Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Materials Science in Semiconductor Processing 40 (2015) 77–83

Contents lists available at ScienceDirect

Materials Science in Semiconductor Processing


journal homepage: www.elsevier.com/locate/mssp

High performance sol–gel spin-coated titanium dioxide dielectric


based MOS structures
Arvind Kumar, Sandip Mondal, S. Girish Kumar, K.S.R. Koteswara Rao n
Department of Physics, Indian Institute of Science, Bangalore, 560012, India

art ic l e i nf o a b s t r a c t

Article history: High-κ TiO2 thin films have been fabricated using cost effective sol–gel and spin-coating technique on
Received 15 April 2015 p-Si (100) wafer. Plasma activation process was used for better adhesion between TiO2 films and Si. The
Received in revised form influence of annealing temperature on the structure-electrical properties of titania films were in-
27 June 2015
vestigated in detail. Both XRD and Raman studies indicate that the anatase phase crystallizes at 400 °C,
Accepted 27 June 2015
Available online 10 July 2015
retaining its structural integrity up to 1000 °C. The thickness of the deposited films did not vary sig-
nificantly with the annealing temperature, although the refractive index and the RMS roughness en-
Keywords: hanced considerably, accompanied by a decrease in porosity. For electrical measurements, the films were
Anatase TiO2 integrated in metal-oxide-semiconductor (MOS) structure. The electrical measurements evoke a tem-
Sol–gel spin-coating
perature dependent dielectric constant with low leakage current density. The Capacitance–voltage (C–V)
High-κ
characteristics of the films annealed at 400 °C exhibited a high value of dielectric constant ( 34). Fur-
MOS
Surface roughness ther, frequency dependent C–V measurements showed a huge dispersion in accumulation capacitance
due to the presence of TiO2/Si interface states and dielectric polarization, was found to follow power law
dependence on frequency (with exponent ‘s’ ¼ 0.85). A low leakage current density of 3.6  10  7 A/cm2 at
1 V was observed for the films annealed at 600 °C. The results of structure-electrical properties suggest
that the deposition of titania by wet chemical method is more attractive and cost-effective for production
of high-κ materials compared to other advanced deposition techniques such as sputtering, MBE, MOCVD
and ALD. The results also suggest that the high value of dielectric constant ‘κ‘ obtained at low processing
temperature expands its scope as a potential dielectric layer in MOS device technology.
& 2015 Elsevier Ltd. All rights reserved.

1. Introduction and TiO2. The SiOx and TiO2 layer acts like capacitor in series, be-
cause the dielectric constant of SiOx is only 3.9 (considering the
In order to achieve continuous downscaling and improved SiOx dielectric constant equal to SiO2), hence the thicker interfacial
performance of complementary metal oxide semiconductor layer will decrease the effective dielectric constant of titania based
(CMOS) devices, intense research is devoted into probing high- CMOS devices [10]. Thus, much of the research work is widespread
permittivity (κ) dielectric materials to substitute conventional SiO2 on the application of thermodynamically metastable anatase
with its excessive gate leakage current due to direct tunneling at a phase. The deposition of titania by vacuum based processes are
thicknesses of r1.2 nm [1,2]. Among various high-κ dielectrics extensively investigated, but requires additional attention during
reported so far, titania seems to be a suitable alternative, owing to the process [11–14]. Alternatively, solution based process is quite
its exceptional physicochemical and optoelectronic properties [3– attractive, as titania can be obtained with high purity and chemical
7]. In addition, dielectric constant of titania is crystal structure homogeneity [15]. The solution deposition process is a simple and
dependent with rutile form exhibited higher values (10–170) cost effective approach for the fabrication of large area films with
compared to the anatase form (40) [8,9]. However, formation of diverse compositions, and is applicable to flexible and transparent
rutile is generally observed at high temperature ( 4700 °C), which electronic device systems. Recently, the effects of the processing
results in the formation of thicker SiOx interfacial layer between Si variables on the chemical composition, interfacial reactions, and
dielectric properties of TiO2/Si films deposited by a sol–gel route
have been reported [16–20]. Majority of the earlier work on liquid-
n
Corresponding author. based films deposition are on relatively thick TiO2 films, their
E-mail addresses: arvind@physics.iisc.ernet.in (A. Kumar),
sandipmondal@physics.iisc.ernet.in (S. Mondal),
details are lacking though. In addition, these films possess large
girichem@yahoo.co.in (S.G. Kumar), ksrkrao@physics.iisc.ernet.in, leakage currents with poor device performance and non-uni-
raoksrk@gmail.com (K.S.R. Koteswara Rao). formity. We found no work on the studies related to the electrical

