Two high-k dielectric materials (Al2O3 and HfO2) were deposited on n-type (100) and (110) InAs surface orientations to investigate physical properties of the oxide/semiconductor interfaces and the interface trap density (Dit). X-ray photoelectron spectroscopy analyses (XPS) for native oxides of (100) and (110) as-grown n-InAs epi wafers show an increase in As-oxide on the (100) surface and an increase in InOx on the (110) surface. In addition, XPS analyses of high-k (Al2O3 and HfO2) on n-InAs epi show that the intrinsic native oxide difference between (100) and (110) epi surfaces were eliminated by applying conventional in-situ pre-treatment (TriMethyAluminium (TMA)) before the high-k deposition. The capacitance-voltage (C-V) characterization of HfO2 and Al2O3 MOSCAPs on both types of n-InAs surfaces shows very similar C-V curves. The interface trap density (Dit) profiles show Dit minima of 6.1 × 1012/6.5 × 1012 and 6.6 × 1012/7.3 × 1012 cm−2 eV−1 for Al2O3 and HfO2, respectively for (100) and (110) InAs surfaces. The similar interface trap density (Dit) on (100) and (110) surface orientation were observed, which is beneficial to future InAs FinFET device with both (100) and (110) surface channel orientations present.

III–V compounds, such as GaAs, InGaAs or InAs, have been intensively studied to replace Si as channel material because their high electron mobilities may enable low power and high performance applications for future CMOS. InAs is a promising candidate for MOS high-electron-mobility transistor devices1,2 because it has the highest electron mobility among in the arsenide-based III–V compounds. In addition, InAs is a binary compound that could provide a much simpler interface structure and property as compared to InGaAs. Although there are many studies on high-k/GaAs and high-k/InGaAs structures,3–7 only a few high-k/InAs structures have been investigated.8,9 Moreover, most of studies for high-k/InAs have focused on (100) surface orientation,10,11 with few (110) InAs investigations published.12 

Besides the introduction of high electron mobility channels, the device architectures are also very important for performance enhancement. Figures 1(a) and 1(b) show the conventional planar FET and nonplanar FinFET architecture. The employment of a FinFET architecture can extend the gate scaling beyond the planar transistor limitations to obtain better off-state performance, including subthreshold swing and DIBL (drain induced barrier lowering). FinFET architecture can also offer better device matching and variability due to the low doping concentration in the channel.13–15 However, unlike planar devices, FinFET channels consist of two surface planes, including the (100) and (110) planes, and the interface properties of both surfaces have to be considered. In this way, investigating both InAs (100) and (110) can help to understand the interface properties of InAs channel FinFET device architectures. In this article, we study the interfacial properties of Al2O3 and HfO2 on (100) and (110) n-InAs epi surfaces by physical and electrical characterization, respectively.

FIG. 1.

(a) Schematic diagrams of Planar FET structure and (b) schematic diagrams of FinFET structure.

FIG. 1.

(a) Schematic diagrams of Planar FET structure and (b) schematic diagrams of FinFET structure.

