Click here to load reader

INTRODUCTION - research.manchester.ac.uk€¦  · Web viewThe optimized Ta 2 O 5 /OTS IGZO TFTs operate at 1 V with saturation field-effect mobility larger than 2.3 cm2/Vs, threshold

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

8

> REPLACE THIS LINE WITH YOUR PAPER IDENTIFICATION NUMBER (DOUBLE-CLICK HERE TO EDIT) <

[footnoteRef:1] [1: The authors would like to thank the UGC-UKIERI for financial support (Grant numbers 2017/18-015 and 184-15/2018(IC)). N. Mohammadian is with the Department of Electrical and Electronic Engineering, The University of Manchester, Manchester, M13 9PL, UK(e-mail: [email protected]).B. C. Das is with the Department of Physics, Indian Institute of Science Education and Research, Thiruvananthapuram, Maruthamala PO, Vithura, Kerala 695551, India (e-mail: [email protected]).L. A. Majewski is with the is with the Department of Electrical and Electronic Engineering, The University of Manchester, Manchester, M13 9PL, UK (e-mail: [email protected]).]

Low-voltage IGZO TFTs using solution-deposited OTS-modified Ta2O5 dielectric

N. Mohammadian, Member, IEEE, B. C. Das, and L. A. Majewski, Senior Member, IEEE

Abstract— Low voltage, high performance thin film transistors (TFTs) that use amorphous metal oxide (MO) semiconductors as the active layer have been getting tremendous attention due to their essential role in future portable electronic devices and systems. However, reducing the operating voltage of these devices to or below 1 V is a very challenging task because it is very difficult to obtain low threshold voltage (VTH), small subthreshold swing (SS) MO TFTs. In this paper, indium gallium zinc oxide (IGZO) TFTs that use solution-deposited Ta2O5 operating at 1 V are demonstrated. To enhance the dielectric properties of the fabricated ultra-thin (d ~ 22 ± 2 nm) tantalum pentoxide films, n-octadecyltrichlorosilane (OTS) self-assembled monolayer (SAM) was used. The morphology and electrical properties of both pristine and OTS-treated Ta2O5 films have been studied. The optimized Ta2O5/OTS IGZO TFTs operate at 1 V with saturation field-effect mobility larger than 2.3 cm2/Vs, threshold voltage of around 400 mV, subthreshold swings below 90 mV/dec, and current on-off ratios well above 105. The performance of the presented TFTs is high enough for many commercial applications such as disposable sensors or throwaway, low-end electronics significantly reducing the cost of their production.

Index Terms— Low voltage thin film transistors (TFTs), Tantalum pentoxide, Anodization, Self-assembled monolayer (SAM), Indium gallium zinc oxide (IGZO).

INTRODUCTION

M

etal oxides (MOs) play a very important role in many areas of chemistry, physics, and materials science. Since the metal elements can form a large diversity of oxide compounds these elements can adopt many structural geometries with an electronic structure that can exhibit metallic, semiconductor, or insulator character. In practical applications, MOs are used in sustainable wastewater treatment and energy production [1], solar-driven conversion of CO2 to hydrocarbons [2], solar-driven water splitting [3], as well as the fabrication of supercapacitors [4], aqueous solar cells [5], gas and pressure sensors [6, 7], piezoelectric devices, fuel cells [8], and microelectronic circuits. Indeed, most of the modern electronic chips used in electronic devices contain an oxide component. Amongst MOs, high-k oxide dielectrics such as Al2O3, ZrO2, HfO2, TiO2, Y2O3, CeO2, and Ta2O5 have drawn a lot of attention in recent years because they can be used in the fabrication of high-performance thin film capacitors (TFCs) [9], low voltage thin film transistors (TFTs) [10] and TFT-based bio- and chemical sensors [11]. Typically, to deposit high quality metal oxide dielectric films sophisticated deposition techniques such as atomic layer deposition (ALD), pulsed laser deposition (PLD), chemical vapor deposition (CVD) and radio frequency (RF) magnetron sputtering are used. However, these techniques require expensive equipment and depend on controlled environment and temperature which significantly increases the material and device processing costs. The next generation of thin film electronic devices and systems will have limited power budgets, and as result, the development of high quality, solution-processed dielectric materials that can be cost-efficiently deposited on large areas as ultra-thin films is urgently required. Recently, anodic oxidation (so-called anodization) has attracted a lot of attention because it is a low-cost, solution-based deposition process that can be used to fabricate high quality, ultra-thin metal oxide films [12]. Anodization is a self-limiting and self-healing process that gives pinhole-free, homogenous oxide layers that can be grown in an ambient atmosphere at room temperature using low-cost set-ups and environmentally friendly chemicals. Considering high-k oxide dielectrics, tantalum pentoxide (Ta2O5) is a promising candidate in terms of optical, physical, chemical and electrical properties [13]. Ta2O5 has high transparency, high melting point (1785 °C), high permeability for hydrogen gas, and high dielectric constant, which can be exploited in a wide range of electronic applications. As a result, tantalum pentoxide has been abundantly used in metal-insulator-metal (MIM) capacitors, metal-insulator-semiconductor (MIS) and Schottky diode sensors, DRAM devices and TFTs [14–17]. The dielectric constant (k) of Ta2O5 depends on its thickness and deposition techniques and varies from ~ 35 in the bulk to ~25 in a thin film [13]. This is at least two and six times larger than k of Al2O3 (k = 9) and SiO2 (k = 3.9), respectively, which makes it a good candidate for gate insulator in TFTs [18-19]. To date, low threshold voltage (VTH = 0.25 V) IGZO TFTs operating at 3 V gated with 200 nm thick, sputtered Ta2O5 have been reported [20]. However, the demonstrated devices displayed subthreshold swings larger than 600 mV/dec that is aboutten times bigger than the theoretical limit of SS at 300 K(~ 60 mV/dec), and as result, could not be effectively operated with voltages below 3 V. This shows that the fabrication of metal oxide semiconductor TFTs that can be operated at or below 1 V is not trivial.

