化学蚀刻的全外延钙钛矿氧化物微米级 FET

时间:2024-04-24 14:10:25 浏览量:0


Advent of high mobility perovskite oxide semiconductor BaSnO3 (BSO) has enabled all-perovskite oxide heterostructures such as 2DEGs and FETs. To date all-perovskite oxide device demonstrations have been focused on  finding and integrating the compatible perovskite dielectric oxides such as polar LaInO3 and LaScO3 and nonpolar BaHfO3 and SrHfO3. For these demonstrations the length scale of BSO-based heterostructure devices has  been about 100 µm, primarily due to the use of stencil masks for patterning. In order to further reduce the length  scale, we employed a top-down approach using both photolithography and chemical etching techniques to  pattern FETs made entirely of perovskite oxide materials: Ba0.997La0.003SnO3 channel layer, degenerately doped  Ba0.96La0.04SnO3 contact layer, and SrHfO3 gate oxide layer. FETs of 3 µm channel length were fabricated using  hydrofluoric acid and aqua regia as etchants. The FET exhibits a mobility of 38.8 cm2 /Vs, an on/off ratio of 5.06  × 107 , and a drain current density of 6.05 × 10− 2 mA/μm, consistent with our expectation. These findings  demonstrate the feasibility of patterning BSO through photolithography and chemical etching while maintaining  the subsequent epitaxial growth, suggesting that BSO can be employed in a broader range of applications as well  as for more precise studies of its intrinsic properties.


Introduction  

Epitaxial perovskite oxide of ABO3 is the most abundant form of  oxides in earth and has been extensively studied for its diverse properties during the last few decades. Recently, BaSnO3 (BSO), a  perovskite oxide semiconductor, has been attracting attention as a  semiconducting channel material due to the fact that there are  many interesting compatible perovskite oxides to be used as gate oxides.  Furthermore, it shows the following inherent excellent properties: First,  BSO is a cubic perovskite with a lattice constant of 4.116 Å and a wide  bandgap semiconductor with a bandgap of 3.1 eV. This means it is  suitable for high power devices operating at high voltages. Second, BSO  has high mobility. For La-doped BSO (BLSO) single crystals, the  high doping region results in high mobility up to 320 cm2 /Vs. Affected  by factors such as dislocation, films typically have low mobilities in the  range of 20 to 100 cm2 /Vs, but this is comparable or superior to other  oxides such as ZnO and SrTiO3 (STO). Third, BSO has high oxygen stability . BSO’s high oxygen stability prevents structural degradation  during various manufacturing processes. Fourth, BSO can be easily  doped with La . Intrinsically an insulator, BSO can be easily doped with La to degenerate levels, enabling its use as both a channel  and contact material. The ability to differentiate the channel and electrode through doping concentrations allows for full perovskite  material-based device stacking and low contact resistance between the  channel and contact. Due to these characteristics, BSO can produce  various FETs by epitaxially depositing other compatible perovskites as  gate oxide on the BSO-based metallic contact layer and channel layer  with high mobility.


Method 

 All epitaxial layers used in the FET fabrication were grown using  pulsed laser deposition (PLD) at 750 ◦C. BSO and BLSO (Ba1− xLaxSnO3)  were deposited under 100 mTorr, while SHO was deposited at 30 mTorr.  The laser fluence was uniformly set at 1.43 J/cm2 . The target-sample  distance was maintained at 55 mm for BSO and 63 mm for SHO. The  targets used for deposition were all produced by Toshima manufacturing  Co., Japan. STO substrates were employed, and a 150 nm BSO buffer  layer was deposited to minimize the impact of the threading dislocations  in BSO coming from the lattice constant mismatch. To  define the shapes of the source, drain, channel, and gate electrode,  photolithography techniques and wet chemical etching methods were  employed. For photolithography, a maskless aligner (Nano system solutions, Inc, DL-1000A1) was utilized. After etching, heated dimethyl  sulfoxide (DMSO) at 80 ◦C was used to remove the photoresist.


