the widest part (d and e) in the projection view. |
Direct resonant Raman scattering reveals additional bands compared to non-resonant scattering, as previously reported.31–42 The most intensive Raman bands were observed at 179 cm−1, 419 cm−1, 456 cm−1, 464 cm−1, and 642 cm−1. The absence of bands around 230 cm−1 indicates a low concentration of defects in the sample.31–35 On the low-frequency side of the E2g1 band at 383 cm−1, its Davydov pair, the Raman-inactive E1u2 mode at 379 cm−1, is observed. The small frequency split of the Davydov pairs indicates a weak interlayer interaction.36,37 The intensity of the 379 cm−1 band exceeds the intensity of the peak at 383 cm−1. The 419 cm−1 band was previously observed at 430 cm−1 in MoS2 single crystals,36 and it was reported to downshift to 420 cm−1 in nanoparticles.32,36 This peak is attributed to a two-phonon Raman process involving the successive emission of a dispersive quasi-acoustic (QA) phonon and a dispersionless TO phonon.36 The most intense band is observed around 460 cm−1, which is a superposition of the two bands at 456 and 464 cm−1.32 In the non-resonant Raman scattering, these bands at 456 cm−1 and 464 cm−1 were barely visible.
The intensity of some bands varies across the NR. Raman mapping (Fig. 3b) shows an increased intensity of the band at 409 cm−1 at the edges (red) of the NR (Fig. 3c) compared to the central area (green). Such variation in intensity between the edges and basal plane of MoS2 crystals has been reported previously,43 though using a laser with a wavelength of 785 nm.
Fig. 4a shows the CCIS images of the full width of the NR at +1.6 V, revealing distinct regions oriented parallel to the NR edge with varying conductivity. At the right-hand edge, two regions of maximum conductance (Rinner and Router) are separated by a valley (Rmiddle), where the conductance is similar to that of the central NR (Rcentral). This can be seen in the cross-section profiles in Fig. 4b. A similar pattern is observed at the left edge, where two maxima (Louter and Linner) are separated by a valley (Lmiddle).
Averaged I–V spectra for the left and right edges are shown in Fig. 4c and d, respectively. The conductance of the NR is semi-conductive and slightly higher at positive biases. It is significantly greater at the right edge compared to the left one. This difference is likely due to variations in the contact interaction between the NR and the substrate, with the right edge being more exposed to the AFM probe, thereby revealing its electronic structuration. The local conductance at the right edge more than doubled at Rinner and tripled at the very edge (Router) relative to the central region, whereas the left edge showed lower conductance compared to the central area.
The electronic properties of the NR's edge were further analysed by CAFM, applying alternating biases to the CAFM tip while simultaneously measuring the current between the sample and the tip. A c-AFM image of the NR edge is shown in Fig. 5a. The current images recorded at positive (+0.3 V) and negative (−0.6 V) voltages are shown in Fig. 5b and c, respectively. The thickness of the NR, extracted from the topography profiles (Fig. 5d, black and grey), was approximately 34 nm at both polarities, revealing a good interaction strength with the substrate, which prevented any shift of the NR due to electric force. Electronic details of the edge became visible in the current images, with cross-section profiles shown in Fig. 5d as red and blue lines for positive and negative voltages, respectively. At the border between the NR's central part and its edge (at −25 nm in Fig. 5d), a belt of zero conductivity is visible for both polarities. Between this belt and the very edge, additional structuring is observed as lines of increased conductance. Horizontal modulations (Fig. 5b and c) are artifacts of the scanning process.
The c-AFM image and the corresponding CCIS images of the NR's central part are presented in Fig. 6a and b, respectively. The surface is not completely flat but exhibits slight rippling. These ripples, with an irregular periodicity of around 100 nm, are shallow, with a height difference between the elevated and depressed areas of less than 0.5 nm. The presence of these ripples influences the local conductance such that the elevated areas have higher electric conductance than the depressed ones, as demonstrated in the CCIS image and corresponding line profiles shown in Fig. 6c. Averaged I–V profiles derived from the CCIS image are shown in Fig. 6d. Conductivity in both elevated (gold) and depressed (green) regions is greater at positive voltages. The relative difference in conductivity is more pronounced at voltages lower than −150 mV, where elevated regions exhibit 21% greater conductivity compared to the depressed areas. As the voltage increased from −150 mV to 250 mV, the relative difference in conductivity decreased linearly from 21% at −150 mV to approximately 6% at 250 mV (ESI 1†).
