Introduction

Transition Metal Dichalcogenides (TMDCs) are non-carbon layered materials and in single layer form they are direct bandgap semiconductors, overcoming graphene’s lack of energy bandgap, vital for optoelectronic applications1,2. Atomically thin TMDCs such as MoS2 offer unique optical, electronic and physical properties such as quantum size effects3,4, mobilities exceeding theoretical values of 410 cm2 V−1 s−15,6, on/off ratios of up to 10107,8,9, subthreshold slopes down to 74 mV/dec4,10, beyond the thermal transport limit switching performance9,11 and mechanical flexibility. Altogether this indicates such material’s high potential for replacing and exceeding current material technologies. Weak van der Waals forces between each atomic layer led to “exfoliation and transfer” being one of the first methods used to obtain monolayer TMDCs7,10,12. Whilst the “exfoliation and transfer” method offers high quality single crystal layers with excellent electronic and optoelectronic properties, the micron scale material size limits the applicability of the method to very few non-manufacturable applications. Large area growth methods for TMDCs that are scalable and can be used in traditional top-down fabrication processes or transferred on application appropriate substrates are therefore under intense research. Scalable methods have recently included RF sputtering13,14,15, CVD16,17,18,19,20,21, ALD22,23,24,25,26,27,28 and thermal decomposition29,30 techniques, however the number of layers, crystallinity, stoichiometry and performance uniformity over large areas are still challenging to accurately control.

In this work, we present a robust large-area MoS2 direct growth method, compatible with a variety of substrates, with decoupled layer count, stoichiometry and crystallisation offering excellent control over the film properties. The process employs atomic layer deposition (ALD) to form a template MoO3 layer by which the thickness of the final MoS2 layer is defined. In addition, the ALD process creates a protective interface with the substrate which enables non-chemical exfoliation for subsequent steps such as transfer. The film is subsequently converted to MoS2 by vapour sulfurization at an intermediate temperature, where the stoichiometry of the film is defined, followed by a high temperature anneal that controls the crystallinity independently from layer number. To highlight the offered process control over the film characteristics we present results from ellipsometry, x-ray photoelectron spectroscopy (XPS), transmission electron microscopy (TEM) and Raman spectroscopy for films grown directly on silica substrates. We demonstrate the retention of the electronic quality and performance uniformity of the ALD 2D MoS2 layer by characterising an array of field effect transistors (FETs) fabricated by transferring the 2D material onto a flexible substrate.

Finally, we employ a ferroelectric FET (FeFET) structure to demonstrate the performance of the films as flexible in-computing memories and synaptic transistors highlighting the potential for advanced novel applications. Ferroelectric field effect transistors (FeFETs) are an attractive choice for memory transistors, logic-in-memory and computing-in-memory (CIM) architectures31. In recent years, MoS2 based FeFETs have shown promising performance as memory transistors32,33,34, negative capacitance transistors35 and synaptic weight elements36. Here we demonstrate that the ALD MoS2 of this work can be used to create state-of-the-art FeFET devices.

Results and discussion

Film characterisation

Ellipsometry is used to ensure the growth of a uniform MoO3 layer across a 6-inch wafer, which in addition provides complex refractive index data. Thickness and optical data, mapped across the wafer, allow the intra-wafer (variation across a single wafer) and inter-wafer (wafer-to-wafer) variability to be monitored, ensuring the uniformity and repeatability of our process. Figure 1a, b shows the ellipsometry results from the MoO3 ALD layer and Fig. 1c when converted to MoS2, both were fitted using three Tauc-Lorentz oscillators. As it can be seen, the MoO3 layer is homogeneous over a 6-inch wafer as also seen in the pictures of Supplementary Fig. 1. The thickness non-uniformity over the entire wafer is 6.3% and over the 80 × 80 mm2 central area 4%, calculated as 100*Tmax−Tmin/Tave, where Tmax, Tmin and Tave are the maximum, minimum and average thicknesses of the MoO3 across the wafer. The uniformity of the starting MoO3 film is limited by the configuration of the ALD system used as the inlets and outlets are located at opposite sides in the wafer plane as shown in Fig. 9. In an optimised system the self-limiting nature of the ALD process defines the uniformity of the starting film. The resulting MoS2 film uniformity and continuity is further affected by the parameters of the anneal (temperature of H2S introduction, pressure, and partial pressure of H2S) but also by the configuration of the annealing system (temperature uniformity, gas distribution uniformity). In this work, we used a tube furnace as shown in Fig. 9 to anneal chips of 2.5 × 2.5 cm2 the uniformity of which was assessed by Raman mapping as presented in Supplementary Fig. 3 showing the peak separation and the FWHM of E2g and A1g peaks for a 1 × 1.2 cm2 area.

