Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2024 Aug 26;16(36):48556–48564. doi: 10.1021/acsami.4c09688

Contact Resistance Engineering in WS2-Based FET with MoS2 Under-Contact Interlayer: A Statistical Approach

Małgorzata Giza †,*, Michał Świniarski , Arkadiusz P Gertych , Karolina Czerniak-Łosiewicz , Maciej Rogala , Paweł J Kowalczyk , Mariusz Zdrojek †,*
PMCID: PMC11403553  PMID: 39186441

Abstract

graphic file with name am4c09688_0005.jpg

One of the primary factors hindering the development of 2D material-based devices is the difficulty of overcoming fabrication processes, which pose a challenge in achieving low-resistance contacts. Widely used metal deposition methods lead to unfavorable Fermi level pinning effect (FLP), which prevents control over the Schottky barrier height at the metal/2D material junction. We propose to harness the FLP effect to lower contact resistance in field-effect transistors (FETs) by using an additional 2D interlayer at the conducting channel and metallic contact interface (under-contact interlayer). To do so, we developed a new approach using the gold-assisted transfer method, which enables the fabrication of heterostructures consisting of TMDs monolayers with complex shapes, prepatterned using e-beam lithography, with lateral dimensions even down to 100 nm. We designed and demonstrated tungsten disulfide (WS2) monolayer-based devices in which the molybdenum disulfide (MoS2) monolayer is placed only in the contact area of the FET, creating an Au/MoS2/WS2 junction, which effectively reduces contact resistance by over 60% and improves the Ion/Ioff ratio 10 times in comparison to WS2-based devices without MoS2 under-contact interlayer. The enhancement in the device operation arises from the FLP effect occurring only at the interface between the metal and the first layer of the MoS2/WS2 heterostructure. This results in favorable band alignment, which enhances the current flow through the junction. To ensure the reproducibility of our devices, we systematically analyzed 160 FET devices fabricated with under-contact interlayer and without it. Statistical analysis shows a consistent improvement in the operation of the device and reveals the impact of contact resistance on key FET performance indicators.

Keywords: contact resistance, van der Waals heterostructure, interlayer, transition metal dichalcogenides, field-effect transistor, gold-assisted transfer

Introduction

Transition metal dichalcogenides (TMDs) are a broad family of materials with intriguing properties, making them highly attractive for a wide range of applications. In recent years, TMDs have been used as one of the building blocks in van der Waals heterostructures, leading to the development of efficient semiconductor components like highly scaled field-effect transistors,1 wide-range photodetectors,2 sensors,3 and flash memories.4 Although all mentioned types of devices are promising for next-generation electronics and optoelectronics, they need further development due to difficult-to-overcome processing steps, which affect the properties of thin TMD layers.

The critical problem in producing any 2D material-based devices is obtaining low-resistance contacts. One of the primary methods of controlling the current flowing through a metal/semiconductor junction is to align the metal work function to the semiconductor’s electron affinity to minimize the Schottky barrier height. In bulk semiconductors, low contact resistance is often facilitated by methods that can locally dope the area under the contact. This strategy is challenging to implement in the case of two-dimensional TMDs due to their atomic thickness. Moreover, widely used methods of producing contacts (e.g., resistive evaporation5 or e-beam evaporation)6 damage the structure of TMD materials and lead to the Fermi level pinning (FLP) at the metal/semiconductor junction. This effect makes Schottky barrier height, in fact, independent from metal work function.7

Several approaches have recently been investigated to address this issue, and the FLP effect has even been used to create new strategies for contact engineering in TMD-based devices. One of the most promising is weakening the FLP at the metal/TMDs interface by suppressing metal-induced gap states (MIGS) with semimetal contacts. For instance, bismuth has near-zero DOS at the Fermi level, which is also positioned close to the conduction band of tungsten disulfide (WS2) and molybdenum disulfide (MoS2), resulting in ultralow contact resistance at the Bi/WS2 and Bi/MoS2 junctions.8 Another example involves traditional substitutional doping techniques with Cl-dopants9 or chemical surface molecular n-dopants,10 which can passivate sulfur vacancies at the interface and thus move the position of the Fermi level. Few layer graphene11 and indium12 van der Waals contacts also exhibit a highly depinning nature due to a lack of interface states and covalent bonds to MoS2 and WS2. A similar strategy is applied by using buffer layers (interlayers), which can reduce contact resistance by preventing interaction between the metal and 2D semiconductor. For example, a dielectric layer of transition metal oxide TiO2 has been used to unpin the Fermi level in field-effect transistors based on WS2,13 reducing the contact resistance by over 18 times. Recently, even thin layered semiconductors (WSe2, MoSe2)14 and insulators (hBN)15,16 have been used as interlayers, creating van der Waals heterostructure-based devices with a lowered contact resistance from 55% even up to 90% in comparison to devices without interlayer. Usually, in reported studies, the interlayer is located both under the contact and in the FET’s channel. Without comparing the devices in which the interlayer is present only in the contact area and not in the channel, it is difficult to determine the actual impact of the interlayer on the current flowing through the metal/2D material junction. The lack of such studies results from challenges in the fabrication of complex devices (e.g., with heterostructures only in the contact area) caused by the necessity of using selective etching, which removes individual layers from specific areas of the heterostructure. However, selective etching of heterostructures with atomic precision, which can address each layer of the heterostructure individually, is still in development. Currently, selective etching methods for layered materials are mainly applied to graphene17,18 and hBN19 layers, there is a lack of studies performed on TMDs. There have been two alternative methods reported for selective etching. In one, layers of Nb-doped WSe2 with a thickness of 21 nm were patterned and etched into suitable shapes, then transferred using PDMS onto WSe2.20 In another, 40 nm thick wheel-shaped MoS2 layers were transferred onto the MoS2 monolayer using a silicon nitride membrane with a gold adhesion layer.21 Moreover, other studies show heterostructures based on TMDs monolayers with defined shapes with laterals sizes from 10 to 100 μm.22,23 However, to the best of our knowledge, no studies have been published where selective etching is replaced by a method that allows the fabrication of heterostructures composed of TMDs monolayers with arbitrary shapes down to 100 nm in lateral dimensions.