Close modal

Firstly, 400 nm InAs epilayers were grown by MOVPE (metal-organic vapor phase epitaxy) on (100) and (110) oriented Se-doped n-type InAs substrates. AFM (atomic force microscopy) scans on both the (100) and (110) n-InAs epi-layers show flat surfaces with root-mean-square roughness less than 2 Å in a 9 μm2 (3μm × 3μm) area. Both as-grown (100) and (110) n-InAs epitaxial wafers were analyzed by XPS (x-ray photoelectron spectroscopy) to study the surface native oxide composition formed when the layers were removed from the deposition chamber. Prior to the high-k oxide deposition, the samples were first cleaned in acetone and isopropanol, and a standard ex-situ surface clean solution to remove native oxides followed by rinsing in de-ionized (DI) water. For each experimental condition samples of both (100) and (110) orientations were loaded together into a Picosun 100 ALD (atomic layer deposition) system. The deposition started with TMA (tri-methyl aluminum) pre-treatment before depositing films of Al2O3 and HfO2. 8 nm and 2 nm of Al2O3 and HfO2 sample pairs were deposited for metal-oxide-semiconductor capacitor (MOSCAP) fabrication and XPS analysis, respectively. The TMA pre-treatment prior to the high-k deposition on III–V channel materials has been widely employed and proven to eliminate native oxide on III–V surface.16,17 MOSCAPs were fabricated by evaporation of Pd gate metal and Ti/Au backside ohmic contact, followed by annealing in forming gas (FG) ambient. TEM (transmission electron microscopy), EDX (energy dispersive X-ray spectroscopy), and XPS were used to investigate the physical and chemical structure of the high-k/III–V interface. C-V measurement was performed using an Agilent HP4294A impedance analyzer with a voltage sweep between −2 V to +2 V and frequencies varied from 1 kHz to 1 MHz at room temperature to characterize the behavior of the MOSCAP and extract interface trap density (Dit).

The As3d and In3d XPS spectra were collected to reveal the composition of native oxide for both surface planes, as shown in Figures 2(a) and 2(b). Figure 2(a) shows that there is more As-oxide on (100) than on (110), which is consistent with the As-rich condition at the (100) surface.18 Figure 2(b) shows slightly stronger In-oxide intensity on (110) surface indicating that the (110) surface may be In-rich.19 The intrinsic difference of the native oxides on the two different surface orientations was eliminated or minimized by TMA pre-treatment before the high-k (Al2O3 and HfO2) deposition. Figures 3(a)–3(d) show the In3d and As3d XPS spectra of Al2O3 on (100) and (110) n-InAs epi, respectively. The Al2O3 samples for XPS were pretreated with TMA. Figures 3(a) and 3(b) show that the XPS spectra of In-oxide on (100) and (110) after Al2O3 deposition are comparable and the peak location of the indium-oxide bond energy is similar on both surface orientations. Figures 3(c) and 3(d) show that the As-oxide bonds were suppressed on both (100) and (110) samples after TMA pre-treatment. As-As dimers were detected with comparable intensities and binding energy on (100) and (110) surface after Al2O3 deposition. Samples with HfO2 high-k were processed for XPS using the same TMA pre-treatment before HfO2 deposition. Figures 4(a) and 4(b) show that HfO2 samples on (100) and (110) n-InAs have similar In-oxide bond intensities on the surface and the same bond energy peak location in the energy spectrum. Similarly, As-oxide bonds were suppressed on both surface orientations. A similar intensity of As-As dimers and the binding energies were found, as shown in Figures 4(c) and 4(d). The intensity and binding energy of Hf5p in the HfO2 film on (100) and (110) were also found to be similar, as shown in Figures 4(c) and 4(d). This implies comparable HfO2 film quality on both surface orientations.

FIG. 2.

(a) The XPS spectra of As 2p3/2 on as-grown (100) and (110) InAs epi and (b) The XPS spectra of In 3d5/2 on as-grown (100) and (110) InAs epi.

FIG. 2.

(a) The XPS spectra of As 2p3/2 on as-grown (100) and (110) InAs epi and (b) The XPS spectra of In 3d5/2 on as-grown (100) and (110) InAs epi.

Close modal
FIG. 4.

The In3d and As3d XPS spectra of the HfO2 on (100) and (110) n-InAs epi: (a) and (b): In 3d5/2 on HfO2/InAs epi, (c) and (d): As3d spectra on HfO2/InAs epi.

FIG. 4.

The In3d and As3d XPS spectra of the HfO2 on (100) and (110) n-InAs epi: (a) and (b): In 3d5/2 on HfO2/InAs epi, (c) and (d): As3d spectra on HfO2/InAs epi.

Close modal
FIG. 3.

The In3d and As3d XPS spectra of the Al2O3 on (100) and (110) n-InAs epi: (a) and (b): In 3d5/2 on Al2O3/InAs epi, (c) and (d): As3d spectra on Al2O3/InAs epi.