Fig. 1: MoO3 Ellipsometry results on SiO2/Si.
figure 1

a 6-inch wafer thickness mapping. b Real and imaginary part of the refractive index of MoO3. c Real and imaginary part of the refractive index of MoS2.

The refractive index and extinction coefficients of MoO3 are n = 1.96 and k = 0.08 and that of the resulting MoS2 film n = 5.17 and k = 1.27 at 633 nm. The optical properties of both materials are typical37,38 and are used as a confirmation for wafer-to-wafer reproducibility. After 15 cycles of MoO3 ALD, the thickness of the MoO3 layer as fitted by ellipsometry is 1.31 ± 0.13 nm resulting in 0.87 Å/cycle growth rate.

Monitoring the stoichiometry of MoO3 is important to examine the oxidation level ratios, account for ALD process drifts and adjust for the sulfurisation process, all key when upscaling a process to a fabrication line39. Figure 2a, b shows high resolution XPS results from the molybdenum and oxygen core levels of the MoO3 film grown by ALD. Mo d5/2 core levels corresponding to MoO3 were fitted at 233.31 eV and MoO2 at 232.33 eV at an atomic ratio of 2.9. At the O1s core level, oxygen bound as MoO3 was detected at 531.19 eV with a clear distinction from oxygen bound to silicon at 532.84 eV. Grazing incidence diffraction (GI-XRD) measurements (Supplementary Fig. 5) verified the presence of MoO3 suboxides but also the presence of both α and β phases characteristic of a film grown at 250 oC40,41,42

Fig. 2: XPS analysis of the MoO3 film.
figure 2

a Mo 3d, b O 1s and the MoS2 c Mo 3d and d S 2p core levels.

XPS analysis of the sulfurized MoO3 reveals the stoichiometry of the resulting MoS2 but also any residual oxides in the film. Figure 2c, d shows the core levels of molybdenum and sulfur for the converted sample. Mo 3d5/2 core levels for MoS2 were detected at 229.64 eV and MoO3 at 232.12 eV with no MoO2 detected after deconvolution. MoO3 accounts for 8% of the Mo atomic ratio for the as deposited films. High angle annular darkfield—scanning transmission electron microscopy (HAADF-STEM) and selective area electron diffraction (SAED) indicated a very small fraction of MoO3 remaining on the transferred suspended films as shown in Supplementary Figs. 6 and 7. MoO3 was long considered to act as an impurity due to its insulating nature and the overall lattice distortions effect to the MoS2 crystal by the replacement of sulfur atoms by oxygen43. However, it has recently been shown that for certain scenarios MoO3 can be viewed as a mobility enhancer owing to its two-dimensional phase and its high acoustic phonon-limited carrier mobility of (>3000 cm2 V−1 s−1)44,45. Like what was observed in46 which had a similar level of oxide content, we did not observe a degradation of FET performance linked to MoO3 content. Sulfur 2p3/2 was fitted at 162.44 eV. By preserving the full width at half maximum between the spin orbit splits and abiding by the 3/2 area ratio for 3d and 1/2 for 2d orbitals the atomic ratio of sulfur to molybdenum was found to be 2.1, revealing the stoichiometric nature of the deposited film47.

To assess the surface quality of the layer we performed AFM measurements and the results are presented in Fig. 3. To minimise background noise MoS2 samples were grown on sapphire for its superior smoothness to SiO2 on silicon. The MoS2 on sapphire was prepared alongside the samples on SiO2 in the exact same process. Samples grown on sapphire had similar Raman signatures and did not present any other structural difference apart from lower roughness. As it can be seen, the MoS2 layer is highly uniform with an average roughness Ra of 146 pm. The AFM and STEM measurements did not reveal any ripples, normally present on monolayer films, as expected for a multilayer film48. The thickness of the 2D layer was found to be 1.48 nm on average over the entire 1 μm2 area of measurement, indicating a bilayer film. The real non-contact mode of the AFM provided us with information on the lateral size of what appears to be distinct material sections which are between 20 and 35 nm. The material sections are of small size when compared to CVD grown MoS2 crystal grains from the literature, typically reaching up to 400 μm49,50. As discussed in the electrical results section the electron mobility is at the high end of CVD grown films while the consistent small-scale variation of the crystal reduces device variations. It is also apparent from the measurement that the individual sections are in close contact without discontinuities composing this way a continuous highly crystalline layer, a merit often absent in larger crystal sizes51,52. These characteristics constitute the ALD grown MoS2 directly applicable to large scale processes without the requirement for precise device placement and with the advantage of performance uniformity.

