Next Article in Journal
Low-Platinum-Content Exchange-Coupled CoPt Nanoalloys with Enhanced Magnetic Properties
Next Article in Special Issue
Self-Powered Broadband Photodetector Based on NiO/Si Heterojunction Incorporating Graphene Transparent Conducting Layer
Previous Article in Journal
Comparative Studies of the Structural and Physicochemical Properties of the First Fullerene Derivative FD-C60 (Fullerenol) and Second Fullerene Derivate SD-C60 (3HFWC)
Previous Article in Special Issue
Asymmetric Schottky Barrier-Generated MoS2/WTe2 FET Biosensor Based on a Rectified Signal
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Multi-Layer Palladium Diselenide as a Contact Material for Two-Dimensional Tungsten Diselenide Field-Effect Transistors

1
Department Physics, Mechanics and Electrical Engineering, Montanuniversität Leoben, Franz Josef Strasse 18, 8700 Leoben, Austria
2
Laboratory of Organic Matter Physics, University of Nova Gorica, Vipavska 13, SI-5000 Nova Gorica, Slovenia
3
Department of Inorganic Chemistry, University of Chemistry and Technology Prague, Technická 5, 166 28 Prague, Czech Republic
4
Research School of Chemistry & Applied Biomedical Sciences, Tomsk Polytechnic University, Lenina ave. 30, 634034 Tomsk, Russia
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2024, 14(5), 481; https://doi.org/10.3390/nano14050481
Submission received: 15 January 2024 / Revised: 28 February 2024 / Accepted: 1 March 2024 / Published: 6 March 2024

Abstract

:
Tungsten diselenide ( WSe 2 ) has emerged as a promising ambipolar semiconductor material for field-effect transistors (FETs) due to its unique electronic properties, including a sizeable band gap, high carrier mobility, and remarkable on–off ratio. However, engineering the contacts to  WSe 2  remains an issue, and high contact barriers prevent the utilization of the full performance in electronic applications. Furthermore, it could be possible to tune the contacts to  WSe 2  for effective electron or hole injection and consequently pin the threshold voltage to either conduction or valence band. This would be the way to achieve complementary metal–oxide–semiconductor devices without doping of the channel material.This study investigates the behaviour of two-dimensional  WSe 2  field-effect transistors with multi-layer palladium diselenide ( PdSe 2 ) as a contact material. We demonstrate that  PdSe 2  contacts favour hole injection while preserving the ambipolar nature of the channel material. This consequently yields high-performance p-type  WSe 2  devices with  PdSe 2  van der Waals contacts. Further, we explore the tunability of the contact interface by selective laser alteration of the  WSe 2  under the contacts, enabling pinning of the threshold voltage to the valence band of  WSe 2 , yielding pure p-type operation of the devices.

1. Introduction

Two-dimensional (2D) materials, and especially 2D semiconductors, are emerging as ever-more promising platforms to be added into very-large-scale integration (VLSI) technologies [1,2]. This is driven by the shrinking pitch sizes required to achieve higher integration density, energy efficiency, and speed of electronic circuits [3,4]. To achieve this feat, comprehensive studies have been undertaken to offer 2D channel materials and insulators [5,6] with performance parameters comparable to those of silicon-based technologies. Both traditional ( Al 2 O 3 /HfO 2 ) and 2D insulators have shown promising results in developing complex architectures [7,8]. Similarly, a huge library of 2D semiconductors is available for the purpose of choosing p-type, n-type, or ambipolar channel materials [9,10]. In this regard, ambipolar  WSe 2  has garnered keen interest in the scientific community due to its potential applications towards complementary metal–oxide–semiconductor (CMOS) technology, solar cells, water splitting, light emitting, and gas sensing [11,12,13,14,15]. Moreover, patterned nanoribbons of  WSe 2  have been shown to offer high electrical performance and the possibility to be coupled with metallic nanoparticles, which offers exciting possibilities in optoelectronic applications and tunable catalysis [16,17]. Like other 2D materials, the properties of  WSe 2  can be tuned via thickness [18], plasma treatment [19], strain [20], and choice of contacts [21,22].
However, the development of technology-relevant metal–semiconductor interfaces remains a significant bottleneck for the integration of 2D semiconductors into VLSI [1,23]. This is also true for achieving high-quality contacts to ambipolar  WSe 2 . The existing metal electrode deposition technologies cause the degradation of the 2D materials at the contact interface by the formation of metal-induced gap states (MIGS) and defect-induced gap states. In turn, these gap states result in the formation of large barriers at the junctions, consequently lowering the device performance and increasing energy consumption [24,25]. In addition to the creation of potential barriers, MIGS also alter transport fundamentally by changing transmission around the transport gap. An example of this can be seen in metallization-induced change of the quantum limits of contact resistance in one-dimensional contacts to semiconducting graphene nanoribbons [26]. There have been several efforts to find suitable contact materials and contact deposition methods to realise the full potential of 2D material-based circuits. These include the use of edge contacts [27,28], low-work-function metals [29], ultra-high vacuum evaporation [30], buffer layers [31], self-assembled dipolar monolayers [32], and dry stamping of metal electrodes [33,34]. In particular, metallised edge contacts are commonly the best-performing technology in large-area 2D material-based devices [27,28]. However, this is not the case for one-dimensional (1D) or quasi-1D nanostructures of 2D materials such as nanoribbons and nanowires, as phosphorene nanodevices with edge contacts [35,36].
More recently, semi-metallic contacts such as bismuth, antimony, and graphene have shown promising results [37,38,39]. However, semi-metal depositions involve heating of the substrate up to 100 °C to achieve a particular orientation of the metal (Sb 011 2 ¯  on  MoS 2 ), which tends to introduce defects into heat-sensitive 2D semiconductors with ambipolar functionality, such as  WSe 2  [40,41] and black phosphorus [42]. In case of graphene contacts, the growth of graphene films as top electrodes would require very high temperatures [43], hindering direct growth as the incorporation path to the back-end-of-line (BEoL) processes. However, ambipolar 2D materials are critical for CMOS electronic applications. It is essential to develop contacts which preserve the ambipolar behaviour while providing sufficient on-state currents and  I ON / I OFF  ratios relevant for technological aspects. To address this challenge, significant efforts have been made, including the use of In and Pd contacts to  WSe 2  [44,45].
The use of  PdSe 2  as a contact material was first demonstrated by Oyedele et al., who employed defective  Pd 17 Se 15  as contacts with  PdSe 2  to demonstrate a low Schottky barrier [46] and later by Seo et al. for the realization of  PdSe 2 -based CMOS devices [47].  PdSe 2  has also been used to contact  MoS 2  in a junction field-effect transistor as a top gate due to its promising optoelectronic properties which include long-wavelength infrared photo responsivity [48,49,50,51]. With a layer-dependent bandgap in the infrared region,  PdSe 2  itself is a unique member of the transition metal dichalcogenide family with potential uses in optoelectronic devices [52]. It behaves as a semi-metal for thicknesses above 20 nm and transforms to a semi-conducting state for thinner layers [53,54]. Large-area  PdSe 2  can be grown at temperatures as low as 250 °C [55], unlike graphite, which makes it critical for BEoL integration as a van der Waals electrode material. Moreover, graphite contacts dope the  WSe 2  towards a dominant n-type electrical response, therefore disrupting the ambipolar nature of  WSe 2  [56].
In this work, we propose  PdSe 2  contacts to  WSe 2  which demonstrate high  I ON / I OFF  ratio and high on-state currents while maintaining the intrinsic ambipolar behaviour of the channel material. To further tune the behaviour of our devices, we propose localised laser treatment of  WSe 2  at the contact regions to demonstrate dominant p-type FETs with high threshold voltage stability. This can allow for the co-integration of p-type and ambipolar devices in a circuit without the need to change the contact material nor the need to introduce any dopants into the channel.

2. Materials and Methods

2.1. PdSe 2  Crystal Growth

PdSe 2  crystals were synthesised by direct reaction of elements in a quartz glass ampoule. Powder palladium (99.99%, −100 mesh, Safina, Prague-Vestec, Czech Republic) and selenium granules (99.9999%, 2–4 mm granules, Wuhan Xinrong New Material Co., Wuhan, China) corresponding to 3 g of  PdSe 2  were placed in a quartz ampoule (25 × 100 mm) with additional selenium corresponding to 1% in excess. The ampoule was melt sealed under a high vacuum (1 ×  10 3  Pa) using an oxygen–hydrogen torch and placed in muffle furnace. The ampoule was heated at 850 °C using a heating rate of 1 °C/min, and after 12 h it was cooled to room temperature at a cooling rate of 0.1 °C/min. The ampoule with formed  PdSe 2  crystals was opened in an argon-filled glove box.