http://dx.doi.org/10.1016/j.mssp.2015.06.073
1369-8001/& 2015 Elsevier Ltd. All rights reserved.
78 A. Kumar et al. / Materials Science in Semiconductor Processing 40 (2015) 77–83

20

19

Grain Size (nm)


18

17

16

15
400 500 600 700 800 900 1000

Temperture ( 0 C)
Fig. 1. (a) GIXRD patterns of 36 nm as deposited and annealed TiO2 films at different annealing temperatures. (b) Dependence of grain size on annealing temperature,
estimated from FWHM of (101) peaks.

properties of the films as a function of annealing temperature. In respectively. The Raman spectra were obtained in back-scattering
this work, we report 36 nm thin film of anatase TiO2 deposited on configuration using on a Horiba JobinYvon Lab RAM HR instru-
Si-wafer by combined sol–gel spin-coating method. The relation ment. The excitation wavelength was 514.5 nm line using argon
between structure-electrical properties with annealing tempera- ion laser. The acquisition time was 10 s, and the data was recorded
ture of the film is focused in detail. using o3 mW of laser power (at the laser head) with a spectral
resolution of 0.4 cm  1. The film thickness and ellipsometry data
(refractive index vs wavelength) was measured by variable angle
2. Experimental spectroscopic ellipsometry (M-2000, J.A. Woollam Co., Inc., USA),
and later the thickness was verified by cross-sectional Scanning
2.1. Sample preparations and MOS capacitor fabrication Electron Microscopy using Ultra 55, Carl Zeiss (Fig. S2). The film
roughness was measured by AFM (NDMDT Russia). X-ray photo-
TiO2 films of  36 nm thick were fabricated by the sequential electron Spectroscopy measurements were performed on an AXIS
steps of solution preparation, spin-coating, drying, and finally ULTRA from AXIS 165 using monochromatized Al Kα as the ex-
annealing at different temperatures. The TiO2 precursor solution citation source. All binding energies were calibrated to the C 1 s
was prepared by dissolving 0.6 ml titanium isopropoxide peak at 284.8 eV of the surface adventitious carbon. Area of the
[C12H28O4Ti] in 8.5 ml isopropyl alcohol [CH3CHOHCH3], and 20 μl device is measured by optical microscope and it was different from
HCl was added to avoid rapid hydrolysis of the precursor. The the mask area due to shadow effect. Metal-oxide-semiconductor
solution was then subjected to magnetic mixer for 3 h to ensure (MOS) capacitors were used to investigate the dielectric properties
homogeneity, followed by 12 h aging at room temperature. As of the TiO2 film. Capacitance–voltage (C–V) and leakage current
obtained transparent solution was spin coated on oxygen plasma density–voltage (JG–V) characteristics were estimated using an
treated p-type Si (100) substrates at 4000 rpm for 30 s and then Agilent 4294A Precision Impedance Analyzer in the range of 1 kHz
dried on a hot plate at 100 °C for 5 min. The resistivity of Si wafer to 1 MHz with 5 mV AC excitation.
supplied by Micro Vision Corporation, USA was 1–10 Ω-cm. Prior
to the deposition of TiO2 thin films, wafer was cleaned with pir-
anha solution (H2SO4:H2O2 is 3:1) followed by 1% HF dipping for 3. Discussion
10 s to remove any native oxide layer and dried under nitrogen
ambience. Thus, the coated films were found to have poor adhe- 3.1. Structural characterization
sion to silicon substrate. To overcome this problem, the plasma
activation process was carried out in an oxygen environment for The films as deposited were amorphous and the calcination at
20 min at 0.008 mbar prior to the TiO2 film deposition. The films 200 °C did not induce any crystallization (Fig. 1a). The anatase
so obtained were subjected to annealing in the temperature range phase was formed at 400 °C, and stabilized even at high tem-
of 200–1000 °C for 1 h in a preheated horizontal furnace. The perature of 1000 °C. It is widely accepted that anatase to rutile
electrical investigation was carried out using metal-oxide-semi- phase transition occurs at a temperature 4700 °C. Such extreme
conductor (MOS) structure with Ag as a back contact (Fig. S1). The stability in our case is attributed to silicon substrate, which sup-
Al metal was evaporated to a thickness of 200 nm by using ther- presses rutile formation. The diffusion of Si4 þ at high temperature
mal evaporation method through a metal mask with a circular into TiO2 film substitutes Ti4 þ lattice sites and hinders structure
hole (d ¼288 m) on the samples as front contacts. reorganization into rutile nucleus [21]. The (101) reflection was the
only prominent peak and other diffraction planes in the pattern
2.2. Characterization could not be observed. The substrate peak changed its position
randomly during the thermal treatment. Fig. 1(b) shows the grain
The X-ray diffraction patterns were recorded over a 2θ range of size (D) dependence on annealing temperature. The grain size of
20–80° at a scan rate of 2° per min on a Rigaku X-ray dif- the films was calculated from the full width at half maximum
fractometer under Grazing Incidence X-ray Diffraction (GIXRD) intensity (β) of the prominent (101) peak using Debye–Scherrer's
mode. The operating voltage and current were 40 kV and 30 mA, equation
A. Kumar et al. / Materials Science in Semiconductor Processing 40 (2015) 77–83 79

