DOI:
10.1039/C5RA07542D
(Paper)
RSC Adv., 2015,
5, 52602-52611
Role of substrate effects on the morphological, structural, electrical and thermoelectrical properties of V2O5 thin films
Received
25th April 2015
, Accepted 18th May 2015
First published on 27th May 2015
Abstract
The present work focuses on the influence of different substrates on the morphological, compositional, phase purity, structural and transport properties of vanadium pentaoxide (V2O5) thin films. Thin films of V2O5 were fabricated on different substrates: glass, quartz, Si, and alumina (hereafter these films are referred to as V2O5:G, V2O5:Q, V2O5:Si and V2O5:A, respectively) using inorganic sol–gel with V2O5 powder and hydrogen peroxide (H2O2) as precursors by spin coating. Films deposited on glass substrates were found to be amorphous in nature with smooth surfaces, whereas films deposited on quartz, silicon, and alumina substrates exhibited a polycrystalline nature, having an orthorhombic structure with space group Pmmn. The crystallinity improves from quartz to silicon and the best crystalline films were fabricated on alumina. Electrical measurements as a function of temperature and substrate are investigated and characterized by measuring the resistivity, Hall and Seebeck coefficients. Negative values of Hall and Seebeck coefficients reveal that all the films are of n-type semiconductors. Electrical resistivity as well as charge carrier density decreases from the films on glass to quartz to silicon and to alumina. Carrier mobility decreases in the following order V2O5:G > V2O5:A > V2O5:Si > V2O5:Q, whereas the Seebeck coefficient varies in the reverse order. Variation of these transport parameters has been understood on the basis of scattering and trapping of charge carriers along the grain boundaries. Furthermore, a model based on thermodynamics is proposed to explain the effect of substrates on the crystallinity of thin films. Interactions between sol and substrate (adhesive forces) determine the thickness, phase purity, structural and morphological properties of thin films. As the magnitude of adhesive forces increases, both film thickness and crystallinity increase.
1. Introduction
Vanadium oxides offer many convenient physical effects, which mean they can be used for different microelectronic device applications.1 Vanadium is a multivalent element, and as a result forms a variety of oxides upon reacting with oxygen, such as VO, V2O3, VO2, V6O11, V2O5 etc. V2O5 is the most saturated (highest oxidation state) in the V–O system and is consequently the most stable among these vanadium oxides.2,3 It has a lamellar, or sheet-like structure.4 It is a distorted orthorhombic structure, and this deformation creates its sheet formation. V2O5 in the form of thin films has attracted much attention due to its unique electronic, chemical and optical properties.5–8 Thin films of V2O5 have been prepared by different techniques including vacuum evaporation,9 sputter deposition,10–13 thermal oxidation,14,15 pulsed laser deposition,16 chemical vapor deposition,17–19 sol–gel processes20 and spray pyrolysis.21 Due to the advantages of high purity, homogeneity, stoichiometry, simplicity of equipment involved and cost effectiveness for large scale production, the sol–gel process has become one of the best approaches to deposit metal oxide thin films.
Recently, it was reported that the metal to insulator transition (MIT) at approximately 280 °C in V2O5 is due to lattice distortion and a structural inhomogeneity due to the vanadyl–oxygen vacancies,22–24 and this makes V2O5 a promising material for thermoelectric devices. V2O5 exhibits highly anisotropic electrical and optical properties due to its orthorhombic structure25 and these properties have been extensively investigated by many researchers.25–28 On the other hand, the thermo-physical properties, including thermal conductivity (κ), Seebeck coefficient (s), and dynamic heat capacity, have received less attention.29–31 One of the important parameters for characterizing the thermo-physical properties of materials is a power factor (s2/ρ), where s is the thermoelectric power (Seebeck coefficient) and ρ is the electrical resistivity. Thus, to be a better thermoelectric material, a large value for s and small values of ρ are required. Surprisingly, the s and electrical resistivity (ρ) vary in opposite directions, preventing the power factor from improving excellently, and both the values are also dependent on each other with carrier density (n). Generally, the s value of a semiconductor or an insulator decreases with increasing temperature. The variation of s with temperature is very important for thermoelectric applications.
In the present work, we investigate the effect of substrates on the microstructural evolution, crystal phase, and surface morphology, electrical and thermo-electrical properties of V2O5 thin films fabricated by sol–gel method. Crystallite size, surface morphology and film thickness were observed to vary from substrate to substrate, which in turn affects the transport properties of thin films.