Fig. 3: AFM measurements of MoS2 bi-layer on sapphire.
figure 3

a 3D representation. b MoS2 thickness profile.

To investigate the layered structure and crystallinity of the MoS2 layer, we created a lamella using a Focused Ion beam (FIB) and inspected the cross-section using TEM. The TEM cross-section shown in Fig. 4a as well as the STEM Z-contrast top view of Fig. 4b indicate a high degree of crystallinity composed by sections of 2 and 3 layers. It is notable here that the lateral size of the layer sections matches with the sections size of the AFM measurement, which further verifies that the crystallinity is layer uniformity limited at 20–35 nm. To obtain the top view of the layer, the film was transferred onto a fully perforated silicon nitride TEM grid. The sample was inspected by STEM and a HAADF top-view image is shown in Fig. 4b. The image verifies the grain size seen from the cross-sectional view and the AFM. It also confirms sections of different thickness. Monolayer areas were observed, near sporadically present holes in the MoS2 film. A rhombohedral stacking is indicated by the presence of an intensity maximum within the projected hexagons. Rhombohedral 3R stacking of the layers breaks the inversion symmetry which results in non-linear optical properties, tuneable bandgap, piezoelectricity, ferroelectricity and two-dimensional confinement of valley electrons53,54,55,56,57,58,59,60. There were also regions of hexagonally stacked MoS2 present. The occurrence of moiré fringes in the HAADF-STEM images may also indicate either a relative rotation between MoS2 layers or a slight variation in crystal orientation due to bending of the layer. However, as no additional peaks appeared in the power spectrum a rotation is less likely. Analysing the TEM and AFM data showed that the film is constituted by areas of 2 and 3 layers. The third layer appears on top of the other two at certain areas and at the bottom at other areas. Although the continuity of the third layer is broken, we did not observe the two layers being completely interrupted at any point. For this reason, a plausible assumption is that at least one layer of MoS2 forms a continuous crystal significantly larger than the distinct areas we see in the AFM. This assumption is consistent with the electrical performance of the fabricated FET and FeFET devices.

Fig. 4: TEM images of the ALD MoS2 film.
figure 4

a Cross-sectional view by conventional TEM and b top view HAADF-STEM image (noise filtered), the insets show magnifications of regions with 3R stacking (yellow) and 2H stacking (green) along the [0001] zone axis.

Throughout our fabrication process Raman scattering spectroscopy provided the most direct non-destructive method to monitor the quality of the films. This method shows the changes that the film undergoes during transfer and patterning and it serves as a quality control index during the fabrication process. Raman spectroscopy was used to characterise the directly grown MoS2 on silica substrates. The standard MoS2 recipe, with 15 ALD cycles, demonstrates two distinct Raman peaks associated with MoS2, E2g and A1g indicating high quality and crystalline behaviour, with an average across three measurements of one sample of 23.9 ± 0.1 cm−1 peak separation, depicting a three-layer film61. To demonstrate the ability to tune the layer number, the ALD cycle number was reduced from 15 down to 8 and Raman scattering measurements were performed on the resulting MoS2 films (Supplementary Fig. 4). The results of Supplementary Fig. 4 show a blue-shift of the E2g peak and a red-shift of the A1g consistent with a decreasing layer number62. A clear trend between the ALD cycle number and peak separation can be seen in Fig. 5a, highlighting the ability to accurately control the number of layers, down to 2 layers, or less, with 8 ALD cycles63 with a peak difference of 22.5 ± 0.1 cm−1. The films had a clear contrast between them, to the naked eye, as shown for the 8, 10 and 15 cycles in Supplementary Fig. 2. The full width half maximum (FWHM) of the E2g peaks is used to investigate the effect of ALD cycles on the film quality. As shown in Fig. 5b, the E2g FWHM of both peaks is unaffected with cycle number, with no clear trend seen with ALD cycle number, confirming the ability to tune the number of MoS2 layers can be performed without reducing the quality.

