Interface traps (ITs) and border traps (BTs) in Al2O3/GeOx/p-Ge gate stacks were characterized using deep-level transient spectroscopy. Through evaluating the gate stacks with different GeOx thicknesses, the respective BTs in Al2O3, the Al2O3/GeOx interface region, and GeOx were detected. The density of ITs (Dit) near the midgap is lower in the metal-oxide-semiconductor (MOS) capacitors with thicker GeOx, while Dit near the valence band is lower in the MOS capacitor with thinner GeOx. The density of BTs (Nbt) in Al2O3 (6–9 × 1017 cm−3) is lower than those in GeOx (∼2 × 1018 cm−3), and the highest Nbt (∼1 × 1019 cm−3) was found in the Al2O3/GeOx interface region. Ge p-channel metal-oxide-semiconductor field-effect-transistors (MOSFETs) with Al2O3/GeOx/p-Ge gate stacks were fabricated and analyzed. We confirmed that the ITs and the BTs near the valence band edge of Ge affect the effective mobility of Ge p-MOSFETs in the high-field region.

Germanium (Ge) is one of the candidate materials for the next generation metal-oxide-semiconductor (MOS) field-effect transistors (MOSFETs) due to its high carrier mobility and high compatibility with the silicon (Si) platform.1,2 Besides, the quasi-direct bandgap of Ge can make a possibility for near infrared optical applications such as on-chip optical interconnect.3,4 Other Ge-based novel devices such as spintronics5 and thin-film transistors6,7 are also attracting great attention. Even though Ge has many advantages, there are many issues such as the leakage current in bulk Ge and the low dopant solubility of Ge. In order to suppress the leakage current and integrate Ge devices with the conventional Si process, the Ge-on-insulator (GOI) technology8–11 is developed. To increase the active dopant concentration, the epitaxial growth of in situ doped Ge is also studied.12,13

In order to bring out the high potential of Ge, the formation of high quality dielectric films on Ge as a gate insulator and passivation layer is necessary. Particularly, the quality of the gate insulator directly affects the current drivability of Ge MOSFETs or Ge spin MOSFETs. Many research groups have focused on the fabrication of Ge gate stacks and developed the characterization method of its interface state. About fabrication, nowadays, it is well known that GeOx/Ge has good interfacial quality.14,15 Therefore, Ge gate stacks with a thin Ge oxide interlayer, such as SiO2/GeO216 and Al2O3/GeOx,17,18 have been studied actively. About characterization, a low temperature conductance method19 and deep level transient spectroscopy (DLTS)20 are widely used for the quantitative evaluation of interface traps (ITs). On the other hand, in the case of the Ge MOS, not only ITs but border traps (BTs) located in a gate dielectric are also problematic because these traps degrade MOS characteristics and mobility. For deep understanding and performance improvement of the Ge MOS device, both the density and the spatial position of BTs in a stacked gate dielectric should be clarified. However, an in-depth study about BTs in a Ge stacked gate dielectric is limited.21 Recently, we succeeded in separating BT and IT signals through the DLTS measurement for SiO2/GeO2/Ge gate stacks grown by thermal and plasma oxidation.22 

In this study, the densities of ITs (Dit) and BTs (Nbt) in Al2O3/GeOx/p-Ge gate stacks grown by post-plasma oxidation (PPO) were evaluated. In addition, to study the relationship between BTs and the mobility of devices, Ge p-MOSFETs with Al2O3/GeOx/Ge gate stacks were also fabricated and analyzed.

In Sec. II, the fabrication procedures of MOS capacitors and MOSFETs are described. In Sec. III, the structural analyses and electrical properties of Al2O3/GeOx/Ge gate stacks are investigated. The electrical properties of Ge p-MOSFETs with the same gate stacks are also reported. In Sec. IV, the origin of the flat band voltage (VFB) shift and position of BTs are discussed. The effect of BTs on the mobility of MOSFETs is also discussed. Section V presents the conclusions.