2.2. Device Fabrication

Using laser lithography (DaLi, Cerklje, Slovenia) and thermal evaporation, 45 nm/5 nm stripe-like Au/Cr electrodes were patterned onto a 300 nm  SiO 2 /Si substrate. hBN flakes were used as a bottom gate oxide on top of one the Au pads. Multi-layer  PdSe 2  or crystal (kish) graphite flakes were placed on top of the  WSe 2  flakes as source and drain electrodes. Flakes of 2D materials were mechanically exfoliated from bulk single crystals using commercially available Nitto tape and polydimethylsiloxane (PDMS) Gel–Pak–DGL–X4. The flakes were selected based on optical contrast and transferred one by one to build up the devices. The thickness of the hBN used for the devices was approximately 20 nm, considering a value for the relative dielectric constant of   ϵ r  = 3.5 [57]. Further, optical microscopy, atomic force microscopy (AFM), and Raman spectroscopy measurements were performed to confirm the layer thickness, uniformity, and exact device geometries.

2.3. Electrical Characterization

Room temperature (RT) and low-temperature (78 K) electrical characterizations were performed using a Keithley 2636A Source Meter (Tektronix GmbH, Koeln, Germany) attached to an Instec probe station (Boulder, CO, USA). The samples were contacted via Au-coated Ti electrical cantilever microprobes. The Instec’s mK2000 temperature controller was used to monitor the temperature with a resolution of 0.01 K. The cooling and heating rates were 20 °C/min and 10 °C/min, respectively.

2.4. FET Figures of Merit (FOM) Extraction and Device Modelling

The off-state current ( I OFF ) was defined as the minimum in  I D ( V G )  curves, while the on-state current ( I ON ) was defined as the maximum obtained in  I D ( V G )  for the electron or hole branch. The maximum current is limited by the amount of the electrostatic field that can be applied through the back gate, and  I ON  was estimated 5 to 8 V away from the threshold voltage ( V th ). The threshold voltage was estimated by extrapolation of the linear fit to the point of intersection of the  I D  = 0 A line. The linear fit was performed in the  V G  region shifted by 2 V from the onset voltage point ( V on ) and by 4 V to 5 V from the  V on . The onset voltage point was defined as the  V G  point from which the  I D  continuously increases from the gate leakage levels (usually 0.5–2 × 10 11  A). The middle of the region between  V on  and  V th  was used to estimate the sub-threshold swing (SS) values.
Modelling of the FET output curves was performed using the ideal transistor operating in the linear regime, shifted by the  V th  via a capacitor at the gate. To model the non-ideal and non-linear behaviour of the contacts, a linear resistor and a Schottky diode were added in series to the ideal transistor. The current through the transistor was described as:  I D = ( μ C ox W / L ) · ( ( V G V th ) · V FET + ( V FET 2 / 2 ) ) . Here,  V FET  corresponds to the fraction of the total  V D  bias that is experienced by the ideal transistor,  C ox  is the area-specific gate dielectric capacitance, and  μ  is the intrinsic mobility. Upon reaching the limit of the linear regime — defined as the maximum  I D ( V FET )  value of the model — the maximum current level was kept independent of the  V FET , describing the saturation of the device. The ohmic component of the contact resistance is defined by a linear resistor, with its corresponding potential drop described as  V ohmic = R ohmic · I D . The non-linear component of the contact resistance is described by:  V junction = V thermal ln ( 1 + I D / I 0 ) ) . Here,  V thermal = k B T / e  and  k B  stands for the Boltzmann’s constant, T is fixed to the set temperature of the experiment, and e is the unit charge.  I 0  represents the reverse current of the Schottky diode. Considering that  V D = V FET + V ohmic + V junction  the system is solved in a self-consistent manner using three fitting parameters:  μ R ohmic , and  I 0 . In the first fitting iteration, the parameters are assumed to be independent of  V G  and are fitted to the sequence of the electrical output curves for the hole or the electron branch. In the second iteration, for each  V G  the parameters are allowed to vary by ±20% from the previously determined values. The contact resistance is further expressed as a device width-scaled (W) value:  W R C = W ( ( V ohmic + V junction ) / I D ) .

2.5. Laser Treatment of   WSe 2

The freshly exfoliated channel  WSe 2  was laser-treated (532 nm, 100× objective) under ambient conditions using a motorised sample stage. The laser modification of  WSe 2  was performed prior to the transfer of  PdSe 2  contacts. The laser power was set to 50 mW. A point-to-point scan was carried out with a resolution of 0.2 µm, and a fixed exposure time of 0.1 s for each point.

2.6. AFM and In Operando KPFM Measurements

Horiba/AIST-NT Omegascope (Lille, France) AFM system was used for the AFM topography measurements, with Nunano SPARK 350 Pt probes (spring constant of 42  Nm 1 , resonant frequency 330 kHz, and tip radius of 30 nm). Topography images were processed in the open-source software Gwyddion v2.56 [58], applying zero-order line correction and three-point plane averaging.
In operando Kelvin Probe Force Microscopy (KPFM) measurements were carried out on  PdSe 2 -contacted devices under the ambient conditions. To prevent device degradation during prolonged ambient operation, the devices for the KPFM experiments were top capped by an additional 10 nm thick hBN flake. For the device biasing during the KPFM measurements, a Keithley 2636A Source Meter was used, and the device ground (source) was connected to the ground of the KPFM feedback loop. KPFM was operated in a frequency-modulated two-pass regime with a second-pass lift height of 8 nm, yielding a total of about 18 nm distance between the probe and the hBN capped channel. To extract electrostatic potential drops across the channel of an operating device, a single line in the middle of the device was repeatedly scanned while the external bias was applied. To compensate for the work function and stray field differences, each potential drop is normalised to the cross-sections recorded with  V D  = 0 V, following the procedure detailed in Ref. [32].

2.7. Raman Spectroscopy

Raman spectroscopy measurements were performed using a Horiba LabRam HR Evolution confocal Raman spectrometer (Lille, France) with 1800 lines/mm gratings. A 532 nm laser was used with an excitation power in the range of 0.1–3.2 mW. The laser spot was focused by a 100×, 0.9 NA objective.

3. Results and Discussions

3.1. Electrical Characteristics of  WSe 2  FETs with Graphite and  PdSe 2  Electrodes

Figure 1a depicts a schematic representation of the 2D layer stacks along with optical images of the typical devices with graphite (Gr) and  PdSe 2  contacts. Figure 1b,c present in a semi-log scale the device width-scaled electrical transfer curves, source–drain current as a function of the applied local back gate bias  I D ( V G ) , comparing the two different types of van der Waals contacts (Gr and  PdSe 2 ) to multilayered  WSe 2 . For each transfer curve, five subsequent forward and backward  V G  sweeps were carried out at 2 V/s. In both cases, a small hysteresis of 200 mV was observed. The measurements were carried out at 78 K to minimise charge-trap-related effects and unintentional doping effects from the trapped water and air at the interfaces [16,59,60]. Graphite-contacted devices showed a dominant n-type behaviour which was previously reported and attributed to the band alignment that favours electron injection from the graphite towards the channel material [56,61]. In Figure 1b, the second device (Device 2) also exhibits notably high current in the hole branch, however, the threshold voltage remains closer to the electron branch, as expected for the efficient electron injection from graphite electrodes [56].
The main difference in the electrical transfer curves between Gr and  PdSe 2 -contacted devices occurs at the negative  V G  values, i.e., in the hole branch. In contrast to graphite-contacted devices, when  PdSe 2  is used as a contact, the FETs were found to exhibit dominant p-type behaviour and an increased device performance for both electron and hole branches. This is explained by the favoured level alignment of the  PdSe 2  with the hole branch of the  WSe 2  due to the higher work function of  PdSe 2  in comparison to graphite. The  I ON / I OFF  ratio for  PdSe 2  (∼4 ×  10 4 ) was one order of magnitude better than that of graphite contacts. Horizontal dashed lines that interconnect Figure 1b,c serve as a guide to help compare the current levels. For the  WSe 2  devices reported in the literature, the  I ON / I OFF  ratio varies over several orders of magnitude [11]: from  10 2  (e.g.,  NbSe 2  contacts to the n-branch [56]) up to  10 9  with more elaborate device architectures and high-k dielectrics [18]. With respect to the electrode engineering to access the p-branch,   NbSe 2 - and Pt-contacted  WSe 2  were reported to reach the values in the range  10 4 10 7  [18,56].
Furthermore,  PdSe 2 -contacted devices maintained an intrinsic behaviour which is evident by an almost equidistant  V on  for both electron and hole branches with reference to  V G =  0 V. This was not the case for graphite-contacted devices where larger  V G  was required to reach the on state of the p-branch compared to the n-branch, therefore indicating a disruption in the intrinsic doping levels. On average, we observe a  V on  for the hole branch to be at (−4.5 ± 0.9) V and at (−1.9 ± 1.3) V respectively for the Gr and  PdSe 2  contacts; similar values for the  V on  were observed for the electron branch: (2.5 ± 1.7) V and (3.0 ± 0.8) V respectively for the Gr and  PdSe 2  contacts.
A comparison between 300 K and 78 K width-scaled transfer curves of a  PdSe 2 -contacted device is presented in Figure 1d. The temperature primarily impacts the phonon-related carrier scattering in the channel, the Schottky junction-related potential drop, and gate dielectric interface charge trap states. Consequently, at lower temperatures we observe an overall increase in the drain currents and mobilities for both branches (by a factor of ∼2 comparing 78 K and 300 K), quenching of the hysteresis with respect to the forward and backward  V G  sweeping, and a minor reduction in the  V th  values.
Figure 1e represents the electrical output curves for the  PdSe 2 -contacted channel; the source–drain current as a function of the applied source–drain bias is  I D ( V D ). Especially at more negative  V G  values (on state of the p-branch), the electrical output curves of the p-branch exhibit linear behaviour. For the n-branch, the overall  I D  values are about one order of magnitude lower than that of the p-branch and show significant deviation from the linear behaviour at lower  V D  values regardless of the applied  V G . All of these observations indicate that a significantly larger barrier exists for the electron than for the hole injection from  PdSe 2  into  WSe 2 . At low temperatures, within the applied  V D  range and for  V G  more than 0.5 V away from the  V th , we did not observe the current saturation. However, within the same bias range at room temperature, saturation can be achieved (see Figure 2).