Fig. 5: Raman scattering of MoS2 by varying ALD cycles.
figure 5

a Peak difference between E2g and A1g for MoS2/SiO2/Si with ALD cycles from 15 down to 8. Each data point is an average of 3 different areas per sample, with the error bars originating from the standard deviation of the 3 measurements. A linear fit with peak difference and cycle number is seen. b Full Width Half Maximum of E2g peaks for MoS2/SiO2/Si samples with ALD cycles from 15 down to 8. Each data point is an average of 3 different areas per sample, with the error bars originating from the standard deviation of the 3 measurements. No trend with E2g FWHM and ALD cycle number is seen.

As presented in this section the primary advantage over other processes that rely on the sulfurisation of a starting film (Mo, MoO3, MoO2 and molybdate)46,51,52,64,65,66, is that ALD allows for excellent control over the material thickness down to a monolayer while maintaining excellent uniformity over large areas. In addition, ALD offers excellent control of the stoichiometry and results in atomically flat and smooth surfaces. The process presented here allows for the control of the thickness of the film separately to its stoichiometry and crystallinity. The layer count is defined uniformly on the wafer by the ALD process while the stoichiometry is defined by the low temperature H2S anneal and the crystallinity by the high temperature anneal. The post anneal in H2S is necessary for most ALD MoS2 processes to passivate defects and improve crystallinity. This decoupled control gives significantly more freedom in the control of the film properties when compared to other methods. The resulting films are continuous over the entire converted area and the demonstrated bilayer crystal structure facilitates the fabrication of FETs presenting amongst the highest mobilities for wafer scale continuous films with a tight performance uniformity as shown in the following electrical measurements.

Electrical measurements – flexible FET devices

To characterise the electrical performance of the MoS2 films, we implemented two different device architectures, bottom-gated FETs (Fig. 6a) and top-gated FeFETs (Fig. 6b). All devices were fabricated on polyimide substrates supported by a silicon wafer, to test their performance as logic transistors and as in-computing memory components for flexible applications. The FET and FeFET devices are using the same design geometry with the addition of a top ferroelectric gate for the FeFETs. The process from growth wafer to polyimide substrate was performed using a chemical free removal and transfer process. The ease of lifting the ALD films is exceptional when compared to other methods we have used. The lack of need for chemicals is attributed to the interface created between SiO2 and MoO3 during ALD deposition. This significantly improves the quality of the transfer with no other adverse effects. This clean transfer of MoS2 can be seen in a microscope image (Fig. 6c) of a device structure used to measure different device lengths, whilst also doubling up as TLM structures. The area marked as MoS2 in Fig. 6c shows the patterned MoS2 channel and highlights the typical clean transfer of the layer that can be achieved with our process, with an absence of visible defects in large areas providing devices with high yield and scalable production. Figure 6d illustrates the scalable nature of this process, with multiple devices shown.

Fig. 6: Flexible device schematics and pictures.
figure 6

a FET schematic, b FeFET schematic, c transmission line measurement FET, d multiple bottom-gate FETs over 5 × 5 mm2 area of polyimide substrate.

Figure 7a shows a transfer curve (Id-Vg) of a typical back-gated FET device when measured with a drain-to-source voltage (Vds) of 100 mV and a Subthreshold Slope of 200 mV/dec. Figure 7b shows output characteristics (Id-Vd) of a typical FET when back gate voltage (Vg) varies between +4 V to −0.5 V with a step of 0.5 V. From the linear part of the Id-Vd curves, the on-state resistance (Ron) of the devices was calculated for various applied gate biases and for different channel lengths. The Ron value, obtained from the output characteristics, revealed that the transferred MoS2 layer has a sheet resistance (Rs) of ~10 kOhm/µm and a contact resistance (Rc) of 8.5 kOhm for a 20 × 20 μm2 contact (Fig. 7c). The variation of sheet resistance and contact resistance is <10% over an area of 5 × 5 mm2. The contact resistance between the 2D semiconductor and metal electrodes is a performance defining parameter67,68,69, whereby contaminants on the surface create an additional Schottky barrier, increasing Rc. To minimise these effects, we use long repeated washing cycles for the MoS2 layer, followed by vacuum annealing at 300 oC and pumping cycles to reach a base pressure of 10–8 mbar before metal deposition. To improve the contact resistance further, by reducing the reactivity with Ti, we used a Au/Ti/Au electrode stack8. Our contact resistance is defined by the injection barrier height for electrons from the work function of the selected contact metals, but also by the Van der Waals gaps between the MoS2 interlayers and the contact metal layer. We expect that an edge-contact device and a contact metal work function modification can lead to significant improvement in device performance.