The fabrication process of MOS capacitors is as follows. The p-type (100) Ge substrate with a doping concentration of 2.3 × 1016 cm−3 was used. After chemical cleaning by acetone and HF solution, the first layer of Al2O3 was deposited at 300 °C by atomic layer deposition (ALD) with precursors of water and trimethylaluminum for 3, 9, 14, and 20 cycles, corresponding to four samples, followed by PPO using electron cyclotron resonance (ECR) at room temperature with a microwave power of 500 W and Ar/O2 flow rates of 18/3 sccm for 1 min. The second layer of Al2O3 was deposited at 300 °C by ALD for 25 cycles to avoid current leakage during electrical measurements. For comparison, one sample with 45 cycles of Al2O3 deposition by ALD without PPO was also fabricated. 400 °C post-deposition annealing (PDA) was performed for 30 min. Then, the TiN layer was deposited by sputtering, followed by 350 °C post-metallization annealing (PMA) for 20 min. After that, the Al layer was also deposited and the gate electrode was patterned by lithography. To improve the electrical contact between TiN and Al layers, contact annealing (CA) was performed at 300 °C for 10 min. Finally, an InGa back contact was formed. All annealing in this study was performed in N2 ambient.

Ge p-MOSFETs were fabricated on an n-type (100) Ge substrate with a doping concentration of 9.3 × 1015 cm−3. We used a gate-first process. Both the process flow and the device structure are shown in Fig. 1. After wafer cleaning, SiO2 field oxide was deposited by ECR sputtering for isolation. Then, windows were opened by standard lithography and wet etching to define the active areas. To eliminate the influences of sputtering and etching on the surface of the active area, sacrificial GeO2 was formed by thermal oxidation at 450 °C for 30 min in O2 ambient and removed by rinsing in a dilute HF solution. Next, an Al2O3/GeOx/Ge gate stack was fabricated by the same method as MOS capacitors, as mentioned above. After 400 °C PDA, the Al/TiN gate electrode was deposited and patterned. To form the source/drain (S/D), B ion implantation was carried out with an acceleration energy of 30 keV and a dose of 1015 cm−2. The Al/Pt contact electrode for S/D was formed by the lift-off method, followed by CA and activation annealing (for S/D) at 300 °C for 10 min in N2 ambient. Subsequently, an InGa back contact was formed. The active hole density of 4.3 × 1019 cm−3 was confirmed by Hall effect measurement. Both Dit and Nbt were characterized using DLTS.22 

FIG. 1.

Fabrication process and device structure of Ge p-MOSFETs with Al2O3/GeOx/Ge gate stacks.

FIG. 1.

Fabrication process and device structure of Ge p-MOSFETs with Al2O3/GeOx/Ge gate stacks.

Close modal

To clarify the formation of the GeOx interlayer, x-ray photoelectron spectroscopy (XPS) measurement was carried out. Figure 2 shows the Ge 3d XPS spectra of Ge gate stacks with and without PPO at a photoelectron take-off angle of 90°. The spectra were calibrated with a Ge 3d core level (29.3 eV) and normalized by the signal intensity of the Ge bulk. The Ge oxide signal intensity decreases with increasing cycle number of the first Al2O3 layer. Simultaneously, the Ge oxide peak shifts to the low-energy direction, indicating that there is a change in the stoichiometry of Ge and oxygen. In other words, the x in GeOx becomes less than 2 with decreasing energy of the Ge oxide peak (the lower the energy, the smaller the x in GeOx). The spectrum of the 20-cycle one is well overlapped with that of the one without oxidation.

FIG. 2.

Normalized Ge 3d XPS spectra for Al2O3/GeOx/Ge structures with and without PPO.

FIG. 2.

Normalized Ge 3d XPS spectra for Al2O3/GeOx/Ge structures with and without PPO.

Close modal

The electrical properties of Al2O3/GeOx/Ge gate stacks were examined by using MOS capacitors. Figure 3 shows the capacitance–voltage (C–V) curves of Al/TiN/Al2O3/GeOx/Ge MOS capacitors with varied cycle numbers of the first Al2O3 layer deposition. With increasing cycle number, the VFB shifts to the positive direction and a bit stretch-out and small bumps can be seen. However, for the MOS capacitor without PPO, VFB is even more positive and the bump is bigger than the others. The thickness and equivalent oxide thickness (EOT) for each oxide layer is listed in Table I. Here, Al2O3 thickness was measured by ellipsometry. To calculate the EOT of each dielectric layer, permittivities of Al2O3 and GeOx were assumed to be 8.47 and 5.7, respectively. With increasing cycle number of the first Al2O3 layer, the GeOx layer becomes thinner. Figure 4 shows the energy distribution of Dit. In the region close to the midgap, lower Dit can be achieved by the thicker GeOx layer. However, in the region close to the valence band, the thicker GeOx gives higher Dit. Figure 5 shows the Nbt for all capacitors. The details of the measurement have been shown elsewhere.22 The highest and lowest Nbt were observed in the 9- and 20-cycle capacitors, respectively. The temperature dependence of Nbt is obvious for the 9-cycle capacitor.