FIG. 3.

The In3d and As3d XPS spectra of the Al2O3 on (100) and (110) n-InAs epi: (a) and (b): In 3d5/2 on Al2O3/InAs epi, (c) and (d): As3d spectra on Al2O3/InAs epi.

Close modal

TEM micrographs for the (100) and (110) oriented structures consisting of Pd/Al2O3/n-InAs epi/n-InAs substrate/Ti/Au MOSCAP are shown in Figures 5(a) and 5(b). The physical thickness for Al2O3 is 8.9 nm on both crystal planes was measured by TEM. Figures 6(a)–6(d) show the LAADF (low-angle annular dark field) TEM micrographs and EDX analysis for the HfO2 MOSCAP on both (100) and (110) InAs epi. The physical thickness extracted from TEM for HfO2 (Figures 6(a) and 6(b)) is 7.1 nm with 7 Å from the TMA (Al2O3) pre-treatment layer at the interface between HfO2 and InAs for both (100) and (110) planes. EDX analysis (Figures 6(c) and 6(d)) shows strong Al and O signals due to the pre-treatment at the interface of HfO2 and InAs. A sharp transition from InAs to the HfO2 layer is indicated by the fast decay of the In and As signals. The TEM analysis shows that the Al2O3 and HfO2 MOSCAPs on (100) and (110) have good thickness uniformity over the examined area. The Al2O3 and HfO2 dielectric layers are amorphous without any crystallite formation. In addition, there are no visible defects and no surface roughening at the Al2O3 interface with InAs and the Al2O3/HfO2 interface, indicating good interface quality on both surface planes. No interfacial native oxide is observed in the Al2O3 layer, while a thin layer is observed in HfO2 samples, which can be attributed to the TMA pre-treatment step.

FIG. 5.

(a) TEM micrographs for Al2O3 MOSCAP on (100) n-InAs and (b) TEM micrographs for Al2O3 MOSCAP on (110) n-InAs.

FIG. 5.

(a) TEM micrographs for Al2O3 MOSCAP on (100) n-InAs and (b) TEM micrographs for Al2O3 MOSCAP on (110) n-InAs.

Close modal
FIG. 6.

(a) LAADF TEM micrographs for HfO2 MOSCAP on (100) InAs, (b) LAADF TEM micrographs for HfO2 MOSCAP on (110) InAs, (c) EDX elemental analysis for HfO2 MOSCAP on (100) n-InAs, and (d) EDX elemental analysis for HfO2 MOSCAP on (110) n-InAs.

FIG. 6.

(a) LAADF TEM micrographs for HfO2 MOSCAP on (100) InAs, (b) LAADF TEM micrographs for HfO2 MOSCAP on (110) InAs, (c) EDX elemental analysis for HfO2 MOSCAP on (100) n-InAs, and (d) EDX elemental analysis for HfO2 MOSCAP on (110) n-InAs.

Close modal

Figures 7(a) and 7(b) show the multi-frequency C-V responses of Al2O3 and HfO2 MOSCAPs on (100) and (110) oriented n-InAs epitaxial layers, respectively. In the accumulation regime, the multi-frequency responses of Al2O3 on (100) and (110) n-InAs epi show similar behavior in frequency dispersion and Cacc (accumulation capacitance), with only a slight Cacc difference due to variation in physical oxide thickness. The frequency dispersion for Al2O3 on (100) and (110) n-InAs are 1 %/dec and 1.1 %/dec, respectively. The HfO2 samples also show similar multi-frequency responses for both orientations. The values of frequency dispersion for HfO2 on (100) and (110) n-InAs epi are 2.9%/dec and 2.6%/dec, respectively. The larger frequency dispersion in HfO2 samples is attributed to the presence of more border traps close the InAs/HfO2 interface. For Al2O3 MOSCAPs on (100) and (110) InAs in the depletion regime, the minimum capacitance Cdep, and the corresponding Vg are similar, ranging from -0.45 V (1 kHz) to −0.75 V (1 MHz). For HfO2 MOSCAPs on (100) and (110) InAs Cdep and the corresponding Vg are also similar, ranging from 0 V (1 kHz) to −0.5 V (1 MHz). As shown in the Figures 7(a) and 7(b), the capacitance at negative bias and C-V stretch-out of Al2O3 and HfO2 MOSCAPs on (100) and (110) InAs shows similar behavior in depletion and inversion. However, the visible difference in frequency dispersion in negative bias between the Al2O3 and HfO2 MOSCAPs implies that the high-k traps on the (100) and (110) surfaces have a different frequency response. Larger frequency dispersion was observed in the (110) InAs samples with both Al2O3 and HfO2 deposition.