Fig. 7: Electrical measurements results of the flexible MoS2 FETs.
figure 7

a Transfer and b output characteristics of a typical FET with channel length of 5 µm and width of 20 µm. c Contact resistance calculated from the on-state resistance of the FETs with different channel lengths. Histograms showing distribution of device d on/off ratio, e field-effect mobility, and f subthreshold slopes (SS) of several measured devices over a 5 × 5 mm2 area.

To characterise the performance distribution of the fabricated FETs we performed Id-Vg measurements on a range of single FET transistors over a 5 × 5 mm2 area. The measurements revealed an average on/off ratio of 107 (Fig. 7d) while operating between −1 V to +5 V. The field-effect mobility of the devices, was calculated using the equation, \({{{\mathrm{\mu }}}} = \frac{1}{{C_g}}\frac{L}{W}(\frac{{dI_{ds}}}{{dV_{bg}}})\frac{1}{{V_{ds}}}\) where L is the channel length, W is the channel width, (dlds /dVbg) is the slope of transfer characteristic of the devices with drain-to-source voltage Vds = 100 mV, and Cg is the gate capacitance calculated for 30 nm Al2O3 as the gate dielectric. The value of (dlds/dVbg) is estimated by fitting the linear regime of the transfer characteristic curves of MoS2 FETs. The results show that the mobility value of non-passivated devices is up to 32 cm2/Vs while that of passivated devices with P(VDF-TrFE) is up to 55 cm2/Vs (Fig. 7e). The Gaussian distribution of mobilities measured from 15 devices shows a peak around 10 cm2/Vs for non-passivated devices, while for the passivated devices, the mobility distribution peaks at 40 cm2/Vs, indicating a significant improvement in mobility values and variability ought to device passivation. The increased mobility is consistent with the observation of the quenching of the normal to the film homopolar phonon mode which is a characteristic of dual gate and passivated devices70,71 at room temperature, but also more importantly due to the screening of the charge impurities and traps achieved by the passivating fluoropolymer layer72,73. An additional advantage of the encapsulation is that it reduces the exposure of MoS2 to the moisture of the environment and the consequent chemisorption of oxygen; as a result, the TFTs demonstrate a reduced hysteresis74,75. The high field effect mobility of the fabricated FETs presented here is on par with the best of CVD grown MoS2 devices76,77,78,79. The subthreshold slope (SS) values measured from multiple devices also present a Gaussian distribution with a peak at 180 mV/dec (Fig. 7f), with best devices showing SS values down to 80 mV/dec, equivalent to the best reported results for non-negative capacitance FETs of 74 mV/dec10,69. The use of a bilayer MoS2 film is a key element in achieving this performance due to inherent advantages over monolayer films such as the smaller bandgap80, higher mobility81, higher on state current82, lower contact Schottky barrier height83. Supplementary Table 1 and Supplementary Table 2 offer a comparison of growth parameters and electrical performance of similar and pure ALD methods.

Our second electronic device is a FeFET by using a ferroelectric P(VDF-TrFE) layer on top of our MoS2 layer. In these MoS2/P(VDF-TrFE) FeFET devices, when the polarization direction points toward the MoS2 channel, electrons accumulate at the interface between the MoS2 and the P(VDF-TrFE) layer. This downward polarization, thus, results in increased conductivity in the MoS2 channel, representing the device on state. Conversely, when the polarization direction points away from the semiconductor channel, a depletion of electrons occurs at the interface resulting in the decrease in drain-source current, leading to a device off state. A proper screening of the ferroelectric polarization charges can lead to non-volatile retention of the device conductance states over extended periods of time. This feature constitutes FeFETs as one-transistor non-volatile memories that have the potential to increase density in array-level integration.