FIG. 3.

C–V characteristics of Al/TiN/Al2O3/GeOx/p-Ge MOS capacitors with different cycle numbers of first-Al2O3. The measurement was performed at RT with a frequency of 1 MHz.

FIG. 3.

C–V characteristics of Al/TiN/Al2O3/GeOx/p-Ge MOS capacitors with different cycle numbers of first-Al2O3. The measurement was performed at RT with a frequency of 1 MHz.

Close modal
TABLE I.

Thickness and EOT of each dielectric layer for Al/TiN/Al2O3/GeOx/p-Ge MOS capacitors with different cycle numbers of first-Al2O3. Al2O3 thickness was measured by ellipsometry, and the permittivities of Al2O3 and GeOx were assumed to be 8.47 and 5.7, respectively.

First-Al2O3First-Al2O3Second-Al2O3First + second Al2O3MOSGeOx EOT (nm)GeOx
cyclethickness (nm)thickness (nm)EOT (nm)EOT (nm)(MOS EOT–Al2O3 EOT)thickness (nm)
3-cycle 1.02 3.2 1.94 2.54 0.60 0.87 
9-cycle 1.65 3.2 2.23 2.52 0.29 0.42 
14-cycle 2.14 3.2 2.46 2.5 0.04 0.06 
20-cycle 2.73 3.2 2.73 2.84 0.11 0.16 
First-Al2O3First-Al2O3Second-Al2O3First + second Al2O3MOSGeOx EOT (nm)GeOx
cyclethickness (nm)thickness (nm)EOT (nm)EOT (nm)(MOS EOT–Al2O3 EOT)thickness (nm)
3-cycle 1.02 3.2 1.94 2.54 0.60 0.87 
9-cycle 1.65 3.2 2.23 2.52 0.29 0.42 
14-cycle 2.14 3.2 2.46 2.5 0.04 0.06 
20-cycle 2.73 3.2 2.73 2.84 0.11 0.16 
FIG. 4.

Energy distribution of Dit for Al/TiN/Al2O3/GeOx/p-Ge MOS capacitors with different cycle numbers of first-Al2O3.

FIG. 4.

Energy distribution of Dit for Al/TiN/Al2O3/GeOx/p-Ge MOS capacitors with different cycle numbers of first-Al2O3.

Close modal
FIG. 5.

Temperature dependence of Nbt for Al/TiN/Al2O3/GeOx/p-Ge MOS capacitors with different cycle numbers of first-Al2O3.

FIG. 5.

Temperature dependence of Nbt for Al/TiN/Al2O3/GeOx/p-Ge MOS capacitors with different cycle numbers of first-Al2O3.

Close modal

Figure 6 shows drain current (ID) vs drain voltage (VD) characteristics of p-MOSFETs with the Al2O3/GeOx/Ge gate stack. Here, the channel length (L) and width (W) are 120 μm and 390 µm, respectively. The p-MOSFETs with a higher cycle number of the first Al2O3 layer showed higher current drivability. Figure 7 shows ID and source current (IS) vs gate voltage (VG) characteristics with VD = −10 mV. High off-states IDs at positive VGs are caused by the drain/substrate leakage current.23,24 Here, the threshold voltage (VTH) was obtained from the x-axis interception of IS/gm1/2 vs VG plots, where gm is the transconductance, and VTH for each sample are indicated as solid circles in Fig. 7. VTH shifts to the positive direction with increasing cycle number, which is consistent with the VFB shift obtained in C–V characteristics. Figure 8 shows the mobility of p-MOSFETs calculated by the split C–V method. All p-MOSFETs showed similar peak mobility (170–200 cm2/V s) at a low inversion carrier density region. On the other hand, in the high inversion carrier density region, the mobility of the 20-cycle one is higher than that of the others, of which the reason will be discussed later.

FIG. 6.

IDVD characteristics of Ge p-MOSFETs with different cycle numbers of first-Al2O3. All the devices have Al2O3/GeOx/Ge gate stacks with the same fabrication process as the MOS capacitors. L and W were 120 μm and 390 µm, respectively.

FIG. 6.

IDVD characteristics of Ge p-MOSFETs with different cycle numbers of first-Al2O3. All the devices have Al2O3/GeOx/Ge gate stacks with the same fabrication process as the MOS capacitors. L and W were 120 μm and 390 µm, respectively.