3.2. Contact Resistance of the  PdSe 2 /WSe 2  Interface

The contact resistance of the interface between  PdSe 2  and  WSe 2  was evaluated independently by two approaches: parameter extraction via device modelling and direct measurements by in operando KPFM. In the first approach, we have modelled the sequence of the electrical output data by applying an equivalent electrical scheme as shown in Figure 2a (see also Section 2). The system was solved in a self-consistent manner and fitted to the set of output curves either for the hole or for the electron branch, as presented in Figure 2b,c. Parameters of the ohmic ( R ohmic ) and non-linear Schottky component ( I 0 ) of the contact resistance were extracted, and width-scaled contact resistance ( W R C ) was expressed considering specific points of operation (fixed  V D V G , and, consequently,  I D  values). We obtain  W R C  = (2.84 ± 0.53) ×  10 6   Ω µm for the hole branch and  W R C  = (3.72 ± 0.69) ×  10 8 Ω µm for the electron branch. The values are reported for the operation at 300 K, with  V G  set 5 V away from the  V th  in both cases of the hole and the electron branches, and under 1.5 V of source–drain bias. In particular, the need to include the non-linear Schottky element in the model is evident in a strong downward bending of the output curves at lower  V D , as pointed out by the red arrows in Figure 2c. Especially in the electron branch case, at lower  I D  the contact resistance and the entire device operation is Schottky junction-dominated, and almost all of the applied  V D  is taken by this junction as the most resistive element in the circuit. At higher  I D , the  V junction  still dominates over  V ohmic  by a factor of 5 to 10.
In the second approach to evaluating the contact resistance of the  PdSe 2 /WSe 2  interface, we have used in operando KPFM. This technique measures the electric potential several nanometres above the channel during device operation. Therefore, it resolves the potential drops between the electrodes, and allows independent distinguishing of the potential drops that correspond to the drain (not observed in our case), the channel, and the source [16,32]. An example of the potential drop profiles is presented in Figure 2d in the hole branch on state and for varied  V D  between 0.5 V and 1.5 V. Four regions are clearly distinguishable in the potential drop profiles: flat potential values corresponding to the source and drain regions of the scan, a monotone drop of the potential along the channel, and a much steeper drop at the contact to the source electrode. Linear fits to these elements are presented by dashed black lines. The steeper drop connected to the transition between the channel-related potential drop and the source contact region is directly related to the  V junction + V ohmic  in the device model. Knowing the  I D  values during the potential drop profile measurements and the width of the device, it is possible to express the observed junction-related potential drop as the width-scaled contact resistance.
Figure 2e presents the potential drop profiles when the connections between the source and the drain are exchanged, effectively reversing the current flow direction. We observe that the contact resistance associated potential drop is connected to the grounded source electrode, i.e., that the  PdSe 2 /WSe 2  interface is rectifying. This proves the predominant Schottky nature of the contact resistance, as also suggested by the model.
Lastly, when biased under very similar operation points as in the case of the contact resistance extraction from the electrical output data sets (Figure 2f), we obtain the following device width-scaled contact resistance values obtained from in operando KPFM:  W R C  = 2.78 ×  10 6   Ω µm for the hole-branch and  W R C  = 3.44 ×  10 8 Ω µm for the electron branch.
The obtained  W R C  values imply that  PdSe 2  is an effective hole injector. This is seen from the two orders of magnitude larger contact resistance of the electron branch under similar operation conditions. Furthermore, the contact resistance of the  PdSe 2 /WSe 2  interface for p-type operation performs similar to the commonly employed evaporated metallic contacts [62,63] while preserving the intrinsic doping levels and the ambipolar nature of the  WSe 2 . Reported values for  WSe 2  contact resistance range from  10 8   Ω  µm to  10 5   Ω  µm with electrostatic gating and down to  10 4 Ω  µm for electrolyte gating that can induce very high density states in  WSe 2  [45,56,62]. Some of the lowest values reported for the contact resistance (1.1 ×  10 5 Ω  µm) are with Pt electrodes, where MIGS cannot be excluded at the electrode interface [45].

3.3. Optimizing Contact Interface via Laser-Driven Oxidation of  WSe 2

Recent work has shown that the application of mild oxygen plasma can be an effective way to reduce the Schottky barrier in multilayer  WSe 2  FETs [64,65,66]. The plasma treatment causes the formation of a conductive tungsten selenium oxide ( WSe y O x ). The oxide was found to form in a layer-by-layer manner [64,65], effectively generating a  WSe y O x /WSe 2  heterostructure that acts as a facilitator for the hole injection [66,67]. However, it is important to protect the channel active area during the plasma treatment to avoid device degradation. We wanted to investigate if a laser-based approach could open a way to achieve similar modification of  WSe 2 , as with the mild plasma treatment. An advantage of the laser-driven oxidation approach is straightforward patterning by laser scanning. Using laser irradiation (532 nm, 50 mW) under ambient conditions, we have observed a similar oxidation process of  WSe 2 .
To explore the influence of the  WSe y O x  layer on the contact properties between  WSe 2  and  PdSe 2 , we have irradiated an area of the  WSe 2  flake that is slightly larger than the contact area with  PdSe 2 . After the laser treatment,  PdSe 2  flakes were transferred and used as contacts. Figure 3a(i i i i ) show the schematic representation of the laser treatment and the device assembly process for the  WSe y O x -modified contacts. Figure 3b(i i i i ) represent the corresponding optical images of the flake and the final device, where only one side of the flake was treated by the laser. Figure 3c presents a zoomed-in region of the interface to highlight the parts of the ablated layers, oxidised layers, and remaining pristine  WSe 2  layers.
AFM was performed to observe the morphological and height changes due to laser treatment. The results are presented in Figure 4a,b. AFM image before laser treatment shows large bubbles formed at the interface between  WSe 2  and hBN, as well as between hBN and  SiO 2  interface. This is expected for 2D material heterostructures assembled under ambient conditions and using PDMS stamps due to the entrapment of air and water [68,69,70]. Such interfaces result in localised charge-trap and scattering centres, and a flat interface is desired to achieve better performance [71,72]. Interestingly, laser treatment resulted in the removal and migration of these bubbles from the scan area, even at the regions not directly exposed to the laser irradiation. This is illustrated in Figure 4b. Such behaviour can be attributed to the self-cleaning property of 2D materials under a systematic sweep of the laser spot which allows local heating and migration of the trapped water/air bubbles at the interfaces [71,72]. The arrows in Figure 4b represent the direction of laser sweeping, and the dashed rectangle indicates the laser-exposed area. Figure 4a,b (bottom) show the change in height of the flake before and after the treatment. The resultant height corresponds to a thickness of 3.9 nm. This indicates the ablation of about five mono-layers of  WSe 2 , and the remaining flake effectively forms a  WSe y O x /WSe 2  heterostructure. Combined with Raman spectroscopy data (Figure 4c) we estimate that after the laser treatment, about three layers of  WSe 2  remain, with about 2 nm of  WSe y O x  formed on top [64,65,66]. Raman spectroscopy was also performed to verify the crystal quality of the laser-modified flakes. Figure 4c presents the Raman spectra before and after the laser treatment of a  WSe 2  flake. The increase in the Raman intensity of the  A 1 g  mode (shown in the inset) after the treatment validates the thinning of  WSe 2  with the oxidation of top layers. Such an increase in the intensity of the peaks is related to the thinning of  WSe 2  and an increase in the phonon lifetime [45,73]. An increase in phonon lifetime should be also observed in the according change in the device-apparent field-effect mobility. However, for both treated and the untreated devices, the apparent hole mobilities were within the sample-to-sample variation.