The FeFETs fabricated in this work show an on/off current ratio of 107, memory window (MW) of 3 V at an operation voltage range of ±5 V (Fig. 8a), 4–5 pA level of gate leakage current in the on state and multiple programmable states in response to single gate pulse writing. Our MoS2 FeFETs outperform earlier flexible memory transistors with organic semiconductors as the channel material. The much lower voltage of operation of our devices is mainly attributed to the thinner 30 nm P(VDF-TrFE) layer compared to typical 100 nm ferroelectric layers used in previous reports36. The key performance matrices presented in ref. 84 place our devices among the most efficient memory transistors reported on flexible substrates. Furthermore, ultrafast switching can be expected from these devices, based upon successful nanosecond pulsed switching demonstrated in our previous work using P(VDF-TrFE)85,86. Devices evaluated at scaled nodes will reveal further performance merits for their integration in memory and in-memory computing arrays.

Fig. 8: Electrical measurements results of the flexible MoS2 FeFETs.
figure 8

a Transfer characteristics of a typical FeFET with channel length of 2 µm and width of 5 µm, demonstrating a memory window of 3 V. b Retention characteristics of the programmed states measured after programming with +5 V or −3 V once followed by repeated reading cycles at 0.5 V. c Repeatability of programming of the FeFETs to multiple conductance states in response to gradually increasing magnitude of gate voltage pulses of millisecond duration. d Distribution of the memory window (MW), e Ion and f Ioff of 14 measured devices from the 5 × 5 mm2 area.

The FeFETs demonstrate stable, non-volatile data retention (Fig. 8b), when device conductance states are programmed once and read many times with 0.5 V read bias. Further reproducibility of different programmed states was tested with gradually increasing pulse amplitudes of the gate voltage. The drain current (Id) as a function of pulsed gate voltage (Vg) is shown in the top and bottom panels of Fig. 8c, respectively. The results demonstrate that multiple intermediate conductance states, arising from the mixed polarization phase of the FE, can be achieved in our MoS2 FeFETs with high repeatability. Devices were measured over thousands of cycles to multiple conductance states (although few hundred cycles are shown here for clarity) without any significant sign of degradation and breakdown, showing their high endurance to the bias stress. These features make these MoS2 FeFETs suitable, not only for non-volatile data retention, but also as electronic synaptic transistors where post-synaptic currents can be modified in a controllable analog manner, enabling their application in pseudo-crossbar based analog accelerators.

For confirming the homogeneity of device performance over large area, we plotted histograms of FeFET off current (Fig. 8d), on current (Fig. 8e) and MW (Fig. 8f) over 5 × 5 mm2 area. The majority of the devices have similar performance with very few deviations. The off current distribution shows a Gaussian distribution with a peak at around 200 pA while that for the on current shows a peak at around 15 µA. The memory window distribution peaks at around 3.5 V when the FeFET gate voltage is swept in the −5 V to +5 V range. Like the FETs the FeFETs show excellent performance uniformity, demonstrating the potential for large functional device arrays.

The FETs and FeFETs shown here demonstrate high performance properties comparable or exceeding current technology published in literature for comparable devices, highlighting our MoS2 layer’s superior electronic device compatibility but also the performance advantage of an optimised ferroelectric layer. The uniformity of these devices also demonstrates the scalability of this high-performing material for commercial electronic applications.

In this work we have demonstrated the large area uniform growth of 2D TMDCs by a novel ALD process. Highly crystalline MoS2 films were grown on Si/SiO2 substrates using our scalable 2-step process. The first step, growing MoO3 via ALD, results in a film uniformity of 6% over a 6-inch wafer, with a 0.87 Å/cycle growth rate. This stable self-terminated MoO3 growth rate provides a decoupled, accurate and repeatable monitoring and control method for tuning the resulting MoS2 films’ layer number down to a monolayer as verified by Raman, AFM and TEM. The second step, an anneal in H2S is used to control the stoichiometry of the layer at the conversion temperature while the crystallinity is defined at the higher crystallisation temperature of the anneal process. The optimised stoichiometry, imperative for upscaling, has been demonstrated via XPS, resulting in a molybdenum to sulfur atomic ratio of 2.1. AFM and TEM characterisation, highlights the MoS2 films’ low roughness of 146 pm and its continuous nature, with 35 nm sized distinct crystal areas. Thus, this process is ideal for upscaling as it produces smooth continuous layers for subsequent fabrication steps and low device-to-device variation from smaller crystal sizes, without a need for accurate device placement as seen for larger crystals. MoS2 was successfully transferred onto flexible substrates using a chemical free transfer process resulting in two types of electronic devices, FETs and FeFETs. MoS2 FETs over a 5 × 5 mm2 produced an on/off ratio of 107, with best devices exhibiting a SS of down to 80 mV/dec. The MoS2 FeFETs showed a current ratio of 107, alongside a 3 V memory window with an ideal low operation voltage of 5 V, outperforming other flexible FeFETs, by utilising a thin P(VDF-TrFE) layer. Mixed polarisation has been demonstrated, enabling switching between different states to be performed over thousand times with no degradation, highlighting the MoS2 films’ ability to endure high levels of bias stress.