2. Experimental details
A clear yellow solution, containing 0.3 g V2O5 (Sigma-Aldrich, purity >99.9%) dissolved in 30 ml, 30% hydrogen peroxide H2O2 (Sigma-Aldrich), was formed at room temperature with vigorous stirring. This yellow solution was then heated at 60 °C with continuous stirring to evolve excess oxygen by decomposition of H2O2. Immediately, the solution turned into a red brown viscous gel. After aging for 24 hours, the V2O5 gel was ready for coating. The mechanism of the reaction is given in reactions (1) and (2) following Ren et al.32| | |
V2O5 + 2H2O2 → 2HVO4 + H2O
| (1) |
| | |
2HVO4 + (n − 1)H2O → V2O5·nH2O + O2
| (2) |
The spin coater (Spin-NXG-P1: made by Apex Instruments, India) was used to deposit thin films on glass, quartz, Si, and alumina substrates. Prior to deposition, these substrates were cleaned in dilute sulphuric acid for 30 minutes and then thoroughly rinsed with ethanol, acetone and de-ionized water. Five coatings of V2O5 gel were performed on each substrate at a rate of 3000 rpm for 30 seconds and after each coating the films were dried at 80 °C for 15–20 minutes. Finally these V2O5 gel films were crystallized by annealing at 400 °C for 4 hours under ambient atmosphere in a programmable tubular furnace (Nabertherm GmbH Tube furnace: RHTC80), with a heating and cooling rate of 3 °C per minute. Rutherford backscattering (RBS) with He+ ions of energy 2 MeV was used to characterize the stoichiometry, thickness and uniformity of the films. X-ray diffraction (XRD) measurements at room temperature in the 2θ range 20–60° was carried out to identify the crystalline phases and structure of the films using a Bruker D8 advance diffractometer with Cu Kα (0.15406 nm) X-ray source at a scan speed of 0.5° min−1. Raman studies of these films were carried out and the spectra were measured in back scattering geometry using an Ar excitation source, with a wavelength of 488 nm coupled with a Labram-HR800 micro-Raman spectrometer equipped with a 50× objective, appropriate notch filter, and a Peltier cooled charge-coupled device detector. The surface morphology of films was characterized by SEM (model: MIRA II LMH, TESCAN). The electrical resistivity, Seebeck coefficients of the films were measured in the temperature range 300–400 K using a standard DC four probe technique and bridge method, respectively. The Hall effect measurements were carried out in the same temperature range by Ecopia HMS-3000 Hall Measurement System to evaluate charge carrier density and mobility.
3. Results and discussion
3.1. Morphological studies
Fig. 1 shows the surface morphology of V2O5 thin films deposited on glass, quartz, Si, and alumina substrates investigated by SEM, and these films hereafter will be referred to as V2O5:G, V2O5:Q, V2O5:Si, and V2O5:A, respectively. As evident, there is a significant change in the surface morphologies. Films deposited on glass substrates have relatively smooth surfaces, with few compact structures revealing the amorphous nature of these films; also some cracks are observed on the surface, which may be due to shrinkage of sol during the annealing of these film. Films deposited on quartz possess irregular nanostructures with varying sizes. The bigger particles grow due to the agglomeration of smaller ones to reduce the surface energy under a thermodynamic driving force, when two particles merge at high temperature. The surfaces of films deposited on Si and alumina possess cubic growth characteristics resembling spiral-like features, which may have emanated from screw dislocations. The average crystallite size was found to be increased while moving from quartz through silicon to alumina. Further grain size distribution was found to be more random for V2O5 films deposited on quartz than those deposited on silicon and alumina substrates. This may be due to the difference in sticking coefficients and strains developed in the films grown on an amorphous background. The average grain sizes of V2O5:Q, V2O5:Si and V2O5:A are approximately 70 nm, 135 nm and 180 nm, respectively. Larger grain size was observed in V2O5:A compared to V2O5:Si due to lattice mismatches of V2O5 and Al2O3 then V2O5 and Si.
 |
| | Fig. 1 SEM images of (a) V2O5:G, (b) V2O5:Q, (c) V2O5:Si and (d) V2O5:A. Films deposited on glass are amorphous in nature with smooth surfaces. Films deposited on quartz, Si and alumina are crystalline in nature and average crystallite size increases from quartz to Si to alumina. | |
3.2. Compositional analysis
As mentioned above, oxides of vanadium can exist in a wide range of stoichiometries. The RBS measurements with SIMNRA Program were used to determine composition and thickness of films. RBS spectra (with SIMNRA simulation) of the four samples (V2O5:G, V2O5:Q, V2O5:Si, and V2O5:A) are shown in Fig. 2(a)–(d), respectively. The simulated results are summarized in Table 1. This simulation suggests that the V/O ratios for all the samples are 2/5 within simulation errors, indicating good stoichiometries of all these four thin films. The thickness was found to be ∼46.7 nm, 85.5, 118.4 and 130.3 nm for films V2O5:G, V2O5:Q, V2O5:Si, V2O5:A, respectively. Increase in film thickness from glass to quartz to Si to alumina may be due to the increase in magnitude of cohesive forces in comparison to adhesive forces from V2O5:G, V2O5:Q, V2O5:Si, and V2O5:A. It is also clear that the RBS results indicated that there was no contamination during film deposition.
 |
| | Fig. 2 RBS spectra of (a) V2O5:G, (b) V2O5:Q, (c) V2O5:Si, and (d) V2O5:A. Thickness of film increases from glass (46.7 nm) to quartz (85.5 nm) to silicon (118.4 nm) to alumina (130.3 nm). | |
Table 1 Simulation results on composition of thin films on various substrates
| Substrates |
Composition of film |
Thickness (nm) |
| V |
O |
| Glass |
0.286 |
0.714 |
46.7 |
| Quartz |
0.286 |
0.714 |
85.5 |
| Si |
0.286 |
0.714 |
118.4 |
| Alumina |
0.286 |
0.714 |
130.3 |
3.3. Phase study
Fig. 3 shows the XRD pattern of the V2O5 thin films deposited on glass, quartz, Si, and alumina substrates. The X-ray diffraction spectrum of V2O5 films deposited on a glass substrate indicates amorphous nature, as the diffraction pattern is diffused and non characteristic. Films fabricated on quartz, Si, and alumina substrates are comprised of the V2O5 phase within the XRD detection limit, without any other phases of vanadium oxide. The intensity of the diffraction peaks demonstrates that crystallinity increases from quartz through Si to alumina. The peaks are indexed according to the standard pattern [JCPDS file no. 85-0601] for polycrystalline orthorhombic V2O5. The relatively high intensity of the (001) peaks demonstrates the growth of films oriented along the c-axis perpendicular to the surface of the substrate. In addition to the (001) Bragg reflection, the subsequent appearance of other characteristic orientations, such as (101), (110) and (002), reveals the existence of in-plane organization of V–O–V chains. The evaluated lattice parameters on the basis of measured d spacings for films on quartz, Si, and alumina are found to be a = 0.3564 ± 0.001, b = 1.151 ± 0.001, and c = 0.433 ± 0.001 nm; these are in good agreement with previously reported values in the literature.33 Even though the (001) peak is the strongest in a polycrystalline XRD powder pattern of V2O5, the intensity ratios (I(001)/I(hkl)) in the sol–gel prepared films are larger than in polycrystalline V2O5 powder, indicating a strong (001) texture. This preferred orientation in the sol–gel films can be understood from the properties of the starting material. The gel is formed by the hydrolysis and condensation of molecular precursors. The chemical control of these reactions allows the formation of V2O5 gels directly from the solutions at lower temperature than by the standard solid state process.34–36 Therefore, the sol–gel films are comprised of V2O5·nH2O, before annealing at 400 °C.37 These have a V2O5 layered structure with trapped water molecules and are characterized by a strong structural anisotropy.
 |
| | Fig. 3 XRD Spectra of (a) V2O5:G, (b) V2O5:Q, (c) V2O5:Si, and (d) V2O5:A. The ‘*’ indexed peaks correspond to substrates. | |
3.4. Raman measurements
Fig. 4 displays the Raman spectra of V2O5 thin films fabricated on glass, quartz, silicon, and alumina substrates in the wavelength range 100–1100 cm−1. The broad and non-characteristic Raman spectrum reveals the amorphous nature of V2O5 when deposited on the glass substrate. The Raman spectra of films deposited on quartz, silicon and alumina substrates exhibit distinguishable and characteristic assignable Raman peaks, indicating the polycrystalline nature of these films. Peak intensities clearly demonstrate that crystallinity increases from films fabricated on quartz through silicon to alumina. Raman measurements of the films can be described using the shape and frequency of the peaks.
 |
| | Fig. 4 Raman scattering spectra of (a) V2O5:G, (b) V2O5:Q, (c) V2O5:Si and (d) V2O5:A. The ‘*’ indexed peaks correspond to the substrate. | |
V2O5 crystallizes in the orthorhombic system with the group Pmmn (D132h) according to Bachmann et al.33 The structure of V2O5 is built up from VO5 square pyramids sharing edges to form (V2O4)n zigzag double chains along [001] and cross linked along [100] by corner sharing and thus forming sheets in the xz plane.38,39 Thus in each layer V is five-fold coordinated; with three V–O bonds involving three fold coordinated oxygen (Oc) belonging to (V2O4)n chains, one V–O bond involving two fold coordinated oxygen (OB) constituting bridges between two chains and one involving vanadyl oxygen (OV). The successive layers are kept together by an equal number of weak van der Waals bonds and much stronger double bonds.40 The unit cell contains two formula units (V4O10) yielding a total number of 39 optical zone centers. Due to the inversion symmetry and according to D2h factor group analysis, the modes are split into IR (odd) and Raman (even) active modes. All 21 g modes are Raman active while 15 Bu modes are IR active. Beattie and Gilson41 indicated that a convenient way to describe vibrations in a binary oxide lattice is to treat the oxygen atom as vibrating against an array of immobile metal atoms. Each oxygen species is then considered first as vibrating according to its site symmetry and, following this, the unit cell modes are generated by symmetry operations.
Table 2 depicts the principle Raman active modes with the corresponding motion of V2O5 thin films. Each spectrum consists of two groups of peaks located in the high-frequency region, known as internal modes, and in the low frequency region, known as external modes. The internal modes that are observed in the high frequency region are assigned to the stretching and bending of V–O bonds. The high-frequency Raman peak at 1006 cm−1 gives the structural quality of the films and can be ascribed to the stretching mode related to the Ag symmetry vibrations of the shortest vanadium oxygen bond, which is V
OV. Unlike the other O atoms this atom is strongly bonded to only one V atom and for this reason is called terminal oxygen.42 The frequency shift of this mode measures the deviations from stoichiometry. The frequency shift to lower values of this mode is due to softening of the V5+
O bond in oxygen deficient V2O5 films, resulting from vacancies created by removing OV, with some of the V5+ reduced to V4+ in order to balance the charge. Negligible frequency shift of this mode to 995 cm−1 demonstrates good stoichiometry of all three crystalline samples, which is in good agreement with the RBS and XRD results. This is also evident as no peak corresponding to V4+
O near 932 cm−1 is observed as reported by Lee et al.43 The second peak at 702 cm−1 is assigned to the doubly coordinated oxygen (V2–OB) stretching mode, which results from the corner-shared oxygen common to two pyramids. The third peak at 529 cm−1 is assigned to the triply coordinated oxygen (V3–OC) stretching mode, which results from the edge-shared oxygen atoms common to three pyramids. The two peaks located at 404 and 283 cm−1 are assigned to the bending vibration of the V
OV bonds. The peaks located at 481 cm−1 and 304 cm−1 are assigned to the bending vibrations of the bridging V–OB–V (doubly coordinated oxygen) and V3–O (triply coordinated oxygen) bonds, respectively.
Table 2 Raman active modes of orthorhombic V2O5 thin filma
| Wave number (cm−1) |
Symmetry |
Assignment |
Remarks |
| Symmetry and assignation of modes are mainly based on ref. 42 and 43. Frequency values are derived from present measurements. |
| 146 |
B3g, B2g |
Ty, Rz |
Relative motions of V2O5 layers with respect to each other |
| 198 |
B1g |
Tx, Ry |
Relative motions of V2O5 layers with respect to each other |
| 283 |
B2g, B3g |
δ(V O)b |
Bending vibration of the V OV bonds |
| 304 |
Ag |
δ(V3–O)b |
Bending vibrations of the V3–O (triply coordinated oxygen) bond |
| 404 |
Ag |
δ(V O)b |
Bending vibration of the V OV bonds |
| 481 |
Ag |
δ(V–O–V)b |
Bending vibrations of the bridging V–OB–V (doubly coordinated oxygen) |
| 529 |
Ag |
ν(V3–O)s |
Triply coordinated oxygen (V3–OC) stretching mode |
| 710 |
B2g, B3g |
ν(V–O–V)s |
Doubly coordinated oxygen (V2–OB) stretching mode |
| 1006 |
Ag |
ν(V O)s |
Stretching mode vibrations of the shortest vanadium oxygen bond (V OV) |
The external modes can be viewed as relative motions of structural units with respect to each other, i.e. translations T1x, T1y, T1z and librations R1x, R1y, R1z. These vibrations occur at low frequencies because each unit is considered heavier than its constituent atoms, while the restoring force has the same order of magnitude.44 The external low frequency Raman modes at 146 and 198 cm−1 correspond to the relative motions of V2O5 layers with respect to each other.42 These two peaks at 146 cm−1 and 198 cm−1 are strongly associated with the layered structure and only appear when there is long range structural order. The presence of these low frequency modes in the films deposited on quartz, silicon and alumina substrates suggest that these films possess a layered structure and are well crystallized. The films grow preferentially, with the c-axis oriented perpendicular to the substrate plane.45
3.5. Transport measurements
The temperature dependence of resistivity (ρ) of V2O5:G, V2O5:Q, V2O5:Si and V2O5:A thin films in the temperature range 300–400 K is shown in Fig. 5(a). All the samples show semiconducting behavior as electrical resistivity decreases with temperature. Further electrical resistivity decrease occurs in the following order V2O5:G > V2O5:Q > V2O5:Si > V2O5:A. Hall measurements of samples demonstrate that all these films are n type semiconductors. Fig. 5(b) displays variation of carrier density (n) and carrier mobility (μ) in the temperature range 300–400 K of the samples. Carrier concentration increases with temperature but mobility decreases with temperature. Decrease in μ values may be caused by increase in carrier scattering due to thermal phonons. Comparison of n and μ of V2O5:G, V2O5:Q, V2O5:Si, and V2O5:A films demonstrates that carrier concentration increases in the following order V2O5:G < V2O5:Q < V2O5:Si < V2O5:A, whereas mobility increases in the following order: V2O5:Q, V2O5:Si, and V2O5:A to V2O5:G in the whole temperature range. Furthermore the differences become more enhanced with temperature. Variation of resistivity, carrier concentration and mobility in these orders can be explained on the basis of thickness and crystallinity differences of the films. Both film thickness and crystallinity increase from V2O5:G to V2O5:Q to V2O5:Si to V2O5:A. As film thickness decreases, stacking defects increase, and these defects trap the carriers and lower the free carrier concentration,46,47 and thus carrier concentration increases as we go from V2O5:G to V2O5:Q to V2O5:Si to V2O5:A. In addition, the film grown on the glass substrate is amorphous in nature, whereas films deposited on alumina, silicon and quartz are crystalline in nature. The crystallites formed in V2O5:A, V2O5:Si and V2O5:Q produce a number of grain boundaries and, these boundaries acts as scattering centers for the flow of charge carriers and thus cause reduction in the μ values, as we go from amorphous film fabricated on glass to the crystalline films fabricated on alumina, silicon and quartz. Furthermore, as one goes from V2O5:A to V2O5:Si to V2O5:Q, grain size decreases and thus the number of grain boundaries increases, which results in increase in carrier mobility from V2O5:A to V2O5:Si to V2O5:Q. Fig. 5(c) displays the Seebeck coefficient (s) of V2O5:G, V2O5:Q, V2O5:Si and V2O5:A thin films in the temperature range 300–400 K. The negative value of s also demonstrates that all films are n type semiconductors. The s value increases significantly over the entire temperature range from amorphous V2O5:G to crystalline V2O5:Q, V2O5:Si and V2O5:A thin films. The s value increases with decrease in crystallite size from V2O5:A to V2O5:Si to V2O5:Q. This can be understood as follows: the energy barrier formed at the grain boundaries acts as an additional trap center and traps low energy charge carriers; this filtering of charge carriers by grain boundaries enhances the average energy of carriers taking part in the transport mechanism.48 The value of thermoelectric power depends on the mean carrier energy relative to the Fermi level.49 Therefore, this phenomenon of filtering of charge carriers also leads to an increase in s value from V2O5:A to V2O5:Si, to V2O5:Q. Furthermore, with increase in temperature, the average energy of charge carriers taking part in the transport mechanism increases, resulting in an increase in s value with temperature.
 |
| | Fig. 5 (a) Electrical resistivity, (b) carrier mobility and concentration, and (c) Seebeck coefficient of V2O5:glass (b) V2O5:Q (c) V2O5:Si and (d) V2O5:alumina thin films as a function of temperature. | |
4. Model for crystallization of V2O5 thin films on different substrates
Lattice mismatch between the substrates and V2O5 and the underlying surface strongly affects nucleation and growth processes. In this section we model for the effect of substrates on the crystallization of V2O5 thin films. Fig. 6 shows a schematic diagram of spherical nuclei nucleated on the hetero-substrate. The driving force for nucleation is Gibbs free energy.50,51 The Gibbs free energy change for the nucleation of spherical nuclei of radius r on the hetero-substrate50 is given by:| | |
ΔGHet. = −VΔGv + ∑AJγJ
| (3) |
| | |
= −VΔGv + (AELγEL + AESγES − ASLγSL)
| (4) |
where V is the volume of the spherical cluster, ΔGv the Gibbs free energy of molar volume, AEL and γEL the interface area and the interface energy of the embryo–liquid interface, AES and γES the interface area and energy of the embryo–substrate interface, and ASL and γSL the interface area and the energy between the substrate–liquid interfaces. According to geometry| |
 | (6) |
where r is the radius of the spherical cluster and m describes the interaction between the liquid embryo and the solid substrate and this has been described in the classical model with the help of the Young equation:51| |
 | (9) |
where θ is the contact or wetting angle between the substrate and the liquid embryo as shown in Fig. 6. Thus, we can obtain| |
 | (10) |
| |
 | (11) |
where the factor| |
 | (12) |
is called heterogeneous factor, and its value lies in the range 0–1. Its value is 1, when clusters nucleated on the homo-substrates. The critical nucleus radius r* found by setting ∂ΔG/∂r = 0, with the result| |
 | (13) |
 |
| | Fig. 6 Schematic diagram displaying the spherical nuclei nucleated on the hetero-substrate. | |
It is important to note that the critical radius r* remains unchanged for heterogeneous nucleation and homogeneous nucleation. However, the volume (V) can be significantly lower for heterogeneous nucleation, due to the wetting angle affecting the shape of the nucleus. The associated energy barrier for nucleation is found by substituting r* in eqn (8)
| |
 | (14) |
| |
 | (15) |
| |
 | (16) |
The nucleation rate J, which is the number of critical nuclei formed per unit time per unit volume, is usually written in Arrhenius activation form as
| |
 | (18) |
where
D is a kinetic pre-exponential factor, correlated to encounter frequency between the molecules, and is consequently dependent on their diffusion coefficients,
KB is the Boltzmann constant, and Δ
G* is the reversible work of formation of the critical nucleus. The diffusion coefficient
D is observed to depend upon temperature as
52| |
 | (19) |
where
Ed is the activation energy for diffusion,
KB is Boltzmann’s constant,
T is the absolute temperature, and
D0 is a temperature-independent factor that depends upon the choice of material and deposit.
Eqn (18), together with
eqn (17) and
(19), defines the rate of heterogeneous nucleation of liquid embryos on the heterogeneous substrate. Thus the rate of nucleation depends upon the nature of material of the substrate, the temperature of the substrate and the contact angle between the substrate and the liquid embryo. Hence, maintaining all the substrates at fixed temperature during the deposition of V
2O
5 and post annealing at the same temperature, nucleation of V
2O
5 varies from substrate to substrate. The main parameter that may determine the growth of V
2O
5 thin film on a substrate at fixed temperature is the contact angle, and this contact angle characterizes the interactions between the sol and substrate known as adhesive forces, measuring the wettability of the solid surface by liquid. The work of adhesion,
Wad, is defined as
53| |
Wad = σlv(1 + cos θ)
| (20) |
A high work of adhesion indicates good wetting, whereas a low work of adhesion indicates poor wetting. Fig. 7(a)–(d) show the wetting and spreading phenomena of water on glass, quartz, silicon, and alumina substrates, respectively. Glass substrates are the most wetted and alumina substrates are the least wetted. The magnitude of the contact angle increases from 19° for glass to 35.4° for quartz to 38.5° for silicon to 46.5° for alumina substrates. A similar trend of increasing contact angle from glass to quartz to silicon to alumina substrate is expected for V2O5 sol. This indicates that the magnitude of adhesive forces between sol and the substrate decreases from glass to quartz to Si to alumina substrate. Thus V2O5 sol spreads more on glass and the least on alumina. As a consequence of this, the nucleation barrier increases from glass to quartz to silicon to alumina, as shown in Fig. 8, and thus the nucleation rate J, which is the number of critical nuclei formed per unit time per unit volume, decreases from glass to quartz to Si to alumina substrate. Thus, the number of critical nuclei that is formed on the surface of the substrates decreases from glass to quartz to Si to alumina, and hence the thickness and crystallinity of the V2O5 thin film increases from glass to quartz to silicon to alumina substrate.
 |
| | Fig. 7 Images of water contact angle on (a) glass, (b) quartz, (c) Si and (d) alumina substrates. | |
 |
| | Fig. 8 Schematic diagram showing variations of Gibbs free energy with radius of nuclei. | |
5. Conclusion
The nature of substrates determines the properties of V2O5 thin films. SEM images showed a change in surface morphology of V2O5 films from substrate to substrate. Films deposited on glass substrates are amorphous in nature with a smooth surface, whereas films deposited on quartz, Si and alumina substrates are crystalline. The degree of crystallization increases from quartz to Si to alumina. RBS results show that the thickness of the film increases from glass to quartz to silicon to alumina substrate. XRD and Raman results show that films fabricated on a glass substrate are amorphous in nature, whereas films fabricated on quartz, silicon, and alumina are crystalline, and crystallinity increases from quartz to silicon to alumina. Basically, these are the interactions between sol and substrate (adhesive forces) that play a significant role in elucidating the thickness and degree of crystallization of V2O5 thin films. As the magnitude of adhesive forces between the sol and substrate increases, both film thickness and crystallinity are enhanced. The electrical and thermoelectrical properties were found to be functions of crystallization and thickness. As film thickness and crystallization increase from V2O5:G to V2O5:Q to V2O5:Si to V2O5:A, both carrier concentration and conductivity increase. Carrier mobility decreases from amorphous film to crystalline film and increases with increasing crystallite size. The Seebeck coefficient shows a strong dependence on crystallization. The Seebeck coefficient increases from amorphous film to crystalline film and then decreases with increase in crystallite size. This behavior of μ and s with crystallization is due to the scattering and trapping of charge carriers along the grain boundaries and stacking defects.
Acknowledgements
One of the authors (B.A.) acknowledges the director, NIT-Srinagar and Director, IUAC, New Delhi for their support and encouragements. He also thanks Materials Science group for all characterization facilities.
References
- R. B. Darling and S. Iwanaga, Structure, properties, and MEMS and microelectronic applications of vanadium oxides, Sadhana, 2009, 34(4), 531–542 CrossRef CAS PubMed.
- F. J. Morin, Oxides which show a metal-to-insulator transition at the Neel temperature, Phys. Rev. Lett., 1959, 3(1), 34–36 CrossRef CAS.
- R. L. Smith, G. S. Rohrer, K. S. Lee, D. K. Seo and M. H. Whangbo, A scanning probe microscopy study of the (001) surfaces of V2O5 and V6O13, Surf. Sci., 1996, 367(1), 87–95 CrossRef CAS.
- F. Hulliger, in Structural Chemistry of Layer-Type Phases, ed. F. Lévy and D. Reidel, Dordrecht, 1976, Fig. 100, with kind permission of Springer Science and Business Media Search PubMed.
- Y. Wang and G. Cao, Synthesis and enhanced intercalation properties of nanostructured vanadium oxides, Chem. Mater., 2006, 18(12), 2787–2804 CrossRef CAS.
- V. Petkov, et al. Structure beyond Bragg: Study of V2O5 nanotubes, Phys. Rev. B: Condens. Matter Mater. Phys., 2004, 69.8, 085410 CrossRef.
- G. T. Kim, J. Muster, V. Krstic, J. G. Park, Y. W. Park, S. Roth and M. Burghard, Field-effect transistor made of individual V2O5 nanofibers, Appl. Phys. Lett., 2000, 76(14), 1875–1877 CrossRef CAS PubMed.
- C. V. Ramana, R. J. Smith, O. M. Hussain, C. C. Chusuei and C. M. Julien, Correlation between growth conditions, microstructure, and optical properties in pulsed-laser-deposited V2O5 thin films, Chem. Mater., 2005, 17(5), 1213–1219 CrossRef CAS.
- Z. Lu, M. D. Levi, G. Salitra, Y. Gofer, E. Levi and D. Aurbach, Basic electroanalytical characterization of lithium insertion into thin, well-crystallized V2O5 films, J. Electroanal. Chem., 2000, 491(1), 211–221 CrossRef CAS.
- K. West, B. Zachau-Christiansen, S. V. Skaarup and F. W. Poulsen, Lithium insertion in sputtered vanadium oxide film, Solid State Ionics, 1992, 57(1), 41–47 CrossRef CAS.
- Y. J. Park, K. S. Ryu, K. M. Kim, N. G. Park, M. G. Kang and S. H. Chang, Electrochemical properties of vanadium oxide thin film deposited by RF sputtering, Solid State Ionics, 2002, 154, 229–235 CrossRef.
- S. C. Mui, J. Jasinski, V. J. Leppert, M. Mitome, D. R. Sadoway and A. M. Mayes, Microstructure effects on the electrochemical kinetics of vanadium pentoxide thin-film cathodes, J. Electrochem. Soc., 2006, 153(7), A1372–A1377 CrossRef CAS PubMed.
- A. Benayad, H. Martinez, A. Gies, B. Pecquenard, A. Levasseur and D. Gonbeau, XPS investigations achieved on the first cycle of V2O5 thin films used in lithium microbatteries, J. Electron Spectrosc. Relat. Phenom., 2006, 150, 1–10 CrossRef CAS PubMed.
- T. Szörényi, K. Bali and I. Hevesi, Structural characterization of amorphous vanadium pentoxide thin films prepared by chemical vapour deposition/CVD, J. Non-Cryst. Solids, 1980, 35, 1245–1248 CrossRef.
- K. Inumaru, T. Okuhara, M. Misono, N. Matsubayashi, H. Shimada and A. Nishijima, EXAFS analysis of vanadium oxide thin overlayers on silica prepared by chemical vapour deposition, J. Chem. Soc., Faraday Trans., 1992, 88(4), 625–630 RSC.
- R. Lindström, V. Maurice, S. Zanna, L. Klein, H. Groult, L. Perrigaud, C. Cohen and P. Marcus, Thin films of vanadium oxide grown on vanadium metal: oxidation conditions to produce V2O5 films for Li-intercalation applications and characterisation by XPS, AFM, RBS/NRA, Surf. Interface Anal., 2006, 38(1), 6–18 CrossRef PubMed.
- J. Światowska-Mrowiecka, V. Maurice, S. Zanna, L. Klein, E. Briand, I. Vickridge and P. Marcus, Ageing of V2O5 thin films inducedby
Li intercalation multi-cycling, J. Power Sources, 2007, 170(1), 160–172 CrossRef PubMed.
- G. J. Fang, Z. L. Liu, Y. Wang, Y. H. Liu and K. L. Yao, Synthesis and structural, electrochromic characterization of pulsed laser deposited vanadium oxide thin films, J. Vac. Sci. Technol., A, 2001, 19(3), 887–892 CAS.
- A. Mantoux, H. Groult, E. Balnois, P. Doppelt and L. Gueroudji, Vanadium oxide films synthesized by CVD and used as positive electrodes in secondary lithium batteries, J. Electrochem. Soc., 2004, 151(3), A368–A373 CrossRef CAS PubMed.
- D. B. Le, S. Passerini, A. L. Tipton, B. B. Owens and W. H. Smyrl, Aerogels and xerogels of V2O5 as intercalation hosts, J. Electrochem. Soc., 1995, 142(6), L102–L103 CrossRef CAS PubMed.
- A. Bouzidi, N. Benramdane, A. Nakrela, C. Mathieu, B. Khelifa, R. Desfeux and A. Da Costa, First synthesis of vanadium oxide thin films by spray pyrolysis technique, J. Mater. Sci. Eng. B, 2002, 95(2), 141–147 CrossRef.
- M. Kang, I. Kim, S. W. Kim, J. W. Ryu and H. Y. Park, Metal–insulator transition without structural phase transition in V2O5 film, Appl. Phys. Lett., 2011, 98(13), 131907 CrossRef PubMed.
- R. P. Blum, H. Niehus, C. Hucho, R. Fortrie, M. V. Ganduglia-Pirovano, J. Sauer and H. J. Freund, Surface metal–insulator transition on a vanadium pentoxide (001) single crystal, Phys. Rev. Lett., 2007, 99(22), 226103 CrossRef.
- A. L. Pergament, G. B. Stefanovich and A. A. Velichko, Oxide Electronics and Vanadium Dioxide Perspective: A Review, Journal on Selected Topics on Nano Electronics and Computing, 2013, 1, 24 CrossRef.
- C. V. Ramana, O. M. Hussain, B. S. Naidu and P. J. Reddy, Spectroscopic characterization of electron-beam evaporated V2O5 thin films, Thin Solid Films, 1997, 305(1), 219–226 CrossRef CAS.
- M. Zhu, Z. Zhang and W. Miao, Intense photoluminescence from amorphous tantalum oxide films, Appl. Phys. Lett., 2006, 89(2), 021915 CrossRef PubMed.
- M. Losurdo, G. Bruno, D. Barreca and E. Tondello, Dielectric function of V2O5 nanocrystalline films by spectroscopic ellipsometry: Characterization of microstructure, Appl. Phys. Lett., 2000, 77(8), 1129–1131 CrossRef CAS PubMed.
- M. I. Kang, I. K. Kim, E. J. Oh, S. W. Kim, J. W. Ryu and H. Y. Park, Dependence of optical properties of vanadium oxide films on crystallization and temperature, Thin Solid Films, 2012, 520(6), 2368–2371 CrossRef CAS PubMed.
- R. Santos, J. Loureiro, A. Nogueira, E. Elangovan, J. V. Pinto, J. P. Veiga and I. Ferreira, Thermoelectric properties of V2O5 thin films deposited by thermal evaporation, Appl. Surf. Sci., 2013, 282, 590–594 CrossRef CAS PubMed.
- S. Iwanaga, M. Marciniak, R. B. Darling and F. S. Ohuchi, Thermopower and electrical conductivity of sodium-doped V2O5 thin films, J. Appl. Phys., 2007, 101(12), 123709 CrossRef PubMed.
- E. Lazaro, M. Bhamidipati, M. Aldissi and B. Dixon, Thermoelectric Organics, USAR Office, East Falmouth, Massachusetts, USA, 1998 Search PubMed.
- X. Ren, Y. Jiang, P. Zhang, J. Liu and Q. Zhang, Preparation and electrochemical properties of V2O5 submicron-belts synthesized by a sol–gel H2O2 route, J. Sol-Gel Sci. Technol., 2009, 51(2), 133–138 CrossRef CAS PubMed.
- H. G. Bachmann, F. R. Ahmed and W. H. Barnes, The crystal structure of vanadium pentoxide, Z. Kristallogr., 1961, 115(1–6), 110–131 CrossRef CAS PubMed.
- K. S. Hwang, J. H. Ahn and B. H. Kim, Vanadium-doped nanocrystalline TiO2 layer prepared by spin coating pyrolysis, J. Mater. Sci., 2006, 41(10), 3151–3153 CrossRef CAS PubMed.
- J. Livage, Vanadium pentoxide gels, Chem. Mater., 1991, 3(4), 578–593 CrossRef CAS.
- Sol-gel science: the physics and chemistry of solgel processing, ed. C. J. Brinker and G. W. Scherer, Gulf Professional Publishing, 1990 Search PubMed.
- B. Alonso and J. Livage, Synthesis of Vanadium Oxide Gels from Peroxovanadic Acid Solutions: A51V NMR Study, J. Solid State Chem., 1999, 148(1), 16–19 CrossRef CAS.
- C. V. Ramana, O. M. Hussain, B. S. Naidu and P. J. Reddy, Spectroscopic characterization of electron-beam evaporated V2O5 thin films, Thin Solid Films, 1997, 305(1), 219–226 CrossRef CAS.
- L. Abello, E. Husson, Y. Repelin and G. Lucazeau, Vibrational spectra and valence force field of crystalline V2O5, Spectrochim. Acta, Part A, 1983, 39(7), 641–651 CrossRef.
- J. Haber, M. Witko and R. Tokarz, Vanadium pentoxide I. Structures and properties, Appl. Catal., A, 1997, 157(1), 3–22 CrossRef CAS.
- I. R. Beattie and T. R. Gilson, Oxide phonon spectra, J. Chem. Soc. A, 1969, 2322–2327 RSC.
- P. Clauws, J. Broeckx and J. Vennik, Lattice Vibrations of V2O5. Calculation of Normal Vibrations in a Urey–Bradley Force Field, Phys. Status Solidi B, 1985, 131(2), 459–473 CrossRef CAS PubMed.
- S. H. Lee, H. M. Cheong, M. J. Seong, P. Liu, C. E. Tracy, A. Mascarenhas and S. K. Deb, Raman spectroscopic studies of amorphous vanadium oxide thin films, Solid State Ionics, 2003, 165(1), 111–116 CrossRef CAS PubMed.
- L. Abello, E. Husson, Y. Repelin and G. Lucazeau, Vibrational spectra and valence force field of crystalline V2O5, Spectrochim. Acta, Part A, 1983, 39(7), 641–651 CrossRef.
- G. J. Fang, Z. L. Liu, Y. Q. Wang, H. H. Liu and K. L. Yao, Orientated growth of V2O5 electrochromic thin films on transparent conductive glass by pulsed excimer laser ablation technique, J. Phys. D: Appl. Phys., 2000, 33(23), 3018 CrossRef CAS.
- S. S. Lin, J. L. Huang and D. F. Lii, The effect of thickness on the properties of Ti-doped ZnO films by simultaneous rf and dc magnetron sputtering, Surf. Coat. Technol., 2005, 190(2), 372–377 CrossRef CAS PubMed.
- Y. Qu, T. A. Gessert, K. Ramanathan, R. G. Dhere, R. Noufi and T. J. Coutts, Electrical and optical properties of ion beam sputtered ZnO:Al films as a function of film thickness, J. Vac. Sci. Technol., A, 1993, 11(4), 996–1000 CAS.
- M. Bala, S. Gupta, T. S. Tripathi, S. Varma, S. K. Tripathi, K. Asokan and D. K. Avasthi, Enhancement of thermoelectric power of PbTe:Ag nanocomposite thin films, RSC Adv., 2015, 5(33), 25887–25895 RSC.
- J. Martin, L. Wang, L. Chen and G. S. Nolas, Enhanced Seebeck coefficient through energy-barrier scattering in PbTe nanocomposites, Phys. Rev. B: Condens. Matter Mater. Phys., 2009, 79(11), 115311 CrossRef.
- G. W. Yang and B. X. Liu, Nucleation thermodynamics of quantum-dot formation in Vgroove structures, Phys. Rev. B: Condens. Matter Mater. Phys., 2000, 61(7), 4500 CrossRef CAS.
- D. Y. Kwok, C. N. C. Lam, A. Li, A. Leung, R. Wu, E. Mok and A. W. Neumann, Measuring and interpreting contact angles: a complex issue, Colloids Surf., A, 1998, 142(2), 219–235 CrossRef CAS.
- C. Kittel, Introduction to solid state physics, Wiley, New York, 1968 Search PubMed.
- A. W. Neumann and R. J. Good, Surface and Colloid Science, ed. R. J. Good and R. R. Stromberg, Plenum Press, New York, 1979, vol. II Search PubMed.
|
| This journal is © The Royal Society of Chemistry 2015 |
Click here to see how this site uses Cookies. View our privacy policy here.