Fig. 1 shows how HF responds to BSO and SHO. Fig. 1a presents the  outcome of line profiling, comparing regions of a BSO film that were  exposed to HF with those that were not. The exposed area reacted with  HF and was etched. To achieve a stable etching rate with hydrofluoric  acid, buffered oxide etchant 6:1, a mixture of 40% ammonium fluoride  (NH4F) in water and 49% HF in water at a volume ratio of 6:1, was used.  This solution was diluted to 1% in distilled water for appropriate etching  rates and surface conditions. Etching was performed for 30 s, the etching  depth was 23.2 nm, and the etching rate calculated from this was 0.772  nm/s. Fig. 1b is an AFM image of the surface of BSO exposed to HF. The  roughness is 1.28 nm, which is almost the same as 1.23 nm before  etching. This means that stable etching was achieved. Fig. 1c is the result  of line profiling between the exposed and unexposed areas of the SHO  film, which was partially exposed to hydrofluoric acid. The etching  conditions were the same as for BSO, but the results were completely  different from BSO; SHO exposed to hydrofluoric acid showed a higher  height than SHO not exposed. Fig. 1d is an optical microscope image of  SHO film partially etched with hydrofluoric acid. It was confirmed that a  clear boundary was created due to etching. These mean that the product  formed by the reaction between SHO and hydrofluoric acid does not  dissolve in water and remains on the film surface in the form of a salt.  The generated salt remains on the film and has a higher height than the  unexposed portion. Hafnium combines with fluorine to form HfF4 (Hf + 2 F2 → HfF4 (s)), which is insoluble in water, so the salt is assumed to be  HfF4. If the salt was on the film surface, it would easily come off physically, so we wiped the entire film using a cotton swab to confirm. After  wiping, it was confirmed through an optical microscope image that some  of the salt could be wiped off. 


图片1

Fig. 1. Etching by HF. (a) and (c) are the depth profiling between exposed and unexposed areas when parts of BSO and SHO films were exposed to HF, respectively.  In both cases, the etching time is 30 s and 1/100 diluted buffered oxide etchant was used. (b) is an AFM image of the surface of BSO exposed to HF. (d) is an optical  microscopy image of a SHO film partially exposed to HF.


Fig. 2 shows how aqua regia reacts to BSO and SHO. Fig. 2a depicts  the line profile comparison of a BSO film, showing differences between  areas exposed to aqua regia and those that were not. As in the case of  hydrofluoric acid, BSO was also etched by aqua regia. Aqua regia was  prepared by mixing 35% HCl and 60% HNO3 in a volume ratio of 3:1  immediately before etching. And to control the etching rate, it was  diluted to 1/4 with distilled water. Etching was performed for 60 s, the  etching depth was 70.5 nm, and the etching rate was 1.175 nm/s.  However, the etching rate of aqua regia showed a large variation,  ranging from 0.079 to 1.53 nm/s, even when aqua regia was produced  using the same recipe. This is thought to be because the reaction of aqua  regia itself varies greatly over time. Due to the etching rate that changes  in time, aqua regia is not suitable for etching that requires precise depth  control. However, it can be used when etching the entire thickness of the  BSO film. Fig. 2b is an AFM image of the surface of BSO etched with aqua  regia. The roughness is 868 pm, showing a flat surface. On the other  hand Fig. 2c shows the line profiling results for a SHO film, contrasting  the sections exposed to aqua regia with the unexposed areas. Although  the etching conditions are the same as for BSO, unlike BSO, SHO exposed  to aqua regia did not change height, meaning that SHO does not react  with aqua regia. Fig. 2d is an AFM image of the surface of SHO exposed  to aqua regia. The roughness shows a low value of 310 pm. The fact that  SHO is not etched by aqua regia could be an advantage in fabrication.  Especially in situations where it is difficult to etch BSO with aqua regia  to the exact desired thickness, SHO can be used as an etching stop layer.  In addition, the XRD data of the FET in Fig. S1 suggest that epitaxial  growth can be maintained on the etched surfaces. Fig. 3 shows the shape  and fabrication process of the FET using chemical etching. Fig. 3a presents the schematic top view of the entire device.

图片2

Fig. 2. Etching by aqua regia. (a) and (c) are the depth profiling between exposed and unexposed areas when parts of BSO and SHO films were exposed to aqua regia,  respectively. In both cases, the etching time is 60 s and 1/4 diluted aqua regia was used. (b) and (d) are AFM images of the surface of BSO and SHO exposed to aqua  regia, respectively.