Methods

The substrates used are 6-inch p-type silicon wafers. As a first step 285 nm of thermal SiO2 is grown at 1000 oC in a tube furnace. The wafers are then inserted in a UV/O3 reactor for 10 min to improve the chemical termination of the surface oxide. The next step is to grow the MoO3 using a thermal ALD process in the Cambridge Nanotech Savannah S200 system. The process is based on ref. 87 and uses the precursor bis(tert-butylimido)bis(dimethylamido) molybdenum as a molybdenum source and ozone for the oxidation. Fifteen cycles at 250 oC result in a film of 1.31 ± 0.13 nm of MoO3 as shown in Fig. 1, showing high uniformity over the entire 6-inch wafer. This initial step enables us to have high control over the subsequent number of MoS2 layers, further on in the process.

After the growth, the MoO3 layer is converted into MoS2, via annealing in a tube furnace. The tube dimensions require the wafer to be diced into 2.5 × 2.5 cm2 chips. The chips are placed onto a crucible and then inserted in the tube furnace for film sulfurisation as depicted in Fig. 9. The sulfurisation of the film is performed in an H2S/Ar environment and it involves two steps, one at 550 oC which is used to convert the MoO3 into MoS2 and one at 970 oC which is used to crystallise the film. After conversion and crystallisation, the furnace is left to naturally cool down to room temperature. The described process results in the direct growth of a uniform film of MoS2 directly on SiO2. The film can be used directly to process high-quality rigid electronic devices.

Fig. 9: Schematic of the experimental setup for the growth of the 2D film.
figure 9

MoO3 is grown in an ALD chamber while sulfurisation and crystallisation take place in a tube furnace.

The ellipsometer measurements were performed with a Whoollam M-2000DI spectroscopic ellipsometer at angles of 65, 70 and 75o for five seconds at each angle. XPS was performed using a Thermofisher Theta Probe with an x-ray spot size of 400 μm at 15 keV and no flood gun. Molybdenum and sulfur peaks were measured at 50 eV pass energy with a dwell time of 50 ms and 20 averaged scans with a step size of 20 meV. The pressure during the measurement was at 2.4 × 10–9 mBar. All XPS results were referenced to adventitious carbon at 284.8 eV. AFM was performed with a Park XE7 AFM in non-contact mode on a 1 μm2 area with a 2048 × 2048 pixels resolution at a 200 mHz scan rate. The AFM employs a very sharp tip with a tip radius of 2 nm and real non-contact tapping mode preserving the sharpness of the tip throughout the measurement. It can therefore offer very high lateral resolution down to just a few nanometers. For the TEM inspection we created a lamella using a Zeiss FIB. The cross-section was analysed using a FEI Tecnai F20 TEM at 200 kV acceleration voltage with a Gatan Multiscan CCD Camera. For the top view HAADF-STEM inspection a film was transferred to a holey silicon nitride TEM grid and analysed using a probe corrected Thermo Fisher Titan Themis TEM at 120 kV. Raman spectroscopy was done with a Renishaw Invia system using a 532 nm laser and a ×50 objective.

As the high reaction temperature of bottom up MoS2 film requires growth on a SiO2/Si substrate, a subsequent transfer to polyimide substrate is made for the formation of flexible devices. The polyimide substrate is formed onto a silicon wafer by spin coating and annealing to offer support for the subsequent processing steps. To form electrical contacts for both gate and source/drain, the substrates are coated with the positive tone photoresist AZ 5214E and the patterns are written with a mask-less lithography process, using a Microtech LW 405 laser writer. To form the gate dielectric, 30 nm of alumina is grown on the polyimide using thermal ALD in a Beneq TFS-500, which is subsequently patterned by laser lithography and etched in a 20:1 buffered HF solution.