Fig. 2. (a) Raman spectrum of as deposited and annealed TiO2 films at different annealing temperatures on Si substrate. (b) Raman spectrum of TiO2 films on Quartz
substrate annealed at 600 °C.

0.89λ lower energy side. This concludes that the defect concentration
D= ,
β cos θ (1) with Ti3 þ is extremely low. The difference of 5.79 eV between the
peaks indicates the presence of titanium in its tetravalent state
where λ is the wavelength of the X-rays and θ the Bragg diffraction [23,24]. The binding energy of Ti core levels remained constant up
angle. The average grain size showed only marginal increase from to the thermal treatment at 800 °C, and shifted to lower energy
15.8 to 18.8 nm with increase in thermal treatment up to 1000 °C. side of  467.8 and 457.8 eV respectively, at a higher temperature
The Raman spectra of TiO2 films calcined at various temperatures (1000 °C). This shift might be attributed to the vacancies created
are shown in Fig. 2a. It is obvious that all the prepared samples by oxygen because of partial breakage of Ti–O–Ti bonds in octa-
showed similar Raman spectra regardless of the temperature of hedral environment or lowering of valence state of Ti4 þ to Ti3 þ or
the treatment and could be ascribed to the allowed vibrational Ti2 þ accompanied by the formation of oxygen vacancies [25].
frequencies of the anatase form of titania. According to Factor Fig. 4(c) shows the O1s XPS core level spectra of 400 °C annealed
group analysis, anatase has six Raman active modes and as-deposited TiO2 films. The strong photoelectron signals
(A1g þ2B1g þ3Eg), three infrared active modes (A2u þ 2Eg) and one  530.0 eV is assigned to bulk lattice oxygen [25].
vibration of B2u, which is inactive in both infrared and Raman Fig. 5 shows refractive index (n) and porosity (P) dependence
spectra [22]. The film annealed at 400 °C showed intense band at on annealing temperatures. The porosity of the films was calcu-
143 cm  1 followed by weak bands at 396 and 637 cm  1, which lated using Eq. (2) [26].
were assigned to Eg (v6), B1g (v4), and Eg (v1) modes of anatase,
respectively. The position of these bands did not vary with n2 − 1
Porosity = 1 − ,
calcination temperature, although intensity was drastically re- nd2 − 1 (2)
duced as a result of reduced crystallinity. The results of Raman
analysis were in excellent agreement with XRD data. The promi- where n is the refractive index of the as deposited and annealed
nent band of anatase at 198 and 516 cm  1 could not be observed, films calculated from refractive index as a function of wavelength
as they overlap with Raman peaks of Si substrate. This fact was at 550 nm (Fig. S3), nd is the refractive index of pore free anatase
confirmed by the appearance of all the Raman peaks when TiO2 TiO2 with a refractive index of 2.52 [26,27]. It can be observed that
was deposited on glass substrate (Fig. 2b). the porosity decreases with increase in annealing temperature.
Figs. 3(a)–(f) show the AFM images of as deposited and an- Thus, the increase in refractive index is mainly due to the decrease
nealed TiO2 thin films, it can be observed that both the root mean in porosity and densification of the films in the course of anneal-
square (RMS) and average roughness are enhanced with annealing ing, which is in consistent with our AFM studies. It is to be pointed
temperature. The RMS roughness in the case of as-deposited film out here that in our studies, we observed a high value of refractive
is 0.16 nm, while it was 0.2, 0.43, 0.53, 0.98 and 2.43 nm for the index 2.48 for 1000 °C annealed films that is very close to the
films annealed at 200, 400, 600, 800 and 1000 °C, respectively anatase TiO2 pore free films.
(Fig. 3g). The increase in the RMS roughness is due to the change
in crystallinity as a result of aggregation of the smaller grains into 3.2. Electrical characterization
the larger grains during the course of annealing.
Fig. 4(a) shows XPS survey scan spectra of the as-deposited and 3.2.1. Capacitance–voltage characteristics
400 °C annealed TiO2 films. The XPS measurements were carried Fig. 6(a) shows the Capacitance–voltage (C–V) characteristics of
out to investigate the valence state of titanium as well as oxygen as deposited and annealed films at 1 MHz AC frequencies. The
vacancy, which are the major defect sites in TiO2. Fig. 4(b) shows schematic diagram of MOS device is shown in the inset of Fig. 6(a).
the Ti 2p XPS core level spectra of as-deposited and 400 °C an- The C–V shows proper accumulation, depletion and inversion re-
nealed films. The core level binding energies of Ti 2p3/2 and Ti 2p1/ gion with narrow hysteresis, while sweeping the voltage from  3
2 are  459.71 and 465.50 eV, respectively. The Ti 2p signals are to 2 and again back to  3. The films annealed at 400 °C showed
highly symmetric, and no shoulders were observed towards the the highest accumulation capacitance compared to other films
80 A. Kumar et al. / Materials Science in Semiconductor Processing 40 (2015) 77–83

