Study of the structural, thermal, optical, electrical and nanomechanical properties of sputtered vanadium oxide smart thin films

Deeksha Porwal ac, A. Carmel Mary Estherad, I. Neelakanta Reddyb, N. Sridharaa, Nagendra Prasad Yadavc, Dinesh Rangappad, Parthasarathi Berae, Chinnasamy Anandane, Anand Kumar Sharmaa and Arjun Dey*a
aThermal Systems Group, ISRO Satellite Centre, Bangalore 560017, India. E-mail: arjundey@isac.gov.in; arjun_dey@rediffmail.com; Fax: +91 80 2508 3203; Tel: +91 80 2508 3214
bCentre for Nanoscience and Nanotechnology, Sathyabama University, Chennai 600119, India
cBundelkhand Institute of Engineering and Technology, Jhansi 284128, India
dCenter for Nanotechnology, Center for Post Graduate studies, Visvesvaraya Technological University, Belgaum 590018, India
eSurface Engineering Division, CSIR–National Aerospace Laboratories, Bangalore 560017, India

Received 3rd February 2015 , Accepted 18th March 2015

First published on 18th March 2015


Abstract

Vanadium oxide thin films were grown on both quartz and Si(111) substrates, utilizing a pulsed RF magnetron sputtering technique at room temperature with the RF powers at 100 W to 700 W. The corresponding thicknesses of the films were increased from 27.5 nm to 243 nm and 21 nm to 211 nm as the RF power was increased from 100 W to 700 W for the quartz and silicon substrates, respectively. X-ray diffraction and field emission scanning electron microscopy were carried out to investigate the phase and surface morphology of the deposited films. The electronic structure and the vanadium oxidation states of the deposited films were investigated thoroughly by X-ray photoelectron spectroscopy. The as-grown films show only stoichiometric vanadium oxide, where vanadium is in V5+ and V4+ states. The phase transitions of the vanadium oxide films were investigated by the differential scanning calorimetric technique. The reversible i.e. smart transition was observed in the region from 337 °C to 343 °C. The average hemispherical infrared emittance of the deposited vanadium oxide films was evaluated by an emissometer in the wavelength range of 3 μm to 30 μm. The sheet resistance of the deposited films was measured by a two-probe method and the data were in the range of 106 to 105 Ω per square. The optical properties of the films, such as solar transmittance, solar reflectance and solar absorptance, as well as optical constants e.g. optical band gap, were also evaluated. Finally, mechanical properties such as the hardness and the Young’s modulus at the microstructural length scale were evaluated by employing a nanoindentation technique with a continuous stiffness mode.


1. Introduction

Vanadium oxide shows a drastic reversible change in the solar transmittance, infrared (IR) emittance and electrical resistance with an alteration in either the temperature (i.e., thermochromic behaviour) or the potential difference (i.e., electrochromic behaviour). This reversible change in properties occurs primarily due to a metal-to-insulator transition (MIT) and was first reported by Mott1,2 which is commonly referred to as a smart transition. According to Mott,1,2 vanadium oxide undergoes a reversible change in the crystal structure (i.e., an insulator to a metallic state) which leads to a change in the energy band gap while increasing and subsequently decreasing in temperature above or below its transition point. At the transition point, a lattice distortion takes place in vanadium oxide and thereby leads to an energy gap separation from the empty state. Furthermore, an increase in the temperature diminishes the lattice distortion due to the excitation of electrons which results in a reduction in the energy gap and hence it acts as a metallic state.

Among all oxides of vanadium, VO2 and V2O5 possess a phase transition in a positive temperature range, while remaining oxides show transitions at the sub-zero temperature range.3 Due to this, VO2 and V2O5 are well studied in the literature in comparison with the other oxides of vanadium.3,4 The research interest of these oxides has been increased in the last few years due to their potential application in a wide variety of optical modulation devices.5 The smart reversible MIT makes vanadium oxide a promising material for various applications, e.g. intelligent energy conserving window coatings,6 IR light switching or bolometric devices,7 electrochromic devices,6,7 colour memory devices5 and variable emitting smart surfaces for spacecraft applications.8,9

Several techniques have been utilized to deposit vanadium oxide films such as sputtering,10–14 pulsed laser deposition,15,16 thermal evaporation,17–20 electron beam evaporation,21 spray pyrolysis22,23 and sol–gel method.24–26 The sputtering technique is more advantageous than the other methods due to its high deposition rate and good uniformity.13 Study of the systematic variation in the RF power during the sputtering of vanadium oxide, and further detail on the structural, thermo-optical and electrical behaviours have not yet been reported in the literature. Moreover, the phase transition behaviour of vanadium oxide thin films/coatings has not been attempted so far, by the calorimetric technique in particular; although the phase transition behaviour of both VO227 and V2O528,29 either in powder form27 or nanostructured powder dispersed in solution28 has been studied. In addition, works related to the mechanical properties of vanadium oxide films/coatings, especially at the microstructural length scale, are scarce in the literature and the reported hardness and Young’s modulus values are unexpectedly scattered and ambiguous.30,31

Hence, in the present work, the development and systematic study of vanadium oxide thin films has been attempted. Vanadium oxide films were deposited on quartz glass and Si(111) substrates by pulsed RF magnetron sputtering by varying the RF power from 100 W to 700 W. Furthermore, the microstructure, electronic structure, phase transition behavior, electrical properties, thermo-optical, and nanomechanical properties were investigated in detail.