3.4. Electrical Characteristics of  WSe 2  FETs with  WSe 2 /WSe y O x /PdSe 2  Electrode Interface

Figure 5a represents the electrical transfer curves for the  WSe 2  device with both source and drain electrode interfaces modified by the laser treatment. The device showed a notable decrease in  I ON . However, highly stable p-type devices were realised with respect to  V th  variations between subsequent sweeps at room temperature operation and also under varied  V G  sweeping rates.  V th  for the hole branch of the  WSe y O x -modified contacts was found to be at (−0.42 ± 0.06) V, which is about four times lower than the  PdSe 2 /WSe 2  interface. More importantly, the sample-to-sample-, forward/backward sweep-, and multiple sweep-related variations in the  V th  value are almost completely reduced. The dominant p-type behaviour with the quenching of the electron branch can be associated with  WSe y O x , which acts as an efficient hole injection layer [66,67]. The  WSe y O x /WSe 2  layer also extends beyond the contact regions into the channel (for about 1 µm) to ensure that the contact is not made directly with the unmodified  WSe 2 . Consequently, WSe y O x  could also introduce interface traps in the channel active area. To test this, we have probed the stability of the devices by examining the hysteresis voltage ( V H ), as a difference in the  V th  between the forward and backward sweeping electrical transfer curves.  V H  values for the varied  V G  sweep rates are shown in Figure 5b. The device remained stable at high sweeping rates (up to 15 V/s) with a negligible hysteresis of 55 mV. An increase in the hysteresis of up to 150 mV was noted at low sweeping rates. The hysteresis values correspond well to the pristine  PdSe 2 /WSe 2 /PdSe 2  devices, indicating that the  WSe y O x  layers did not affect device stability. The observed hysteresis is likely related to the interface between hBN and  WSe 2  or is inherent to the  WSe 2  layers.
Two-dimensional semiconductors commonly show large variations in the  V th  at varying drain voltages which also impacts the device stability [6,74,75]. To test this, we subjected our device to a  V D  ranging from 0.5 V to 2.0 V. The device maintained the same  V th  for forward and backward sweeps with a  V D -independent hysteresis of 98 mV. The subthreshold swing (SS) values were also significantly improved from 200 mV/dec for the non-treated to 100 mV/dec for the treated devices. It is worth mentioning that these devices operated without a top encapsulation and therefore, a high-k dielectric encapsulation, optimization of the hBN thickness, and integration into dual-gate geometries can further improve  WSe 2  FET performance. The obtained SS values, especially for the laser-treated devices, perform better than commonly reported back-gate implemented FETs, where the SS values range from 2500 mV/dec to 400 mV/dec [11,18,20]. Some of the lowest reported SS values for  WSe 2  devices (95 mV/dec) were achieved by utilizing 20 nm of  HfO 2  as the gate insulator and p-branch matching Pt contacts [45].
We have also fabricated asymmetric devices where only one side of the channel was treated with the laser (as also shown in Figure 3). An example of the electrical transfer curves observed for such devices is presented in Figure 5d. A large hysteresis window was observed, which directly depends on the choice of the drain electrode, i.e., the device exhibits rectifying behaviour with respect to the induced hysteresis. For the case where the laser-treated region was used as a drain, a stable p-branch was realised with negligible hysteresis, represented by a solid black curve in Figure 5d. A pronounced hysteresis of 2 V was only present for the n-branch. This was observed to be in stark contrast to the behaviour of the same device when the non-treated region was used as the drain. In that case, a p-branch hysteresis of 4 V was observed. In both cases, the hysteresis was stable for multiple forward–backward sweeps as indicated by the arrows. Such behaviour can be associated with large differences between the number of carriers available underneath the contact regions. As one end of the channel is intentionally favouring hole injection and prohibiting electron injection, the other stays in its original form. Such large and stable rectifying hysteretic behaviour could be employed in novel concepts as computing in memory and self-reconfiguring electronics [76,77].

4. Conclusions

In summary, we have introduced  PdSe 2  contacts to  WSe 2  FETs that enable effective hole injection, enhanced p-type performance, and preserve the intrinsic ambipolar response of  WSe 2 PdSe 2  contacts allow essentially hysteresis-free electrical response while maintaining high on-state currents and  I ON / I OFF  ratio enhancement by one order of magnitude in comparison to graphite-contacted devices. Considering the low temperatures required for the  PdSe 2  growth, it is a promising electrode candidate especially when considering the potential that  PdSe 2  brings for the upscaling of 2D-material-based electronics and the incorporation of ambipolar  WSe 2  to post-CMOS architectures.
We extended the study to also contact laser-treated  WSe 2 , where the laser irradiation induces the formation of a conductive  WSe y O x  layer at the electrode interface. In this case, we observed highly stable p-type behaviour of the devices with a two-fold improvement in the subthreshold swing, stabilization of the  V th  for the hole branch. Interestingly, if only one electrode interface is modified by the laser treatment, asymmetric  WSe 2  FETs were achieved, which exhibited pronounced and stable hysteretic behaviour of only one (electron or hole) branch. The hysteresis was dependent on the direction of applied drain voltage. Such device response can be used to design in-memory computing and reconfigurable electronic concepts based purely on 2D interfaces.

Author Contributions

G.M. and M.A.A. conceived the idea for the study. M.A.A. and A.M. performed data analysis and wrote the manuscript. G.M. and S.L. prepared the 2D material heterostacks and carried out the experiments. I.P. and Z.S. provided the  PdSe 2  crystals. V.T. and E.P. fabricated the gold pad substrates. R.D.R. supervised Raman and AFM measurements. S.L. and A.M. performed the experiments and data analysis/modelling related to the extraction of the contact resistance. A.M. acquired the main funding for the study. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the Austrian Science Fund (FWF) under grants no. I4323-N36 and Y1298-N. A.M. acknowledges the support provided by the ERC grant POL_2D_PHYSICS (101075821). Z.S. was supported by ERC–CZ program (project LL2101) from Ministry of Education Youth and Sports (MEYS) and used large infrastructure from project reg. No. CZ.02.1.01/0.0/0.0/15_003/0000444 financed by the EFRR. R.D.R. acknowledges the Russian Science Foundation grant 23-42-00081. V.T. and E.P. acknowledge funding by the Slovenian Research and Innovation Agency (ARIS) under research program P1–0055. The authors acknowledge infrastructure support of the Montanuniversitaet Leoben (Raman AFM TERS lab).