(a) (c)

(b) (d)

Fig. 1. 2D and 3D AFM topographical images of (a, b) untreated and (c, d) OTS-treated Ta2O5 films.

In a TFT, the gate voltage, VG, required to switch the transistor “on” scales with the insulator thickness, d, and inversely with the insulator dielectric constant, k, i.e. VG ∼ d/k. Consequently, high k at low d is desirable. However, both minimizing d and maximizing k leads to practical problems: a thin, fragile dielectric may lead to increased leakage currents through the insulation layer, and in consequence rise in power consumption and shorten the device life-time, whereas high-k insulators introduce undesirable effects at the semiconductor-insulator interface which generally leads to increased trapping which causes a rise of VTH and SS, as well as lowers the field-effect mobility of charge carriers in TFTs [21]. To overcome both issues in low voltage organic thin film transistors (OTFTs) passivation of the high-k oxide surfaces with self-assembled monolayer (SAM) compounds is routinely used [22]. For example, modifying Al2O3 with n-octadecyltrichlorosilane (OTS) SAM results in considerably improved dielectric performance of alumina films and significant increases of the field-effect mobility in OTFTs [23]. Although modification of ultra-thin (d ≤ 20 nm) metal oxide gate dielectrics with SAMs is currently a common process in organic TFTs, this approach has not yet been fully explored in metal oxide semiconductor transistors. Here, we show that treating anodically oxidized Ta2O5 films with OTS results in the same benefits to IGZO TFTs as it is in the case of OTFTs. As a result, high performance, low voltage IGZO TFTs gated with OTS-passivated Ta2O5 that can be operated at 1 V are demonstrated. The optimized transistors show highly reproducible characteristics with virtually no voltage-current hysteresis, saturation field-effect mobilities higher than 2 cm2/Vs, threshold voltages around 400 mV, subthreshold swings below 90 mV/dec, and current “on/off” ratios in excess of 105. As a result of this combination of favorable properties, we have demonstrated high performance IGZO TFTs that can be operated with voltages well below 1 V.

(a) (c)

(b) (d)

Fig. 1. 2D and 3D AFM topographical images of (a, b) untreated and (c, d) OTS-treated Ta2O5 films.