2. Materials and methods

A pulsed RF magnetron sputtering system (SD20, Scientific Vacuum Systems, UK) was utilized to deposit the vanadium oxide thin films on quartz glass and Si(111) substrates at room temperature. The quartz substrate (40 mm × 40 mm × 0.2 mm) was obtained from Astro Optics, India and boron doped p type Si(111) (diameter: 125 mm and thickness: 0.6 mm) was procured from Silicon Valley Microelectronics Inc., USA. The V2O5 target (99.999%, diameter: 200 mm and thickness: 3 mm, Vin Karola Instruments, USA) was used in the present study. The distance between the target and the substrate was kept at 140 mm. The RF power during the film deposition was varied from 100 W to 700 W, with a constant increment of 100 W, while the duration of the deposition time was kept constant at 1 h. The pulse frequency was 100 Hz with a duty cycle of 57%. The vacuum chamber was evacuated to a pressure of 5 × 10−6 mbar prior to the deposition, utilizing a combination of both rotary (i.e. roughing) and turbo molecular pumps. However, the working pressure was set as a constant 1.5 × 10−2 mbar after introducing ultra pure argon gas. Pre-sputtering was carried out for 10 min prior to the deposition of films to reduce the contamination.

The thicknesses of the deposited films were measured using a surface profilometer (Nanomap 500 LS 3D, USA). The phase analysis of the deposited films was investigated by the X-ray diffraction (XRD) technique using a commercial diffractometer (X’pert Pro, Philips, The Netherlands). Cu Kα1 radiation was used at a glancing incident angle of 2° with a very slow step size of 0.03°. The surface topography of the deposited thin films was observed by field emission scanning electron microscopy (FESEM:SupraVP35 Carl Zeiss, Germany). The energy dispersive X-ray (EDX:X-Max, USA) spectra of the deposited films were acquired utilizing a customary unit (Oxford Instruments, UK) attached to the FESEM. XPS of the deposited films on the silicon substrate were recorded with a SPECS spectrometer using non-monochromatic Al Kα radiation (1486.6 eV) as an X-ray source run at 150 W (12 kV, 12.5 mA). The binding energies reported here were referenced with the O 1s peak at 530.0 eV.32 All survey spectra were obtained with a pass energy of 70 eV with a step increment of 0.5 eV and the individual spectra were recorded with a pass energy and a step increment of 25 and 0.05 eV, respectively. V 2p and O 1s core level spectra were curve-fitted into their several components with Gaussian–Lorentzian peaks after Shirley background subtraction using the CasaXPS program.

The phase transition temperatures of the vanadium oxide films were investigated by employing the differential scanning calorimetry (DSC) technique (Q100, TA Instruments, USA) in helium environment. The average hemispherical IR emittance (εIR-H) of the deposited vanadium oxide films was measured by an emissometer (AE, Devices and Services Co., USA.) in the wavelength range of 3 μm to 30 μm using both high and low emitting standard surfaces as per ASTM C1371-04a.

The sheet resistances (Rs) of the deposited vanadium oxide films were measured by the two-probe resistance meter (Trek Model 152-1, Trek Inc., USA) as per ASTM D 257-9.

The optical properties of the vanadium oxide thin films deposited on the quartz substrate were obtained by the UV-VIS-NIR spectrophotometer (Cary 5000, Agilent Technologies, USA) in the entire solar region (i.e. 200 nm to 2.3 μm) of the spectral window. The average solar absorptance (αs) was evaluated by a solar spectrum reflectometer (SSR-E, Devices and Services Co., USA) as per ASTM C1549-09. The absorption coefficient (α) of the vanadium oxide films were calculated from the transmittance data received from a UV-VIS-NIR spectrophotometer using the following relations (eqn (1) to (3)). Furthermore, a ‘Tauc extrapolation plot’ was utilized to evaluate the optical band gap of the deposited films. The ‘α’ of the film can be represented by the following, eqn (1):33,34

 
αhν = A(EiE0)m (1)
where, E0 and Ei are the initial energy of the photon and the energy of the incident photon, respectively, while the proportionality constant and Planck’s constant are represented as A and h, respectively. α and ν are known as the absorption coefficient and frequency, respectively. The magnitude of m is 1/2 for direct optical band gap.34 The quantities α and the total transmission (T) are correlated by the following relations (eqn (2) and(3)):
 
T = eαt (2)

or

 
α = (ln[thin space (1/6-em)]T)/t (3)
where, ‘t’ is the thickness of the deposited film.