Fig. 3. AFM images of as deposited and annealed TiO2 films; (a) as deposited; (b) 200 °C; (c) 400 °C; (d) 600 °C; (e) 800 °C; (f) 1000 °C (g): Root mean square (RMS)
roughness of the as-deposited and annealed TiO2 films.

annealed at various temperatures. The dielectric constant was permittivity of free space, and A is the area of top gate electrode of
calculated from accumulation capacitance using Eq. (3) [1] MOS structure. The flat band voltage (VFB) was   1.6 V for as
deposited films and shifted to  0.6 to 0.5 V for all other an-
Coxtox nealing temperatures. The negative flat-band voltage reflects the
k= ,
ε0A (3) positive charges at the interface of TiO2/Si [1]. Relatively large
where Cox is the accumulation capacitance of MOS structure, tox is changes of VFB for the annealed TiO2 films are due to reduction in
the thickness of TiO2 films incorporated in MOS structure, ε0 is the charge density of fixed oxide present in TiO2 films. It is also seen
A. Kumar et al. / Materials Science in Semiconductor Processing 40 (2015) 77–83 81

Fig. 4. (a) XPS survey scan spectra of the as-deposited and 400 °C annealed TiO2 films. (b) Ti 2p XPS core level spectra of 400 °C annealed and as-deposited TiO2 films. (c) O1s
XPS core level spectra of 400 °C annealed and as-deposited TiO2 films.