Some bulbs were also visible in the topography image of one MoS2 NR, but they were not visible in the KFPM image. The results of a detailed investigation of these bulbs are presented in ESI 2.†
![]() | ||
Fig. 7 Topography (top) and CPD (bottom) profiles, extracted over the NR, showing changes in topography and WF due to injection of electrons. |
Fig. 8 shows the result of four injections of holes at +8 V bias, performed subsequently with 40 min time delays. The nc-AFM and KPFM images of the NR before the first injection, after the last injection, and 24 h after the last injection are shown in ESI 4.† Changes in topography are mainly visible at the NR edges, which appear wider as a new belt appeared on the right-hand side of the NR after the 2nd injection, as seen in the topography profile (Fig. 8, +8 V (2)). The number of wrinkles on the NR surface decreased from 5 in the initial state to 4, with an additional one attributed to the newly occurred belt. With additional injections, it appears that the NR rotated slightly around its longitudinal axis in a clockwise direction. This is evidenced by the decrease in height of the belt on the right-hand side, which disappeared completely after the fourth injection (Fig. 8, +8 V (4)). However, during this injection, a new belt formed on the left-hand side and remained visible until the following day.
![]() | ||
Fig. 8 Topography (top) and CPD (bottom) profiles, extracted over the NR, showing changes in topography and WF due to injection of holes. |
In contrast to the gradual decrease of WF with the injection of electrons, the WF did not increase monotonically with the injection of holes at +8 V bias, as one might have expected. After the first injection, the CPD decreased, while the second injection caused an increase in the CPD to a maximum value of 450 mV. A depression in the CPD profile deepened to ≈70 mV with a simultaneous increase in CPD at the NR edges. The 3rd and the 4th injections caused a decrease in CPD, while the depression in the CPD profile became less pronounced. In the next 24 h, the CPD increased, and the depression in the centre of the NR became visible again.
The overall conductivity of the NR was found to be semi-conductive at room temperature and higher for positive voltages. This confirms a pristine n-type semiconducting nature of MoS2,44 but with no open gap at room temperature all over the central part of a NR. The collapse strongly curved and consequently strained the molecular layers at the NR edges, causing significant differences in conductivity in comparison with the central part. The NR edges were found to be semi-metallic, which can be explained by the incorporated strain as it was theoretically predicted by E. Scalise et al.,45 and experimentally demonstrated for 2D MoS2 layers where moderate strain values (∼2%) can already trigger an indirect bandgap transition, induce a finite charge carrier density,46 and enhance carrier mobility.47 From the absence of vibrational modes at 154 cm−1, 219 cm−1 and 327 cm−1, it can be concluded that the presence of metallic 1T-MoS2 is less likely.48 At the transition between the central part and the edges, where a strong conductance modulation was observed, some narrow belts showed zero conductance (Fig. 6d). The difference between the left and right edges can be explained by different contacts between the NR and the substrate, which can alter their electrical characteristics.49 A relatively weak vdW contact interaction with the substrate can enable a buckling of the NR to one side influenced by a torsional strain of the chiral lattice structure or a torsional component of a reverse piezoelectric effect, which will be discussed later.23
All the microscopic techniques used (SEM, c-AFM and nc-AFM) revealed that the surface of the NRs is not flat but slightly rippled with a roughness of less than one molecular layer. The CCIS investigation of these ripples revealed that the elevated parts are more conductive than the depressed ones. This is in accordance with the KPFM investigation of atomically thin MoS2, where space-dependent surface potential and a non-uniform charge distribution were attributed to local strain in the ripples.50 A different differential conductivity in valleys, hills, and flat parts of MoS2 monolayers (MLs), investigated by STM, was explained by bandgap changes resulting from substrate-induced local bending strain.51 For a relatively thick MoS2 NR (17 ± 1 nm) compared to an MoS2 ML, interaction with the Si substrate is less likely. However, one must also consider the interaction with the second wall of a former NT before its collapse into an NR shape. A contact between both walls in the central part of the NR must satisfy one of the possible polytypic stackings of MoS2 (2H, 3R), which can induce a local strain at the interface and lead to the formation of ripples.