Data Availability Statement

The data generated within this study and the samples related to this study are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Cao, W.; Bu, H.; Vinet, M.; Cao, M.; Takagi, S.; Hwang, S.; Ghani, T.; Banerjee, K. The future transistors. Nature 2023, 620, 501–515. [Google Scholar] [CrossRef]
  2. Zhu, K.; Wen, C.; Aljarb, A.A.; Xue, F.; Xu, X.; Tung, V.; Zhang, X.; Alshareef, H.N.; Lanza, M. The development of integrated circuits based on two-dimensional materials. Nat. Electron. 2021, 11, 775–785. [Google Scholar] [CrossRef]
  3. Ahmad, W.; Gong, Y.; Abbas, G.; Khan, K.; Khan, M.; Ali, G.; Shuja, A.; Tareen, A.K.; Khan, Q.; Li, D. Evolution of low-dimensional material-based field-effect transistors. Nanoscale 2021, 10, 5162–5186. [Google Scholar] [CrossRef] [PubMed]
  4. Rawat, A.; Gupta, A.K.; Rawat, B. Performance projection of 2D material-based CMO inverters for sub-10-nm channel length. IEEE Trans. Electron. Devices 2021, 68, 3622–3629. [Google Scholar] [CrossRef]
  5. Fiori, G.; Bonaccorso, F.; Iannaccone, G.; Palacios, T.; Neumaier, D.; Seabaugh, A.; Banerjee, S.K.; Colombo, L. Electronics based on two-dimensional materials. Nat. Nanotechnol. 2014, 9, 768–779. [Google Scholar] [CrossRef]
  6. Li, J.; Chen, X.; Zhang, D.W.; Zhou, P. Van-der-Waals heterostructure based field effect transistor application. Crystals 2017, 8, 8. [Google Scholar] [CrossRef]
  7. Illarionov, Y.Y.; Knobloch, T.; Jech, M.; Lanza, M.; Akinwande, D.; Vexler, M.I.; Mueller, T.; Lemme, M.C.; Fiori, G.; Schwierz, F.; et al. Insulators for 2D nanoelectronics: The gap to bridge. Nat. Commun. 2020, 118, 3385. [Google Scholar] [CrossRef]
  8. Arora, A.; Ganapathi, K.L.; Dixit, T.; Miryala, M.; Masato, M.; Rao, M.S.R.; Krishnan, A. Thickness-Dependent Nonlinear Electrical Conductivity of Few-Layer Muscovite Mica. Phys. Rev. Appl. 2022, 17, 064042. [Google Scholar] [CrossRef]
  9. Zhou, J.; Lin, J.; Huang, X.; Zhou, Y.; Chen, Y.; Xia, J.; Wang, H.; Xie, Y.; Yu, H.; Lei, J.; et al. A library of atomically thin metal chalcogenides. Nature 2018, 556, 355–359. [Google Scholar] [CrossRef]
  10. Di Bartolomeo, A. Emerging 2D materials and their van der Waals heterostructures. Nanomaterials 2020, 10, 579. [Google Scholar] [CrossRef]
  11. Cheng, Q.; Pang, J.; Sun, D.; Wang, J.; Zhang, S.; Liu, F.; Chen, Y.; Yang, R.; Liang, N.; Lu, X.; et al. WSe2 2D p-type semiconductor-based electronic devices for information technology: Design, preparation, and applications. InfoMat 2020, 102, 656–697. [Google Scholar] [CrossRef]
  12. Kumar, R.; Goel, N.; Hojamberdiev, M.; Kumar, M. Transition metal dichalcogenides-based flexible gas sensors. Sens. Actuator A Phys. 2020, 303, 111875. [Google Scholar] [CrossRef]
  13. Sumesh, C.K.; Peter, S.C. Two-dimensional semiconductor transition metal based chalcogenide based heterostructures for water splitting applications. Dalton Trans. 2019, 48, 12772–12802. [Google Scholar] [CrossRef]
  14. Maniyar, A.; Choudhary, S. Visible region absorption in TMDs/phosphorene heterostructures for use in solar energy conversion applications. RSC Adv. 2020, 10, 31730–31739. [Google Scholar] [CrossRef]
  15. Wang, C.; Yang, F.; Gao, Y. The highly-efficient light-emitting diodes based on transition metal dichalcogenides: From architecture to performance. Nanoscale Adv. 2020, 2, 4323–4340. [Google Scholar] [CrossRef]
  16. Aslam, M.A.; Tran, T.H.; Supina, A.; Siri, O.; Meunier, V.; Watanabe, K.; Taniguchi, T.; Kralj, M.; Teichert, C.; Sheremet, E.; et al. Single-crystalline nanoribbon network field effect transistors from arbitrary two-dimensional materials. npj 2D Mater. Appl. 2022, 6, 76. [Google Scholar] [CrossRef]
  17. Murastov, G.; Aslam, M.A.; Tran, T.H.; Lassnig, A.; Watanabe, K.; Taniguchi, T.; Wurster, S.; Nachtnebel, M.; Teichert, C.; Sheremet, E.; et al. Photoinduced edge-specific nanoparticle decoration of two-dimensional tungsten diselenide nanoribbons. Commun. Chem. 2023, 6, 166. [Google Scholar] [CrossRef]
  18. Pudasaini, P.R.; Oyedele, A.; Zhang, C.; Stanford, M.G.; Cross, N.; Wong, A.T.; Hoffman, A.N.; Xiao, K.; Duscher, G.; Mandrus, D.G.; et al. High-performance multilayer WSe2 field-effect transistors with carrier type control. Nano Res. 2018, 11, 722–730. [Google Scholar] [CrossRef]
  19. Nan, H.; Zhou, R.; Gu, X.; Xiao, S.; Ostrikov, K.K. Recent advances in plasma modification of 2D transition metal dichalcogenides. Nanoscale 2019, 11, 19202–19213. [Google Scholar] [CrossRef] [PubMed]
  20. Kozhakhmetov, A.; Stolz, S.; Tan, A.M.Z.; Pendurthi, R.; Bachu, S.; Turker, F.; Alem, N.; Kachian, J.; Das, S.; Hennig, R.G.; et al. Controllable p-type doping of 2D WSe2 via vanadium substitution. Adv. Funct. Mater. 2021, 31, 2105252. [Google Scholar] [CrossRef]
  21. Grützmacher, S.; Heyl, M.; Nardi, M.V.; Koch, N.; List-Kratochvil, E.J.W.; Ligorio, G. Local Manipulation of the Energy Levels of 2D TMDCs on the Microscale Level via Microprinted Self-Assembled Monolayers. Adv. Mater. Interf. 2023, 10, 2300276. [Google Scholar] [CrossRef]
  22. Pang, Y.-D.; Wu, E.-X.; Xu, Z.-H.; Hu, X.-D.; Wu, S.; Xu, L.-Y.; Liu, J. Effect of electrical contact on performance of WSe2 field effect transistors. Chin. Phys. B 2021, 30, 068501. [Google Scholar] [CrossRef]
  23. Liu, Y.; Duan, X.; Shin, H.-J.; Park, S.; Huang, Y.; Duan, X. Promises and prospects of two-dimensional transistors. Chin. Phys. B 2021, 519, 43–53. [Google Scholar] [CrossRef] [PubMed]
  24. Liao, W.; Zhao, S.; Li, F.; Wang, C.; Ge, Y.; Wang, H.; Wang, S.; Zhang, H. Interface engineering of two-dimensional transition metal dichalcogenides towards next-generation electronic devices: Recent advances and challenges. Nanoscale Horiz. 2020, 5, 787–807. [Google Scholar] [CrossRef] [PubMed]
  25. Rai, A.; Movva, H.C.P.; Roy, A.; Taneja, D.; Chowdhury, S.; Banerjee, S.K. Progress in contact, doping and mobility engineering of MoS2: An atomically thin 2D semiconductor. Crystals 2018, 8, 316. [Google Scholar] [CrossRef]
  26. Poljak, M.; Matić, M. Metallization-induced quantum limits of contact resistance in graphene nanoribbons with one-dimensional contacts. Materials 2021, 14, 3670. [Google Scholar] [CrossRef] [PubMed]
  27. Jain, A.; Szabó, Á.; Parzefall, M.; Bonvin, E.; Taniguchi, T.; Watanabe, K.; Bharadwaj, P.; Luisier, M.; Novotny, L. One-dimensional edge contacts to a monolayer semiconductor. Nano Lett. 2019, 19, 6914–6923. [Google Scholar] [CrossRef]
  28. Cheng, Z.; Yu, Y.; Singh, S.; Price, K.; Noyce, S.G.; Lin, Y.-C.; Cao, L.; Franklin, A.D. Immunity to contact scaling in MoS2transistors using in situ edge contacts. Nano Lett. 2019, 19, 5077–5085. [Google Scholar] [CrossRef] [PubMed]
  29. Das, S.; Chen, H.-Y.; Penumatcha, A.V.; Appenzeller, J. High performance multilayer MoS2 transistors with scandium contacts. Nano Lett. 2013, 13, 100–105. [Google Scholar] [CrossRef] [PubMed]
  30. English, C.D.; Shine, G.; Dorgan, V.E.; Saraswat, K.C.; Pop, E. Improved contacts to MoS2 transistors by ultra-high vacuum metal deposition. Nano Lett. 2016, 16, 3824–3830. [Google Scholar] [CrossRef]
  31. Kwon, G.; Choi, Y.; Lee, H.; Kim, H.; Jeong, J.; Jeong, K.; Baik, M.; Kwon, H.; Ahn, J.; Lee, E.; et al. Interaction-and defect-free van der Waals contacts between metals and two-dimensional semiconductors. Nat. Electron. 2022, 5, 241–247. [Google Scholar] [CrossRef]
  32. Matković, A.; Petritz, A.; Schider, G.; Krammer, M.; Kratzer, M.; Karner-Petritz, E.; Fian, A.; Gold, H.; Gärtner, M.; Terfort, A.; et al. Interfacial band engineering of MoS2/gold interfaces using pyrimidine-containing self-assembled monolayers: Toward contact-resistance-free bottom-contacts. Adv. Electron. Matter. 2020, 6, 2000110. [Google Scholar] [CrossRef]