Finally, the nanomechanical properties, e.g., nanohardness (H) and Young’s modulus (E), of the vanadium oxide films were evaluated by nanoindentation (G200, MTS-Agilent technologies, USA) technique with a continuous stiffness measurement (CSM) mode. The Berkovich diamond indenter with a tip radius of ∼20 nm was used in the present study. A thick (∼3.8 μm) vanadium oxide coating deposited at RF power of 700 W on the silicon substrate was judiciously utilized in the present nanoindentation study to avoid substrate influence. Generally, a recommended depth of penetration for the indenter is ∼10% of the film thickness during conducting nanoindentation experiment. The slackness can also be attributed to the influence of the mechanical properties of the substrate.35 Therefore, to avoid the substrate’s mechanical properties, the final depth of penetration was wisely chosen as 300 nm, which is well below 10% of the film thickness (i.e. ∼3.8 μm). Furthermore, a CSM mode was employed in the present study to investigate any depth dependence of the mechanical properties in the deposited vanadium oxide films.

3. Results and discussion

The thicknesses of the deposited films, measured by a nanoprofilometric technique, are 27.5 nm, 40 nm, 102 nm, 145 nm, 200 nm, 230 nm and 243 nm for the quartz substrate and 21 nm, 36 nm, 89 nm, 133 nm, 156 nm, 201 nm, and 211 nm for the silicon substrate as the RF power is increased from 100 W to 700 W with a constant increment of 100 W (Fig. 1). The increase in RF power leads to an increase in the deposition rate as the deposition duration is kept constant, which ultimately results in an increase in film thickness. A marginally lower thickness in deposited films is observed on the silicon substrates than on the quartz substrates.
image file: c5ra02092a-f1.tif
Fig. 1 The variation of deposited film thicknesses on quartz and Si(111) substrates as a function of RF power.

All as-grown films are confirmed to be amorphous in nature by XRD analysis. The RF sputtering technique often offers amorphous vanadium oxide film deposited at room temperature.36,37 Crystallinity can be achieved at elevated substrate temperatures,38–41 by post annealing36,40–43 or an increase in the oxygen partial pressure.37,41

The FESEM photomicrographs of the thin films deposited on a silicon substrate with thicknesses of 21 nm, 156 nm, and 211 nm grown at 100 W, 500 W and 700 W, respectively, for a constant duration of 1 h are shown in Fig. 2(a–c), respectively. Featureless, smooth and compact surfaces are observed in Fig. 2(a–c). The corresponding EDX spectra are appended in the insets of Fig. 2(a–c). Only the presence of V and O are observed in the corresponding EDX spectra. However, the presence of the substrate peak is also prominent as the thickness of the film is very less.34 The intensities of the vanadium and oxygen peaks are increased as the thickness of the film increased, as expected.


image file: c5ra02092a-f2.tif
Fig. 2 Typical FESEM photomicrographs of vanadium oxide thin films grown on a silicon substrate at different RF powers e.g., (a) 100 W, (b) 500 W and (c) 700 W. (Insets: corresponding EDX spectra.)

Fig. 3a shows typical XPS survey spectra of vanadium oxide thin films grown on silicon substrates at 100 W and 600 W. It clearly shows the presence of V and O species in the deposited films. Typical XPS spectra of the V 2p core levels in the vanadium oxide films deposited with different powers are shown in Fig. 3b. Both the V 2p and O 1s core level spectra are shown together in Fig. 3b as the V 2p and O 1s core level regions are close. No significant alteration of the characteristic peaks in terms of position and intensity are observed with the variation of the RF power. The spectral envelopes of the V 2p core levels of the as-grown films contain a long tail in the lower binding energy region, along with the main peak, indicating that V is present in different oxidation states and it can be curve-fitted into sets of spin–orbit doublets. Accordingly, the observed V 2p3/2 peaks at 516.2 and 517.2 eV in all films are assigned to V4+ (VO2) and V5+ (V2O5) species. Three component peaks are fitted in the O 1s core level region (Fig. 3c). The component peak at 530.1 is associated with the oxide species related to the vanadium oxides, whereas the peaks at 531.5 and 533.1 eV correspond to adsorbed oxygen and water, respectively, present in the films.44 The peak areas (A) of the V5+ and V4+ components are used to estimate their relative concentrations (C) in the films using the following equation:

 
image file: c5ra02092a-t1.tif(4)
According to eqn (4), the concentration of V4+ in the film deposited with a RF power of 100 W is estimated to be 20% with respect to the total amount of V species. The binding energies and relative surface concentrations of the different V species, as obtained from the V 2p core levels of the vanadium oxide films grown at different RF powers, are summarized in Table 1. It is observed from Table 1 that the concentrations of the V4+ species are in the range of 18% to 25%, with the remaining amount from V5+ species. It is to be noted that the characteristic binding energies of the V 2p3/2 core level peaks for V2+, V3+, V4+ and V5+ are reported as 513.5–513.7 eV,45–48 515.15–515.7 eV,45,48–51 515.8–516.2 eV45,48–51 and 516.9–517.3 eV,45,49–51 respectively. In the present study, the fitted curves clearly show that the characteristic peaks correspond to V4+ and V5+ possessing binding energies of V 2p3/2 of 516.1–516.2 eV and 517.1–517.3 eV, respectively (Table 1).


image file: c5ra02092a-f3.tif
Fig. 3 (a) Typical XPS survey spectra of the vanadium oxide thin films deposited at RF powers of 100 W and 600 W. (b) V 2p and O 1s core level spectra of vanadium oxide thin films deposited with different RF powers. (c) A typical curve-fitted V 2p and O 1s core level spectrum of a vanadium oxide thin film deposited at 600 W.
Table 1 Binding energies and relative peak areas of V species evaluated from XPS studies of the vanadium oxide thin films
RF power (W) V species Binding energy of V 2p3/2 (eV) Relative peak area (%)
100 V4+ 516.1 20
V5+ 517.1 80
200 V4+ 516.1 18
V5+ 517.2 82
300 V4+ 516.1 25
V5+ 517.3 75
400 V4+ 516.0 25
V5+ 517.2 75
500 V4+ 516.0 25
V5+ 517.3 75
600 V4+ 516.2 20
V5+ 517.2 80
700 V4+ 516.1 25
V5+ 517.1 75


The present XPS data demonstrate that there is insignificant variation in the electronic structure and oxidation state of vanadium oxide as a function of the RF power. Pure phases of V2O5 and VO2 generated by the sputtering technique in particular are seldom reported in the literature.42,52 In general, sputtered vanadium oxide films showed mixed phases with one or more vanadium species that includes VO2, V2O5, V3O7, V4O7, and V6O13.53,54 However, in the present study, no other V species were found except V4+ and V5+.

The variation in the derivative heat flow of bare quartz and the vanadium oxide thin films grown with 100 W–700 W on quartz glass as a function of the temperature is shown in Fig. 4 to investigate the phase transition of the present vanadium oxide films. Both in heating and cooling cycles, no signature of a phase transition is observed for the bare quartz glass as expected. However, the vanadium oxide thin films show prominent peaks appearing in both heating and cooling cycles around the region from 337 °C to 343 °C which indicates the phase transition temperature. Furthermore, in both the heating and cooling cycles, the curves of the derivative heat flow follow almost the same path which implies that the hysteresis loss is insignificant and proves a reversible or smart characteristic transition. In the present study, the transition temperatures remain almost constant despite the increase in RF power. In other words, the phase transition temperatures of the vanadium oxide films vary insignificantly as a function of the RF power. This observation is corroborated with the XPS investigation, where it is seen that the oxidation states of vanadium oxides are altered insignificantly with increases in the RF power, as summarized in Table 1.


image file: c5ra02092a-f4.tif
Fig. 4 The variation of the derivative heat flow of the bare quartz and the vanadium oxide thin films on quartz as a function of the temperature.

DSC studies of V2O5 thin films are not yet reported in literature. However, a DSC investigation of nano-structured V2O5 powder formed via the thermal decomposition of vanadyl oxalate (VOC2O4·nH2O) indicates that the phase transition occurs in the range of 267 °C to 353 °C, giving an endothermic peak.28 Three exothermic peaks at 183 °C, 261 °C and 418 °C are observed which are attributed to the evaporation of absorbed organic compounds on the V2O5 foam surface, decomposition of the organic molecules (e.g. hexadecylamine) intercalated among the V2O5 layers and recrystallization of the V2O5 foam, respectively.29 The phase transition of β-V2O5 powder heated in an oxygen atmosphere is reported in the range of 220 °C to 425 °C, as studied through DSC.55 V2O5 nanobelts grown by a xerogel process show a transition at ∼465.6 °C measured by the DSC technique.56 In addition, V2O5 powder and V2O5 powder dissolved in H2O2 show transitions in the range of 239 °C to 344 °C57 and 357 °C,58 respectively. Hence, the aforesaid DSC studies on powder V2O5 samples reveal that the range for the transition temperature of V2O5 is widely scattered and depends on the processing technique of V2O5 as well as its phase purity.