FIG. 7.

(a) C-V curves of Al2O3 MOSCAP on (100) and (110) n-InAs and (b) C-V curves of HfO2 MOSCAP on (100) and (110) n-InAs with frequencies from 1kHz to 1MHz.

FIG. 7.

(a) C-V curves of Al2O3 MOSCAP on (100) and (110) n-InAs and (b) C-V curves of HfO2 MOSCAP on (100) and (110) n-InAs with frequencies from 1kHz to 1MHz.

Close modal

At sufficiently low frequency all charges (inversion and accumulation carriers and traps) are assumed to respond to the ac signal. To extract Dit from the measured low-frequency C-V curve we can solve Poisson's equation:

(1)

where V(y) is the position-dependent electrostatic potential, q the electronic charge, Nd and Na the ionized donor and acceptor concentrations, n(y) and p(y) the electron and hole concentrations and εr the relative permittivity of the semiconductor. Following,20 we account for non-parabolicity of the conduction band Γ-valley by writing the conduction band density of states as:

(2)

where |$m_\Gamma $|mΓ is the gamma-valley electron effective mass, Ec is the conduction band minimum and α = (1/Eg) · (1 − mΓ/m0) is the non-parabolicity factor where Eg is the semiconductor band gap energy; we take α∼2.69 eV−1 for InAs. Fermi-Dirac statistics are applied for both CB (conduction band) and VB (valence band).

Subsequently, by setting V(y) = 0 at y = infinity and applying a Cauchy boundary condition at the semiconductor/dielectric interface, we can calculate the intrinsic (without interface traps) MOSCAP Vgi(Ef) and Qsc(Ef), where Vgi is the intrinsic gate voltage, Qsc the integrated semiconductor charge and Ef the Fermi energy at the semiconductor/dielectric interface. To calculate Vgi the work function difference between gate and semiconductor ϕms is arbitrarily chosen to be equal to zero.

Now we introduce a fitting function Qit(Ef) which represents the total trapped charge at the semiconductor/dielectric interface at a given interface Fermi energy, such that Cgate(Vg) = −d(Qsc + Qit)/dVg matches the experimental data. Here,

(3)

is the applied gate voltage with Cox = ɛ0κ/Tox the gate dielectric capacitance. The effective interface trap density is then given by Dit(E) = −1/q dQit(E)/dE. In this procedure, uncertainty in the metal gate effective work function ϕm,eff appears as a constant offset in the fitted Qit-function, and is eliminated when the derivative is taken to calculate the effective Dit.

In the above analysis, input parameters Na and Nd are well-known from layer growth calibration. Despite the good layer thickness control of the ALD process, this is not the case for Cox since the dielectric constant κ is a priori unknown for low-temperature ALD. Therefore, we chose to treat Cox as fitting parameter. In our analysis, it was found that the condition of a monotonously decreasing Qit(Ef) function (i.e. positive Dit) in combination with a good fit with experimental data could only be met for a small Cox window - in practice we can estimate Cox with an uncertainly of about 10%.