Moreover, those new transfer strategies and, thus, the reduction of contact resistance, are often evaluated based on the performances of only a few devices. This may hide issues related to the performance reproducibility and potential inhomogeneity of 2D materials. There is a lack of more extensive studies on this topic, with conclusions based on statistical measures of a large number of devices, which is particularly crucial when investigating complex 2D material-based structures.

In this work, we are the first to demonstrate a Au/MoS2/WS2 junction to effectively reduce contact resistance in monolayer WS2-based field-effect transistors. To enable this, we propose a new fabrication approach using a gold tape based transfer method, allowing the creation of numerous monolayer-based van der Waals heterostructures in one transfer process with lateral dimensions even down to 100 nm. In our devices fabricated with gold-assisted transfer, the MoS2 monolayer, acting as an under-contact interlayer (UCI), which is a layer directly between the channel (WS2) and metal contact (Au), enables using FLP to our advantage for achieving favorable band alignment at the junction. To access reproducibility and thoroughly investigate the influence of the UCI on the device’s performance, we examined key FET performance indicators of 160 devices (80 devices with and 80 devices without MoS2 UCI). In the designed devices with a UCI, the channel of the FET consists only of the WS2 monolayer, while the MoS2/WS2 van der Waals heterostructure is located only below the contact area of the device. This architecture of the FET allows for a reliable determination of the contact resistance, which decreased by over 60% due to the influence of the MoS2 UCI.

Results and Discussion

Devices with and without under-contact interlayer (UCI) were fabricated on SiO2/Si substrate using monolayers of MoS2 and WS2 prepared with gold-assisted mechanical exfoliation (as described in the “Methods” section). To create the unique heterostructures consisting of MoS2 stripes on the WS2 monolayer, we developed an approach using the gold-assisted transfer method (Figure 1a), which allows for the transfer of prepatterned monolayers with shapes defined by e-beam lithography and etching (details provided in the “Methods” section). The transfer method is possible because of the higher binding energy of MoS2 to Au than to SiO2 substrate.24 Moreover, the nobility of Au and interfacial strain between Au and MoS225 facilitates the detachment of the MoS2 monolayer from SiO2. In our work, the MoS2 monolayer was patterned and etched to form stripes with varying spacings for contact resistance investigation using the transfer length method (TLM). MoS2 stripes were transferred with gold tape on a continuous WS2 monolayer. The optical image presented in Figure 1b illustrates the capability of gold-assisted transfer to fabricate numerous heterostructures during a single transfer (on the single substrate), covering a total area of structure exceeding 500 000 μm2. With our method, in which we are using layers shaped by the e-beam lithography technique, it is possible to create heterostructures with precisely defined edges and lateral sizes even down to 100 nm (Figure 1c), making it suitable for scale-down devices. The Raman and photoluminescence spectra of one of the heterostructures consisting of monolayer MoS2 and WS2 are shown in Figures 1d,e. More extensive analysis, including Raman mapping, AFM, and SEM images, showcasing the transfer of monolayers is presented in Supporting Information 1.

Figure 1.

Figure 1

a) Simplified schematic representation of the gold-assisted transfer process. #n represents different SiO2/Si substrates (the detailed process is described in the Methods section). b) An example of heterostructure with monolayer MoS2 stripes transferred on continuous WS2 monolayer. The scale bar on the left picture is 150 μm, and on the right picture is 10 μm. c) SEM image of patterned and etched MoS2 monolayer on SiO2/Si substrate (Top image) and MoS2 monolayer transfer on WS2 monolayer on SiO2/Si substrate (Bottom image). The images show that the edges are well-preserved after the transfer and that a 100 nm-wide MoS2 stripe was successfully transferred. d) Raman spectrum and e) photoluminescence spectrum of MoS2/WS2 heterostructure with marked peaks corresponding to the individual component layers.

In order to fabricate field-effect transistors, Au contacts were thermally evaporated on MoS2/WS2 heterostructure areas, creating Au/MoS2/WS2 junctions (Figure 2a,b). In the designed devices with UCI, the MoS2 stripes have a width of 1.3 μm, while the gold contacts have a width of 1 μm to ensure no direct contact between Au and WS2. Channel lengths (L) are 0.5, 0.75, 1.5, 2, 3, 5, and 8 μm, which are strictly related to the arrangement of transferred MoS2 stripes. The channel width (W) is 6 μm and remains the same in all devices. To see the influence of the UCI on the performance of devices, we also prepared devices only with the WS2 monolayer (without the MoS2 as a UCI) with the same dimensions of the channel and contacts. Considering that both types of devices were not prepared on exactly the same WS2 monolayer, it was necessary to exclude differences in the fabrication processes. In the case of devices with UCI, gold was used to both exfoliate and transfer, so it was important to ensure that the gold was removed entirely after fabrication. If the etching of the gold layer is performed poorly, then it can leave behind residues that dope the TMDs monolayer,26 leading to a change in FET properties. To prove the elimination of the effects related to the fabrication, we performed XPS analysis (see Supporting Information 2), which revealed that the presence of the Au 4f doublet line in the spectra was below the detection limit estimated for 0.01% of atom concentration in the surface region. This demonstrates that the cleaning process is effective, and any expected differences in the performance of devices with and without UCI should not be attributed to gold residues.