Experimental procedure

MIM capacitors with Ta/Ta2O5/Al and Ta/Ta2O5/OTS/Al structures were fabricated on 2×1.5 cm ultra-flat quartz coated glass substrates (Ossila, UK). Firstly, the substrates were rinsed with deionized water in order to remove any water-soluble contaminants and dried with compressed N2. Next, the slides were sonicated for 30 min in acetone, methanol and isopropyl alcohol (IPA). Then, the glass slides were taken from the sonic bath and again dried with pure N2. To remove any residual organic contaminants, the substrates were put into UV Ozone cleaner for 30 min. The fabrication of the MIM capacitors and IGZO TFTs started with depositing a 100-nm thick layer of Ta through a shadow mask by using radio frequency (RF) magnetron sputtering with a 2-inch diameter Ta target (Kurt J. Lesker, 99.9% pure) at a power of 70 W in a pure Ar atmosphere with a total pressure of 0.5 Pa at room temperature. Next, the patterned Ta was oxidized using anodization process to form a thin Ta2O5 film. The anodization process was carried out in 1 mM citric acid which served as the electrolyte. A pure Au wire (99.999%) and the patterned Ta film were used as the cathode and the anode, respectively [24]. To form Ta2O5 films with different thicknesses, various anodization voltages VA = 20, 15, 10, and 5 V and a constant current density JA = 0.01 mA/cm2 were applied using Keithley 2400 Source Meter. The anodization was carried out until JA dropped to 0.005 mA/cm2. Then, the samples were rinsed with deionized water and dried using pure N2. Based on the anodization ratio of tantalum reported in the literature ([c]Ta = 2.2 ± 0.2 nm/V) [24], the Ta2O5 thicknesses for 20, 15, 10, and 5 V anodization voltages were approximately 44 ± 4, 33 ± 3, 22 ± 2, and 11 ± 1 nm, respectively. To study the effect of OTS on the dielectric properties of the anodically oxidized Ta2O5 films, some of the as-prepared Ta/Ta2O5 samples were treated by O2 plasma for 2 min and then immersed in 5 mM of fresh OTS solution (Sigma-Aldrich, ≥ 99.5%) in anhydrous toluene for 30 min in ambient conditions. Afterwards, the substrates were taken from the OTS solution and promptly rinsed with fresh toluene, acetone and IPA and blow-dried with compressed N2. Finally, the OTS-treated samples were placed on a hot plate at ~150 °C under flow of N2 for 2 h. To finish MIM capacitors, a 100-nm thick, 1000 m x 100 m layer of Al was thermally evaporated through a shadow mask to form the top electrode of the devices (the total area of capacitors including the connection to the contact pads was 1.70×10-3 cm2). To confirm the reproducibility of the measured dielectric parameters, 100 Ta/Ta2O5/Al and 100 Ta/Ta2O5/OTS/Al MIM capacitors for each Ta2O5 film anodized with VA = 20, 15, 10, and 5 V were fabricated and 20 of each lot were randomly chosen for the measurements. To finish IGZO TFTs, a 25-nm thick layer of IGZO was firstly deposited onto the as-prepared Ta/Ta2O5 and OTS-treated Ta/Ta2O5 samples through a shadow mask by RF magnetron sputtering using a 2-inch diameter IGZO target (In:Ga:Zn = 1:1:1) at the power of 50 W at room temperature. Then, a 100 nm (1000 m x 100 m) Al film was deposited onto the IGZO films to form the source and drain contacts through a shadow mask. The transistor channel widths (W) and lengths (L) were 1 mm and 30 µm, respectively. In order to test the TFT fabrication reproducibility, 50 Ta2O5/IGZO and 50 Ta2O5/OTS/IGZO TFTs for each Ta2O5 film anodized with VA = 20, 15, and 10 V were fabricated and 20 of each TFTs were arbitrarily chosen for the measurements. The electrical characterization of capacitors and TFTs were measured by an Agilent E4980A LCR meter and an Agilent E5270B semiconductor analyzer at room temperature, respectively.

(a) (b)

(c) (d)

Fig. 2. (a) Capacitance density vs. frequency, (b) leakage current density vs. applied voltage, (c) loss tangent vs. frequency, and (d) capacitance-voltage (C-V) characteristics of the studied Ta/Ta2O5/Al capacitors.

Results and Discussion

Fig. 3. 3D schematic of the fabricated Ta/Ta2O5/IGZO/Al TFTs using pristine Ta2O5 gate dielectric.

Fig. 1 shows 500 nm × 500 nm 2D and 3D atomic force microscopy (AFM) topography images of Ta/Ta2O5 (a/b) and Ta/Ta2O5/OTS (c/d) surfaces, respectively. AFM images were obtained with a PSIA XE100 non-contact AFM, with 100-micron scanner. Measurements were performed in the tapping mode and a NT-MDT ultra-sharp cantilever was used for the data collection. As can been seen, the pristine and OTS-modified Ta2O5 films are relatively flat and featureless. The root-mean-square (RMS) roughness of the unmodified Ta2O5 is found to be 0.92 nm. It appears that the anodization process is capable of delivering as smooth Ta2O5 films as it is in the case of Ta2O5 deposited using ALD [25]. The RMS roughness of the OTS-modified Ta2O5 is significantly reduced and is found to be 0.55 nm. This result is in a good agreement with the previously published reports where it was shown that the modification of metal oxide dielectrics with self-assembled monolayers significantly reduces their surface roughness [26].