The variation of εIR-H (using both high and low emitting standard surfaces) for the vanadium oxide thin films on a quartz glass substrate as a function of the RF power is shown in Fig. 5. The εIR-H data of bare quartz is around 0.78, which is not significantly altered after being coated with vanadium oxide. Furthermore, no significant change in εIR-H is observed when it is measured with a low or high emitting standard surface. This data indicates that the deposited vanadium oxide is transparent in the IR region, which agrees well with the literature.59


image file: c5ra02092a-f5.tif
Fig. 5 The variation in the average hemispherical IR emittance (εIR-H) of the vanadium oxide thin films on a quartz substrate as a function of RF power.

The variation of Rs values is insignificantly altered with increases in the RF power. The values of Rs are in the range of 106 to 105 Ω per square which agrees well with Rs data reported for vanadium oxide films.60

The solar transmittance spectra of the different vanadium oxide thin films deposited on a quartz substrate at 100 W to 700 W as a function of wavelength in the entire solar regime of the spectral window are shown in Fig. 6a. Transmittance spectra shown in Fig. 6a is further blown up and shown in Fig. 6b–d for the ultra violet (UV: 200–340 nm), visible (VIS: 340–780 nm) and near infrared (NIR: 780–2300 nm) regions, respectively. The bare quartz glass shows an almost constant 94% transmittance value. The transmittance of the vanadium oxide film decreases as the thickness of the film is increased from 27.5 nm to 243 nm due to the increase in the RF power. In the UV regime, the transmittance decreases considerably from ∼75% to ∼1%. In the VIS regime, a similar trend is also observed, however the transmittance value is not constant throughout the VIS regime. The corresponding reflectance spectra are also shown in Fig. 7. The reflectance is found to increase with increases in the thickness of the film, particularly in the NIR regime except for the film deposited at 700 W.


image file: c5ra02092a-f6.tif
Fig. 6 (a) The transmittance spectra of vanadium oxide thin films on quartz substrates at 200–2300 nm of the spectral window, blown up view of (a): (b) UV region (200–340 nm), (c) VIS region (340–780 nm) and (d) NIR region (780–2300 nm).

image file: c5ra02092a-f7.tif
Fig. 7 The reflectance spectra recorded in 200–2300 nm of the spectral window for vanadium oxide thin films deposited on quartz surfaces at 100–700 W.

The decrement of the average transmittance value in the NIR regime is observed from ∼0.2% to ∼30% when the thickness of the film increases from only 27.5 nm to 243 nm (Fig. 8a). The variation of αs as a function of the RF power is shown in Fig. 8b. An increase in αs is observed with an increase in the thickness of the film as expected. The phenomenon of the decrease in transmittance with the increase in the thickness of the vanadium oxide films, while the reverse trend observed for the reflectance is well studied in literature.61,62 In the present work, a significant decrease in the transmittance is observed in the UV, VIS and NIR regions with an increase in the thickness, which is only in a nanometric range i.e., from ∼27.5 nm to ∼243 nm. Thus, tuning the transmittance of the vanadium oxide film can be achieved for the desired application by altering only the thickness without a trade-off with its structural phase, electronic configuration, and phase transition behaviour.


image file: c5ra02092a-f8.tif
Fig. 8 (a) Average transmittance in the NIR region with respect to bare quartz. (b) Average solar absorptance (αs) of vanadium oxide thin films coated on quartz as a function of the RF power.

The variation of α as a function of the photon energy, calculated for optical band gap, is shown in Fig. 9. The optical band gap can be represented by the value of energy for which α is equal to zero. Thus, the linear portions of the corresponding absorption coefficient data plot (Fig. 9) are extrapolated to intersect the energy axis (i.e.x’ axis) at α = 0. The corresponding values of the energy data give the optical band gaps of the vanadium oxide films. The optical band gaps of the vanadium oxide films grown at 100–700 W are summarized in Table 2. The optical band gap data is in the range of 2.4 eV to 2.8 eV. The present optical band gap data agree well with the data reported in the literature.63


image file: c5ra02092a-f9.tif
Fig. 9 The variation of the absorption coefficient (α) as a function of the photon energy of the vanadium oxide films on quartz substrates developed at 700 W for the calculation of optical band gaps.
Table 2 Optical band gaps of the vanadium oxide thin films
RF power (W) Optical band gap (eV)
100 2.8
200 2.4
300 2.4
400 2.6
500 2.5
600 2.6
700 2.6