45 the latter case [28]. The dielectric constant was 34 at 1 MHz was
2.5
refractive index @ 550 nm estimated for 400 °C annealed films.
Porosity 40
Fig. 7(a) shows the frequency-dependent accumulation capa-
Refractive index @ 550 nm

2.4 35 citance of 400 °C annealed TiO2 films at AC frequencies from 1 kHz


Porosity (%)

30 to 1 MHz. The accumulation capacitance varies from 1141 to 410 pF


2.3 as the frequency varies from 1 kHz to 1 MHz. The dispersion in the
25
accumulation capacitance is observed. This kind of behavior is
2.2 20 frequently observed in non-native/ high-κ oxide based MOS ca-
15 pacitor due to the dielectric polarization and presence of interface
2.1 states at the interface of TiO2/Si [10,28]. At low frequencies the
10
interface states can respond and contribute to the total capaci-
2.0
5 tance, however, at high frequencies (1 MHz) the response of the
0 interface traps is relatively weak and cannot contribute to the
0 200 400 600 800 1000 accumulation capacitance [29]. Therefore, we observed a higher
Temperature ( 0C) accumulation capacitance for lower frequencies and vice versa.
Fig. 5. Refractive index (n) and porosity (P) dependence on annealing Fig. 7(b) shows the dielectric constant estimated from accu-
temperatures. mulation capacitance in Fig. 7(a) for 400 °C annealed films. The
dielectric constant decreased from 94 to 34 as the frequencies
that the hysteresis loops [change in flat band voltage (ΔVFB) varied from 1 kHz to 1 MHz. We observed a power-law depen-
during bidirectional C–V sweep)] in the case of as-deposited is dence of frequency on dielectric constant of TiO2 films annealed at
42 mV, while it was 18, 6, 6 and 12 for the films annealed at 200, 400 °C. This result follows the modeling as reported by A. Ghosh
400, 600, and 800 °C, respectively. A narrow hysteresis loop of the [30]. In this model, the hopping mechanism is liable for the di-
film annealed at 400–600 °C indicates a very small amount of electric polarization, and with the help of Kramers–Kronig (KK)
trapping charges within the films and interface of TiO2/Si. relations, it was verified that the power-law follows the AC-con-
Fig. 6(b) shows the dielectric constant dependence on anneal- ductivity of an insulator that leads its capacitance to the following
ing temperature. The dielectric constant increases up to 400 °C frequency dependence [31]:
and decreases for temperature beyond that. The crystallization of
anatase from amorphous was accounted in the former case, while C (ω) = ωs − 1, (4)
increase in the thickness of interlayer (SiOx) was responsible for where C is capacitance, and ω ¼2πf. The frequency exponent ‘s’
82 A. Kumar et al. / Materials Science in Semiconductor Processing 40 (2015) 77–83

Fig. 6. (a) C–V characteristics of the as-deposited and annealed TiO2 films (36 nm) and at AC frequency of 1 MHz. Inset: A schematic diagram of Al/TiO2/p-Si MOS structure.
(b) Annealing temperature dependent dielectric constants of thick TiO2 film (36 nm) at frequency of 1 MHz.

Fig. 7. (a) Frequency-dependent accumulation capacitance of 400 °C annealed TiO2 films (36 nm) and at AC frequencies from 1 kHz to 1 MHz. (b) Dielectric constants
estimated from accumulation capacitance.

was found to be 0.85 from the slope of linear fitting of log (C)¼
s log (ω). This value is very close as estimated by various research
groups [30–32].

3.2.2. Current–voltage characteristics


Fig. 8 shows the leakage current density of the 36 nm TiO2 films
as a function of applied voltage. The leakage current density de-
creased with annealing temperature up to 600 °C. This improve-
ment is the combined effect of increase in interlayer (SiOx) thick-
ness [we consider interlayer having a large conduction band offset
(ΔEC)] [1], and due to the densification of TiO2 films as supported
by AFM and ellipsometry studies (Figs. 3 and 4) [10]. The increased
leakage current for 800 °C annealed films may be due to the im-
proved crystallization and increased grain size, which might pro-
vide the conduction paths for carriers through the grain bound-
aries [32]. We observed a small leakage current density
9.4  10  7 A/cm2 at 1 V for our films annealed at 200 and 400 °C.
Fig. 8. Leakage current density of as-deposited and annealed TiO2 films (36 nm)
The lowest leakage current density 3.6  10  7 was observed for versus applied voltage.
600 °C annealed TiO2 films.
The Table 1 shows a comparison of the physical and electrical
properties of the anatase TiO2 films obtained by various methods very low leakage currents with excellent surface morphology. This
like DC sputtering, PECVD, ALD and LTAVD with the present work also highlight that the sol–gel spin-coating method can be a good
[33–36]. The results demonstrate that the TiO2 films obtained by alternative to other deposition techniques for fabrication of high-κ
the present protocol showed comparable dielectric constant and materials.
A. Kumar et al. / Materials Science in Semiconductor Processing 40 (2015) 77–83 83

Table 1
The comparison of the TiO2 thin films properties with their counterparts fabricated by different technique on Si substrate.