Close modal
FIG. 7.

ID– and ISVG characteristics of Ge p-MOSFETs with different cycle numbers of first-Al2O3.

FIG. 7.

ID– and ISVG characteristics of Ge p-MOSFETs with different cycle numbers of first-Al2O3.

Close modal
FIG. 8.

Effective hole mobility in Ge p-MOSFETs with different cycle numbers of first-Al2O3.

FIG. 8.

Effective hole mobility in Ge p-MOSFETs with different cycle numbers of first-Al2O3.

Close modal

Even though the Ge oxide signal was not detected by XPS for the 20-cycle one, the difference in C–V curves (Fig. 3) of 20-cycle and w/o PPO can be seen. Therefore, we believe that there is a Ge oxide layer in the 20-cycle one, or at least, the interface property was changed during PPO. A similar clue can be found in the Dit result and will be discussed in a later part.

The VFB shift is found in Fig. 3 even though the EOTs for all MOS capacitors are nearly the same. This is considered as an effect of fixed charges in first Al2O3 or GeOx. To clarify the density of fixed charges (Qf), another set of MOS capacitors with varied cycles of second Al2O3 deposition were fabricated, while the cycle numbers of first Al2O3 deposition were fixed. Their VFBs are plotted in Fig. 9. Since VFB’s are linearly dependent on the EOT while the cycle numbers of first Al2O3 deposition were fixed, the second Al2O3 is free of volume charges within the detection limit.25VFB is given by the following expression:

(1)

where ΦMS is the work function difference between the metal gate electrode and semiconductor, εSiO2 is the dielectric constant of SiO2, and ε0 is the permittivity in vacuum. Since the electrode material and substrate doping were not changed, Qf can be derived from the slope of plots without the information of ΦMS. The fitting of data with Eq. (1) revealed positive Qf with a density of 8.2 × 1010 cm−2 in the 9-cycle capacitor and negative Qf with densities of 2.9 × 1012 cm−2 and 4.6 × 1012 cm−2 in the 14- and 20-cycle capacitors, respectively. It was reported that ALD-grown Al2O3 contributes the negative polarity of interface fixed charges,26–28 while GeO2 contributes the positive fixed charges.14,29,30 Therefore, Qf became more and more positive with decreasing cycle number of the first Al2O3 due to thicker GeOx. Accordingly, VFB shifted negatively when the GeOx layer was formed, which coincides with the finding of Matsuoka et al.28 

FIG. 9.

VFB–EOT plots for Ge MOS capacitors with different cycle numbers of first-Al2O3.

FIG. 9.

VFB–EOT plots for Ge MOS capacitors with different cycle numbers of first-Al2O3.

Close modal

It is plausible that the thinner GeOx interlayer causes a higher Dit near the midgap. Many groups have reported similar findings.16,31,32 According to the XPS result (Fig. 2), when the GeOx interlayer is thinner, GeOx is composed of a higher portion of suboxides due to the shift of the Ge oxide peak. This could imply an increase in imperfect Ge–O or Ge–Ge bonds at the interface, causing the increase in Dit. By contrast, the Dit near the valence band decreases with decreasing GeOx thickness. This phenomenon is similar to the defect termination during Al post-metallization annealing reported previously.33 The thinner the GeOx interlayer, the higher the chance of Al oxide existence at the interface, thus decreasing the Dit near the valence band. Because the 20-cycle sample contains an ultrathin GeOx interlayer, Al2O3 possibly exists at the GeOx/Ge interface. Instead, to be more precise, the interface could be considered as a mixture of Al oxide and GeOx, and this Al-bonded oxide acts as the defect terminator, as we have found in the previous work.33 