The variations of H and E as a function of depth are shown in Fig. 10(a) and (b), respectively. Both the H and E values are almost constant at ∼0.2 GPa and 8.5 GPa, respectively. Large variations of both the hardness and the modulus of vanadium oxide films are reported in literature.30,31 For instance, with increase in the deposition temperature, the V2O5 films by DC sputtering show a transition from amorphous at room temperature to a polycrystalline growth with a preferred (200) orientation, as demonstrated by Fateh and coworkers.30 This phenomenon leads to an increase in the H and E values of the films from 3.2 ± 0.1 GPa and 79.4 ± 3.2 GPa at 26 °C to 4.8 ± 0.6 GPa and 129.2 ± 6.4 GPa at 300 °C, respectively. Furthermore, Zhu et al. show an ambiguous variation of the E value of as-grown V2O5 nanobelts between 5.6 to 98 GPa.31 They suggest that such scattered values are attributed to the different amounts of water molecules in the nanobelts. However, after annealing in vacuum, the nanobelts are converted to a polycrystalline α-V2O5 phase, which show a consistent value of the modulus at ∼28.9 GPa. Thus, it is clearly observed that both hardness and modulus values of vanadium oxide films depend upon the crystallinity, growth axis and annealing temperature.30,31 In the present study, comparatively lower data for the hardness and modulus are found, possibly due to lack of crystallinity of the vanadium oxide films, as discussed earlier. It is also important to note that the measured nanoindentation data depicts that both hardness and modulus values vary insignificantly as a function of the depth.


image file: c5ra02092a-f10.tif
Fig. 10 (a) The nanohardness and (b) the modulus of the vanadium oxide films as a function of depth.

4. Conclusions

The pulsed RF magnetron sputtering technique was utilized to grow amorphous, featureless and dense vanadium oxide thin films on both quartz glass and silicon substrates by varying the RF powers from 100 W to 700 W to give a film thickness in the range of ∼21 nm to ∼243 nm. The deposited films only show a major presence of V5+ (75–82%) and a minor presence of V4+ (18–25%), without any other species of vanadium. The reversible, i.e., smart, transition of vanadium oxide films is observed in the temperature range of 337 °C to 343 °C. The present XPS data indicate that there is insignificant variation in the electronic structure and vanadium oxidation states as a function of the RF power and further DSC results also support this. The hemispherical IR emittance of the quartz glass is not altered after deposition of vanadium oxide. Furthermore, the sheet resistance value of the vanadium oxide films is found to be in the range of 106 to 105 Ω per square. A significant decrease in the solar transmittance and increase in the solar absorptance of the vanadium oxide films on quartz is observed when the thickness increases from 27.5 nm to 243 nm. The optical band gap of the vanadium oxide films is found to be in the range of 2.4 eV to 2.8 eV. Finally, comparatively lower nanohardness and modulus values of vanadium oxide are measured as ∼0.2 GPa and ∼8.5 GPa, respectively, which is plausibly due to the lack of crystallinity of the deposited vanadium oxide films. In addition, the present nanohardness and modulus data are found as almost independent on depth.