  33. Liu, Y.; Guo, J.; Zhu, E.; Liao, L.; Lee, S.-J.; Ding, M.; Shakir, I.; Gambin, V.; Huang, Y.; Duan, X. Approaching the Schottky–Mott limit in van der Waals metal–semiconductor junctions. Nature 2018, 557, 696–700. [Google Scholar] [CrossRef] [PubMed]
  34. Liu, G.; Tian, Z.; Yang, Z.; Xue, Z.; Zhang, M.; Hu, X.; Wang, Y.; Yang, Y.; Chu, P.K.; Mei, Y.; et al. Graphene-assisted metal transfer printing for wafer-scale integration of metal electrodes and two-dimensional materials. Nat. Electron. 2022, 5, 275–280. [Google Scholar] [CrossRef]
  35. Poljak, M.; Matić, M.; Župančić, T.; Zeljko, A. Lower limits of contact resistance in phosphorene nanodevices with edge contacts. Nanomaterials 2022, 12, 656. [Google Scholar] [CrossRef] [PubMed]
  36. Poljak, M.; Matić, M. Optimum Contact Configurations for Quasi-One-Dimensional Phosphorene Nanodevices. Nanomaterials 2023, 13, 1759. [Google Scholar] [CrossRef] [PubMed]
  37. Shen; Su, P.C.; Lin, C.; Chou, Y.; Cheng, A.S.; Park, C.C.; Chiu, J.H.; Lu, M.H.; Tang, A.Y.; Tavakoli, H.L.; et al. Ultralow contact resistance between semimetal and monolayer semiconductors. Nature 2021, 593, 211–217. [Google Scholar] [CrossRef] [PubMed]
  38. Mootheri, V.; Arutchelvan, G.; Banerjee, S.; Sutar, S.; Leonhardt, A.; Boulon, M.; Huyghebaert, C.; Houssa, M.; Asselberghs, I.; Radu, I.; et al. Graphene based Van der Waals contacts on MoS2 field effect transistors. 2D Mater. 2020, 8, 015003. [Google Scholar] [CrossRef]
  39. Ryu, H.; Kim, D.; Kwon, J.; Park, S.K.; Lee, W.; Seo, H.; Watanabe, K.; Taniguchi, T.; Kim, S.; van der Zande, A.M.; et al. Fluorinated Graphene Contacts and Passivation Layer for MoS2 Field Effect Transistors. Adv. Electron. Matter. 2022, 8, 2101370. [Google Scholar] [CrossRef]
  40. Li, Z.; Wang, Y.; Jiang, J.; Liang, Y.; Zhong, B.; Zhang, H.; Yu, K.; Kan, G.; Zou, M. Temperature-dependent Raman spectroscopy studies of 1–5-layer WSe2. Nano Res. 2020, 13, 591–595. [Google Scholar] [CrossRef]
  41. Liu; Tan, Y.; Chou, C.; Nayak, H.; Wu, A.; Ghosh, D.; Chang, R.; Hao, H.Y.; Wang, Y.; Kim, X.; et al. Thermal oxidation of WSe2 nanosheets adhered on SiO2/Si substrates. Nano Lett. 2015, 15, 4979–4984. [Google Scholar] [CrossRef]
  42. Illarionov, Y.Y.; Waltl, M.; Rzepa, G.; Knobloch, T.; Kim, J.-S.; Akinwande, D.; Grasser, T. Highly-stable black phosphorus field-effect transistors with low density of oxide traps. npj 2D Mater. Appl. 2017, 1, 23. [Google Scholar] [CrossRef]
  43. Wang, J.-B.; Ren, Z.; Hou, Y.; Yan, X.-L.; Liu, P.-Z.; Zhang, H.; Zhang, H.-X.; Guo, J.-J. A review of graphene synthesis at low temperatures by CVD methods. New Carbon Mater. 2020, 35, 193–208. [Google Scholar] [CrossRef]
  44. Liu, W.; Kang, J.; Sarkar, D.; Khatami, Y.; Jena, D.; Banerjee, K. Role of metal contacts in designing high-performance monolayer n-type WSe2 field effect transistors. Nano Lett. 2013, 13, 1983–1990. [Google Scholar] [CrossRef] [PubMed]
  45. Zhang, L.; Zhang, Y.; Sun, X.; Jia, K.; Zhang, Q.; Wu, Z.; Yin, H. High-performance multilayer WSe2 p-type field effect transistors with Pd contacts for circuit applications. J. Mater. Sci. Mater. Electron. 2021, 32, 17427–17435. [Google Scholar] [CrossRef]
  46. Oyedele, A.D.; Yang, S.; Feng, T.; Haglund, A.V.; Gu, Y.; Puretzky, A.A.; Briggs, D.; Rouleau, C.M.; Chisholm, M.F.; Unocic, R.R.; et al. Defect-mediated phase transformation in anisotropic two-dimensional PdSe2 crystals for seamless electrical contacts. J. Am. Chem. Soc. 2019, 141, 8928–8936. [Google Scholar] [CrossRef] [PubMed]
  47. Seo, J.-E.; Park, E.; Das, T.; Kwak, J.Y.; Chang, J. Demonstration of PdSe2 CMOS Using Same Metal Contact in PdSe2 n-/p-MOSFETs through Thickness-Dependent Phase Transition. Adv. Electron. Mater. 2022, 8, 2200485. [Google Scholar] [CrossRef]
  48. Long, M.; Wang, Y.; Wang, P.; Zhou, X.; Xia, H.; Luo, C.; Huang, S.; Zhang, G.; Yan, H.; Fan, Z.; et al. Palladium diselenide long-wavelength infrared photodetector with high sensitivity and stability. Adv. Electron. Mater. 2019, 13, 2511–2519. [Google Scholar] [CrossRef] [PubMed]
  49. Gu, Y.; Cai, H.; Dong, J.; Yu, Y.; Hoffman, A.N.; Liu, C.; Oyedele, A.D.; Lin, Y.C.; Ge, Z.; Puretzky, A.A.; et al. Two-dimensional palladium diselenide with strong in-plane optical anisotropy and high mobility grown by chemical vapor deposition. Adv. Mater. 2020, 32, 1906238. [Google Scholar] [CrossRef]
  50. Sun, J.; Shi, H.; Siegrist, T.; Singh, D.J. Electronic, transport, and optical properties of bulk and mono-layer PdSe2. Appl. Phys. Lett. 2015, 107, 153902. [Google Scholar] [CrossRef]
  51. Liang, Q.; Wang, Q.; Zhang, Q.; Wei, J.; Lim, S.X.; Zhu, R.; Hu, J.; Wei, W.; Lee, C.; Sow, C.; et al. High-performance, room temperature, ultra-broadband photodetectors based on air-stable PdSe2. Adv. Mater. 2019, 31, 1807609. [Google Scholar] [CrossRef]
  52. Wang, Y.; Pang, J.; Cheng, Q.; Han, L.; Li, Y.; Meng, X.; Ibarlucea, B.; Zhao, H.; Yang, F.; Liu, H.; et al. Applications of 2D-layered palladium diselenide and its van der Waals heterostructures in electronics and optoelectronics. Nano–Micro Lett. 2021, 13, 143. [Google Scholar] [CrossRef]
  53. Liang, Q.; Chen, Z.; Zhang, Q.; Wee, A.T.S. Pentagonal 2D transition metal dichalcogenides: PdSe2 and beyond. Adv. Funct. Mater. 2022, 32, 2203555. [Google Scholar] [CrossRef]
  54. Oyedele, A.D.; Yang, S.; Liang, L.; Puretzky, A.A.; Wang, K.; Zhang, J.; Yu, P.; Pudasaini, P.R.; Ghosh, A.W.; Liu, Z.; et al. PdSe2: Pentagonal two-dimensional layers with high air stability for electronics. J. Am. Chem. Soc. 2017, 139, 14090–14097. [Google Scholar] [CrossRef] [PubMed]
  55. Withanage, S.S.; Khondaker, S.I. Low pressure CVD growth of 2D PdSe2 thin film and its application in PdSe2-MoSe2 vertical heterostructure. 2D Mater. 2022, 9, 025025. [Google Scholar] [CrossRef]
  56. Sata, Y.; Moriya, R.; Masubuchi, S.; Watanabe, K.; Taniguchi, T.; Machida, T. n-and p-type carrier injections into WSe2 with van der Waals contacts of two-dimensional materials. Jpn. J. Appl. Phys. 2017, 56, 04CK09. [Google Scholar] [CrossRef]
  57. Laturia, A.; Van de Put, M.L.; Vandenberghe, W.G. Dielectric properties of hexagonal boron nitride and transition metal dichalcogenides: From monolayer to bulk. npj 2D Mater. Appl. 2018, 2, 6. [Google Scholar] [CrossRef]
  58. Nečas, D.; Klapetek, P. Gwyddion: An open-source software for SPM data analysis. Open Phys. 2012, 10, 181–188. [Google Scholar] [CrossRef]
  59. Zhou, L.; Ge, C.; Yang, H.; Sun, Y.; Zhang, J. A high-pressure enhanced coupling effect between graphene electrical contacts and two-dimensional materials thereby improving the performance of their constituent FET devices. J. Mater. Chem. C 2019, 7, 15171–15178. [Google Scholar] [CrossRef]
  60. Watson, A.J.; Lu, W.; Guimarães, M.H.D.; Stöhr, M. Transfer of large-scale two-dimensional semiconductors: Challenges and developments. 2D Mater. 2021, 8, 032001. [Google Scholar] [CrossRef]
  61. Pan, Y.; Rahaman, M.; He, L.; Milekhin, I.; Manoharan, G.; Aslam, M.A.; Blaudeck, T.; Willert, A.; Matković, A.; Madeira, T.I.; et al. Exciton tuning in monolayer WSe2 via substrate induced electron doping. Nanoscale Adv. 2022, 4, 5102–5108. [Google Scholar] [CrossRef]
  62. Allain, A.; Kang, J.; Banerjee, K.; Kis, A. Electrical contacts to two-dimensional semiconductors. Nat. Mater. 2015, 14, 1195–1205. [Google Scholar] [CrossRef]
  63. Wang, Y.; Chhowalla, M. Making clean electrical contacts on 2D transition metal dichalcogenides. Nat. Rev. Phys. 2022, 4, 101–112. [Google Scholar] [CrossRef]