After the sulfurisation of the films, the chips are spin coated at 3500 rpm for 60 seconds with a 2% polystyrene/Toluene solution (PS molecular weight: 280,000 g/mol) and baked at 90 oC for 10 min88. They are then dipped into a deionised water bath and within a minute the film is lifted off the substrate and floats on the surface. The MoS2 layer is then transferred onto the patterned polyimide and baked at 80 oC for 10 min to remove the bulk of the water and subsequently at 120 oC. The substrates are then left in a toluene bath for 72 h. Upon removal they are rinse in DI water and dried with a nitrogen gun before they are left in ambient cleanroom conditions to dry for a day. This results in a clean MoS2 surface ready to be patterned.

The FET transistors and FeFET devices are fabricated on flexible polyimide substrates that are mechanically supported on Si backplates. For the FET transistors, a patterned back-gate metal layer of 2 nm Ti and 50 nm Au is deposited using e-beam evaporation followed by lift-off. Next, a 30 nm thick Al2O3 layer is grown by ALD at 300 oC. Vias are opened using wet oxide etching followed by oxygen plasma cleaning. Next, the MoS2 layer is transferred, the carrier film is washed, and the channel area is defined using a mask-less laser writer lithography process by a Microtech Laser Writer 405. MoS2 areas outside of the channel are etched using an SF6/O2 plasma. Drain-source contacts are patterned on top of MoS2 using a lithography step and Au (3 nm), Ti (2 nm), and Au (50 nm) electrodes are deposited by e-beam evaporation at room temperature followed by lift-off. The devices are then annealed at 200 °C in a vacuum furnace for 2 h to remove resist residues and to improve the contact resistance. To protect the MoS2 layer from atmospheric adsorbates, we fabricated a batch of devices with a passivation layer of 100 nm thick spin-coated fluoropolymer P(VDF-TrFE) followed by a drying step at 90 °C in a furnace for 2 h. Although this is the same material we use as a ferroelectric layer, P(VDF-TrFE) does not attain its polar β-phase when annealed at 90 °C and therefore remains as a purely amorphous, dielectric layer.

For the FeFET devices, two additional processing steps are implemented on top of the FET devices. A ferroelectric (FE) layer consisting of P(VDF-TrFE) (70:30 mol%) copolymer and a top gate metal are deposited on top of the channel. The FE layer is made from ferroelectric polymer powder (from Piezotech) dissolved in Methyl Ethyl Ketone (MEK) at 0.5 wt% dilution and the solution is spin coated on top of MoS2. First, the spin-coated films are dried at 100 °C for 10 min and then they are annealed in vacuum at 135 °C for 2 h to improve their crystallinity and obtain the ferroelectric β-phase. The FeFETs consist of a 30 nm thick P(VDF-TrFE) gate insulator layer directly between the MoS2 channel and the Au gate metal electrode (30 nm thin Au patterned by lithography and etched using wet etching). The thickness of the ferroelectric films was characterised by profilometry using a Si/SiOx/P(VDF-TrFE)/Au capacitor structure.

The size of the MoS2 channel for both the FETs and the FeFETs varied between ≈ 2 (L) × 5 (W)−20 (L) × 40 (W) µm2.

All measurements were performed using a Keithley 4200A-SCS parameter analyzer in a probe station. Continuous transfer characteristics measurements were performed by applying sweeping voltage at the gate terminal under a continuous bias voltage between drain and source contacts. When operating the double gated transistors as FeFETs, the bottom gate contact can either be grounded or can be set at a fixed bias that can modulate the operating point of the transistors. Retention measurements were performed by programming the devices once with +5 V or −3 V pulse followed by repeated reading at 0.5 V. Repeatability of multiple programmed states of the FeFETs were tested using gradually increasing magnitude of gate voltage pulses of millisecond duration over several thousands of cycles. These measurements also provided information on the endurance of the devices in response to repeated bias stress. All measurements were carried out under atmospheric conditions and room temperature without illumination. In the case of the FET transistors, the P(VDF-TrFE) layer on top of MoS2 acts as the encapsulation layer. Devices were stored in a N2 filled glove box between measurements.