References

  1. N. F. Mott, Rev. Mod. Phys., 1968, 40, 677 CrossRef CAS.
  2. S. N. Mott, Phys. Today, 1978, 31, 42 CrossRef PubMed.
  3. P. Kiri, G. Hyett and R. Binions, Adv. Mater. Lett., 2010, 1, 86 CrossRef CAS PubMed.
  4. A. L. Pergament, G. B. Stefanovich, N. A. Kuldin and A. A. Velichko, ISRN Condens. Matter Phys., 2013, 960627 Search PubMed.
  5. L. J. Meng, R. A. Silva, H. N. Cui, V. Teixeira, M. P. dos Santos and Z. Xu, Thin Solid Films, 2006, 515, 195 CrossRef CAS PubMed.
  6. M. A. Sobhan, M. R. Islam and K. A. Khan, Appl. Energy, 1999, 64, 345 CrossRef CAS.
  7. R. Mustafa Oksuzoglu, P. Bilgiç, M. Yildirim and O. Deniz, Opt. Laser Technol., 2013, 48, 102 CrossRef CAS PubMed.
  8. R. V. Druzelecky, E. Haddad, W. Jamroz, M. Soltani and M. Chaker, Proceeding of 2003 SAE Conference, 2003-01-2472, Vancouver, BC, Canada, 2003 Search PubMed.
  9. M. Benkahoul, M. Chaker, J. Margot, E. Haddad, R. Kruzelecky, B. Wong, W. Jamroz and P. Poinas, Sol. Energy Mater. Sol. Cells, 2011, 95, 3504 CrossRef CAS PubMed.
  10. X. Chen and J. Dai, Optik, 2010, 121, 1529 CrossRef CAS PubMed.
  11. Q. Jiang, Y. Li, S. Hu, B. Wu, X. Yu and H. Wang, Proceeding of SPIE 7279, Photonics and Optoelectronics Meetings POEM 2008, Optoelectronic Devices and Integration, 2009.
  12. E. Cazzanelli, G. Mariotto, S. Passerini, W. H. Smyrl and A. Gorenstein, Sol. Energy Mater. Sol. Cells, 1999, 56, 249 CrossRef CAS.
  13. M. Jiang, X. Cao, S. Bao, H. Zhou and P. Jin, Thin Solid Films, 2014, 562, 314 CrossRef CAS PubMed.
  14. Y. Lv, M. Hu, M. Wu and Z. Liu, Surf. Coat. Technol., 2007, 201, 4969 CrossRef CAS PubMed.
  15. C. V. Ramana, R. J. Smith, O. M. Hussain, M. Massot and C. M. Julien, Surf. Interface Anal., 2005, 37, 406 CrossRef CAS PubMed.
  16. M. Maaza, K. Bouziane, J. Maritz, D. S. McLachlan, R. Swanepool, J. M. Frigerio and M. Every, Opt. Mater., 2000, 15, 41 CrossRef CAS.
  17. R. Santos, J. Loureiro, A. Nogueira, E. Elangovan, J. V. Pinto, J. P. Veiga, T. Busani, E. Fortunato, R. Martins and I. Ferreira, Appl. Surf. Sci., 2013, 282, 590 CrossRef CAS PubMed.
  18. L. Derbali and H. Ezzaouia, Sol. Energy, 2012, 86, 1504 CrossRef CAS PubMed.
  19. X. Wu, F. Lai, L. Lin, Y. Li, L. Lin, Y. Qu and Z. Huang, Appl. Surf. Sci., 2008, 255, 2840 CrossRef CAS PubMed.
  20. A. Axelevitch, B. Gorenstein and G. Golan, Microelectron. Reliab., 2011, 51, 2119 CrossRef CAS PubMed.
  21. C. V. Ramana, O. M. Hussain, B. S. Naidu, C. Julien and M. Balkanski, Mater. Sci. Eng., B, 1998, 52, 32 CrossRef.
  22. A. Ashour and N. Z. El-Sayed, J. Optoelectron. Adv. Mater., 2009, 11, 251 CAS.
  23. A. Bouzidi, N. Benramdane, S. Bresson, C. Mathieu, R. Desfeux and M. E. Marssi, Vib. Spectrosc., 2011, 57, 182 CrossRef CAS PubMed.
  24. D. Q. Liu, W. W. Zheng, H. F. Cheng and H. T. Liu, Adv. Mater. Res., 2009, 79–82, 747 CrossRef CAS.
  25. J. Wu, W. Huang, Q. Shi, J. Cai, D. Zhao, Y. Zhang and J. Yan, Appl. Surf. Sci., 2013, 268, 556 CrossRef CAS PubMed.
  26. H. K. Chen, H. C. Hung, T. C. K. Yang and S. F. Wang, J. Non-Cryst. Solids, 2004, 347, 138 CrossRef CAS PubMed.
  27. Y. Zhang, M. Fan, M. Zhou, C. Huang, C. Chen, Y. Cao, G. Xie, H. Li and X. Liu, Bull. Mater. Sci., 2012, 35, 369 CrossRef CAS.
  28. A. Pan, J. G. Zhang, Z. Nie, G. Cao, B. W. Arey, G. Li, S. Q. Liang and J. Liu, J. Mater. Chem., 2010, 20, 9193 RSC.
  29. C. H. Kim, Y. J. Yun, B. H. Kim, W. G. Hong, Y. Y. Kim, W. L. Jang, N. E. Lee and H. Y. Yu, J. Nanomater., 2013, 807895 Search PubMed.
  30. N. Fateh, G. A. Fontalvo and C. Mitterer, J. Phys. D: Appl. Phys., 2007, 40, 7716 CrossRef CAS.
  31. Y. Zhu, Y. Zhang, L. Dai, F. C. Cheong, V. Tan, C. H. Sow and C. T. Lim, Acta Mater., 2010, 58, 415 CrossRef CAS PubMed.