In order to obtain high performance thin film transistors, the optimization of the gate dielectric properties is essential. Consequently, different thicknesses of Ta2O5 films were grown by varying the anodization voltage from 20 to 5 V in step of 5 V. To confirm the reproducibility of the measured dielectric parameters, 20 Ta/Ta2O5/Al MIM capacitors for each Ta2O5 film anodized with VA = 20, 15, 10, and 5 V were randomly chosen from 100 fabricated devices. Fig. 2(a) illustrates the capacitance density vs. frequency of the capacitors from 100 Hz to 100 kHz. As shown, decreasing the anodization voltage VA from 20 V to 5 V increases the capacitance density (Ci) from 585 nF/cm2 to 1840 nF/cm2 with a maximum standard deviation of 20 nF/cm2 at 1 kHz. However, as the VA decreases from 20 to 5 V, the measured values of Ci from 100 Hz to 100 kHz become unstable especially at high-end frequency range. This is particularly obvious for Ta2O5 anodized at 5 V. This behaviour of Ci is very likely due to the increase of the leakage current through the dielectric. Indeed, as shown in Fig. 2(b), the leakage current density (JLEAK) rises sharply by at least three orders of magnitude when measured at 1 V from 2.8 × 10-5 A/cm2 for 20 V to 0.1 A/cm2 for 5 V anodized Ta2O5, respectively. High JLEAK is extremely undesirable because it strongly influences the characteristics of MIM capacitors, as well as the operation of TFTs. The negative effect of the leakage current on the capacitor characteristics is reflected in Fig. 2(c) where Ta2O5 anodized at 5 V shows a considerable deterioration of the loss tangent (tan(δ)) at both ends of the probed frequency range. Typically, all dielectrics have two types of losses, namely, conduction loss and dielectric loss. Conduction loss occurs when a charge flows through the dielectric from one electrode to another and dielectric loss is defined as a rotation or movement of atoms or molecules within the dielectric medium in an alternating electric field. An effective approach to defining these losses in detail is to express them by a dielectric constant as a complex number as shown in (1):

(1)

where ε' is the ac capacitivity, ε'' dielectric loss factor and δ dielectric loss angle, respectively. However, loss tangent (tan δ) is usually expressed by ε' and ε" and δ as shown in (2):

(2)

Fig. 3 presents a cross-section of the fabricated bottom-gate, top-contact IGZO TFTs (Ta/Ta2O5/IGZO/Al). Representative output and transfer characteristics of the devices using the Ta2O5 anodized at 10 V (~22 ± 2 nm, device A), 15 V (~33 ± 3 nm, device B), and 20 V (~44 ± 4 nm, device C) are shown in Fig. 4 (a-f), respectively. As demonstrated in Fig. 4, all three types of transistors operate at 1 V in n-channel enhancement mode and show a clear linear, pinch-off and saturation regimes with a very small hysteresis between forward (-0.1 to 1 V) and backward (1 to -0.1 V) sweeps. The drain current (ID) measured at the drain voltage (VD) and the gate voltage (VG) equal to 1 V for the device A is found to be about 18 µA. This is almost two times higher than for the device B and more than three times higher than for the device C. Although high ID is one of the most important parameters when considering the performance of TFTs, other transistor parameters such as leakage current (ILEAK), field-effect mobility (), threshold voltage (VTH), subthreshold swing (SS), onset voltage (VO), off current (IOFF), and on/off current ratio (Ion/off) should also be considered. The field-effect mobility in the saturation regime has been estimated using the drain current equation by applying a linear fit to the ID1/2 vs. VG as shown in:

(3)

where VG is the gate voltage, VT is the threshold voltage, Ci is the gate capacitance density, is saturation field-effect mobility, and W and L are channel width and length, respectively. In addition, to be able to compare the quality of the semiconductor-insulator interface of the fabricated transistors, the interfacial trap density (Nit) can be calculated using (4) [27]:

(a) (b)

(c) (d)

(e) (f)

Fig. 4. Representative output and transfer characteristics of IGZO TFTs using (a, b) 10 V anodized Ta2O5 (22 ± 2 nm, device A), (c, d) 15 V anodized Ta2O5 (33 ± 3 nm, device B), and (e, f) 15 V anodized Ta2O5 (44 ± 4 nm, device C).

(a) (b)

(c) (d)

Fig. 5. (a) Capacitance density vs. frequency, (b) leakage current density vs. applied voltage, (c) loss tangent vs. frequency, and (d) capacitance-voltage (C-V) characteristics of the studied Ta/Ta2O5/OTS/Al capacitors.

TABLE I

Electrical properties of the fabricated IGZO TFTs using pristine, Anodically oxidized Ta2 O5 as the gate dielectric.

(4)

where CG is the gate capacitance density, q is the electron charge, k is the Boltzmann’s constant, T is the temperature, and SS is subthreshold swing. Table I shows all key parameters of the fabricated IGZO TFTs.