Figure 2.

Figure 2

a) Schematic of the device with UCI. b) An optical image of the heterostructure with various spacing between MoS2 stripes placed on the WS2 layer (Top image). An optical image of the device shows gold contacts evaporated on the heterostructure area, creating a configuration that enables the use of the TLM method (Bottom image). c) Schematic representation of band alignment at Au/WS2 junction. d) Schematic representation of band alignment at Au/MoS2/WS2 junctions. e) Normalized transfer characteristics plotted on a linear scale, comparing representative devices with and without UCI. The inset shows the same results on a semilogarithmic scale. Measurements were performed with Vds = 1 V. f) Normalized output characteristics for different gate voltages from 0 to 100 V, with a 20 V step for a representative device without a UCI, and g) for a representative device with a UCI. h) Total contact resistance Rt dependence on channel length L calculated for carrier concentration of ns = 2 × 1012 cm–2 for two representative TLM sets with fitted linear function for Rc extraction. (i) Mean value and standard deviation of contact resistance (Rc) as a function of carrier concentrations. j) Mean value and standard deviation of sheet resistance (Rsh) as a function of carrier concentrations.

Fabricated devices were electrically characterized in a back-gate configuration under high vacuum conditions. Transfer characteristics for representative devices with and without a UCI are shown in Figure 2e. Characteristics were normalized by the channel length and width. Devices with a UCI exhibit a lower threshold voltage and higher on-state current, which indicates that the MoS2 UCI enhances the performance of the devices. In Figure 2f,g, output characteristics demonstrate source-drain current three times higher for devices with UCI. Meanwhile, the shape of characteristics of both types of devices remains nonlinear and corresponds to the typical current–voltage dependence in FETs with Schottky contacts. It is commonly observed that the Fermi level pinning is present at the metal/2D-TMD junctions, affecting the Schottky barrier height at the interface. The level of pinning in a system consisting of specific TMD and various metals is almost independent of the metal’s work function,27 which makes FLP a crucial issue in designing 2D FETs. However, when junctions with a specific metal and various TMDs are compared, the work function of the combined system (metal and TMD layer) changes significantly with different TMDs. Consequently, the pinning level also varies considerably depending on the used TMD.28 In such a case, the Schottky barrier height (ϕSB) at the junction will depend on the work function of the combined system (level of pinning, EFP) and the electron affinity of the TMD layer (Χ),14 as in eq 1.

graphic file with name am4c09688_m001.jpg 1

Previously in the literature, it was confirmed that the FLP effect is occurring at the Au/MoS2 and Au/WS2 interfaces, but the level of pinning is preset at different energy levels in those junctions.29,30 In Figure 2c,d, we schematically show band diagrams of Au/MoS2/WS2 and Au/WS2. The work function of the combined system of Au/MoS2 has a value lower than the work function of the Au/WS2 system. Moreover, the FLP effect is layer-number dependent and strongly affects only the first layer of the material in contact with the metal.31 Therefore, we expect that by using the MoS2 interlayer and creating the Au/MoS2/WS2 junction, we are mitigating FLP and achieving lower Schottky barrier height, which leads to reduced Rc. Furthermore, we are anticipating that the mismatch between the conduction bands of MoS2 and WS2 is creating a step-like offset at the interface of those materials (Figure 2d), which can reduce contact resistance and improve the efficiency of the mechanism responsible for the thermionic transport occurring through the Au/MoS2/WS2 junction.

To verify our predictions, we investigated 80 devices with UCI and 80 devices without UCI. We calculated the total resistance of devices based on data from the linear operation regime of transfer characteristics of 13 TLM structures of each type and used it to extract contact resistance and sheet resistance (from channel resistance, Rch) based on eq 2 and 3.

graphic file with name am4c09688_m002.jpg 2
graphic file with name am4c09688_m003.jpg 3

Representative results of total resistance Rt calculated for the gate voltage corresponding to the carrier concentration of ns = 2 × 1012 cm–2, at which devices are in on-state (linear operation regime), are presented in Figure 2h for representative TLM sets of both types of devices (formulas and detailed graphs are demonstrated in Supporting Information 3). Figure 2i,j shows extracted mean values and standard deviations of Rc and Rch for devices with and without UCI for different carrier concentrations ns. Rc and Rch strongly depend on ns. Carrier concentration in the FET’s channel is modulated by gate voltage, which also affects the interface between channel material and contact. This effect, known as contact-gating, occurs in devices with global back-gate geometry and lowers the Schottky barrier width at the junction by applied back-gate voltage, thus allowing carrier tunneling through the barrier and reducing Rc.32 For all values of carrier concentration, the contact resistance of the Au/MoS2/WS2 junction is around 60% lower than the contact resistance of the Au/WS2 junction. This supports our prediction that using MoS2 UCI can create a more favorable band alignment, reducing contact resistance. Comparing the values for ns = 2 × 1012 cm–2, at which devices are in on-state, we achieved mean Rc values of 242 kΩ·μm with a standard deviation of 72 kΩ·μm for devices without UCI, which are comparable to other studies33,34 and 79 kΩ·μm with a standard deviation of 27 kΩ·μm for devices with a UCI. A similar reduction of Rc, but based on a different Schottky barrier-lowering mechanism, was observed in WS2 devices with graphene interlayer and Ni contacts34 (more extended comparison is presented in Supporting Information 4). However, in the case of our devices with UCI, we are not unpinning the Fermi level, but by using the MoS2 monolayer, we create a junction with a more advantageous energy landscape using the FLP effect in our favor. Moreover, to ensure that the decrease in Rc is not attributed to the extension of contact length and simultaneous reduction of channel length, we considered the possible impact of an additional 150 nm long MoS2 monolayer extending beyond each side of the contact on Rc. We performed the analysis presented in Supporting Information 5, which confirms that the observed reduction cannot be attributed to changes in the effective geometry of the device.