Deposition method Thickness (nm) Dielectric constant (k) Leakage current (at VG ¼ 1.5 V) A/cm2 RMS roughness (nm) Refractive index Ref.

DC sputtering 80–90 6–20 1.8  10  6–5.4  10  8 0.6–3.5 2.2–2.37 [33]


PECVD 87–105 13.7–18.4 – – 2–2.13 [34]
ALD  20 30 1.22  10  4 0.48–2.2 – [35]
LTAVD 25 36 3.2  10  4 – – [36]
Spin-coating 36 18–34 8.33  10  6–3.7  10  7 0.16–2.43 2.03–2.41 (Present work)

Note: PECVD – Plasma enhanced chemical vapor deposition; ALD – Atomic layer deposition; LTAVD – Low temperature arc vapor deposition.

4. Conclusions V. Thomas, Thin Solid Films 550 (2014) 121.


[6] B. Kınacı, S. Şebnem Çetin, A. Bengi, S. Özçelik, Mater. Sci. Semi. Proc. 15 (2012)
531.
Thermally stable TiO2 films on Si substrate were obtained by [7] K. Jiang, X. Ou, X.X. Lan, Z.Y. Cao, X.J. Liu, W. Lu, C.J. Gong, B. Xu, A.D. Li, Y.D. Xia,
combined sol gel spin coating technique. Good quality films having J. Yin, Z.G. Liu, Appl. Phys. Lett. 104 (2014) 263506.
anatase phase with low surface roughness and high dielectric [8] L.D. Zhang, H.F. Zhang, G.Z. Wang, C.M. Mo, Phys. Stat. Sol. (a) 483 (1996) 483.
[9] O. Carp, Prog. Solid State Chem. 32 (2004) 33.
constant of 34 at 1 MHz AC frequency were observed for films [10] W. Yang, C.A. Wolden, Thin Solid Films 515 (2006) 1708.
annealed at 400 °C having thickness 36 nm. The leakage current [11] H.Ã. Altuntas, A. Bengi, U. Aydemir, T. Asar, S.S. Cetin, I. Kars, S. Altindal,
densities were also influenced by annealing temperature; it was S. Ozcelik, Mater. Sci. Semi. Proc. 12 (2009) 224.
[12] S. Dutta, A. Pandey, O.P. Thakur, R. Pal, J. Vac. Sci. Technol. A 33 (2015) 021507.
3.6  10  7 A/cm2 at 1 V for 600 °C annealed films. The high value [13] P. Laha, S.S. Dahiwale, I. Banerjee, S.K. Pabi, D. Kimd, P.K. Barhai, V.
of dielectric constant ‘κ‘ obtained at low processing temperature is N. Bhoraskar, S.K. Mahapatra, Nucl. Instrum. Methods Phys. Res. B 269 (2011)
potentially useful in device applications, as it is in the range of 2740.
many other High-κ materials. The results thus obtained in our low [14] T. Nabatame, A. Ohi, T. Chikyo, M. Kimura, H. Yamada, T. Ohishi, J. Vac. Sci.
Technol. B 121 (2014) 3.
cost sol–gel spin-coating technique, are competitive to the films [15] S.G. Kumar, K.S.R.K. Rao, Nanoscale 6 (2014) 11574.
fabricated by expensive vacuum based techniques. [16] M. Kumar, D. Kumar, Micro. Eng. 87 (2010) 447.
[17] S. Aksoy, Y. Caglar, J. Alloys Compd. 613 (2014) 330.
[18] M.Z.R. Khan, D.G. Hasko, M.S.M. Saifullah, M.E. Welland, J. Phys: Condens.
Matter 215902 (2009) 215902.
Acknowledgment [19] N.B. Chaure, A.K. Ray, R. Capan, Semicond. Sci. Technol. 20 (2005) 788.