图片3

Fig. 3. FET fabricated using chemical etching. (a) Schematic top view of the FET. (b) Optical microscope image of the FET. (c) Schematic fabrication process of FET.  4% BLSO for source-drain, 0.3% BLSO for channel and BSO for buffer are deposited in-situ. Source, drain, and channel structure is fabricated using HF. Gate electrode  is patterned using aqua regia.


Fig. 3b is the optical microscope view of the device’s source, drain,  and gate structure. Fig. 3c is the fabrication process of FET using  chemical etching on an STO substrate. A BSO buffer layer of 150 nm is  first deposited to reduce the dislocation density, followed in situ by a  20 nm layer of 0.3% BLSO for the channel and a 55 nm layer of 4% BLSO  for the contact layer. Subsequently, HF is used to etch 55 nm, exposing  the 0.3% BLSO layer, except where the source and drain pads will be  formed. The gap between these pads, set at 3 µm, defines the channel  length. Another etching step with HF is then performed to expose the  BSO buffer layer up to 24 nm, except for the areas designated for the  channel and contact pads. These two etching steps result in a source,  drain, and channel structure with a channel length of 3 µm and a width  of 10 µm. Next, using a stencil mask and PLD, SHO is deposited as the  gate oxide to a thickness of 197 nm, followed by a 50 nm layer of 4%  BLSO for the electrode. The gate oxide and electrode widths are set at  860 µm and 120 µm, respectively. To reduce leakage current by minimizing the overlap between the source, drain, and gate electrode, an  etching process is conducted to reduce the width of the gate electrode to  4 µm, using aqua regia. During this process, SHO acts as an etching stop  layer for the aqua regia.


Results and discussion

We have measured the electrical characteristics of our fully epitaxial  FET. Fig. 4a shows the output characteristics of our device, with drain  voltage (VDS) measured from 0 to 15 V, and gate voltage (VGS) sequentially measured from 0 to 26 V in 2 V intervals. An increase in the drain  current was observed as both VDS and VGS increased, and saturation of  the drain current at high drain voltages was noted. It can be observed  that the voltage at which the drain current saturates does not significantly differ from that of previous FETs with larger lengths. There  are two factors that determine the saturation of the drain current. These  are pinch-off and electron velocity saturation. The drain current  saturated by pinch-off is proportional to the square of VGS, and the drain  current saturated by the saturation velocity of electrons is proportional  to VGS. Based on this, we can see that the saturation of the drain current  of our device is caused by pinch-off. Fig. 4b displays the transfer characteristics, where VGS was swept from − 3 to 26 V while VDS was held at  the saturation region of 15 V. From this transfer curve, the field effect  mobility was calculated. The maximum mobility was measured at  38.8 cm2 /Vs. The drain current density increased up to 6.05 × 10− 2  mA/μm when VGS was at 26 V. The ratio of the highest to the lowest  drain current is 5.06 × 10⁷, and the subthreshold swing is 0.93 V /dec.  The subthreshold swing was calculated from the equation S = [∂ log₁₀  (IDS)/∂VGS]⁻ 1 . The threshold voltage (Vth) of our device is 16.1 V, obtained from a graph plotting ID 0.5 vs. VGS (Fig. 4c).


图片4

Fig. 4. Electrical characteristics of FET. (a) Output and (b) transfer characteristic of the FET. (c) Square root of drain current as a function of gate voltage in  saturation region. Using extrapolation, we obtained a threshold voltage of 16.1 V.


Conclusion  

In summary, we have successfully fabricated a micron-scale, fully  epitaxial accumulation-mode BSO-based FET with a channel length of  3 µm using photolithography and chemical etching. In the saturation  region, the drain current density reached up to 6.05 × 10− 2 mA/μm.  This is lower compared to submicron scale FETs but higher than those  reported for fully epitaxial BSO-based FETs with a longer channel  length. When compared to previously reported SHO FETs, the scaling of  the drain current density seems reasonable. The mobility was measured  at 38.8 cm2 /Vs, which is lower than that of previous all-epitaxial BSObased FETs. We believe further optimization of surface treatment will  lead to improvements in mobility. Our work demonstrates the feasibility  of fabricating smaller-sized epitaxial devices using accessible and  straightforward wet etching methods for BSO. Going forward, this  approach is expected to enable easy fabrication of various heterostructures and to facilitate detailed studies into the intrinsic properties  of BSO.

文件下载请联系管理员: 400-876-8096