To investigate device-to-device variation in a more statistical approach and assess the impact of contact resistance reduction on FETs performance, in Figure 3, we presented crucial parameters such as threshold voltage (Vth), field-effect mobility (μFE), subthreshold swing (SS), and on- to off-state current ratio (Ion/Ioff) (details in Supporting Information 6). Figure 3a shows the distribution of Vth as a function of channel length (L) and corresponding mean values connected with lines as guides to the eye. High Vth values result from 285 nm thick SiO2. For devices without a UCI, values of Vth are close to 70 V with a slight dependence of L compared to Vth of devices with a UCI, which exhibits a significant dependence of L. The negative shift and change in dependency on channel length can be attributed to different Schottky barrier heights in Au/WS2 and Au/MoS2/WS2 junctions. A high Schottky barrier leads to an increased Vth, indicating that the barrier at the Au/WS2 junction is anticipated to be higher than that at the Au/MoS2/WS2 junction. The dependence of Vth on the channel length can be related to the significant Rc value. When the channel length is scaled down, the influence of Rc becomes more substantial, ultimately becoming the primary factor restricting device performance as it dominates the total device resistance,35 resulting in higher Vth for shorter channels.

Figure 3.

Figure 3

Distribution of a) threshold voltage, b) field-effect carrier mobility, and c) normalized on-state current (Ion) for different channel lengths for devices with and without UCI. Lines represent mean values. d) Histograms showing the variation in subthreshold swing (SS) and e) on-state current to off-state current ratio for devices with and without UCI.

A similar influence of Rc is seen in field-effect mobility (μFE) values (Figure 3b), which show strong channel length dependence, commonly observed due to the effect of Fermi level pinning existing in TMDs/metal junctions.36 Field-effect mobility values are higher for devices with a UCI for each channel length, which may initially seem unusual, because the channel area is the same in both types of devices and consists only of the WS2 monolayer. In fact, μFE depends not only on the intrinsic properties of WS2 but also on the properties of the entire device, which is why a strong influence of the contact resistance can be noticed. Higher values of μFE indicate reduced contact resistance in devices with UCI. This observation is supported by Ion values (Figure 3c) that are greater for devices with UCI and strongly dependent on L despite being normalized by channel dimensions. For both types of devices, μFE reaches almost 20 cm2/(V s), which is comparable with other studies performed on WS2-based devices.36,37Figure 3d shows the distribution of SS. The large SS values for both types of devices result from using a 285 nm thick gate oxide layer. Devices without UCI exhibit SS values with the mode at 11 V/dec, whereas for devices with UCI, the SS is over 3 times lower with the mode value of 3 V/dec. The SS parameter is calculated across voltages lower than the threshold voltage and represents the rate at which the device switches from the off-state to the on-state. Hence, when a higher Schottky barrier and consequently increased contact resistance are encountered, a higher SS can be expected, hindering the flow of charges through the junction. Figure 3e shows the distribution of Ion/Ioff with a mode value for devices with UCI at the level of 106 and for devices without UCI at 105. This supports the previous observation that current in devices without UCI is limited. SS and Ion/Ioff ratios show no dependency on channel length for both types of devices with and without UCI (Supporting Information 7).

To further investigate the influence of contact resistance on devices’ performance, we illustrate the contribution of contact resistance and channel resistance (Rch), calculated with mean values using eq 3, to the mean total resistance of the device in Figure 4a,b for different channel lengths.

Figure 4.

Figure 4

a) Stacked bar plots of Rch and Rc contribution to the total resistance of devices with UCI and b) without UCI.

Rc and Rch are independent of the channel length. Simultaneously, Rch is proportional to channel length, which leads to varying impacts of 2Rc on Rt for long and short channels. For both types of devices, the contribution of 2Rc to Rt is much more significant for shorter channels. Furthermore, 2Rc/Rt shows a stronger channel length dependence in devices with UCI. In contrast, devices without UCI, even with the longest channels, exhibit almost half of the total device resistance being caused by effects occurring on Au/WS2 junctions. This observation aligns with the analysis of Vth and μFE in Figure 3a,b, which demonstrated that these parameters vary with different channel lengths, which is directly related to the impact of contact resistance on the operation of devices. Additionally, the channel length dependence for those parameters was weaker in devices without UCI, consistent with the result in Figure 4, which indicates that 2Rc’s contribution to Rt remains substantial across all channel lengths, while for devices with UCI, it drops significantly due to reduced Rc.

Conclusions

In this study, we are the first to report the enhancement of the performance of monolayer WS2-based FETs by using a monolayer of MoS2, which serves as an under-contact interlayer. To create MoS2/WS2 van der Waals heterostructure, we developed a gold-assisted method that enables the transfer of prepatterned TMD monolayers with lateral dimensions even down to 100 nm. With our method, we can fabricate numerous heterostructures with a total area exceeding 500 000 μm2 in one transfer process. Our approach enables us to reduce contact resistance by over 60% due to favorable band alignment at the Au/MoS2/WS2 junction. Statistical analysis of 160 FETs with and without an UCI highlighted the influence of contact resistance on the operation of devices both in the contact area and in the channel. Therefore, the further development of new strategies is needed to reduce this significant factor in order to achieve high-performance TMD-based electronics and optoelectronics.