[20] R. Mechiakh, N. Ben Sedrine, M. Karyaoui, R. Chtourou, Appl. Surf. Sci. 257
(2011) 5529.
The authors would like to thank Mr. S.R.K. Chaitanya Indukuri, [21] K. Okada, N. Yamamoto, Y. Kameshima, A. Yasumori, K.J.D. Mackenzie, J. Am.
Department of Physics, IISc, for his useful discussion about the Ceram. Soc. 96 (2001) 1591.
work. Arvind would like to thank UGC, New Delhi for the research [22] G. Liu, H.G. Yang, C. Sun, L. Cheng, L. Wang, G.Q. Max Lu, H.-M. Cheng, Crys-
tEngComm 11 (2009) 2677.
fellowship. Dr. Girish acknowledges financial support from Kothari [23] J. Fu, S. Cao, J. Yu, J. Low, Y. Lei, Dalton Trans. 43 (2014) 9158.
Post-Doctoral fellowship cell (2012–2015), UGC, New Delhi, and [24] M.V. Sofianou, M. Tassi, V. Psycharis, N. Boukos, S. Thanos, T. Vaimakis, J. Yu,
CeNSE, Indian Institute of Science, India for characterization of the C. Trapalis, Appl. Catal. B: Environ. 162 (2015) 27.
[25] V. Etacheri, M.K. Serry, S.J. Hinder, S.C. Pillai, Adv. Funct. Mater. 21 (2011) 3744.
materials. [26] N.R. Mathews, R. Erik, Morales M.A. Corte ´s-Jacome, Sol. Energy 83 (2009)
1499.
[27] Y.Q. Hou, D.M. Zhuang, G. Zhang, M. Zhao, M.-S. Wu, Appl. Surf. Sci. 218 (2003)
98.
Appendix A. Supplementary material [28] B.H. Lee, Y. Jeon, K. Zawadzki, W.J. Qi, J. Lee, Appl. Phys. Lett. 74 (1999) 3143.
[29] A. Tataroǧ lu, Ş. Altındal, M.M. Bülbül, Micro Eng. 81 (2005) 140.
Supplementary data associated with this article can be found in [30] A. Ghosh, Phys. Rev. B 41 (1990) 1479.
[31] A. Dakhel, J. Alloys Compd. 422 (2006) 1.
the online version at http://dx.doi.org/10.1016/j.mssp.2015.06.073. [32] V.S. Dang, H. Parala, J.H. Kim, K. Xu, N.B. Srinivasan, E. Edengeiser, M. Havenith,
A.D. Wieck, T. de los Arcos, R.A. Fischer, A. Devi, Phys. Stat. Sol. (a), 211, (2014)
416.
[33] M. Chandra Sekhar, P. Kondaiah, S.V. Jagadeesh Chandra, G. Mohan Rao,
References
S. Uthanna, Surf. Interface Anal. 44 (2012) 1299.
[34] W. Gyu, S. Ihl, J. Choul, S. Han, Thin Solid Films 237 (1994) 105.
[1] G.D. Wilk, R.M. Wallace, J.M. Anthony, J. Appl. Phys. 89 (2001) 5243. [35] D. Wei, T. Hossain, N.Y. Garces, N. Nepal, H.M. Meyer, M.J. Kirkham, C.R. Eddy, J.
[2] J. Robertson, Rep. Prog. Phys. 69 (2006) 327. H. Edgar, ECS J. Solid State Sci. Technol. 2 (2013) N110.
[3] K. Lv, B. Cheng, J. Yu, G. Liu, Phys. Chem. Chem. Phys. 14 (2012) 5349. [36] K. Shubham, P. Chakrabarti, Electron. Mater. Lett. 10 (2014) 579.
[4] J. Park, K.P. Biju, S. Jung, W. Lee, J. Lee, S. Kim, S. Park, J. Shin, H. Hwang, IEEE
Electron Device Lett. 32 (2011) 476.
[5] P.B. Nair, V.B. Justinvictor, G.P. Daniel, K. Joy, V. Ramakrishnan, D. Devraj, P.

You might also like