It can be seen that the device C displays the highest mobility SAT = 6.3 cm2/Vs. At the same time, it shows the highest VTH = 650 mV and the lowest SS = 96 mV/dec. It is believed that charge carrier mobility in TFTs is directly affected by the interfacial trap density [28]. Nit is calculated to be 4.1×1012 cm-2 eV-1, 4.4×1012 cm-2 eV-1, and 2.1×1012 cm-2 eV-1 for device A, B, and C, respectively. Indeed, since the device C has the lowest interfacial trap density it exhibits the highest mobility. Also, since the capacitance density ofthe device C is lowest (Ci = 585 nF/cm2) it shows the highest VTH. On the contrary, it appears that significantly reduced channel capacitance, which is a sum of the depletion capacitance Cdep and Cit being related to the charging of the semiconductor-dielectric interface traps, contributes to the low SS [29]. Indeed, the SS in device C is reduced to 96 mV/dec. Although 1 V operation was achieved for all types of the fabricated IGZO TFTs using pristine Ta2O5, it was believed that the Ta2O5 dielectric needed further optimization to be able to compete with the recently demonstrated state-of-the-art IGZO TFTs [30], [31].

Fig. 6. 3D schematic of the fabricated Ta/Ta2O5/OTS/IGZO/Al TFTs using SAM-modified Ta2O5 gate dielectric.

(a)

(b)

Fig. 7. A representative (a) output and (b) transfer of IGZO TFTs gated with OTS-treated, 10 V anodized Ta2O5.

It has been reported previously that the passivation of metal oxide dielectrics with n-octadecyltrichlorosilane (OTS) SAM is an effective approach to improve the performance of OTFTs [24]. Therefore, to see whether a SAM treatment has the same beneficial effect on the performance of Ta2O5 thin film capacitors, 15 V (33 ± 3 nm) and 10 V (22 ± 2 nm) anodized Ta2O5 films have been modified with OTS. To confirm the reproducibility of the measured dielectric parameters, 20 Ta/Ta2O5/OTS/Al MIM capacitors for each Ta2O5 film anodized with VA = 15, and 10 V were randomly chosen from 100 fabricated devices. Fig. 5(a) demonstrates the capacitance density of Ta/Ta2O5/Al and Ta/Ta2O5/OTS/Al capacitors (cf. inset of Fig. 5(a)) as a function of frequency from 100 Hz to 100 kHz. As can be seen, the Ta/Ta2O5/OTS/Al capacitors are much more stable in the whole studied frequency range than the capacitors with pristine Ta2O5. Although adding a monolayer dielectric layer to the main dielectric appears to make the MIM capacitors more stable this results in an increase of total dielectric thickness, and in consequence, decrease of the capacitance density, because the two dielectric layers (i.e. Ta2O5 + OTS) are equivalent to two capacitors in series. Before the OTS surface modification, the capacitance density of 15 V and 10 V Ta2O5 films at 1 kHz is 875 nF/cm2 and 1020 nF/cm2, respectively. As expected, after the OTS treatment, the capacitance density is reduced to 346 and 433 nF/cm2 with a maximum standard deviation of 10 nF/cm2 at 1 kHz. Fig. 5(b) shows the leakage current density of Ta2O5 and OTS-treated Ta2O5 films vs. applied voltage bias from -1 to 1 V. As it is clear, the leakage current density of OTS-treated samples at ± 1 V has been dramatically reduced and found to be at least three orders of magnitude lower than their untreated counterparts. The leakage current of the OTS-treated Ta2O5 is somewhat asymmetric and slightly lower for positive voltages. This is very likely caused by the differences in the roughness of the Ta/Ta2O5 and Ta/Ta2O5/OTS interfaces and the use of electrodes with different work functions [32]. Fig. 5(c) shows the loss tangent as a function of frequency for both samples. Due to the lower leakage current in OTS-treated samples, tan(δ) is also showing a lower value which implies that the dielectric loss of these capacitors approaches the ideal value for this parameter. This fact is also reflected in the capacitance density vs. voltage characteristics presented in Fig. 5(d). Conversely, the OTS-treated Ta2O5 films show a constant value of Ci in the -1 to 1 voltage bias range the pristine Ta2O5 films appear somewhat unstable.

TABLE II

Electrical properties of the proposed IGZO TFTs usingOTS-modified, Anodically oxidized Ta2 O5 as the gate dielectric.