Based on our measuring condition of Nbt, detected BT signals originated from the position with a depth of 0.4 nm measured from the GeO2/Ge interface when the tunneling barrier is the valence band offset (∼3.6 eV) at the interface.22,34 According to the thickness of GeOx in Table I, the detected BT position can be categorized into three cases, as shown in Fig. 10. Since the GeOx thickness of the 3-cycle capacitor is over 0.4 nm, the detected BT signals are from GeOx [Fig. 10(a)]. The GeOx thickness of the 9-cycle capacitor is ∼0.4 nm, implying that the detected BT signals are from the Al2O3/GeOx interface region [Fig. 10(b)]. The GeOx thickness of 14-cycle and 20-cycle capacitors is less than 0.4 nm, indicating that the detected BT signals are from Al2O3 [Fig. 10(c)]. Here, the actual position of the detected BTs should be less than 0.4 nm when we consider the valence band offset (∼5.4 eV for Al2O3/Ge) of Al2O3/GeOx.35 Even if we take into account the GeOx stoichiometry change, which may slightly reduce the band offset of GeOx/Ge, the measuring position should be still in Al2O3. After matching the BT position with Nbt, Nbt in GeOx (∼2 × 1018 cm−3) is higher than that in Al2O3 (6–9 × 1017 cm−3), and the highest Nbt is in the Al2O3/GeOx interface region (∼1 × 1019 cm−3). The temperature dependence found in Nbt of the 9-cycle capacitor implies the thermal process involved, and this is another clue that the signals are from the interface.

FIG. 10.

Origin of detected BT signals of Al/TiN/Al2O3/GeOx/p-Ge MOS capacitors with first-Al2O3 deposition for (a) 3-cycle, (b) 9-cycle, and (c) 14- and 20-cycle.

FIG. 10.

Origin of detected BT signals of Al/TiN/Al2O3/GeOx/p-Ge MOS capacitors with first-Al2O3 deposition for (a) 3-cycle, (b) 9-cycle, and (c) 14- and 20-cycle.

Close modal

The remarkable results of Ge p-MOSFETs are the effective mobility difference in the high-field region. The mobility degradation in the high-field region is usually considered as an effect of surface roughness scattering.36 In this paper, a potential cause for the difference in surface roughness is the plasma oxidation process. In principle, plasma oxidation should have the greatest impact on the 3-cycle sample and an almost negligible impact on the 20-cycle sample by considering the different thicknesses of the first-Al2O3 layer and the formed GeOx layer. In addition, since the 3-cycle sample shows lower mobility than the 20-cycle one in the high-field region, the 3-cycle sample is supposed to have a rougher interface of GeOx/Ge. To confirm this, scanning transmission electron microscopy (STEM) analysis was performed. Figure 11 shows a STEM high-angle annular dark-field (HAADF) image of the Al/TiN/Al2O3/GeOx/Ge gate stack with 3-cycle first-Al2O3 deposition. The rms roughness value of the interface of GeOx/Ge is ∼0.2 nm, which is nearly the same level as a commercial Ge wafer surface (0.19 nm) measured by using an atomic force microscope (AFM), implying that the surface roughness may not be the factor of the mobility degradation in the high-field region. To reconfirm the interface roughness, we removed the oxide layers by using 0.5% HF solution and measured the surface roughness of two extreme cases, 3-cycle and 20-cycle, by using an AFM.2Figure 12 shows 1 × 1 µm2 AFM images of the GeOx/Ge interface. The rms roughness values of 3- and 20-cycle samples are 0.21 nm and 0.20 nm, respectively, which are entirely consistent with the STEM result. Since the formation of native oxide of Ge is layer-by-layer growth,37 the morphology of the Ge surface can be copied to the surface of its native oxide. Therefore, there is no doubt that these AFM measurements can provide the information of interface roughness. Consequently, the effect of surface roughness scattering cannot explain the mobility degradation in the high-field region.

FIG. 11.

A STEM-HAADF image of Al/TiN/Al2O3/GeOx/Ge gate stacks with 3-cycle first-Al2O3 deposition and without second-Al2O3 deposition.

FIG. 11.

A STEM-HAADF image of Al/TiN/Al2O3/GeOx/Ge gate stacks with 3-cycle first-Al2O3 deposition and without second-Al2O3 deposition.

Close modal
FIG. 12.

AFM image of the GeOx/Ge interface for the Al2O3/GeOx/Ge structures with first-Al2O3 deposition for (a) 3-cycle and (b) 20-cycle. The Al2O3/GeOx was removed by dilute HF, and the AFM measurements were performed in 1 × 1 µm2.

FIG. 12.

AFM image of the GeOx/Ge interface for the Al2O3/GeOx/Ge structures with first-Al2O3 deposition for (a) 3-cycle and (b) 20-cycle. The Al2O3/GeOx was removed by dilute HF, and the AFM measurements were performed in 1 × 1 µm2.