Methods

TMD Monolayer Preparation

Large-area monolayers of MoS2 and WS2 were fabricated using gold-assisted exfoliation.38 Si wafers (Cemat Silicon) were cleaned by using argon plasma treatment (Diener Zepto). The plasma process was held for 10 min with 5 sccm gas flow and power of 4 W. Cleaned Si wafers were vacuum annealed at 300 °C for 3 h in a Kurt J. Lesker Nano 36 chamber followed by thermal evaporation of 100 nm gold layer. A 1 μm thick PMMA protective layer was spin-coated on the gold surface and baked for 2 min at 150 °C before the wafers were cleaved into smaller pieces. Thermal release tape (TRT Revalpha 3195MS) was placed on the PMMA layer. The bulk of chosen TMDs (2D Semiconductors) was prepared by peeling off the oxidized top layer of the crystal using Nitto tape. Gold tape (TRT/PMMA/Au) released from the Si substrate was immediately applied to the surface of the bulk crystal and gently peeled off. The gold tape with an exfoliated layer was placed on a 285 nm SiO2/Si substrate (Process Specialties). Samples were annealed at 150 °C for 10 min to peel off TRT and ensure strong adhesion of the exfoliated layers to the substrate. To remove PMMA and TRT residues, samples were soaked in trichloroethylene (TCE) at 50 °C for 10 min, then in acetone at 50 °C for 10 min, and rinsed in isopropanol. To remove the gold layer, samples were soaked in standard gold etchant (Sigma-Aldrich) for 5 min at room temperature and rinsed in two separate isopropanol solutions (10% IPA in DI water) and isopropanol. To remove left organic residues, WS2 samples were vacuum annealed at 350 °C for 3 h. This step was omitted in MoS2 samples to retain weaker adhesion to the substrate.

Fabrication of van der Waals Heterostructures

Samples with the MoS2 monolayer were patterned with e-beam lithography (Raith e-Line Plus) and etched with an O2 RIE plasma (Oxford Plasmalab 80 Plus). After the removal of the resist layer used for the lithography process, gold tape (fabrication described in the previous section) is placed on the sample, covered with a glass slide, and loaded with a weight of 120 g for 20 min to achieve strong adhesion between the gold layer and patterned MoS2. In the next step, the weight is removed, and the sample with gold tape on top is placed on a hot-plate and annealed at 150 °C for 20 min. TRT peels off under the temperature, and the PMMA/Au stack stays on the substrate. After the annealing, when a sample cools down, another TRT is placed on the sample. A patterned MoS2/Au/PMMA/TRT stack is peeled off using tweezers. The stack is placed on a monolayer WS2 sample, then covered with a glass slide, loaded with a weight of 120 g, and left for 20 min. Next, the weight is removed, and the sample with the stack is placed on a hot-plate and annealed at 150 °C for 20 min. The TRT peels off due to temperature, and other components of gold tape are removed the same way as after exfoliation, as described in the previous section.

Field-Effect Transistors Fabrication

Devices were fabricated in the back-gate configuration on a SiO2/Si substrate (285 nm thick silicon oxide) using e-beam lithography (Raith e-Line Plus) and O2 RIE etching (Oxford Plasmalab 80 Plus). In the case of devices with UCI, contacts were patterned only on the MoS2/WS2 heterostructure area. 70 nm pure Au contacts were thermally evaporated using a Kurt J. Lesker Nano 36, followed by a lift-off process.

Electrical Characterization

Electrical characterization was performed using a DL Instruments 1211 Current Preamplifier, National Instruments DAQ 6366, and source measuring unit Keithley 2450. Measurements were carried out in a vacuum (1e-6 mbar) in an Oxford MicrostatHe2 cryostat at room temperature. We performed electrical characterization using the standard two-probe method, applying voltage and sensing current. Before the measurements, the samples were kept in a vacuum for 19 h, during which they were thermally annealed for 13 h at 200 °C.

Raman Spectroscopy and Photoluminescence Measurements

Raman spectroscopy and photoluminescence measurements were carried out using a Renishaw inVia Qontor Raman spectrometer with a 532 nm excitation wavelength. Spectra were measured using a 100× objective with 1800 lines/mm grating. Laser power was kept below 0.5 mW to avoid damaging the material.

XPS

The XPS system was equipped with a hemispherical energy analyzer EA 125 (Omicron), an RS40B1 (Prevac) source, and an RMC50 monochromator (Prevac); monochromatic radiation of 1486.6 eV (Al Kα) was used. The peak fitting procedure was supported by Casa XPS software.

Acknowledgments

This research was supported by the PRELUDIUM BIS 2 project (2020/39/O/ST5/00416) by the National Science Centre, Poland. M.R. and P.J.K. acknowledge the National Science Centre, Poland, under the 2018/31/B/ST3/02450 grant.

Glossary

Abbreviations

TMDs

Transition metal dichalcogenides

FET

Field-effect transistor

MoS2

Molybdenum disulfide

WS2

Tungsten disulfide

FLP

Fermi level pinning

XPS

X-ray photoelectron spectroscopy

TRT

Thermal release tape

TCE

Trichloroethylene

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsami.4c09688.