To see if the OTS-treated Ta2O5 results in better performing IGZO TFTs, Ta/Ta2O5/OTS/IGZO/Al transistors with 10 V (22 ± 2 nm) anodized Ta were fabricated (Fig. 6). Fig. 7 illustrates a typical output and transfer characteristics of the transistors. The key parameters of the fabricated IGZO TFTs gated with OTS-treated, 10 V anodized Ta2O5 are summarized in Table II. As shown in Fig. 7(a) and 7(b), the devices operate in n-type enhancement mode with a clear linear, pinch-off and saturation regimes. The clear linear increment of ID with VD at low VG implies that low resistance contacts were formed between the IGZO and Al. From Fig. 7(b), the field-effect mobility, the threshold voltage, the subthreshold swing, and Ion/off ratio are found to be 2.3 cm2/Vs, 400 mV, 88 mV/dec, and > 105, respectively. Comparing the transistors with IGZO TFTs using pristine Ta2O5, the mobility is reduced by a factor of three. This is very likely caused by the different structure and composition of IGZO on Ta2O5 and T2O5/OTS surfaces or OTS coating deterioration. The calculated interfacial trap densities for both types of transistors suggest that TFTs using OTS-modified Ta2O5 possess less interfacial traps than TFTs using pristine Ta2O5. However, more detailed studies of the SAM/IGZO interfaces are needed to fully understand this phenomenon. The gate leakage current (IG) is found to be about 0.1 nA at 1 V which is at least two orders of magnitude smaller than in our Ta2O5/IGZO TFTs [cf. ref. 19]. Recently reported IGZO TFTs using 10 nm CVD SiO2 as the gate dielectric display IG < 1 pA [33]. However, the area of the source and drain contacts here is at least 100 times larger than in [25]. Also, it is not obvious if the anodically oxidized Ta2O5 films with thicknesses at or below 20 nm are stoichiometric Ta2O5 which may lead to the increased conductivity of such films [34]. As a result, currently the performance and the application of the demonstrated TFTs is somewhat limited by the gate leakage (i.e. high off current) and a significant reduction of the source/drain contact area is needed for their use in displays. The onset voltage of the OTS-treated Ta2O5 IGZO TFTs is reduced to 100 mV, the threshold voltage is decreased to 400 mV, and the subthreshold swing is below 90 mV/dec. Hence, it appears that the modification of the tantalum pentoxide with SAMs may be a useful approach towards ultra-low voltage, high performance IGZO TFTs.

Conclusions

Low threshold voltage IGZO TFTs gated with pristine and OTS-treated Ta2O5 operating at 1 V are demonstrated. The effect of different anodization voltages on the properties of Ta/Ta2O5/Al and Ta/Ta2O5/OTS/Al MIM capacitors and IGZO TFTs have been studied. It is shown that the best performing capacitors and transistors are realized with 10 V anodized (22 ± 2 nm), OTS-treated Ta2O5. The fabricated Ta2O5/OTS capacitors show good stability in the 100 Hz to 100 kHz and capacitance density in excess of 400 nF/cm2. The fabricated TFTs display relatively high field-effect mobilities (2.3 cm2/Vs), threshold voltages around 400 mV, subthreshold swings below 90 mV/dec, and high current on/off ratios well in excess of 105. It is envisaged that this approach is a very promising alternative to fabricate low voltage, inexpensive TFTs and TFT-arrays for low cost sensors and low-end, disposable electronics.

Acknowledgments

We would like to thank Christopher Muryn from the Department of Chemistry at the University of Manchester who obtained and helped analyse the atomic force microscopy (AFM).

References

[1]A. D. Sekar, T. Jayabalan, H. Muthukumar, N. I. Chandrasekaran, S. N. Mohamed, and M. Matheswaran, “Enhancing power generation and treatment of dairy waste water in microbial fuel cell using Cu-doped iron oxide nanoparticles decorated anode,” Energy, vol. 172, pp. 173–180, 2019.

[2]T. N. Huan et al., “Low-cost high-efficiency system for solar-driven conversion of CO2 to hydrocarbons,” Proc. Natl. Acad. Sci., vol. 116, no. 20, pp. 9735–9740, 2019.

[3]S. Ho-Kimura et al., “Origin of high-efficiency photoelectrochemical water splitting on hematite/functional nanohybrid metal oxide overlayer photoanode after a low temperature inert gas annealing treatment,” ACS omega, vol. 4, no. 1, pp. 1449–1459, 2019.

[4]A. Pedico et al., “High-performing and stable wearable supercapacitor exploiting rGO aerogel decorated with copper and molybdenum sulfides on carbon fibers,” ACS Appl. Energy Mater., vol. 1, no. 9, pp. 4440–4447, 2018.

[5]F. Bella et al., “Boosting the efficiency of aqueous solar cells: A photoelectrochemical estimation on the effectiveness of TiCl4 treatment,” Electrochim. Acta, vol. 302, pp. 31–37, 2019.

[6]F. I. M. Ali, F. Awwad, Y. E. Greish, A. F. S. Abu-Hani, and S. T. Mahmoud, “Fabrication of low temperature and fast response H2S gas sensor based on organic-metal oxide hybrid nanocomposite membrane,” Org. Electron., vol. 76, p. 105486, 2020.