Close modal

Instead of surface roughness, BTs may be a relatively reasonable factor in this case. To discuss the connection between BTs and effective mobility, the band bending and trap states at different MOSFET operating conditions were considered. Figure 13 shows the band diagrams of Al2O3/GeOx/Ge gate stacks at room temperature under weak inversion (ϕs = ΨB), strong inversion (approximately ϕs = ΨB + 0.23 eV in this case), and even stronger inversion (ϕs > ΨB + 0.5 Eg), where ϕs is the surface potential and ΨB is the energy difference of the Fermi level (EF) and intrinsic Fermi level (Ei). It is worth noting that the charge neutrality level is ∼0.23 eV under the intrinsic Fermi level of Ge.38 Under weak inversion [Fig. 13(a)], the acceptor-like ITs near the lower half of the bandgap showing negative charges cause mobility degradation. Under strong inversion [Fig. 13(b)], there is nearly no contribution of neutral charged ITs to mobility degradation. When the inversion gets even stronger [Fig. 13(c)], the donor-like ITs near the valence band showing positive charges cause mobility degradation. Besides, the BTs have a higher chance to capture holes (carriers) at this state due to the high electric field. Therefore, both ITs and BTs near the valence band cause mobility degradation in the high-field region. A similar finding about the effect of traps in the insulating oxide near the interface on mobility degradation in the high-field region was reported in a Si-based memory device study, even though the traps were not called BTs there.39 

FIG. 13.

Band diagrams of a Ge p-MOSFET showing the occupancy of ITs and BTs when (a) ϕs = ΨB, (b) ϕs = ΨB + 0.23 eV, and (c) ϕs > ΨB + 0.5 Eg (corresponding to the high-field region). IT net charges are negative at condition (a), neutral at condition (b), and positive at condition (c).

FIG. 13.

Band diagrams of a Ge p-MOSFET showing the occupancy of ITs and BTs when (a) ϕs = ΨB, (b) ϕs = ΨB + 0.23 eV, and (c) ϕs > ΨB + 0.5 Eg (corresponding to the high-field region). IT net charges are negative at condition (a), neutral at condition (b), and positive at condition (c).

Close modal

The total effect of BTs in all Ge p-MOSFETs with Al2O3/ GeOx/Ge gate stacks is categorized into two and summarized in Fig. 14. Since Nbt in Al2O3 is smaller than that in GeOx or at the Al2O3/GeOx interface, the 20-cycle gate with nearly no GeOx contains fewer BTs in total compared to a 3- or 9-cycle gate. In addition, the 20-cycle gate shows the lowest Dit near the valence band, as shown in Fig. 4. Thus, in the high-field region, the highest effective mobility was achieved from MOSFETs with 20-cycle first Al2O3. Due to the low Nbt near the Ge valence band edge and the defect termination of the GeOx/Ge interface, Al2O3 is an exceptional material for the gate dielectric of Ge p-MOSFETs.

FIG. 14.

Illustration of Nbt in Ge p-MOSFETs with Al2O3/GeOx/Ge gate stacks.

FIG. 14.

Illustration of Nbt in Ge p-MOSFETs with Al2O3/GeOx/Ge gate stacks.

Close modal

By using DLTS, we evaluated Dit and Nbt of Al2O3/GeOx/p-Ge gate stacks with different GeOx thicknesses fabricated by PPO. By changing the thickness of GeOx, we successfully measured Nbts in the GeOx interlayer, the Al2O3/GeOx interface region, and the Al2O3 layer, respectively. Dit near the midgap is lower in the MOS capacitors with a thicker GeOx layer. By contrast, Dit near the valence band is lower in the MOS capacitors with a thinner GeOx layer, owing to the defect termination by Al oxide. BTs are mainly located at the Al2O3/GeOx interface region. ITs and BTs near the valence band edge of Ge affect the effective mobility of Ge p-MOSFETs in the high-field region. Therefore, reduction of ITs and BTs near the valence band is the key to reach high-performance p-MOSFETs. In this sense, Al2O3 is a suitable dielectric material for Ge p-MOSFETs.

The data that support the findings of this study are available within this article.

This work was partially supported by (JSPS) KAKENHI (Grant No. 17H03237 and 18KK0134) and the Advanced Graduate Program in Global Strategy for Green Asia, Kyushu University. XPS measurements were carried out using the facilities of the Center of Advanced Instrumental Analysis of Kyushu University. Ion implantation was carried out in the Center for Microelectronic System, Kyushu Institute of Technology.