By using a low-frequency fitting routine we discussed above on the 1 kHz samples, the Dit profiles for (100) and (110) samples has been extracted and shown in Figure 8. Nearly identical Dit profiles (Figure 8) were obtained for Al2O3 on (100) and (110) n-InAs epi, with Dit minimum of 6.1 × 1012 and 6.5 × 1012 cm−2 eV−1, respectively. HfO2 on (100) and (110) n-InAs epi-layers have slightly higher Dit values with Dit minimum of 6.6 × 1012 and 7.3 × 1012 cm−2 eV−1, respectively. The TMA pre-treatment before high-k deposition can efficiently passivate the surface on (100) and (110) n-InAs and provide similar Dit profiles and Dit_min on both surface orientations, which can be detected and quantified by LAADF TEM and EDX analysis as shown in Figures 6(a)–6(d). The difference of Dit_min location between Al2O3 and HfO2 were attributed to more border traps rather than the difference between Dit in the HfO2 film compared to the Al2O3 film since the interface of both types of samples are almost identical. The process conditions and parameters of all samples (Al2O3 or HfO2 on (100) and (110) InAs) are listed in Table I, including ex-situ pre-treatment, post-treatment (PMA: Post Metalization Anneal), high-k thickness, and Dit_min.

FIG. 8.

Dit (interface state density) profiles vs. E-Ev (eV).

FIG. 8.

Dit (interface state density) profiles vs. E-Ev (eV).

Close modal
Table I.

Summary of the all MOSCAP samples (Al2O3, HfO2 on (100) and (110) n-InAs epi).

DielectricSubstrate, n-InAsPre-treatmentPost-treatmentFilm thickness (Å)Dit_min (cm−2 eV−1)
Al2O3 (100) TMA Forming Gas anneal 89 +/− 1 6.1 × 1012 
  (110)     89 +/− 1 6.5 × 1012 
HfO2 (100)     71 +/− 1a 6.6 × 1012 
  (110)     71 +/− 1a 7.3 × 1012 
DielectricSubstrate, n-InAsPre-treatmentPost-treatmentFilm thickness (Å)Dit_min (cm−2 eV−1)
Al2O3 (100) TMA Forming Gas anneal 89 +/− 1 6.1 × 1012 
  (110)     89 +/− 1 6.5 × 1012 
HfO2 (100)     71 +/− 1a 6.6 × 1012 
  (110)     71 +/− 1a 7.3 × 1012 
a

HfO2 film thickness exclude 7 Å of Al2O3 pre-treatment layer

We have reported the physical and electrical characterization of Al2O3 and HfO2 on (100) and (110) n-InAs epitaxial layers. Both TEM and XPS results show nearly identical interface properties between Al2O3 and HfO2 for both (100) and (110) InAs surface orienations. Similar C-V characteristics in accumulation and depletion regimes as well as Dit profiles were also observed for both orientations. The similar interface trap density (Dit) on (100) and (110) surface orientation was observed, which can be extended to InAs FinFET device architecture with both (100) and (110) surfaces.

The authors would like to thank Y. C. Sun of TSMC and L. Samuelson of Lund University for their support.