[7]M. Jung et al., “transparent and flexible Mayan-pyramid-based pressure Sensor using facile-transferred indium tin oxide for Bimodal Sensor Applications,” Sci. Rep., vol. 9, no. 1, pp. 1–11, 2019.

[8]L. Lefrançois Perreault et al., “Spray‐Dried Mesoporous Mixed Cu‐Ni Oxide@ Graphene Nanocomposite Microspheres for High Power and Durable Li‐Ion Battery Anodes,” Adv. Energy Mater., vol. 8, no. 35, p. 1802438, 2018.

[9]E. Hourdakis, A. Travlos, and A. G. Nassiopoulou, “High-Performance MIM Capacitors With Nanomodulated Electrode Surface,” IEEE Trans. Electron Devices, vol. 62, no. 5, pp. 1568–1573, May 2015, doi: 10.1109/TED.2015.2411771.

[10]W. Tang, J. Li, J. Zhao, W. Zhang, F. Yan, and X. Guo, “High-performance solution-processed low-voltage polymer thin-film transistors with low-k/High-k bilayer gate dielectric,” IEEE Electron Device Lett., vol. 36, no. 9, pp. 950–952, 2015, doi: 10.1109/LED.2015.2462833.

[11]J. T. Smith, S. S. Shah, M. Goryll, J. R. Stowell, and D. R. Allee, “Flexible ISFET Biosensor Using IGZO Metal Oxide TFTs and an ITO Sensing Layer,” IEEE Sens. J., vol. 14, no. 4, pp. 937–938, Apr. 2014, doi: 10.1109/JSEN.2013.2295057.

[12]N. Mohammadian, S. Faraji, S. Sagar, B. C. Das, M. L. Turner, and L. A. Majewski, “One-Volt, Solution-Processed Organic Transistors with Self-Assembled Monolayer-Ta2O5 Gate Dielectrics,” Materials, vol. 12, no. 16, pp. 2563–2575, 2019, doi: 10.3390/ma12162563.

[13]S. Ezhilvalavan and T. Y. Tseng, “Preparation and properties of tantalum pentoxide (Ta2O5) thin films for ultra large scale integrated circuits (ULSIs) application -- A review,” J. Mater. Sci. Mater. Electron., vol. 10, no. 1, pp. 9–31, 1999, doi: 10.1023/A:1008970922635.

[14]S. Dueñas, E. Castán, J. Barbollas, R. R. Kola, and P. A. Sullivan, “Use of anodic tantalum pentoxide for high-density capacitor fabrication,” J. Mater. Sci. Mater. Electron., vol. 10, no. 5-6, pp. 379–384, 1999, doi: 10.1023/A:100890162.

[15]S. Kim, “Hydrogen gas sensors using a thin Ta2O5 dielectric film,” J. Korean Phys. Soc., vol. 65, no. 11, pp. 1749–1753, 2014, doi: 10.3938/jkps.65.1749.

[16]J. Yu et al., “Hydrogen gas sensing properties of Pt/Ta2O5 Schottky diodes based on Si and SiC substrates,” Sensors Actuators, A Phys., vol. 172, no. 1, pp. 9–14, 2011, doi: 10.1016/j.sna.2011.02.021.

[17]L. Zhang, J. Li, X. W. Zhang, X. Y. Jiang, and Z. L. Zhang, “High performance ZnO-thin-film transistor with Ta2O5 dielectrics fabricated at room temperature,” Appl. Phys. Lett. vol. 95, no. 7, pp. 072112–072112-3, 2009, doi: 10.1063/1.3206917.

[18]P. Barquinha et al., “Low-temperature sputtered mixtures of high-κ and high bandgap dielectrics for GIZO TFTs,” J. Soc. Inf. Disp., vol. 18, no. 10, pp. 762–772, 2010, doi: 10.1889/JSID18.10.762.

[19]L. Zhang, W. Xiao, W. Wu, and B. Liu, “Research progress on flexible oxide-based thin film transistors,” Appl. Sci., vol. 9, no. 4, 2019, doi: 10.3390/app9040773.

[20]C. J. Chiu, S. P. Chang, and S. J. Chang, “High-performance a-IGZO thin-film transistor using Ta2O5 gate dielectric,” IEEE Electron Device Lett., vol. 31, no. 11, pp. 1245–1247, 2010, doi: 10.1109/LED.2010.2066951.

[21]J. K. Jeong, “The status and perspectives of metal oxide thin-film transistors for active matrix flexible displays” Semicond. Sci. Technol., vol. 26, no. 3, pp. 034008–034018, 2011, doi: 10.1088/0268-1242/26/3/034008.