Besides ripples, bulbs were observed in the central part of the NR. The conductivity of these bulbs was higher than that of the surrounding flat regions, particularly at positive voltages. Distinct electronic properties of the bulbs and the flat regions between them have already been documented in collapsed carbon NTs, where the central semi-flat region maintains a finite density of states (DOS), displaying metallic behaviour, while the bulbs exhibit a bandgap opening with zero DOS, indicative of semiconductor-like characteristics.25,52 This phenomenon is attributed to quantum confinement and charge transfer interactions between the bilayer graphene-like region and the NT-like edges, which exhibit weaker inter-wall interactions compared to the flattened region.52
Upon investigating charge transport properties and their retention, it was demonstrated that both negative and positive charges could be injected into the NRs, which is consistent with previous research on MoS2 thin films.21,22,53 The injected charge was trapped at the structural and electrical defects and at the interface between the NR and substrate, where a kind of 1D potential well is formed due to the different work functions of two semiconductors, forming a heterojunction. Since the 17 nm NR effectively screens charges from the MoS2–SiO2 interface54 and given that MoS2 exhibits significantly higher in-plane conductivity compared to interlayer conductivity, where the vdW interaction between layers acts as a tunnel barrier introducing resistance,55 the observed change in CPD originates from electrons trapped at the surface of the NR. Taking this into account, the calculated charge density (ESI 5†) after the last injection of electrons was 0.21 mC m−2. Injection of negative charge (electrons) decreases WF and injection of positive charge (holes) increases the NR's WF. A monotonic decrease in WF was observed with subsequent injection of electrons, whereas during injection of holes (Fig. 8), a monotonic increase in WF was not achieved. The reason for this could be the formation of new electron transport paths in the material, caused by the formation of structural defects during the charge injection.56 Changes in work function, as well as in topography, remained stable at least for 24 h, as previously reported.53 As MoS2 is an n-type semiconductor due to electron-donating native defects,44 it would be expected that the retention of holes would be less durable than that of electrons. However, this was not the case. This discrepancy may be attributed to charge trapping, not only at the defects but also at the potential barrier at the NR–substrate interface or between the two walls of the NR. The charge distribution on the surface was not perfectly homogeneous, as a slight depression or elevation formed in the central part. A similar but more pronounced effect was observed during charge injection into non-collapsed MoS2 NTs.23 After injection of electrons, the work function decreased in the central part of the NT, falling below that of the MoS2 single crystal used as the substrate, while at the edges, trapped electrons and/or tensile strain increased the work function, making it higher than that of the substrate. The opposite trend was observed for hole injection. Furthermore, the changes in the NT's contact potential difference were much larger (400 mV) compared to the MoS2 NR (100 mV). A possible reason for this is the much smaller surface area of the NT than that of the NR.
Charge injection also caused a change in the NR's shape observed as wrinkles on the surface and the rotation of the NR around its longitudinal axis. After charge injection, a very stable number (5) of wrinkles appeared oriented along the NR length. They were observed already after the first injection of electrons and did not change during subsequent injections. It is important to note that these wrinkles are not visible in KPFM images and do not affect a local work function. The wrinkling of the surface can be caused by different mechanisms. First, it is possible that the wrinkling occurred during charge injection, where the external electric field drives weakly coupled dichalcogenide layers into an unstable state, making them susceptible to mechanical exfoliation57 and consequently leading to wrinkling. Additionally, surface wrinkling could result from charge-induced lattice deformations58 or unevenly distributed electrons on the surface caused by Coulomb repulsion (ESI 5†). Above all, these shape changes could result from the charge injection-induced inverse piezoelectric effect, which exhibits both radial and torsional components in chiral NTs.59 The extent of deformation depends on the diameter of individual molecular layers, meaning that different layers within the NT's wall, each with distinct circumferences and curvature energies, experience varying degrees of strain. This induces helical strain, leading to the compression of molecular layers with different radii while simultaneously causing their twisting.23
Indications of the reverse piezoelectric effect were observed as topographical modifications at the edges of the NR, which appeared wider due to the formation of a new belt on its right-hand side. With additional charge injections, the belt on the right-hand side diminished, while a new belt emerged on the left-hand side. This suggests that the NR underwent a clockwise rotation around its longitudinal axis. This rotational behaviour was observed twice in two different NRs and is attributed to the rotational component of the reverse piezoelectric effect, which induced rotation due to helical strain.23 A similar phenomenon has been reported for MoS2 NTs, where charge injection altered their shape by forming new shoulders on either side of the NT.23 Additionally, it was possible to control the NT's rotation by adjusting the polarity of the injected charge, as injection of electrons induced clockwise rotation, while injection of holes caused counterclockwise rotation.
The presented results demonstrate that charge injection into MoS2 NRs significantly affects their electrical properties, particularly surface potential and topography. Defect-free MoS2 NTs have already been identified as promising materials for quantum and electronic devices.7–9,11 However, device fabrication remains a challenge,60 as electrical contacts to 2D TMDCs have been a major limiting factor in achieving high device performance due to strong Fermi level pinning and high contact resistance.61 For planar TMDCs, remarkable progress has recently been made in overcoming these barriers.62–64 However, in the case of MoS2 NTs and NRs, where reduced geometry adds further complexity, several approaches have recently been proposed to mitigate Schottky barrier formation.60 Our study demonstrates that charge injection by the AFM tip, which mimics the effect of MoS2 NR contacts with metals, alters the surface potential of MoS2 NRs. Since surface potential has been reported to influence interfacial charge transfer and transport behavior in MoS2–metal contacts,65 this factor should be carefully considered when designing contacts for NRs. Additionally, the curved edges of NRs exhibit different electrical properties than the flat central region, which could be exploited for contact optimization, as edge contacts in 2D-MoS2 have been shown to outperform flat surface contacts.66 Furthermore, we observed that charge injection can alter the shape of NRs, potentially creating gaps between the NR and contact materials. This disruption may prevent the formation of a continuous crystalline interface, ultimately reducing contact quality.60 These findings emphasize the need for further research into the electrical properties of MoS2 NTs and NRs, as variations in their electrical characteristics could have a significant impact on the performance of electronic devices.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5nr00284b |
This journal is © The Royal Society of Chemistry 2025 |