1.
Gerben
Doornbos
and
Matthias
Passlack
,
IEEE Electr. Dev. Lett.
31
(
10
),
1110
(
2010
).
2.
S. W.
Chang
 et al., IEDM 2013.
3.
P. D.
Ye
,
J. Vac. Sci Technol. A
26
(
4
),
697
(Jul/Aug,
2008
).
4.
Han
Zhao
,
Jeff
Huang
,
Yen-Ting
Chen
,
Jung Hwan
Yum
,
Yanzhen
Wang
,
Fei
Zhou
,
Fei
Xue
, and
Jack C.
Lee
,
Appl. Phys. Lett.
95
,
253501
(
2009
).
5.
H. D.
Trinh
,
E. Y.
Chang
,
P. W.
Wu
,
Y. Y.
Wong
,
C. T.
Chang
,
Y. F.
Hsieh
,
C. C.
Yu
,
H. Q.
Nguyen
,
Y. C.
Lin
,
K. L.
Lin
, and
M. K.
Hudait
,
Appl. Phys. Lett.
97
,
042903
(
2010
).
6.
Zuoguang
Liu
,
Sharon
Cui
,
Pini
Shekhter
,
Xiao
Sun
,
Lior
Kornblum
,
Jie
Yang
,
Moshe
Eizenberg
,
K. S.
Chang-Liao
, and
T. P.
Ma
,
Appl. Phys. Letter.
99
,
222104
(
2011
).
7.
R.
Suzuki
,
N.
Taoka
,
M.
Yokoyama
,
S.
Lee
,
S. H.
Kim
 et al.,
Appl. Phys. Lett.
100
,
132906
(
2012
).
8.
H. D.
Trinh
,
G.
Brammertz
,
E. Y.
Chang
,
C. I.
Kuo
,
C. Y.
Lu
,
Y. C.
Lin
,
H. Q.
Nguyen
,
Y. Y.
Wong
,
B. T.
Tran
,
K.
Kakushima
, and
H.
Iwai
,
IEEE Electr. Dev. Lett.
32
(
6
),
752
(
2011
).
9.
R.
Timm
,
A.
Fian
,
M.
Hjort
,
C.
Thelander
,
E.
Lind
,
J. N.
Andersen
,
L.-E.
Wernersson
, and
A.
Mikkelsen
,
Appl. Phys. Lett.
,
97
,
132904
(
2010
).
10.
A. P.
Kirk
,
M.
Milojevic
,
J.
Kim
, and
R. M.
Wallace
,
Appl. Phys. Lett.
96
,
202905
(
2010
).
11.
Jun
Wu
,
E.
Lind
,
R.
Timm
,
Martin
Hjort
,
A.
Mikkelsen
, and
L.-E.
Wernersson
,
Appl. Phys. Lett.
100
,
132905
(
2012
).
12.
Hau-Yu
Lin
,
San-Lein
Wu
,
Chao-Ching
Cheng
,
Chih-Hsin
Ko
,
Clement H.
Wann
,
You-Ru
Lin
,
Shoou-Jinn
Chang
, and
Tai-Bor
Wu
,
Appl. Phys. Lett.
98
,
123509
(
2011
).
13.
Digh
Hisamoto
,
Wen-Chin
Lee
,
Jakub
Kedzierski
,
Hideki
Takeuchi
,
Kazuya
Asano
,
Charles
Kuo
,
Erik
Anderson
,
Tsu-Jae
King
,
Jeffrey
Bokor
, and
Chenming
Hu
,
IEEE Trans. Electron Devices
47
,
2320
2325
, (DEC,
2000
).
14.
Jakub
Kedzierski
,
David M.
Fried
 et al., IEDM 2001, pp.
437
440
.
15.
Bin
Yu
,
Leland
Chang
 et al., IEDM 2002, pp.
251
254
.
16.
Byungha
Shin
,
Jonathon B.
Clemens
,
Michael A.
Kelly
,
Andrew C.
Kummel
, and
Paul C.
Mclntyre
,
Appl. Phys. Letter.
96
,
252907
(
2010
).
17.
H. D.
Trinh
,
E. Y.
Chang
,
Y. Y.
Wong
,
C. C.
Yu
,
C. Y.
Chang
,
Y. C.
Lin
,
H. Q.
Nguyen
, and
B. T.
Tran
,
Jap JAP.
49
,
111201
(
2010
).
18.
Akihiro
Ohtake
,
Surface Science Reports
63
,
295
327
(
2008
).
19.
S.
Nannarone
,
M.
Pedio
,
Surface Science Reports
51
,
1
149
(
2003
).
20.
Viktor
Ariel-Altschul
,
Eliezer
Finkman
, and
Gad
Bahir
,
IEEE Trans. Electron Devices
,
39
,
1312
1316
(JUN,
1992
). H. Ko.