1.
S.
Takagi
,
R.
Zhang
,
J.
Suh
,
S.-H.
Kim
,
M.
Yokoyama
,
K.
Nishi
, and
M.
Takenaka
,
Jpn. J. Appl. Phys.
54
,
06FA01
(
2015
).
2.
A.
Toriumi
and
T.
Nishimura
,
Jpn. J. Appl. Phys.
57
,
010101
(
2018
).
3.
J.
Kang
,
S.
Takagi
, and
M.
Takenaka
,
Opt. Express
26
,
30546
(
2018
).
4.
T.
Maekrua
,
T.
Goto
,
K.
Nakae
,
K.
Yamamoto
,
H.
Nakashima
, and
D.
Wang
,
Jpn. J. Appl. Phys.
58
,
SBBE05
(
2019
).
5.
K.
Hamaya
,
Y.
Fujita
,
M.
Yamada
,
M.
Kawano
,
S.
Yamada
, and
K.
Sawano
,
J. Phys. D.: Appl. Phys.
51
,
393001
(
2018
).
6.
H. A.
Kasirajan
,
W.-H.
Huang
,
M.-H.
Kao
,
H.-H.
Wang
,
J.-M.
Shieh
,
F.-M.
Pan
, and
C.-H.
Shen
,
Appl. Phys. Express
11
,
101305
(
2018
).
7.
K.
Moto
,
K.
Yamamoto
,
T.
Imajo
,
T.
Suemasu
,
H.
Nakashima
, and
K.
Toko
,
Appl. Phys. Lett.
114
,
212107
(
2019
).
8.
T.
Maeda
,
K.
Ikeda
,
S.
Nakaharai
,
T.
Tezuka
,
N.
Sugiyama
,
Y.
Moriyama
, and
S.
Takagi
,
Thin Solid Films
508
,
346
(
2006
).
9.
T.
Hosoi
,
Y.
Suzuki
,
T.
Shimura
, and
H.
Watanabe
,
Appl. Phys. Lett.
105
,
173502
(
2014
).
10.
J.
Kang
,
X.
Yu
,
M.
Takenaka
, and
S.
Takagi
,
Mater. Sci. Semicond. Process.
42
,
259
(
2016
).
11.
K.
Yu
,
F.
Yang
,
H.
Cong
,
L.
Zhou
,
Q.
Liu
,
L.
Zhang
,
B.
Cheng
,
C.
Xue
,
Y.
Zuo
, and
C.
Li
,
J. Alloys Compd.
750
,
182
(
2018
).
12.
Y.
Kumazawa
,
X.
Xu
,
T.
Maruizumi
, and
K.
Sawano
,
Semicond. Sci. Technol.
33
,
124006
(
2018
).
13.
A.
Vohra
,
C.
Porret
,
D.
Kohen
,
S.
Folkersma
,
J.
Bogdanowicz
,
M.
Schaekers
,
J.
Tolle
,
A.
Hikavyy
,
E.
Capogreco
,
L.
Witters
,
R.
Langer
,
W.
Vandervorst
, and
R.
Loo
,
Jpn. J. Appl. Phys.
58
,
SBBA04
(
2019
).
14.
H.
Matsubara
,
T.
Sasada
,
M.
Takenaka
, and
S.
Takagi
,
Appl. Phys. Lett.
93
,
032104
(
2008
).
15.
C.
Lu
,
C. H.
Lee
,
W.
Zhang
,
T.
Nishimura
,
K.
Nagashio
, and
A.
Toriumi
,
J. Appl. Phys.
116
,
174103
(
2014
).
16.
K.
Hirayama
,
K.
Yoshino
,
R.
Ueno
,
Y.
Iwamura
,
H.
Yang
,
D.
Wang
, and
H.
Nakashima
,
Solid-State. Electron.
60
,
122
(
2011
).
17.
R.
Zhang
,
T.
Iwasaki
,
N.
Taoka
,
M.
Takenaka
, and
S.
Takagi
,
Appl. Phys. Lett.
98
,
112902
(
2011
).
18.
R.
Asahara
,
I.
Hideshima
,
H.
Oka
,
Y.
Minoura
,
S.
Ogawa
,
A.
Yoshigoe
,
Y.
Teraoka
,
T.
Hosoi
,
T.
Shimura
, and
H.
Watanabe
,
Appl. Phys. Lett.
106
,
233503
(
2015
).
19.
D.
Kuzum
,
T.
Krishnamohan
,
A. J.
Pethe
,
A. K.
Okyay
,
Y.
Oshima
,
Y.
Sun
,
J. P.
McVittie
,
P. A.