  • SEM, AFM, and Raman characterization of MoS2/WS2 van der Waals heterostructure fabricated with gold-assisted transfer. Moreover, we include an evaluation of the removal of the gold layer after transfer based on XPS analysis. We also present a formula for extracting key FET parameters, a showcase of complete results of Rch and Rc for all devices, and SS and Ion/Ioff ratio dependence on channel length (PDF)

Author Contributions

M.G. designed the experiments. M.G., M.Ś., A.P.G., and K.C.Ł. made important contributions to interpreting the results. M.G. and A.P.G. developed a gold-assisted transfer method. M.G. exfoliated TMD monolayers, prepared heterostructures, fabricated the devices, and performed electrical measurements. A.P.G. and M.G. performed the Raman and PL measurements. M.R. and P.J.K. performed XPS measurements and analysis. M.Z. and M.Ś. supervised the study. M.G. wrote the manuscript. All the authors revised and commented on the paper.

The authors declare no competing financial interest.

Due to a production error, some minor changes were needed after this paper was published ASAP August 26, 2024. The corrected version was reposted August 26, 2024.

Supplementary Material

References

  1. Nourbakhsh A.; Zubair A.; Sajjad R. N.; Tavakkoli K. G. A.; Chen W.; Fang S.; Ling X.; Kong J.; Dresselhaus M. S.; Kaxiras E.; et al. MoS2 Field-Effect Transistor with Sub-10 Nm Channel Length. Nano Lett. 2016, 16 (12), 7798–7806. 10.1021/acs.nanolett.6b03999. [DOI] [PubMed] [Google Scholar]
  2. Flöry N.; Ma P.; Salamin Y.; Emboras A.; Taniguchi T.; Watanabe K.; Leuthold J.; Novotny L. Waveguide-Integrated van Der Waals Heterostructure Photodetector at Telecom Wavelengths with High Speed and High Responsivity. Nat. Nanotechnol. 2020, 15 (2), 118–124. 10.1038/s41565-019-0602-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Hou H.; Anichini C.; Samorì P.; Criado A.; Prato M. 2D Van Der Waals Heterostructures for Chemical Sensing. Adv. Funct Materials 2022, 32 (49), 2207065. 10.1002/adfm.202207065. [DOI] [Google Scholar]
  4. Vu Q. A.; Shin Y. S.; Kim Y. R.; Nguyen V. L.; Kang W. T.; Kim H.; Luong D. H.; Lee I. M.; Lee K.; Ko D.-S.; Heo J.; Park S.; Lee Y. H.; Yu W. J. Two-Terminal Floating-Gate Memory with van Der Waals Heterostructures for Ultrahigh on/off Ratio. Nat. Commun. 2016, 7 (1), 12725. 10.1038/ncomms12725. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Kim B.-K.; Kim T.-H.; Choi D.-H.; Kim H.; Watanabe K.; Taniguchi T.; Rho H.; Kim J.-J.; Kim Y.-H.; Bae M.-H. Origins of Genuine Ohmic van Der Waals Contact between Indium and MoS2. npj 2D Mater. Appl. 2021, 5 (1), 9. 10.1038/s41699-020-00191-z. [DOI] [Google Scholar]
  6. Liu L.; Kong L.; Li Q.; He C.; Ren L.; Tao Q.; Yang X.; Lin J.; Zhao B.; Li Z.; Chen Y.; Li W.; Song W.; Lu Z.; Li G.; Li S.; Duan X.; Pan A.; Liao L.; Liu Y. Transferred van Der Waals Metal Electrodes for Sub-1-Nm MoS2 Vertical Transistors. Nat. Electron 2021, 4 (5), 342–347. 10.1038/s41928-021-00566-0. [DOI] [Google Scholar]
  7. 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 (7707), 696–700. 10.1038/s41586-018-0129-8. [DOI] [PubMed] [Google Scholar]
  8. Shen P.-C.; Su C.; Lin Y.; Chou A.-S.; Cheng C.-C.; Park J.-H.; Chiu M.-H.; Lu A.-Y.; Tang H.-L.; Tavakoli M. M.; Pitner G.; Ji X.; Cai Z.; Mao N.; Wang J.; Tung V.; Li J.; Bokor J.; Zettl A.; Wu C.-I.; Palacios T.; Li L.-J.; Kong J. Ultralow Contact Resistance between Semimetal and Monolayer Semiconductors. Nature 2021, 593 (7858), 211–217. 10.1038/s41586-021-03472-9. [DOI] [PubMed] [Google Scholar]
  9. Yang L.; Majumdar K.; Liu H.; Du Y.; Wu H.; Hatzistergos M.; Hung P. Y.; Tieckelmann R.; Tsai W.; Hobbs C.; Ye P. D. Chloride Molecular Doping Technique on 2D Materials: WS 2 and MoS 2. Nano Lett. 2014, 14 (11), 6275–6280. 10.1021/nl502603d. [DOI] [PubMed] [Google Scholar]
  10. Zhang S.; Chuang H.-J.; Le S. T.; Richter C. A.; McCreary K. M.; Jonker B. T.; Hight Walker A. R.; Hacker C. A. Control of the Schottky Barrier Height in Monolayer WS2 FETs Using Molecular Doping. AIP Advances 2022, 12 (8), 085222. 10.1063/5.0101033. [DOI] [Google Scholar]