[22]S. A. DiBenedetto, A. Facchetti, M. A. Ratner, and T. J. Marks, “Molecular Self-Assembled Monolayers and Multilayers for Organic and Unconventional Inorganic Thin-Film Transistor Applications,” Adv. Mater., vol. 21, no. 14‐15, pp. 1407–1433, 2009, doi: 10.1002/adma.200803267.

[23]P. Sista et al., “Enhancement of OFET performance of semiconducting polymers containing benzodithiophene upon surface treatment with organic silanes,” J. Polym. Sci. Part A Polym. Chem., vol. 49, no. 10, pp. 2292–2302, 2011, doi: 10.1002/pola.24663.

[24]Y.-H. Kim and K. Uosaki, “Preparation of Tantalum Anodic Oxide Film in Citric Acid Solution - Evidence and Effects of Citrate Anion Incorporation,” J. Electrochem. Sci. Technol., vol. 4, no. 4, pp. 163–170, 2014, doi: 10.5229/JECST.2013.4.4.163.

[25]C.-H. Kao, H. Chen, and C.-Y. Chen, “Material and electrical characterizations of high-k Ta2O5 dielectric material deposited on polycrystalline silicon and single crystalline substrate,” Microelectron. Eng., vol. 138, pp. 36–41, 2015, doi: https://doi.org/10.1016/j.mee.2015.01.011.

[26]W. Wang, X. Shi, X. Li, and Y. Zhang, “Improved Electrical Properties of Pentacene MIS Capacitor with OTS Modified Ta2O5 Dielectric,” IEEE Electron Device Lett., vol. 37, no. 10, pp. 1332–1335, 2016, doi: 10.1109/LED.2016.2601626.

[27]J. Raja, K. Jang, H. H. Nguyen, T. T. Trinh, W. Choi, and J. Yi, “Enhancement of electrical stability of a-IGZO TFTs by improving the surface morphology and packing density of active channel,” Curr. Appl. Phys., vol. 13, no. 1, pp. 246–251, 2013, doi: https://doi.org/10.1016/j.cap.2012.07.016.

[28]J. S. Park, W.-J. Maeng, H.-S. Kim, and J.-S. Park, “Review of recent developments in amorphous oxide semiconductor thin-film transistor devices,” Thin Solid Films, vol. 520, no. 6, pp. 1679–1693, 2012, doi: https://doi.org/10.1016/j.tsf.2011.07.018.

[29]L. Feng, W. Tang, X. Xu, Q. Cui, and X. Guo, “Ultralow-Voltage Solution-Processed Organic Transistors With Small Gate Dielectric Capacitance,” IEEE Electron Device Lett., vol. 34, no. 1, pp. 129–131, Jan. 2013, doi: 10.1109/LED.2012.2227236.

[30]R. Yao et al., “Low-temperature fabrication of sputtered high-k HfO2 gate dielectric for flexible a-IGZO thin film transistors,” Appl. Phys. Lett., vol. 112, no. 10, p. 103503, 2018, doi: 10.1063/1.5022088.

[31]J. Yang et al., “High-Performance 1-V ZnO Thin-Film Transistors With Ultrathin, ALD-Processed ZrO2 Gate Dielectric,” IEEE Trans. Electron Devices, vol. 66, no. 8, pp. 3382–3386, Aug. 2019, doi: 10.1109/TED.2019.2924135.

[32]W. S. Lau, D. Q. Yu, X.Wang, H. Wong, Y. Xu, "Mechanism of I-V asymmetry of MIM capacitors based on high-k dielectric". In Proceedings of the 2015 China Semiconductor Technology International Conference, Shanghai, China, 18–19 March 2015, doi: 10.1109/CSTIC.2015.7153403.

[33]Y. Zhang, H. Yang, H. Peng, Y. Cao, L. Qin, and S. Zhang, “ Self-Aligned Top-Gate Amorphous InGaZnO TFTs With Plasma Enhanced Chemical Vapor Deposited Sub-10 nm SiO2 Gate Dielectric for Low-Voltage Applications,” IEEE Electron Device Lett., vol. 40, no. 9, pp. 1459–1462, 2019, doi: 10.1109/led.2019.2931358.

[34]T. V. Perevalov, V. A. Gritsenko, A. A. Gismatulin, V. A. Voronkovskii, A. K. Gerasimova, V. Sh. Aliev, and I. A. Prosvirin, " Electronic structure and charge transport in nonstoichiometric tantalum oxide," Nanotechnology, vol. 29, no. 26, pp. 264001-264010, 2018, doi: 10.1088/1361-6528/aaba4c.

7

> REPLACE THIS LINE WITH YOUR PAPER IDENTIFICATION NUMBER (DOUBLE-CLICK HERE TO EDIT) <