Pianetta
,
P. C.
McIntyre
, and
K. C.
Saraswat
,
IEEE Electron Device Lett.
29
,
328
(
2008
).
20.
D.
Wang
,
S.
Kojima
,
K.
Sakamoto
,
K.
Yamamoto
, and
H.
Nakashima
,
J. Appl. Phys.
112
,
083707
(
2012
).
21.
M.
Ke
,
X.
Yu
,
C.
Chang
,
M.
Takenaka
, and
S.
Takagi
,
Appl. Phys. Lett.
109
,
032101
(
2016
).
22.
W.-C.
Wen
,
K.
Yamamoto
,
D.
Wang
, and
H.
Nakashima
,
J. Appl. Phys.
124
,
205303
(
2018
).
23.
T.
Maeda
,
Y.
Morita
, and
S.
Takagi
,
Appl. Phys. Express
3
,
061301
(
2010
).
24.
K.
Yamamoto
,
R.
Ueno
,
T.
Yamanaka
,
K.
Hirayama
,
H.
Yang
,
D.
Wang
, and
H.
Nakashima
,
Appl. Phys. Express
4
,
051301
(
2011
).
25.
H.
Kamata
and
K.
Kita
,
Appl. Phys. Lett.
110
,
102106
(
2017
).
26.
S.
Swaminathan
,
Y.
Sun
,
P.
Pianetta
, and
P. C.
McIntyre
,
J. Appl. Phys.
110
,
094105
(
2011
).
27.
D. K.
Simon
,
P. M.
Jordan
,
T.
Mikolajick
, and
I.
Dirnstorfer
,
ACS Appl. Mater. Interfaces
7
,
28215
(
2015
).
28.
R.
Matsuoka
,
E.
Shigesawa
,
S.
Miyamoto
,
K.
Sawano
, and
K. M.
Itoh
,
Semicond. Sci. Technol.
34
,
014004
(
2019
).
29.
C. H.
Lee
,
T.
Tabata
,
T.
Nishimura
,
K.
Nagashio
,
K.
Kita
, and
A.
Toriumi
,
Appl. Phys. Express
2
,
071404
(
2009
).
30.
H.
Nakashima
,
Y.
Iwamura
,
K.
Sakamoto
,
D.
Wang
,
K.
Hirayama
,
K.
Yamamoto
, and
H.
Yang
,
Appl. Phys. Lett.
98
,
252102
(
2011
).
31.
R.
Zhang
,
T.
Iwasaki
,
N.
Taoka
,
M.
Takenaka
, and
S.
Takagi
,
IEEE Trans. Electron Devices
59
,
335
(
2012
).
32.
C.
Lu
,
C. H.
Lee
,
T.
Nishimura
, and
A.
Toriumi
,
Appl. Phys. Lett.
107
,
072904
(
2015
).
33.
Y.
Nagatomi
,
T.
Tateyama
,
S.
Tanaka
,
W.-C.
Wen
,
T.
Sakaguchi
,
K.
Yamamoto
,
L.
Zhao
,
D.
Wang
, and
H.
Nakashima
,
Mater. Sci. Semicond. Process.
70
,
246
(
2017
).
34.
W. F.
Zhang
,
T.
Nishimula
,
K.
Nagashio
,
K.
Kita
, and
A.
Toriumi
,
Appl. Phys. Lett.
102
,
102106
(
2013
).
35.
G. D.
Wilk
,
R. M.
Wallace
, and
J. M.
Anthony
,
J. Appl. Phys.
89
,
5243
(
2001
).
36.
S.
Takagi
,
A.
Toriumi
,
M.
Iwase
, and
H.
Tango
,
IEEE Trans. Electron Devices
41
,
2357
(
1994
).
37.
S. K.
Sahari
,
H.
Murakami
,
T.
Fujioka
,
T.
Bando
,
A.
Ohta
,
K.
Makihara
,
S.
Higashi
, and
S.
Miyazaki
,
Jpn. J. Appl. Phys.
50
,
04DA12
(
2011
).
38.
A.
Dimoulas
,
P.
Tsipas
,
A.
Sotiropoulos
, and
E. K.
Evangelou
,
Appl. Phys. Lett.
89
,
252110
(
2006
).
39.
T.
Hori
,
Gate Dielectrics and MOS ULSIs
(
Springer
,
Berlin, Heidelberg
,
1997
).