  11. Murali K.; Dandu M.; Watanabe K.; Taniguchi T.; Majumdar K. Accurate Extraction of Schottky Barrier Height and Universality of Fermi Level De-Pinning of van Der Waals Contacts. Adv. Funct Materials 2021, 31 (18), 2010513. 10.1002/adfm.202010513. [DOI] [Google Scholar]
  12. Wang Y.; Kim J. C.; Wu R. J.; Martinez J.; Song X.; Yang J.; Zhao F.; Mkhoyan A.; Jeong H. Y.; Chhowalla M. Van Der Waals Contacts between Three-Dimensional Metals and Two-Dimensional Semiconductors. Nature 2019, 568 (7750), 70–74. 10.1038/s41586-019-1052-3. [DOI] [PubMed] [Google Scholar]
  13. Park W.; Kim Y.; Jung U.; Yang J. H.; Cho C.; Kim Y. J.; Hasan S. M. N.; Kim H. G.; Lee H. B. R.; Lee B. H. Complementary Unipolar WS 2 Field-Effect Transistors Using Fermi-Level Depinning Layers. Adv. Elect Materials 2016, 2 (2), 1500278. 10.1002/aelm.201500278. [DOI] [Google Scholar]
  14. Andrews K.; Bowman A.; Rijal U.; Chen P.-Y.; Zhou Z. Improved Contacts and Device Performance in MoS 2 Transistors Using a 2D Semiconductor Interlayer. ACS Nano 2020, 14 (5), 6232–6241. 10.1021/acsnano.0c02303. [DOI] [PubMed] [Google Scholar]
  15. Phan N. A. N.; Noh H.; Kim J.; Kim Y.; Kim H.; Whang D.; Aoki N.; Watanabe K.; Taniguchi T.; Kim G. Enhanced Performance of WS 2 Field-Effect Transistor through Mono and Bilayer h-BN Tunneling Contacts. Small 2022, 18 (13), 2105753. 10.1002/smll.202105753. [DOI] [PubMed] [Google Scholar]
  16. Wang J.; Yao Q.; Huang C.; Zou X.; Liao L.; Chen S.; Fan Z.; Zhang K.; Wu W.; Xiao X.; Jiang C.; Wu W. High Mobility MoS 2 Transistor with Low Schottky Barrier Contact by Using Atomic Thick h-BN as a Tunneling Layer. Adv. Mater. 2016, 28 (37), 8302–8308. 10.1002/adma.201602757. [DOI] [PubMed] [Google Scholar]
  17. Son J.; Kwon J.; Kim S.; Lv Y.; Yu J.; Lee J.-Y.; Ryu H.; Watanabe K.; Taniguchi T.; Garrido-Menacho R.; et al. Atomically Precise Graphene Etch Stops for Three Dimensional Integrated Systems from Two Dimensional Material Heterostructures. Nat. Commun. 2018, 9 (1), 3988. 10.1038/s41467-018-06524-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Sun Z.; Peng G.; Bai Z.; Zhang X.; Wei Y.; Deng C.; Zhang Y.; Zhu M.; Qin S.; Li Z.; Luo W. Selective Etching in Graphene–MoS2 Heterostructures for Fabricating Graphene-Contacted MoS2 Transistors. AIP Advances 2020, 10 (3), 035219. 10.1063/1.5141143. [DOI] [Google Scholar]
  19. Lee C.; Rathi S.; Khan M. A.; Lim D.; Kim Y.; Yun S. J.; Youn D.-H.; Watanabe K.; Taniguchi T.; Kim G.-H. Comparison of Trapped Charges and Hysteresis Behavior in hBN Encapsulated Single MoS 2 Flake Based Field Effect Transistors on SiO 2 and hBN Substrates. Nanotechnology 2018, 29 (33), 335202. 10.1088/1361-6528/aac6b0. [DOI] [PubMed] [Google Scholar]
  20. Chuang H.-J.; Chamlagain B.; Koehler M.; Perera M. M.; Yan J.; Mandrus D.; Tománek D.; Zhou Z. Low-Resistance 2D/2D Ohmic Contacts: A Universal Approach to High-Performance WSe 2, MoS 2, and MoSe 2 Transistors. Nano Lett. 2016, 16 (3), 1896–1902. 10.1021/acs.nanolett.5b05066. [DOI] [PubMed] [Google Scholar]
  21. Wang W.; Clark N.; Hamer M.; Carl A.; Tovari E.; Sullivan-Allsop S.; Tillotson E.; Gao Y.; De Latour H.; Selles F.; Howarth J.; Castanon E. G.; Zhou M.; Bai H.; Li X.; Weston A.; Watanabe K.; Taniguchi T.; Mattevi C.; Bointon T. H.; Wiper P. V.; Strudwick A. J.; Ponomarenko L. A.; Kretinin A. V.; Haigh S. J.; Summerfield A.; Gorbachev R. Clean Assembly of van Der Waals Heterostructures Using Silicon Nitride Membranes. Nat. Electron 2023, 6 (12), 981–990. 10.1038/s41928-023-01075-y. [DOI] [Google Scholar]
  22. Li Z.; Ren L.; Wang S.; Huang X.; Li Q.; Lu Z.; Ding S.; Deng H.; Chen P.; Lin J.; Hu Y.; Liao L.; Liu Y. Dry Exfoliation of Large-Area 2D Monolayer and Heterostructure Arrays. ACS Nano 2021, 15 (8), 13839–13846. 10.1021/acsnano.1c05734. [DOI] [PubMed] [Google Scholar]
  23. Nguyen V.; Gramling H.; Towle C.; Li W.; Lien D.-H.; Kim H.; Chrzan D. C.; Javey A.; Xu K.; Ager J.; Taylor H. Deterministic Assembly of Arrays of Lithographically Defined WS2 and MoS2Monolayer Features Directly From Multilayer Sources Into Van Der Waals Heterostructures. Journal of Micro and Nano-Manufacturing 2019, 7 (4), 041006. 10.1115/1.4045259. [DOI] [Google Scholar]