  32. G. Silversmit, D. Depla, H. Poelman, G. B. Marin and R. De Gryse, Surf. Sci., 2006, 600, 3512 CrossRef CAS PubMed.
  33. P. Kumar, M. K. Wiedmann, C. H. Winter and I. Avrutsky, Appl. Opt., 2009, 48, 5407 CrossRef CAS PubMed.
  34. I. N. Reddy, V. R. Reddy, N. Sridhara, V. S. Rao, M. Bhattacharya, P. Bandyopadhyay, S. Basavaraja, A. K. Mukhopadhyay, A. K. Sharma and A. Dey, Ceram. Int., 2014, 40, 9571 CrossRef CAS PubMed.
  35. E. Uchaker, Y. Z. Zheng, S. Li, S. L. Candelaria, S. Hu and G. Z. Cao, J. Mater. Chem. A, 2014, 2, 18208 CAS.
  36. C. R. Cho, S. Cho, S. Vadim, R. Jung and I. Yoo, Thin Solid Films, 2006, 495, 375 CrossRef CAS PubMed.
  37. M. Benmoussa, E. Ibnouelghazi, A. Bennouna and E. L. Ameziane, Thin Solid Films, 1995, 265, 22 CrossRef CAS.
  38. S. C. Mui, J. Jasinski, V. J. Leppert, M. Mitome, D. R. Sadoway and A. M. Mayes, J. Electrochem. Soc., 2006, 153, A1372 CrossRef CAS PubMed.
  39. S. Saitzek, F. Guinneton, G. Guirleo, L. Sauques, K. Aguir and J.-R. Gavarri, Thin Solid Films, 2008, 516, 891 CrossRef CAS PubMed.
  40. Y. Zhang, R. Wang, Z. Qiu, X. Wu and Y. Li, Mater. Lett., 2014, 131, 42 CrossRef CAS PubMed.
  41. J. W. Lee, S. R. Min, H. N. Cho and C. W. Chung, Thin Solid Films, 2007, 515, 7740 CrossRef CAS PubMed.
  42. A. A. Akl, Appl. Surf. Sci., 2007, 253, 7094 CrossRef CAS PubMed.
  43. M. I. Kang, I. K. Kim, E. J. Oh, S. W. Kim, J. W. Ryu and H. Y. Park, Thin Solid Films, 2012, 520, 2368 CrossRef CAS PubMed.
  44. J. Liqiang, W. Baiqi, X. Baifu, L. Shudan, S. Keying, C. Weimin and F. Honggang, J. Solid State Chem., 2004, 177, 4221 CrossRef PubMed.
  45. N. Alov, D. Kutsko, I. Spirovová and Z. Bastl, Surf. Sci., 2006, 600, 1628 CrossRef CAS PubMed.
  46. M. D. Negra, M. Sambi and G. Granozzi, Surf. Sci., 1999, 436, 227 CrossRef.
  47. M. Petukhov, G. A. Rizzi and G. Granozzi, Surf. Sci., 2001, 490, 376 CrossRef CAS.
  48. G. Silversmit, D. Depla, H. Poelman, G. B. Marin and R. De Gryse, J. Electron Spectrosc. Relat. Phenom., 2004, 135, 167 CrossRef CAS PubMed.
  49. J. Mendialdua, R. Casanova and Y. Barbaux, J. Electron Spectrosc. Relat. Phenom., 1995, 71, 249 CrossRef CAS.
  50. G. A. Sawatzky and D. Post, Phys. Rev. B: Condens. Matter Mater. Phys., 1979, 20, 1546 CrossRef CAS.
  51. M. Demeter, M. Neumann and W. Reichelt, Surf. Sci., 2000, 454–456, 41 CrossRef CAS.
  52. M. Jiang, X. Cao, S. Bao, H. Zhou and P. Jin, Thin Solid Films, 2014, 562, 314 CrossRef CAS PubMed.
  53. C. Batista, V. Teixeira and J. Carneiro, J. Nano Res., 2008, 2, 21 CrossRef CAS.
  54. C. Liu, M. P. Jiang, J. H. Li and S. K. Wang, Adv. Mater. Res., 2011, 399–401, 589 Search PubMed.
  55. S. Ji, Y. Zhao, F. Zhang and P. Jin, J. Ceram. Soc. Jpn., 2010, 118, 867 CrossRef CAS.
  56. C. V. S. Reddy, J. Wei, Z. Quan-Yao, D. Zhi-Rong, C. Wen, S. Mho and R. R. Kalluru, J. Power Sources, 2007, 166, 244 CrossRef CAS PubMed.
  57. V. B. Chanshetty, K. Sangshetty and G. Sharanappa, Int. J. Eng. Res. Ind. Appl., 2012, 2, 611 Search PubMed.
  58. F. P. Gokdemir, U. D. Menda, P. Kavak, A. E. Saatci, O. Ozdemir and K. Kutlu, AIP Conf. Proc., 2012, 1476, 279 CrossRef CAS PubMed.
  59. L. Jacques, Coord. Chem. Rev., 1999, 190–192, 391 Search PubMed.
  60. K. Karsli, Dissertation, Middle East Technical University, Ankara, Turkey, 2012.
  61. N. R. Mlyuka, G. A. Niklasson and C. G. Granqvist, Sol. Energy Mater. Sol. Cells, 2009, 93, 1685 CrossRef CAS PubMed.
  62. F. Guinneton, L. Sauques, J. C. Valmalette, F. Cros and J. R. Gavarri, Thin Solid Films, 2004, 446, 287 CrossRef CAS PubMed.
  63. Z. Luo, Z. Wu, T. Wang, X. Xu, W. Li, W. Li and Y. Jiang, J. Phys. Chem. Solids, 2012, 73, 1122 CrossRef CAS PubMed.

Footnote

Both authors worked equally.

This journal is © The Royal Society of Chemistry 2015