  64. Lee, K.; Ngo, T.D.; Lee, S.; Shin, H.; Choi, M.S.; Hone, J.; Yoo, W.J. Effects of Oxygen Plasma Treatment on Fermi-Level Pinning and Tunneling at the Metal–Semiconductor Interface of WSe2 FETs. Adv. Electron. Mater. 2023, 9, 2200955. [Google Scholar] [CrossRef]
  65. Ngo, T.D.; Choi, M.S.; Lee, M.; Ali, F.; Hassan, Y.; Ali, N.; Liu, S.; Lee, C.; Hone, J.; Yoo, W.J. Selective Electron Beam Patterning of Oxygen-Doped WSe2 for Seamless Lateral Junction Transistors. Adv. Sci. 2022, 9, 2202465. [Google Scholar] [CrossRef] [PubMed]
  66. Moon, I.; Lee, S.; Lee, M.; Kim, C.; Seol, D.; Kim, Y.; Kim, K.H.; Yeom, G.Y.; Teherani, J.T.; Hone, J.; et al. The device level modulation of carrier transport in a 2D WSe2 field effect transistor via a plasma treatment. Nanoscale 2019, 11, 17368–17375. [Google Scholar] [CrossRef]
  67. Kang, W.–M.; Lee, S.T.; Cho, I.–T.; Park, T.H.; Shin, H.; Hwang, C.S.; Lee, C.; Hao, H.Y.; Park, B.-G.; Lee, J.-H.; et al. Multi-layer WSe2 field effect transistor with improved carrier-injection contact by using oxygen plasma treatment. Solid-State Electron. 2018, 140, 2–7. [Google Scholar]
  68. Li, Q.; Song, J.; Besenbacher, F.; Dong, M. Two-dimensional material confined water. Acc. Chem. Res. 2015, 48, 119–127. [Google Scholar] [CrossRef] [PubMed]
  69. Jain, A.; Bharadwaj, P.; Heeg, S.; Parzefall, M.; Taniguchi, T.; Watanabe, K.; Novotny, L. Minimizing residues and strain in 2D materials transferred from PDMS. Nanotechnology 2018, 29, 265203. [Google Scholar] [CrossRef]
  70. Wang, W.; Clark, N.; Hamer, M.; Carl, A.; Tovari, E.; Sullivan-Allsop, S.; Tillotson, E.; Gao, Y.; de Latour, H.; Selles, F.; et al. Clean assembly of van der Waals heterostructures using silicon nitride membranes. Nat. Electron. 2023, 6, 981–990. [Google Scholar] [CrossRef]
  71. Purdie, D.G.; Pugno, N.M.; Taniguchi, T.; Watanabe, K.; Ferrari, A.C.; Lombardo, A. Cleaning interfaces in layered materials heterostructures. Nat. Commun. 2018, 9, 5387. [Google Scholar] [CrossRef] [PubMed]
  72. Jeon, D.; Kim, H.; Gu, M.; Kim, T. Imaging Fermi-level hysteresis in nanoscale bubbles of few-layer MoS2. Commun. Mater. 2023, 4, 62. [Google Scholar] [CrossRef]
  73. Zhang, R.; Drysdale, D.; Koutsos, V.; Cheung, R. Controlled layer thinning and p-type doping of WSe2 by vapor XeF2. Adv. Funct. Mater. 2017, 27, 1702455. [Google Scholar] [CrossRef]
  74. Park, W.; Pak, Y.; Jang, H.Y.; Nam, J.H.; Kim, T.H.; Oh, S.; Choi, S.M.; Kim, Y.; Cho, B. Improvement of the bias stress stability in 2D MoS2 and WS2 transistors with a TiO2 interfacial layer. Nanomaterials 2019, 9, 1155. [Google Scholar] [CrossRef] [PubMed]
  75. Ye, M.; Zhang, D.; Yap, Y.K. Recent advances in electronic and optoelectronic devices based on two-dimensional transition metal dichalcogenides. Electronics 2017, 6, 43. [Google Scholar] [CrossRef]
  76. Fei, W.; Trommer, J.; Lemme, M.C.; Mikolajick, T.; Heinzig, A. Emerging reconfigurable electronic devices based on two-dimensional materials: A review. InfoMat 2022, 4, e12355. [Google Scholar] [CrossRef]
  77. Feng, C.; Wu, W.; Liu, H.; Wang, J.; Wan, H.; Ma, G.; Wang, H. Emerging Opportunities for 2D Materials in Neuromorphic Computing. Nanomaterials 2023, 13, 2720. [Google Scholar] [CrossRef]
Figure 1. Electrical characteristics of graphite- and  PdSe 2 -contacted  WSe 2  FETs: (a) Schematic representation of device configuration with optical images of  WSe 2  FETs (scale bar: 10 µm). (b,c) Semi-log electrical transfer curves of devices with graphite (Gr) and  PdSe 2  contacts, respectively. The  I D  in (b,c) is scaled by the mean width of the channels to allow for better comparison of the current values between the different devices. The horizontal dashed lines that interconnect (b,c) serve as a guide to see the reached on- and off-state current levels. The red arrow in (c) indicates over an order of magnitude larger current of the hole branch in the case of  PdSe 2  contacts. (d) Comparison of the device width-scaled electrical transfer curves ( PdSe 2  contacted device) measured at 300 K (orange) and 78 K (purple), presented in linear scale. The arrows indicate the direction of the  V G  sweep, highlighting an increase in the hysteresis observed at 300 K. (e) Output curves for the hole and electron branches at 78 K (2 ×  10 2  mbar) of a device with  PdSe 2  contacts. Note that the current values for the n-branch are approximately one order of magnitude lower than for the p-branch. The different colored lines in (e) represent the curves at the different values of  V G , as indicated in the figure.
Figure 1. Electrical characteristics of graphite- and  PdSe 2 -contacted  WSe 2  FETs: (a) Schematic representation of device configuration with optical images of  WSe 2  FETs (scale bar: 10 µm). (b,c) Semi-log electrical transfer curves of devices with graphite (Gr) and  PdSe 2  contacts, respectively. The  I D  in (b,c) is scaled by the mean width of the channels to allow for better comparison of the current values between the different devices. The horizontal dashed lines that interconnect (b,c) serve as a guide to see the reached on- and off-state current levels. The red arrow in (c) indicates over an order of magnitude larger current of the hole branch in the case of  PdSe 2  contacts. (d) Comparison of the device width-scaled electrical transfer curves ( PdSe 2  contacted device) measured at 300 K (orange) and 78 K (purple), presented in linear scale. The arrows indicate the direction of the  V G  sweep, highlighting an increase in the hysteresis observed at 300 K. (e) Output curves for the hole and electron branches at 78 K (2 ×  10 2  mbar) of a device with  PdSe 2  contacts. Note that the current values for the n-branch are approximately one order of magnitude lower than for the p-branch. The different colored lines in (e) represent the curves at the different values of  V G , as indicated in the figure.
Nanomaterials 14 00481 g001
Figure 2. Contact resistance of the  PdSe 2 /WSe 2  interface: (a) Equivalent electrical scheme used for the self-consistent modelling of the output curves. (b,c) Electrical output curves of a PdSe 2 /WSe 2 /PdSe 2  device measured at 300 K for the hole and electron branches, respectively. Different colored circles represent the measured  I D  values at set different  V G  as indicated in the right corner of the sub-panels (b,c). The dashed lines are a model for the entire data set. Red arrows in (c) indicate a severe downward bending of the output curves at lower  V D . Contact resistance values ( W R C ) extracted by modelling the curves from (b,c) are indicated in each sub-panel. (df) In operando KPFM potential profiles recorded as single lines across the channel, measured under ambient conditions. Solid lines present the work function difference corrected potential drops, and the dashed lines are linear fits to the experimental curves. (d) A sequence of the potential drops with varied  V D . (e) Alternating the source and drain contacts, which demonstrates that the steep potential drop is related to the grounded electrode. (f) Comparison of the potential drops at  V D  = 1.5 V, with  V G  setting the device in an on state of the hole and electron branches, labelled with (1) and (2), respectively. Insets in (f) provide the operation points and the extracted  W R C  values from the KPFM measurements.