  24. Kim H.-J.; Kim D.; Jung S.; Bae M.-H.; Yi S. N.; Watanabe K.; Taniguchi T.; Chang S. K.; Ha D. H. Homogeneity and Tolerance to Heat of Monolayer MoS 2 on SiO 2 and h-BN. RSC Adv. 2018, 8 (23), 12900–12906. 10.1039/C8RA01849A. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Velický M.; Donnelly G. E.; Hendren W. R.; DeBenedetti W. J. I.; Hines M. A.; Novoselov K. S.; Abruña H. D.; Huang F.; Frank O. The Intricate Love Affairs between MoS 2 and Metallic Substrates. Adv. Materials Inter 2020, 7 (23), 2001324. 10.1002/admi.202001324. [DOI] [Google Scholar]
  26. Shi Y.; Huang J.-K.; Jin L.; Hsu Y.-T.; Yu S. F.; Li L.-J.; Yang H. Y. Selective Decoration of Au Nanoparticles on Monolayer MoS2 Single Crystals. Sci. Rep 2013, 3 (1), 1839. 10.1038/srep01839. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Liu X.; Choi M. S.; Hwang E.; Yoo W. J.; Sun J. Fermi Level Pinning Dependent 2D Semiconductor Devices: Challenges and Prospects. Adv. Mater. 2022, 34 (15), 2108425. 10.1002/adma.202108425. [DOI] [PubMed] [Google Scholar]
  28. Chen R.-S.; Ding G.; Zhou Y.; Han S.-T. Fermi-Level Depinning of 2D Transition Metal Dichalcogenide Transistors. J. Mater. Chem. C 2021, 9 (35), 11407–11427. 10.1039/D1TC01463C. [DOI] [Google Scholar]
  29. Tang H.; Shi B.; Pan Y.; Li J.; Zhang X.; Yan J.; Liu S.; Yang J.; Xu L.; Yang J.; Wu M.; Lu J. Schottky Contact in Monolayer WS 2 Field-Effect Transistors. Advcd Theory and Sims 2019, 2 (5), 1900001. 10.1002/adts.201900001. [DOI] [Google Scholar]
  30. Kim C.; Moon I.; Lee D.; Choi M. S.; Ahmed F.; Nam S.; Cho Y.; Shin H.-J.; Park S.; Yoo W. J. Fermi Level Pinning at Electrical Metal Contacts of Monolayer Molybdenum Dichalcogenides. ACS Nano 2017, 11 (2), 1588–1596. 10.1021/acsnano.6b07159. [DOI] [PubMed] [Google Scholar]
  31. Wang Q.; Shao Y.; Gong P.; Shi X. Metal–2D Multilayered Semiconductor Junctions: Layer-Number Dependent Fermi-Level Pinning. J. Mater. Chem. C 2020, 8 (9), 3113–3119. 10.1039/C9TC06331E. [DOI] [Google Scholar]
  32. Ber E.; Grady R. W.; Pop E.; Yalon E. Uncovering the Different Components of Contact Resistance to Atomically Thin Semiconductors. Adv. Elect Materials 2023, 9 (6), 2201342. 10.1002/aelm.202201342. [DOI] [Google Scholar]
  33. Ovchinnikov D.; Allain A.; Huang Y.-S.; Dumcenco D.; Kis A. Electrical Transport Properties of Single-Layer WS 2. ACS Nano 2014, 8 (8), 8174–8181. 10.1021/nn502362b. [DOI] [PubMed] [Google Scholar]
  34. Khan M. F.; Ahmed F.; Rehman S.; Akhtar I.; Rehman M. A.; Shinde P. A.; Khan K.; Kim D.; Eom J.; Lipsanen H.; Sun Z. High Performance Complementary WS 2 Devices with Hybrid Gr/Ni Contacts. Nanoscale 2020, 12 (41), 21280–21290. 10.1039/D0NR05737A. [DOI] [PubMed] [Google Scholar]
  35. English C. D.; Shine G.; Dorgan V. E.; Saraswat K. C.; Pop E. Improved Contacts to MoS 2 Transistors by Ultra-High Vacuum Metal Deposition. Nano Lett. 2016, 16 (6), 3824–3830. 10.1021/acs.nanolett.6b01309. [DOI] [PubMed] [Google Scholar]
  36. Sebastian A.; Pendurthi R.; Choudhury T. H.; Redwing J. M.; Das S. Benchmarking Monolayer MoS2 and WS2 Field-Effect Transistors. Nat. Commun. 2021, 12 (1), 693. 10.1038/s41467-020-20732-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Cui Y.; Xin R.; Yu Z.; Pan Y.; Ong Z.; Wei X.; Wang J.; Nan H.; Ni Z.; Wu Y.; Chen T.; Shi Y.; Wang B.; Zhang G.; Zhang Y.; Wang X. High-Performance Monolayer WS 2 Field-Effect Transistors on High-κ Dielectrics. Adv. Mater. 2015, 27 (35), 5230–5234. 10.1002/adma.201502222. [DOI] [PubMed] [Google Scholar]
  38. Panasci S. E.; Schilirò E.; Greco G.; Cannas M.; Gelardi F. M.; Agnello S.; Roccaforte F.; Giannazzo F. Strain, Doping, and Electronic Transport of Large Area Monolayer MoS 2 Exfoliated on Gold and Transferred to an Insulating Substrate. ACS Appl. Mater. Interfaces 2021, 13 (26), 31248–31259. 10.1021/acsami.1c05185. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials


Articles from ACS Applied Materials & Interfaces are provided here courtesy of American Chemical Society

RESOURCES