Figure 2. Contact resistance of the  PdSe 2 /WSe 2  interface: (a) Equivalent electrical scheme used for the self-consistent modelling of the output curves. (b,c) Electrical output curves of a PdSe 2 /WSe 2 /PdSe 2  device measured at 300 K for the hole and electron branches, respectively. Different colored circles represent the measured  I D  values at set different  V G  as indicated in the right corner of the sub-panels (b,c). The dashed lines are a model for the entire data set. Red arrows in (c) indicate a severe downward bending of the output curves at lower  V D . Contact resistance values ( W R C ) extracted by modelling the curves from (b,c) are indicated in each sub-panel. (df) In operando KPFM potential profiles recorded as single lines across the channel, measured under ambient conditions. Solid lines present the work function difference corrected potential drops, and the dashed lines are linear fits to the experimental curves. (d) A sequence of the potential drops with varied  V D . (e) Alternating the source and drain contacts, which demonstrates that the steep potential drop is related to the grounded electrode. (f) Comparison of the potential drops at  V D  = 1.5 V, with  V G  setting the device in an on state of the hole and electron branches, labelled with (1) and (2), respectively. Insets in (f) provide the operation points and the extracted  W R C  values from the KPFM measurements.
Nanomaterials 14 00481 g002
Figure 3. Laser treatment of  WSe 2 : ((a) i i i i ) Schematic cross-section of the laser-treated devices (not to scale), presenting the laser treatment process of the electrode interface step by step. ((b) i i i i ) Optical micrographs (scale 5 µm) of a representative device corresponding to each fabrication step in ((a) i i i i ). ((a,b) i) The heterostack of  WSe 2 /hBN on a local gate electrode prior to the laser treatment, and ((a,b i i ) after the top part of the  WSe 2  flake was scanned by the laser (exposed part of the  WSe 2  flake is indicated by the dashed lines). ((a,b i i i ) The same device after stamping of  PdSe 2  contacts. In the presented case, only one side of the channel–electrode interface was laser-treated. (c) A zoom in on the schematic in ((a i i i ) highlighting the part of the ablated  WSe 2  layer, part of the oxidised  WSe y O x  layer, and the unmodified part of the  WSe 2  layer.
Figure 3. Laser treatment of  WSe 2 : ((a) i i i i ) Schematic cross-section of the laser-treated devices (not to scale), presenting the laser treatment process of the electrode interface step by step. ((b) i i i i ) Optical micrographs (scale 5 µm) of a representative device corresponding to each fabrication step in ((a) i i i i ). ((a,b) i) The heterostack of  WSe 2 /hBN on a local gate electrode prior to the laser treatment, and ((a,b i i ) after the top part of the  WSe 2  flake was scanned by the laser (exposed part of the  WSe 2  flake is indicated by the dashed lines). ((a,b i i i ) The same device after stamping of  PdSe 2  contacts. In the presented case, only one side of the channel–electrode interface was laser-treated. (c) A zoom in on the schematic in ((a i i i ) highlighting the part of the ablated  WSe 2  layer, part of the oxidised  WSe y O x  layer, and the unmodified part of the  WSe 2  layer.
Nanomaterials 14 00481 g003
Figure 4. Topography changes and Raman investigation of laser-treated  WSe 2 : (a) Atomic force microscopy (AFM) image of a  WSe 2  flake on hBN before laser exposure with the corresponding line profile (height) of the flake. The predominant morphological features are water/air bubbles trapped at the WSe 2 /hBN and  hBN/SiO 2  interfaces. (b) The same area as in (a) treated with a 50 mW 532 nm laser beam. The exposed region is marked with a dashed rectangle, and the laser scanning direction is indicated with an arrow. The corresponding height profiles are presented at the bottom of the topography images. (a,b) Lateral scale bar 2 µm, z-scale 25 nm. (c) Raman spectrum before and after laser irradiation, recorded with 5 mW, 532 nm, and 5 × 10 s acquisition parameters. The main  WSe 2  peaks are preserved and enhanced in intensity after the laser treatment. Inset (b) presents a zoomed-in region of the main  E 1 2g  and A 1 g  modes.
Figure 4. Topography changes and Raman investigation of laser-treated  WSe 2 : (a) Atomic force microscopy (AFM) image of a  WSe 2  flake on hBN before laser exposure with the corresponding line profile (height) of the flake. The predominant morphological features are water/air bubbles trapped at the WSe 2 /hBN and  hBN/SiO 2  interfaces. (b) The same area as in (a) treated with a 50 mW 532 nm laser beam. The exposed region is marked with a dashed rectangle, and the laser scanning direction is indicated with an arrow. The corresponding height profiles are presented at the bottom of the topography images. (a,b) Lateral scale bar 2 µm, z-scale 25 nm. (c) Raman spectrum before and after laser irradiation, recorded with 5 mW, 532 nm, and 5 × 10 s acquisition parameters. The main  WSe 2  peaks are preserved and enhanced in intensity after the laser treatment. Inset (b) presents a zoomed-in region of the main  E 1 2g  and A 1 g  modes.
Nanomaterials 14 00481 g004
Figure 5. Electrical response of  WSe 2 /WSe y O x /PdSe 2  electrode interface: (a) Semi-log scale electrical transfer curves of a  WSe 2  device with both source and drain electrode interfaces treated by a laser prior to stamping  PdSe 2  contacts measured at 298 K, 2 ×  10 2  mbar. (b V H  plot as a function of a scan speed (measured at 298 K, 2 ×  10 2  mbar). (c) Position of the  V th  for both forward and backward  V G  sweeping with varied  V D . The difference indicates the hysteresis ( V H ) is independent of  V D . (d) Semi-log scale electrical transfer curves for an asymmetric  WSe 2  FET with only one contact pad treated by the laser. The dotted lines represent the drain electrode connected to the non-treated  PdSe 2  contact side, while source and drain were swapped for the solid black line. The arrows indicate the  V G  sweeping direction.
Figure 5. Electrical response of  WSe 2 /WSe y O x /PdSe 2  electrode interface: (a) Semi-log scale electrical transfer curves of a  WSe 2  device with both source and drain electrode interfaces treated by a laser prior to stamping  PdSe 2  contacts measured at 298 K, 2 ×  10 2  mbar. (b V H  plot as a function of a scan speed (measured at 298 K, 2 ×  10 2  mbar). (c) Position of the  V th  for both forward and backward  V G  sweeping with varied  V D . The difference indicates the hysteresis ( V H ) is independent of  V D . (d) Semi-log scale electrical transfer curves for an asymmetric  WSe 2  FET with only one contact pad treated by the laser. The dotted lines represent the drain electrode connected to the non-treated  PdSe 2  contact side, while source and drain were swapped for the solid black line. The arrows indicate the  V G  sweeping direction.
Nanomaterials 14 00481 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Murastov, G.; Aslam, M.A.; Leitner, S.; Tkachuk, V.; Plutnarová, I.; Pavlica, E.; Rodriguez, R.D.; Sofer, Z.; Matković, A. Multi-Layer Palladium Diselenide as a Contact Material for Two-Dimensional Tungsten Diselenide Field-Effect Transistors. Nanomaterials 2024, 14, 481. https://doi.org/10.3390/nano14050481

AMA Style

Murastov G, Aslam MA, Leitner S, Tkachuk V, Plutnarová I, Pavlica E, Rodriguez RD, Sofer Z, Matković A. Multi-Layer Palladium Diselenide as a Contact Material for Two-Dimensional Tungsten Diselenide Field-Effect Transistors. Nanomaterials. 2024; 14(5):481. https://doi.org/10.3390/nano14050481

Chicago/Turabian Style

Murastov, Gennadiy, Muhammad Awais Aslam, Simon Leitner, Vadym Tkachuk, Iva Plutnarová, Egon Pavlica, Raul D. Rodriguez, Zdenek Sofer, and Aleksandar Matković. 2024. "Multi-Layer Palladium Diselenide as a Contact Material for Two-Dimensional Tungsten Diselenide Field-Effect Transistors" Nanomaterials 14, no. 5: 481. https://doi.org/10.3390/nano14050481

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop