Processing math: 100%
Review Topical Sections

V2O5 thin films for energy storage and conversion

  • V2O5 is one of the best material for many applications. Progress is currently made to improve its performance for use as a sensor, or an electrode, or smart window, electrochromic device, supercapacitor, photovoltaic applications among others. In this work, we review the progress that has been done these recent years, in relation to the mode of preparation of the V2O5 films. The results outline the complex relationship between the synthesis and the properties, which should serve as a guide for further research on the dependence of the best synthesis process on the type of application that is targeted.

    Citation: Alain Mauger, Christian M. Julien. V2O5 thin films for energy storage and conversion[J]. AIMS Materials Science, 2018, 5(3): 349-401. doi: 10.3934/matersci.2018.3.349

    Related Papers:

    [1] Haridas D. Dhaygude, Surendra K. Shinde, Ninad B. Velhal, G.M. Lohar, Vijay J. Fulari . Synthesis and characterization of ZnO thin film by low cost modified SILAR technique. AIMS Materials Science, 2016, 3(2): 349-356. doi: 10.3934/matersci.2016.2.349
    [2] Nur Jassriatul Aida binti Jamaludin, Shanmugan Subramani . Thermal performance of LED fixed on CVD processed ZnO thin film on Al substrates at various O2 gas flow rates. AIMS Materials Science, 2018, 5(2): 246-256. doi: 10.3934/matersci.2018.2.246
    [3] Johana Gamez, Luis Reyes-Osorio, Oscar Zapata, Roberto Cabriales, Luis Lopez, Miguel Delgado-Pamanes . Study of protective hard coatings of SiO2-TiO2 on aluminum substrates. AIMS Materials Science, 2024, 11(2): 200-215. doi: 10.3934/matersci.2024011
    [4] Xuan Luc Le, Nguyen Dang Phu, Nguyen Xuan Duong . Enhancement of ferroelectricity in perovskite BaTiO3 epitaxial thin films by sulfurization. AIMS Materials Science, 2024, 11(4): 802-814. doi: 10.3934/matersci.2024039
    [5] Keith E. Whitener Jr . Rapid synthesis of thin amorphous carbon films by sugar dehydration and dispersion. AIMS Materials Science, 2016, 3(4): 1309-1320. doi: 10.3934/matersci.2016.4.1309
    [6] Ravipati Praveena, Gottapu. Varaprasada Rao, Karna Balasubrahmanyam, M. Ghanashyam Krishna, Vinjanampati Madhurima . Optical and water repellant properties of Ag/SnO2 bilayer thin films. AIMS Materials Science, 2016, 3(1): 231-244. doi: 10.3934/matersci.2016.1.231
    [7] Sree Sourav Das, Zach Fox, Md Dalim Mia, Brian C Samuels, Rony Saha, Ravi Droopad . Demonstration of ferroelectricity in PLD grown HfO2-ZrO2 nanolaminates. AIMS Materials Science, 2023, 10(2): 342-355. doi: 10.3934/matersci.2023018
    [8] Mihaela Girtan, Laura Hrostea, Mihaela Boclinca, Beatrice Negulescu . Study of oxide/metal/oxide thin films for transparent electronics and solar cells applications by spectroscopic ellipsometry. AIMS Materials Science, 2017, 4(3): 594-613. doi: 10.3934/matersci.2017.3.594
    [9] Nur Jassriatul Aida binti Jamaludin, Shanmugan Subramani, Mutharasu Devarajan . Thermal and optical performance of chemical vapor deposited zinc oxide thin film as thermal interface material for high power LED. AIMS Materials Science, 2018, 5(3): 402-413. doi: 10.3934/matersci.2018.3.402
    [10] Larysa Khomenkova, Mykola Baran, Jedrzej Jedrzejewski, Caroline Bonafos, Vincent Paillard, Yevgen Venger, Isaac Balberg, Nadiia Korsunska . Silicon nanocrystals embedded in oxide films grown by magnetron sputtering. AIMS Materials Science, 2016, 3(2): 538-561. doi: 10.3934/matersci.2016.2.538
  • V2O5 is one of the best material for many applications. Progress is currently made to improve its performance for use as a sensor, or an electrode, or smart window, electrochromic device, supercapacitor, photovoltaic applications among others. In this work, we review the progress that has been done these recent years, in relation to the mode of preparation of the V2O5 films. The results outline the complex relationship between the synthesis and the properties, which should serve as a guide for further research on the dependence of the best synthesis process on the type of application that is targeted.


    1. Introduction

    Vanadium is sparsely distributed in the earth's crust. It is found at an average concentration of 150 mg kg–1 in mineral ores and in the range of 3–300 mg kg–1 in soil. Vanadium pentoxide, the highest oxidation state 5+ of vanadium, has been known for more than a century. In December 1867, as a part of his Bakerian lecture to the Royal Society, the British chemist Sir Henry Roscoe outlined the multiple oxidation states of binary oxides of vanadium and set the end-member "vanadic acid" V2O5 [1]. The first V2O5 gel was described by the French chemist Ditte in 1885 [2]. The natural occurrence (50%) of V2O5 is found in flue-gas deposits from oil-fired furnaces and residues from elemental phosphate plants. However, V2O5 mineral is rare; shcherbinaite V2O5 (orthorhombic crystal system) is found on the walls of volcanic fissures [3] and navajoite is a mineral trihydrate V2O5·3H2O (monoclinic structure). The titanoferous magnetite ore in lump form contains approximately 1.5–1.7% vanadium pentoxide.

    The structure of crystal isolated from fused V2O5 was first elucidated by Katelaar [4] as a rhombic unit cell containing two molecules of V2O5. Refinements were further done by Byström et al. [5] who established the orthorhombic layered structure of V2O5 with Pmmn space group. In 1961, Bachmann et al. [6] redetermined the crystal structure of V2O5 and reported its lattice parameters a = 11.510 Å, b = 4.369 Å, c = 3.563 Å. Their structural description is as follows: "the structure is built up from distorted trigonal bipyramidal coordination polyhedral of O around V, which share edges to form zigzag double chains along [001] and are cross-linked along [100] through shared corners, thus forming sheets in the zz plane". More recently, Enjalbert and Galy [7] proposed the bonding interactions of van der Waals type between two-dimensional V2O5 slabs and stated the presence of "square pyramid" instead of trigonal bipyramid. Ramana et al. [8] and Julien et al. [9] have described the easy cleavage along the (001) plane due to the weak interaction between V2O5 layers. Three V2O5 polymorphs have been identified as α-, β-V2O5, which crystallize with the orthorhombic and tetragonal structure, respectively, and the rutile-type δ-V2O5, which is a modification of β-V2O5 (space group C12/c1) [10]. Recently, Parija et al. [11] evaluated the metastable polymorphs including γ'-, δ'- and ρ'- phase of V2O5 using density functional theory calculations.

    Here, we review the properties of V2O5 thin films employed in energy storage and conversion systems, which were prepared with a variety of deposition options. Numerous works prior 2001 have been devoted to the studies of synthesis, physical characterizations and electrochemical performances of films used as either cathode in all-solid-state batteries [12,13,14,15] or electrode of electrochromic devices [16]. Recently, Mo-doped V2O5 thin film has been studied as an electrode of supercapacitors [17]. This is an illustration of the increasing interest in this material in the recent years for many applications, sustained by numerous works that are reviewed here.

    This review paper is organized as follows. In Section 2, we summarize the techniques used for the preparation of films and discuss the structure and morphology of V2O5 related to the synthesis conditions (temperature, partial pressure, substrate, etc.). Section 3 is devoted to the physical properties of V2O5 thin films; structure, morphology, vibrational spectroscopy, elemental analysis, electrical properties and intercalation process are considered. In the following Sections 4 and 5, doped V2O5 thin films and composite films are treated, respectively. Finally, Section 6 is devoted to the applications in the field of energy, for which V2O5 thin films are used as electrode materials in lithium microbatteries, electrochromic devices, sensors and supercapacitors.


    2. Thin film synthesis

    The structure and morphology of vanadium oxide films are intimately related to the deposition method and the operating conditions. As the direct growth of crystalline V2O5 films is very difficult except in the cases of some sub-stoichiometric VOx oxides, an annealing process of the as-prepared films is required in air and high temperature. In this section we describe the different evaporation methods and provide some typical examples for each technique used. Note that the choice of a deposition method depends of the thin film application. For instance, electrodeposition [18], reactive sputtering [19], sol-gel [20], hydrothermal method [21], doctor-blade route [22] were carried out for electrochromic V2O5 films that requests large surface coatings. The reader will find supplementary data in the review by Beke [23]. During the deposition process in vacuum or in a reducing atmosphere, the removal of oxygen atoms occurs from the film network when V2O5 is heated above its melting point that induces the formation of defects or reduced VOx phases. Consequently, the structural and morphological disorders and phase instability can be controlled using appropriate deposition parameters.


    2.1. Thermal deposition

    The thermal evaporation technique is the simplest method based in the production of flux of vapor in a high vacuum chamber (pressure 100 mPa) to form thin films without the presence of catalyst. In this method, the substance is evaporated by means of resistive heating. Generally, the raw powder is placed in a molybdenum boat heated at Tboat. The modifications of the structure, stoichiometry and morphology of the deposited films are obtained by varying the substrate temperature Ts, the flow of reactive gases (in cm3 min–1 at standard pressure and temperature, expressed as "sccm" hereafter) and the duration of the target evaporation. For example, thermally evaporated polycrystalline V2O5 films (10–20 nm grain size) were fabricated using Tboat = 650 ℃ for 6 h with gas flow 13 sccm Ar + 50% O2 [24]. The nature of the substrate is an important factor to obtain films with preferential orientation [25]. V2O5 films deposited on silicon (111) wafers by vacuum thermal evaporation were amorphous when deposited at Ts ≤ 200 ℃, while polycrystalline at Ts ≥ 300 ℃. This later temperature is optimum for V2O5 films strongly oriented with (001) planes parallel to the substrate [26]. Nanostructured V2O5 thin films (25 nm grain size) thermally deposited onto Ni substrates at Ts = 300 ℃ show preferential (001) orientation. These films have pseudocapacitance of 730 mF cm–2 at current density of 1 mA cm–2 and charge transfer resistance of 7.5 Ω [27]. The thermally evaporated V2O5 films at Ts = 25 ℃, 100-nm thick, crystallized after annealing process at 500 ℃ with grain size 26 nm and exhibited an electrical conductivity of 5.5 S cm–1 (activation energy of 0.16 eV) [28] and optical bandgap of 2.8 eV.


    2.2. Flash-evaporation

    In the flash-evaporation method, the powders are placed in a reservoir and poured drop by drop into the boat heated at Tboat to ensure the vaporization. This process allows the control of the vaporization rate and preserves any decomposition of the starting material before evaporation [29]. It was demonstrated that V2O5 flash-evaporated films are more homogeneous [30]. Flash-evaopration was also used for depositing LixV2O5 [31]. Polycrystalline V2O5 flash-evaporated films have been investigated as a function of the deposition conditions: various substrates, substrate temperature (Ts), oxygen partial pressure (pO2) and post-annealing treatment (Ta) [9]. Figure 1 presents the evolution of the O/V ratio as a function of the temperature of the molybdenum boat [32]. Films with the best stoichiometry O/V = 2.497 were obtained for Tboat in the range 910–980 ℃. However, properties of flash-evaporated films are strongly dependent on the deposition "quenching rate" ΔT, which is the difference in temperature between the melt and the substrate (ΔT = TaTs).

    Figure 1. Evolution of the O/V ratio in flash-evaporated V2O5 films as a function of the boat temperature (From Ref. [32]).

    2.3. Chemical vapor deposition (CVD)

    Chemical vapor deposition (CVD) is a versatile technique that consists of an evaporator and a deposition system. The precursor solution is vaporized into a reactor which acts as a flash evaporation system. The vapor condenses in a cold-wall reactor in which the substrate placed in the central zone is maintained at constant temperature Ts [33]. This technique has been successfully carried out to grow V2O5 films using pure or diluted triisopropoxyvanadium oxide VO(OC3H7)3 precursor [34]. Typical V2O5 film 240 nm thick was formed at Ts = 300 ℃ in the total pressure pt = 2400 Pa with a flow of O2 of 100 sccm. V3O7 and V4O7 films are deposited when pt decreases to 1200 Pa, while V6O13 films are formed at Ts = 350 ℃ and O2 flow rate of 120 sccm. Highly crystallized films (94 nm crystallite size) were obtained after post-annealing at 500 ℃ for 2 h in O2 atmosphere with a slow cooling process to insure a good adherence on substrate. Plasma-enhanced chemical vapor deposition (PE-CVD) is a method for large scale film deposition. In this technique, glow discharge creating high-energy electrons ionize gaseous molecules and generate chemically reactive ions. Heat-treated (500 ℃ for 2 h) films deposited by CVD technique onto Pt foil are highly structured with homogeneous and smooth surfaces as shown in Figure 2a. The coherence domains along the a-and c-axis are L200 = 71 nm and L001 = 36 nm, respectively. The cross-section image (Figure 2b) displays the compactness of annealed V2O5 films [35].

    Figure 2. Morphology of CVD V2O5 films deposited onto Pt substrate. (a) SEM image of film annealed at 500 ℃ for 2 h and (b) cross section (from Ref. [35]).

    Barreca et al. [36] prepared V2O5 thin films by PE-CVD using VO(hfa)2·H2O (Hhfa = 1, 1, 1, 5, 5, 5-hexafluoro-2, 4-pentanedione) as precursor. This precursor was vaporized at 70 ℃ at the rate of 2 × 10–4 mmol m–2 s–1 in a reactor in which argon (constant flow rate ϕ = 40 sccm) and oxygen (flow rate 5 ≤ ϕ ≤ 20 sccm) were plasma sources. V2O5 thin films deposited at Ts = 200 ℃ and 10 sccm of O2 are highly textured (14 nm crystallite size) and grow with the (001) preferential orientation. In the metal organic chemical vapor deposition (MOCVD) the organometallic vapor phase takers place at moderate pressure (10–1000 hPa). Watanabe et al. [37] prepared V2O5 thin films by means of microwave plasma MOCVD on ITO-coated fused silica substrate. Bis-acetylacetonatovanadyl, VO(acac)2, VO(C5H7O2)2 heated at ~600 ℃ was selected as vanadium precursor. Its vapor was injected into the oxygen plasma generated by the microwave discharge. Other experimental conditions were as follows: the substrate was maintained at Ts ≈ 300 ℃, the O2 flow was ϕ = 1.2 dm3 h–1 under pressure of 650 Pa. Typical polycrystalline film, 120 nm thick, was deposited after 15 min. Vanadyl(Ⅳ) β-diketonate has been also used with water in a low-pressure reactor under different conditions [38]. A novel vanadium(Ⅲ) precursor such as vanadium(Ⅲ) alkoxide [V(OCMe2CH2OMe)3] was also used, which presents an appreciable volatility at 55 ℃ under pressure of 200 Pa [39]. Strongly (00l)-oriented film composites of V2O5–V6O13 were grown at temperatures ≥560 ℃ onto fused quartz substrate using vanadyl acetylacetonate as precursor. Single phase was obtained for substrate maintained at Ts = 580 ℃ due to the reentrant-type growth behavior [40]. V2O5 electrochromic films were grown on flexible polymer substrates using plasma-enhanced chemical vapor deposition (PECVD). Films were deposited at high rate of 50 nm min–1 from the decomposition of vanadium oxytrichloride (VOCl3) and O2 [41]. Nandakumar [42] studied the growth rate rG of CVD V2O5 films as a function of Ts and reported an Arrhenius behavior with 0.14 eV activation energy; rG = 50 nm min–1 at Ts ≈ 230 ℃.


    2.4. Magnetron sputtering

    Sputtering deposition method is the most popular technique to grow metal-oxide films, since it allows faster deposition rates. Its main advantage comes from the production of good surface uniformity of as-deposited films. It includes radio-frequency (r.f.) magnetron sputtering [43,44,45,46], direct current (dc) magnetron sputtering [47,48,49] and ion beam sputtering [50]. The stoichiometry of V2O5 films can be tuned by this evaporation method [51,52]. Typical r.f.-magnetron sputtered V2O5 films are formed in a plasma sputtering chamber using V2O5 target bombarded by argon ions in a plasma of power Pw = 150–300 W. For specific application, low sputtering power Pw < 100 W can be used [53]. The reactive deposition is realized by injection of pure argon and oxygen gases, at flow rate ϕ < 10 sccm. Silversmit et al. [51] obtained V2O5 films on Si (100) wafers by reactive dc magnetron sputtering and studied the influence of deposition parameters on their stoichiometry. These experiments were carried out in a vacuum of 5 × 10–5 Pa with a magnetron discharge voltage at ~550 V and a plasma constant current of 150 mA. The O2 flow was maintained at ϕ = 3.5 sccm. Fateh et al. [54] deposited V2O5 films onto (100) oriented Mg substrates and established detailed synthesis-structure relationships showing the best crystallinity for the film deposited at 80 ℃. Lourenço et al. [55] investigated the effect of different oxygen flow rates on the structure of VOx films and determined that V2O5 films are obtained at high oxygen flows (ϕ > 9 sccm). Benmoussa et al. [56,57] sputtered V2O5 thin films on Corning glass and ITO-coated glass substrates for electrochromic application. The films deposited on both substrates are c-axis preferred oriented showing that texture is substrate independent.

    Yoon et al. [58] compared the structure of V2O5 thin films grown on (100) Si wafers by d.c. and r.f. reactive sputtering at room temperature; d.c.-sputtered films are amorphous, while those prepared from reactive r.f.-sputtering crystallized with large grain size that is due to the self-bias effect. Gianneta et al. [46] deposited VOx films using a base pressure 1 mPa, a r.f. plasma power of 200 W and an Ar flow ϕ = 20 sccm. As-deposited films are almost amorphous and exhibit broad XRD reflections matching those of VO0.9, while annealing at 475 ℃ in air for 3 h promotes the well-crystallized V2O5 phase with well-developed orthorhombic shape and grain size > 100 nm (see Figure 3). Ottaviano et al. [59] reported the improved electrochromic properties of rf reactive sputtered films deposited at low O2 flow (2%) showing the highest value of injected charge (49.8 mC cm–2) and the greatest differences in optical density. Lin and Tsai [19,60] prepared V2O5–z films on flexible PET(polyethylene terephthalate)/ITO (indium tin oxide) substrates. Oxygen deficient V2O5–z films (z = 0.13) were formed in a chamber pressure of 6 Pa with Ar and O2 flow rate of 4 sccm that show a transmittance variance od 36.5% after 200 cycles. Kang et al. [61] prepared nanorod-like V2O5 film by electron-beam irradiated amorphous film, which were obtained by r.f. sputtering at power of 200 W at 1 nm min–1 growth rate on alumina substrate.

    Figure 3. XRD diffractogram for r.f. sputtered V2O5 films (a) as-deposited at 25 ℃ and (b) thermally annealed at 475 ℃ in air for 3 h (from Ref. [46]).

    2.5. Pulsed-laser deposition (PLD)

    Extensive works using the PLD technique have been successful for the growth of V2O5 thin films. Advantages include an easily control of the film composition by tuning the deposition parameters and a good reproducible stoichiometry of the target material in the films [62,63,64,65,66,67,68,69,70,71,72,73]. In the PLD technique, a pulsed laser beam (10 ns duration) is focused by a lens to ablate the V2O5 target. The energy of the beam is in the range 100–500 mJ per pulse (laser fluence of 10 J cm–2) with a laser pulse repetition of 10 Hz. Target and substrates are placed inside the deposition chamber evacuated to ~1 mPa. To avoid depletion of deposit at any given spot, the target rotates at 10 rpm. For reactive synthesis, pure oxygen gas is introduced into the chamber with a typical partial pressure (pO2) in the range 0.1–10 Pa to obtain single-phased and stoichiometric V2O5 films. Typical deposition rate is in the range 2–5 Å s–1. Various laser wavelengths are reported in the literature: 532 nm line of a doubled frequency pulsed Nd:YAG laser [63], 248 nm line of a KrF excimer UV laser [62,64] or the 266 nm line of the quadrupled Nd:YAG laser. Zhang et al. [62] prepared PLD crystalline oriented V2O5 films deposited at 200 ℃ in different partial pressure of O2 + Ar mixed gases. The films grown in oxygen-rich, pO2 = 0.1 Pa, environment at low substrate temperature Ts ≈ 200 ℃ crystallizes in the orthorhombic structure. Iida et al. [65] suggested that films deposited at Ts = 350 ℃ in a chamber under partial pressure pO2 = 13.3 Pa using the 238-nm line of a KrF excimer laser have a surface morphology suitable for electrochromic application. The same process was used to grow doped films, with various elements Nb, Ce, Nd, Dy, Sm, Ag [66]. Fang et al. [74] prepared orientated V2O5 electrochromic thin films on glass by pulsed excimer laser ablation. Nanocrystalline films highly oriented along the c-axis were obtained at lower temperature Ts (200 ℃) that that of thermal vacuum deposition (300–500 ℃). McGraw et al. [67] found that the crystallographic texture of PLD V2O5 films depends mainly on temperature and oxygen pressure rather than on the choice of substrate. Films deposited on SnO2/glass substrate are dense and phase pure orthorhombic V2O5. For Ts = 200 ℃ and pO2 = 2.7 Pa, the films are well (200)-textured, while they are (001) and (101) dual-oriented for Ts = 500 ℃ and pO2 > 12 Pa. Ramana et al. [75] investigated the structural patterns of PLD V2O5 films which revealed that stoichiometric specimens are well-crystallized with the layered orthorhombic structure even onto amorphous substrates at Ts = 200 ℃ under partial oxygen atmosphere pO2 = 10 Pa. Atomic force microscopy (AFM) analysis gave a consistent picture of the surface morphology and microstructure: grains of small size, spherical in shape. Figure 4 shows the variation of the grain size as a function of the growth temperature for various substrate materials. The evolution of the grain size follows an exponential law (dashed lines) as [76]:

    Lc=L0exp(Qd/kBT) (1)
    Figure 4. Evolution of the grain size of PLD V2O5 films with growth temperature during deposition onto amorphous glass, ITO-coated glass and silicon substrates [64].

    where Qd is the activation energy, kB is the Boltzmann constant, T is the absolute temperature, and Lo is a preexponential factor that depends on the physical properties of the substrate-deposit. The grain size appears to be 50–300 nm for films deposited onto glass, 50–400 nm for films on ITO-coated glass and 60–600 nm for films evaporated on Si wafer. Qd of the PLD V2O5 film in the range 0.43–0.55 eV reflects the nucleation rate [64]. AFM images show that films deposited onto Si (8 nm) and ITO-coated glass (10 nm) exhibited lower surface roughness. The dependence of the substrate temperature on the gross microstructure of PLD V2O5 films is presented in Figure 5. The domain of optimum controlled morphology is illustrated by the shaded region of substrate temperature range 200 ≤ Ts ≤ 300 ℃.

    Figure 5. Diagram mapping the effect of substrate temperature on the microstructure evolution of PLD V2O5 thin films. The domain of optimum controlled morphology is illustrated by the shaded region of substrate temperature range 200 ≤ Ts ≤ 300 ℃ [76].

    2.6. Electron-beam (e-beam) evaporation

    The e-beam evaporation method implies a high vacuum coating unit with vacuum better that 0.1 mPa, in which an electron beam accelerated by a voltage of several kV (typically 6 kV) and power density of ~1.5 kW cm–2 is scanned on the surface of the target. It is a low-cost technique, which allows high deposition rate (30–50 Å s–1) with good control of the structure and morphology, and allows for sequential and co-deposition with minimum contamination [8,77,78,79,80,81,82,83,84].

    In their early work, Ramana et al. [81] reported the growth of non-stoichiometric V2O5 films deposited on Corning 7059 glass substrates maintained at Ts = 280 ℃ in a vacuum of 0.05 mPa, while the injection of pure oxygen (partial pressure of 0.1 mPa) produced near-stoichiometric films with pale orange yellow color. The XPS experiments revealed that films formed at room temperature are nearly stoichiometric, while the films formed at elevated temperatures are oxygen deficient. These results are supported by the shift of the vanadyl mode in the FTIR and Raman spectra. However, the films deposited at Ts ≈ 420 ℃ in oxygen partial pressure of 10–4 mbar led to vanadium ions in their highest oxidation state [8,79]. Alumina-doped V2O5 thin films, (V2O5)100–x(Al2O3)x with x = 5, 10, 15, 20 wt.%, were fabricated by e-beam evaporation method [85]. Thiagarajan et al. [84] studied the physical properties of highly oriented V2O5 films (0.8–1.2 µm thick, 31–45 grain size) grown by electron beam evaporation at 200 ℃ throughout all depositions. It was shown that the microstrain decreases with the decrease of the film thickness, due to the reduced dislocation density.


    2.7. Spray pyrolysis

    Spray pyrolysis is considered as a cost-efficient technique and has been widely employed to fabricate V2O5 thin films. Currently, the pyrolysis process is performed at relatively low temperature > 500 ℃ [86,87,88,89,90]. The V2O5 films are deposited through thermal decomposition of vanadium precursor such as VCl3 in deionized water [91]. By increasing the spray deposition rates, the crystallite size increases, while the microstrain decreases. Nanostructured flower-like V2O5 (crystallite size of ~25 nm) were prepared by spray pyrolysis from 0.1 mol L–1 ammonium vanadate in aqueous solution. Polycrystalline films deposited at Ts = 300 ℃ exhibit the largest carrier density 3.6 × 1018 cm–3 and the lowest electrical resistivity 5.56 Ω cm–1, which allow to use them as xylene sensors [92]. 10% Mo-doped films (~28 nm crystallite size) prepared by spraying VCl3 and MoCl5 in aqueous solution on substrate maintained at 380 ℃ crystallize with a tetragonal structure [93]. Wei et al. [90] prepared V2O5 films using ultrasonic spray method that revealed a small amount of V4+ species. These films heat-treated at 350 ℃ show good charge storage stability over 10000 electrochromic cycles. A d.c. plasma torch was used to atomize the solution VOCl3 precursor, which is injected externally into the plasma flame and deposited on the substrate [88]. Nanostructural V2O5 thin films deposited by spray pyrolysis technique exhibit the highest transmittance when deposited at 550 ℃. A shift of the absorption edge from 2.5 to 2.8 eV is observed for Ts > 450 ℃ due to modified chemical bonds at the film/substrate interface [89]. The influence of the molarity of the spray pyrolysis precursor on the structure and morphology of V2O5 films shows that vanadium nitrate V(NO3)5·5H2O with concentration ≥0.1 mol L–1 provides orthorhombic structured films with (001) preferred orientation [94].


    2.8. Sol-gel

    Sol-gel method is a low-temperature wet-chemistry method in which the raw materials are mixed at the molecular level. Wang et al. [95] investigated the structural modifications of spray deposited V2O5·nH2O xerogel films as a function of post-annealing temperatures. Films sprayed at 150 ℃ showed broad peaks in their 51V NMR spectra for heat-treatment at Ta ≈ 120 ℃, which gives evidence of a distortion of the square pyramids. The structural feature of V2O5·nH2O (with structural water) favors the lithium intercalation between the slabs with respect to contracted dried films at Ta > 250 ℃. The effect of deposition conditions and post-annealing treatment of xerogel V2O5 films was studied by several groups. Najdoski et al. [96] examined the electrochromic properties of (NH4)0.3V2O5·1.25H2O nanostructured films deposited onto SnO2:F (FTO) as a function of the temperature of preparation. Elongated aggregates 250–500 nm length were obtained at 50 ℃, while shorter particles (50–300 nm) were obtained at 85 ℃. Chen et al. [97] prepared thin films by spin coating of the sol-gel composed by the mixture of V2O5 powders benzyl alcohol (BA) and isobutanol (IB) (molar ratio 1:40:4) heated at 110 ℃ for 5 h. Wang et al. [98] fabricated V2O5·nH2O xerogel films by reacting V2O5 and H2O2 under ambient conditions. V2O5·⅓H2O film exhibited the best Li+ intercalation properties with a stabilized specific capacity of 185 mAh g–1 at 0.1 mA cm–2 current density after 50 cycles. Cazzanelli et al. [99] analyzed the Raman spectra of xerogel films intercalated with lithium. V2O5·nH2O xerogel films (1.8 ≤ n ≤ 0) were spin-coated at angular velocity of 1500 rpm. Spectral features of Li-intercalated samples show significant effect of the water content in the interlayer space inducing new Raman bands at 800 and 950 cm–1. V2O5 films deposited by the doctor-blade route using polyol dispersed in waterand acetyl acetone crystallized with the C2/c structure, which transformed to orthorhombic Pmmn (86 nm crystallite size) after calcination at 170 ℃ for 3 h [22]. Dip-coating method was used to produce V2O5 xerogel films from organic (vanadium tri-isopropoxide into isopropyl alcohol) and inorganic (V2O5 powers in 15 wt.% H2O2) precursors [100,101]


    2.9. Electrodeposition

    A typical electrodeposition experiment takes place in an aqueous solution of vanadium precursor such as vanadyl sulfate hydrate (VOSO4·nH2O) or V2O5 powers dissolved in hydrogen peroxide as electrolyte using a constant potential (1.7 V vs. Ag/AgCl) [18,102,103]. Liu et al. [104] prepared nanostructured V2O5 films by means of cathodic deposition from V2O5 and H2O2 aqueous solution. As-deposited samples contained 14% V4+, while annealing at 500 ℃ in air oxidized all vanadium cations. Electrochromic V2O5 films were synthesized by electrophoretic deposition of a sol formed by the dissolution of V2O5 powers in H2O2 solution with the H2O2/V2O5 ratio of 8:1. The deposition occurred in a Teflon vessel at the voltage of 5 V using a Pt rod as counter electrode [102]. The electrodeposition experiments carried out by Vernardou et al. [18] consisted in the coating of V2O5 powders dissolved in a solution of methanol and water as electrolyte to form an orange suspension. Pt rod was the counter electrode. Experiments were performed using deposition current density of 0.25 mA cm–2.


    2.10. Atomic layer deposition (ALD)

    The atomic layer chemical vapor deposition (ALCVD) was successfully applied to grow V2O5 films [105,106,107,108,109,110,111,112]. This method provides uniform and compact films with a high control of phase formation and the film thickness that include atomic layer epitaxy (ALE) and atomic layer deposition (ALD). ALD process consists in alternative exposition of a substrate to the precursor fluxes, which are timely separated in space. In principle this process is self-limiting as the monomolecular layers are absorbed with molecular excess flushed away by evacuation of the system. The vapor pressure of the precursor must agree with the deposition temperature chosen. Groult et al. [105,106] prepared nanometric V2O5 thin films deposited onto Si wafers (100) by means of ALCVD.

    The atomic layer deposition (ALD) technique is proven suitable for the growth of uniform and pin-hole free V2O5 thin films. ALD is becoming an attractive process to prepare thin films for energy storage, because it allows the control of deposit films of thickness in the range 1–100 nm with remarkable uniformity and control of the thickness. V2O5 has been synthesized by this process by several authors [113,114,115,116,117,118,119]. This scalable method, which allows to do bath processes, is based on self-limited reactions between gas phase precursors and active sites on surfaces. This technique leads to monolayer thickness control. The growth of V2O5 thin films consists of the successive cycles of deposition of atomic layers from the reaction of a vanadium precursor with oxygen. Various vanadium precursors have been proposed such as vanadyl tri-isopropoxide VO(OC3H7)3 (VO(OPri)3, VTIP), vanadyl triethoxide (VO(OEt)3), vanadyl triisobutoxide (VO(OBui)3), vanadyl acetylacetonate (VO(acac)2) and vanadium acetylacetonate (V(acac)3 (acac = C5H7O2)), VTIP being the most popular due to its high vapor pressure (38 Pa at 45 ℃) and its rapid reaction with water [107,108,109]. Examination of the literature shows that the choice of ALD precursors has a significant impact on the crystallinity and morphology of V2O5 films. One ALD cycle is composed of four pulses in the sequence t1/t2/t3/t4, where ti is the duration (in s) of injection of V precursor, purge, reactive gas, purge, respectively. For instance, the global ALD reaction of VTIP precursor with water is:

    VO(OC3H7)3+3H2OV2O5+6HOC3H7 (2)

    The pressure of (VTIP) in the reactor is adjusted to 5 mPa and the water pressure to 0.3 Pa. Typical cycle to grow V2O5 films lasts 17 s (2 s VTIP, 5 s pumping, 5 s reactive gas and 5 second pumping) that makes a growth rate of 0.7 Å per cycle. Several authors demonstrated that the rate-determining step is the oxidant in O3-based or H2O-based ALD reaction [109,110]. Keränen et al. [107] prepared ALD V2O5 films deposited onto various metal oxides for catalytic applications using VO(acac)2 as V precursor with a N2 flow of 3 dm3 h–1. Chen et al. [111] used a water-free system to prepare crystalline as-deposited V2O5 films from VTIP and ozone, using an Al2O3 template at low temperature (170–180 ℃). With this process, a deposition rate of 0.27 Å per cycle onto Si wafer was achieved. Badot et al. [110] reported a growth rate of ~20 ng cm–2 per cycle at 170 ℃ using VTIP + water sequence. With these growth conditions and annealing at 400 ℃, the ALD V2O5 films were well-crystallized and preferentially oriented with the (a, b) planes parallel to the SnO2-coated glass substrate. ALD V2O5 films deposited onto Ti foil using the VTIP-water vapor route in the pulsing sequence—VTIP/N2 purge/H2O vapor/N2 purge of 200/400/2400/1000 ms and annealed at 500 ℃ for 2 h showed a small-polaron drift mobility of 1.84 × 10–2 cm2 V–1 s–1 at room temperature (activation energy of 0.14 eV) [112]. Ostreng et al. [113] prepared nano-structured ALD V2O5 films from vanadyl 2, 2, 6, 6-tetramethylhepta-3, 5-dione (VO(thd)2) and ozone at 215 ℃ with N2 as carrier gas (500 sccm flow) applied as cathode films for lithium microbatteries. 86-nm thick films oriented along a (001)/(010) orientation were thus obtained after 2000 cycles. Chen et al. [114] employed VTIP precursor with both O3 and H2O as oxidant. For ALD cycles operating in the 170–185 ℃ window, the water-VTIP route (growth rate 0.4 Å per cycle) produces amorphous films, while crystalline films are formed with the ozone-VTIP reaction (growth rate 0.20 Å). The transition from amorphous to crystalline structure was characterized by rectangular V2O5 nanograins (20 nm size) with a monomodal distribution starting at 400 ℃.

    A comparison of thermal and plasma-enhanced ALD/CVD has been carried out by Musschoot et al. [109]. Chen et al. [114] determined that areal energy and power density is optimized with V2O5 thickness round 60 nm. Larger thickness results in a decrease of the rate capability due to the small intrinsic diffusion coefficient of lithium, while thinner films suffer from a decrease of material loading. The authors estimated that the device would outperform the gravimetric power density of the current Li-ion batteries by one order of magnitude. Song et al. [114] adopted to deposit V2O5 thin films by ALD using vanadyl tri-isopropoxide and water and showed remarkable electrochemical response and voltage-induced insulator-to-metal transition. Ultrathin V2O5 films (10–90 nm thick) were prepared by ALD using vanadyl acetylacetonate as the vanadium precursor [119]. Crystalline films were obtained below 200 ℃ yielding a growth rate of 4.5 pm per cycle. Heterojunction diodes based on TiO2(p)-(n)V2O5 were fabricated as humidity sensor.


    3. Characterizations of V2O5 films


    3.1. Structure and morphology

    V2O5 is the highest oxidation compound of the vanadium oxide system which crystallizes with a lamellar orthorhombic structure in space group Pmmn (D2h13, No. 59) with lattice parameters a = 11.512 Å, b = 3.564 Å, c = 4.368 Å and V = 179.17 Å3 (see Table 1). It has a layered structure with atomic layers in the [100] × [010] plane and van der Waals-type interlayer coupling. The structure is formed of three types of oxygen atoms: apical-vanadyl (Ovan), bridge (Obri) and chain (Och) (see Figure 6). Oxygen vacancies in V2O5–δ are predominantly of the Ovan type, forming the vanadyl bonds. Atomic positions in V2O5 are summarized in Table 1. Pure V2O5 is a diamagnetic semiconductor with an energy gap (Eg) of 2.2–2.3 eV [120], while oxygen-deficient material is ferromagnetic with spin moment ~2 µB per vacancy [121]. A density functional calcualtion of the electronic structure is given in Refs. [122,123]. The role of octahedral deformations has been explored in [124].

    Figure 6. Structure of orthorhombic V2O5 with netplane stacking along the (010) direction that shows the vanadium and inequivalent oxygen singly coordinated apical vanadyl-Ovan, doubly coordinated bridge-Obri, and triply coordinated chain-Och atoms (from Ref. [123]).
    Table 1. Atomic positions in orthorhombic V2O5 (space group Pmmn) (from Ref. [124]).
    Atom Wyckoff position Parameters
    x y z
    V 4f 0.10118 0.25 –0.1083
    Ovan 4f 0.10430 0.25 –0.469
    Obr 2a 0.25 0.25 0.001
    Och 4f –0.0689 0.25 0.003
     | Show Table
    DownLoad: CSV

    Figure 7 shows the typical XRD patterns of V2O5 films deposited by PLD (a) and (b) by electron beam (EBD) techniques. The spectra of samples deposited at low Ts are rather amorphous, while they can be obtained crystallized by raising the substrate temperature and/or increasing p(O2). The significant feature is the growth of XRD reflections at ca. 2θ = 20.26° and around 2θ = 41.4° that correspond to the (001) and (002) lattice planes of the orthorhombic structure, respectively. The (001) reflection is the dominant line for well-crystallized films showing the preferential c-axis orientation lying perpendicular to the substrate with stacking ab planes [125]. Gradual increase in intensity of the (101) reflection at high Ts is attributed to a re-organization process of the in-plane V–O–V chains in flash-evaporated film. This crystallographic behavior seems to be the general trend for the growth of V2O5 films; it has been found for various deposition techniques: as-prepared r.f. sputtered films under a gas mixture of Ar and O2 [56,126], sol-gel spin-coated films annealed at 400 ℃ [127], pulsed-laser deposited V2O5 films [63], plasma-enhanced CVD method [128] and vapor deposition in high vacuum [26,129] on heated substrates. The c-axis preferred growth was also found for films prepared by ALD [110,113] and EBD [79] processes. Well-ordered V2O5-(001) films were obtained by the oxidation of vacuum deposited vanadium layers with oxygen partial pressure of 50 hPa [130]. Due to the imperfect stacking of ab planes built up of Ovan, the (001) orientation for crystalline V2O5 films deposited by PECVD from a vanadyl(IV) β-diketonate compound appears at Ts as low as 150 ℃ [38], while higher temperature Ts = 240 ℃ for films prepared by CVD of VOCl3 with water [131]. Note that films prepared from sol-gel precursor display complex behavior due to the large fraction of water content. For instance, the XRD spectrum of spin-coated films obtained from V2O5 + H2O2 sol-gel (crystallized at Ta = 150 ℃) displayed the dominant (110) Bragg line, which disappeared and was replaced at Ta = 300 ℃ by the (00l) reflections showing the nucleation parallel to the substrate [121]. Other workers found that spin-coated films crystallize above 350 ℃ [132]. The texture of V2O5 films grown by ALD technique varies with the number of deposition cycles (nc). For nc < 1000, the predominant orientation is along the (001) plane, which is modified for nc ≈ 2000 with the appearance of sharp (0k0) Bragg reflections. Finally (021) and (061) reflections are observed for nc ≈ 5000. This is attributed to the overgrowth of nuclei on the surface of platelets with their edges orientated normal to the substrate [113]. Similar thickness (df) effect was observed on e-beam deposited films [84]. Moreover, the increase of the peak broadening of the (001) lattice place with df is attributed to microstrains inversely proportional to the crystallite size that linearly decreases with the increase in df.

    Figure 7. XRD patterns ox deposited V2O5 films. (a) PLD films grown at various substrate temperatures [75]. (b) EBD films grown using various oxygen partial pressures [79].

    Rajendra-Kumar et al. [133] deposited V2O5 films at various substrate temperatures (25 ≤ Ts ≤ 400 ℃) using vacuum evaporation technique. The orthorhombic lattice parameters "a" and "c" of polycrystalline films determined by Nelson-Rieley function were found to decrease when the deposition temperature increases (Figure 7a). According to the model proposed by Gilles and Boesman [134] from EPR spectroscopy, these structural changes are assigned to the increase of non-stoichiometry (oxygen vacancies) leading to a contraction of the (001) interplanar spacing.

    The crystallite size Lc of composites is calculated using the Scherrer's formula by considering the (002) reflection line:

    Lc=KλBcosθ (3)

    where K is the shape factor, which varies with the crystallite shape (0.89 for spherical crystallites), λ is the wavelength of X-ray radiation (0.15406 nm), B is the full-width at half maximum (FWHM) expressed in radians and θ is the Bragg angle. Generally, the peak intensity and the full-width at half-maximum (FWHM) vary with the deposition conditions, which indicates differences in crystallinity, particle sizes and ordering of local structure between samples. The analysis is obtained by the combination of the Scherrer and Bragg formulæ:

    B2cos2θ=16e2sin2θ+K2λ2L2c (4)

    where < e2 > denotes local strain (defined as ∆d/d with d the interplanar spacing). From the slope 16 < e2 > and intercept K2λ2/Lc2 of the plot of B2cos2θ as a function of sin2θ, one can estimate the strain < e2 > and coherence length Lc. The variation of the local strain as a function of the substrate temperature for V2O5 films deposited by PLD method is reported in Figure 7b.

    The surface morphology and surface topography are currently investigated by scanning electron microscopy (SEM) and atomic force microscopy (AFM) as a function of at various Ts. The AFM images recorded at the scan area of 2 × 2 μm2 for thermally evaporated V2O5 films are displayed in Figure 8 [82]. These patterns show that both the surface roughness and the grain size increase with the increase in Ts. Surface roughness of 27 nm and grain size of 130 nm are found for films deposited at 300 ℃. It was demonstrated that the morphology, structure and optical and electrical properties of dip-coated V2O5 films are affected by the sol aging [135,136,137,138]. Senapati and Panda [131] investigated nanoscale films prepared by spin coating of V2O5·nH2O sol at different stages of aging. Microstrains of films 92–137 nm thick were found to decrease from 4.93 × 10–3 for fresh sol to 2.62 × 10–3 for 192 h for aged sol in relation to the water loss, while the optical bandgaps decrease from 2.66 to 2.36 eV suggesting a decrease in the localized states with aging.

    Figure 8. AFM images of the V2O5 films at various Ts: (a) Ts = 25 ℃, (b) post-annealed film deposited at 25 ℃, (c) Ts = 300 ℃ and (d) Ts = 500 ℃ (from Kumar et al. [82]).

    3.2. Optical properties

    Optical properties of V2O5 thin films were investigated by means of absorption measurements for the determination of the optical bandgap, dispersion parameters and spectral window of electrochromic films. Optical absorption coefficient α is calculated by the relation:

    T=(1R)2exp(αd)1R2exp(αd) (5)

    where T and R are the spectral transmittance and reflectance and d the film thickness. The interference effect due to internal reflection at normal incidence is negligible for higher optical density (αd > 1) that reduce Eq 1 to:

    Texp(αd) (6)

    and the optical absorption coefficient becomes:

    α=(1/d)ln(T) (7)

    The optical bandgap Egopt of a semiconductor is calculated using the interband absorption theory:

    αhVi=A(hViEoptg)n (8)

    where νi is the incident photon energy, A is a probability parameter and n is the transition coefficient with n = 2 for indirect transition and n = 3/2 for direct bandgap that is the case of V2O5.

    Overall, optical absorption spectra, electrical conductivity and electron paramagnetic resonance (EPR) spectra of V2O5 show a polaronic character. The polaron originates from the oxygen vacancies that result in an excess of electrons localized in the empty 3d orbitals of vanadium atoms. Consequently, V5+/V4+ pairs are closed to oxygen vacancies and result in an absorption in the infrared spectrum. The optical transmittance Topt of V2O5 prepared by e-beam deposition at different oxygen partial pressure revealed a sharp increase of Topt in the spectral range 550–600 nm. This is due to the fundamental absorption edge that shifts towards the higher energy with the increase in oxygen partial pressure. Figure 9 shows the plots of (αhνi)2/3 vs. photon energy for V2O5 deposited at Ts = 280 ℃ under various oxygen partial pressure. The high value of the optical bandgap Eg = 2.32 eV determined by extrapolation of the linear part of the graph to zero (α = 0) indicates that the films are stoichiometric for p(O2) = 20 mPa, while oxygen vacancies are created at lower pressure.

    Figure 9. Tauc plots of (αhνi)2/3 vs. photon energy for V2O5 e-beam deposited at Ts = 280 ℃ under various oxygen partial pressures (from Ref. [81]).

    The partial filling of oxygen vacancies results in a decrease of the concentration of electronic localized states and an increase of Eg. The overall process of oxygen loss can be expression by the Kroger notation:

    OxoVxo+12O2VxoVo+eVo+2e (9)

    Lourenço et al. [55] studied the effect of oxygen flow rate ϕ on the optical properties of r.f. sputtered V2O5 films. The highest Eg = 2.56 eV was obtained at low ϕ = 1 sccm while higher f produced films with an optical bandgap in the range 2.41–2.45 eV. Optical bandgap data of the literature for r.f. sputtered films are in the range 2.15–2.40 eV [56,139,140]. The evolutions of both the width of electronic localized states Eels and the bandgap Eg as a function of Ts and as a function of p(O2) for e-beam evaporated V2O5 films are shown in Figure 10a, b respectively. These results show that stoichiometric films were obtained at p(O2) > 5 mPa and Ts < 100 ℃ [79].

    Figure 10. Evolution of both the width of electronic localized states Eels and the bandgap Eg as a function of Ts and as a function of (a) p(O2) and (b) Ts for e-beam evaporated V2O5 films (from Ref. [81]).

    Song et al. [141] determined the work function of the annealed V2O5 films synthesized by VTOP-based ALD method at 135 ℃ using valence-band ultraviolet photoelectron spectroscopy (UPS). The results showed that the valence band maximum (Ev) is located at ∼2.45 eV below the Fermi level (EF) and gave evidence of an indirect bandgap (Eg) of ~2.63 eV for V2O5 film annealed at 500 ℃ for 1 h in air. Irani et al. [89] investigated the structural and optical properties of nanostructured V2O5 thin films deposited by spray pyrolysis technique. The substrate temperature Ts appears to be the key growth along the (001) direction with an orthorhombic structure. It was pointed out that crystallites increased with elevating Ts, reached a maximum at Ts = 450 ℃ and decreased for Ts > 450 ℃. Optical characterizations showed that the highest transmittance occurs for Ts = 550 ℃ and the absorption edge shifts from 2.5 to 2.8 eV due to the formation of chemical bonds at the interface between V2O5 film and substrate.

    Beke et al. [73] determined a bandgap of 2.52 eV for PLD V2O5 films prepared onto amorphous glass substrate with p(O2) = 13.3 Pa and Ts = 220 ℃ but did not distinguish between indirect allowed (n = 1/2) and direct forbidden (n = 3/2) transitions. Luo et al. [142] reported the indirect allowed transition (n = 2) that leads a decrease of Eg from 2.29 to 2.02 eV with increasing Ts from 275 to 320 ℃ for d.c. magnetron sputtered films. Similar trends were obtained on PVD films [143]. Ramana et al. [72] investigated the effect of the grain size on the optical properties of PLD films. The increase of grain size from 0 to 300 nm produces a red shift of the absorption edge (decrease of Eg) from 2.47 to 2.12 eV following a parabolic behavior that is attributed to the imperfections at grain-boundary regions. The optical absorption was explored on e-beam deposited V2O5 films grown at 200 ℃ on glass substrate with various thicknesses (~0.8–1.2 µm) [84]. The authors elucidated the role of film thickness; the indirect optical bandgap Eg = 2.36 eV is found for thicker films. Optical constants of V2O5 films, i.e., refractive index nopt, extinction coefficient k, and plasma frequency ωp have been investigated by Akl [144]. Determination of the optical bandgap of 2.25 eV was obtained after roughness correction for d.c. magnetron reactive-sputtered α-V2O5 films after heating at 500 ℃ for 1 h in oxygen atmosphere [145]. Ellipsometry measurements were performed as a function of incident angles on (001) oriented sputtered films yield the values of the refractive index n = 2.67, optical absorption coefficient k = 0.0011 and thickness d = 165 nm [146].


    3.3. X-ray photoelectron spectroscopy

    X-ray photoelectron spectroscopy (XPS) is the common technique to investigate the composition of the V2O5 surface and the oxidation states of elements [79,97,147,148,149,150]. The typical V 2p and O 1s XPS spectra recorded in the energy range of 510–540 eV for V2O5 thin films deposited by EBD technique at various Ts are displayed in Figure 11. The clear evolution of the spectral response corresponds to the V/O ratio in VOx lattices. Analysis of the XPS spectrum of a V2O5 film deposited at Ts = 400 ℃ under oxidizing atmosphere p(O2) = 0.01 mPa is presented in Figure 12. The binding energies of the V 2p levels are 516.9 and 524.5 eV for V 2p3/2 and V 2p1/2, respectively, with a splitting of 7.6 eV, whereas the binding energy of the O 1s level is 529.6 eV. For more detailed information, the background corrected spectrum was fitted by Gaussian profiles. Different V 2p and O 1s components are detected upon deconvolution by the fit by six Gaussian functions at binding energies of 514.8, 516.9, 520.8, 524.2, 529.7 and 532.5 eV that correspond to V4+ 2p3/2, V5+ 2p3/2, O 1s' satellite of O2–, V5+ 2p1/2, O 1s of O2–, and O, respectively, as shown in Figure 12. From these results, the relative amounts of ions in different chemical states could be calculated by taking into account the atomic sensitivity factors: 1.41 for V 2p3/2 and 0.63 for O 1s. Note that the energy splitting between V 2p3/2 and V 2p1/2 levels is Δ = 7.3 eV for stoichiometric VO2.5 film.

    Figure 11. XPS spectra of V2O5 films prepared by e-beam deposition for various substrate temperatures 25 ≤ Ts ≤ 200 ℃.
    Figure 12. Analysis of the XPS spectrum of a V2O5 film deposited at Ts = 400 ℃ under oxidizing atmosphere p(O2) = 0.01 mPa.

    The chemical composition and topography of V2O5 thin films at different steps of Li insertion/de-insertion have been investigated by quantitative XPS and AFM analyses [151,152,153,154]. From the first cycle, a solid electrolyte interface (SEI) layer has grown on the whole electrode surface, which consists of Li2CO3 aggregates. XPS analysis shows that various chemistries appeared during the cycle of charge/discharge such as partial dissolution of the initial Li2CO3 layer, residual deposit of ROLi (R = radical), PEO and oxalates. The overall dissolution/residue deposit process explains the capacity loss noticed at the end of the first cycle. Time flight secondary ion mass spectroscopy (ToF-SIMS) was carried out to analysize the ion distribution in V2O5 thin films [150,155].


    3.4. Vibrational spectroscopy

    Among the local probes, Raman scattering (RS) and Fourier transform infrared (FTIR) spectroscopy are techniques sensitive to the short-range environment of oxygen coordination around the cations [156]. As a first approximation, the spectra consist of a superposition of the components of all local entities present in the same material in contrast to diffraction data, which give a weighted average. As a rule, the frequency and relative intensity of the bands are sensitive to coordination geometry and oxidation states [157]. Both FTIR and RS spectroscopies are also probing the stoichiometry of films (Figure 13). The vibrational spectra of V2O5, which crystallizes in the orthorhombic system (Pmmn space group) with D2h13 spectroscopic symmetry, are composed of internal modes of the V2O5 layer in the ab crystal plane and the external mode in the low-frequency region (~144 cm–1) [158]. V2O5 has the spectroscopic D132h symmetry with the Wyckoff 4f position for the vanadium, vanadyl and chain oxygen atoms whereas the bridge oxygen occupies the Wyckoff 2a position [124]. The vibration modes associated to the D2h factor group decomposed as: Γopt = 7Ag + 7B1g + 3B2g + 4A3g + 3Au + 3B1u + 6B2u + 6B3u. According this decomposition, 21 modes of vibration are Raman active (7Ag + 7B1g + 3B2g + 4A3g) and 18 modes are infrared active (3Au + 3B1u + 6B2u + 6B3u) [158,159].

    Figure 13. Vibrational spectra of V2O5 crystalline films prepared by pulse laser deposition. (a) FTIR spectrum of film deposited onto Si wafer. (b) Raman spectrum of film deposited onto glass substrate.

    Ex situ and in situ FTIR spectra of V2O5 crystalline films prepared by dip-coating from V-oxoisopropoxide sols have been investigated as a function of the degree of Li+ intercalation and analyzed in terms of dispersion analysis (LO and TO phonons) [160]. A diagnostic test to differentiate the safe region upon intercalation has been established by the red-shift of the V–OA stretching frequency at 1016 cm–1 and the disappearance of the bridging V–OB–V stretching mode at 795 cm–1.

    Numerous studies of the Raman spectroscopy of V2O5 films have been performed on as-deposited, heat-treated, and Li-intercalated samples [9,99,161]. The typical Raman spectrum exhibits more or less well-defined internal and external modes, indicative of the film purity and crystallization (Figure 13b). The internal modes (in the high-frequency region) are the V = O bending vibration at 283 and 405 cm–1 and the bending vibration of the bridging V–O–V bonds at 485 cm–1, the V–O stretching mode at 525 cm–1 that results from the corner-shared oxygen atoms, and the vanadyl terminal oxygen stretching mode at 993 cm–1. The external modes are generated by the motions of two V2O5 layers of the elementary unit cell; these translational vibrations occur at 144 and 196 cm–1 (skeleton vibrational modes) [71,76]. The RS spectra of V2O5 films deposited by flash-evaporation onto silicon substrate show frequency shift of the vanadyl mode at 995 cm–1 (due to the shortest (1.58 Å) V = O double chemical bond stoichiometry) in relation with the stoichiometry of the film [9,161]. Ramana et al. [71] reported a frequency shift of the vanadyl vibration at 993 cm–1 with the substrate temperature for V2O5 films prepared by PLD technique at 200 ≤ Ts ≤ 500 ℃, which is attributed to the stress developed in the film. As XRD patterns indicate that the films deposited at Ts > 200 ℃ are polycrystalline with a preferred orientation along the c-axis, the shift of the vanadyl mode suggests a structural organization that introduces a stress associated to a compression in the layered structure. Raman spectroscopy was utilized to characterize the microstructure of V2O5 films prepared by d.c. magnetron sputtering on Si (111) wafer. Optimized films with a mere desirable layer-like lattice were obtained for a gas flow ratio O2/Ar of 3/2 [162]. Several works investigated the microstructure of amorphous V2O5 thin films by means of Raman [30,163,164] and FTIR [75] spectroscopy. Lee et al. [163] reported the spectral fingerprint of as-deposited amorphous V2O5 films, which exhibit two broad bands at ca. 520 and 650 cm–1 assigned to the stretching modes in a disordered V–O–V lattice. The absence of the skeleton mode at 144 cm–1 is due to the lack of long-range order. Flash-evaporated V2O5 thin films at 25 ≤ Ts ≤ 200 ℃ display a poor crystallization observed by the Raman features, which result in the appearance of the intense external mode at 144 cm–1 and ill-resolved internal Raman modes in the high-frequency region [30].


    3.5. Electrical properties

    V2O5 is a n-type semiconductor. Its semiconducting properties are mainly due to the oxygen non-stoichiometry, which are the consequence oxygen vacancies V2O5–δ. The electrical conductivity of V2O5 thin films have been investigated by means of d.c. four-point probe method and a.c. impedance spectroscopy as a function of the deposition parameters and post-annealing treatment [165]. The temperature dependence of the conductivity follows an Arrhenius law in a large range of temperatures. The general condition for semiconducting behavior is that the transition-metal ion could exhibit several valence states such as V4+ to V5+ in V2O5, so that electron hopping takes place between these two levels. If C = V4+/(V4+ + V5+) is the ratio of the concentration of low valence ions with the total concentration of transition metal ions and A is the average hopping distance, one can estimate the number of charge carriers N = CA–3. Using C ~ 0.02 and A = 0.384 nm, the carrier mobility is quite low lying between 10–7 and 10–2 cm2 V–1 s–1 [166]. Murawski et al. showed that the electrical transport arises from thermally activated hopping that takes place between the electronic states for V2O5 films made by PVD at 840 ℃ [167] and for amorphous V2O5 films obtained by flash evaporation [31]. In V2O5, the V4+/V5+ ionic pair is bound to positively charged defects, i.e., oxygen vacancies. The removal of an electron from the vacancy and produce a free carrier requests an energy e2d. The electrical conductivity is usually described in terms of Mott's theory for electronic transport in transition-metal glasses [168]:

    σ(T)=C(1C)ν0e2AkBTexp(2αA)exp(WkBT) (10)

    where νo is a phonon frequency, and α is the rate of the wave function decay. The thermal activation energy can be expressed as

    W=Wh+Wd/2 (11)

    where Wh corresponds to the small polaron formation and Wd to the disorder energy arising from the energy difference of neighboring sites. The separation of W into a polaron and a disorder term is a rather difficult task, which can be partly solved using data taken at low temperatures. Julien et al. [166] reported that oxygen-deficient V2O5 films grown at high substrate temperature (Ts ≈ 250 ℃) are highly conductive with σe = 10–2 S cm–1 due to the quenching rate ΔT of the films.

    Figure 14a, b present the Arrhenius plots of the electrical conductivity σ and the activation energies of flash-evaporated films as a function of the substrate temperature and annealing conditions. These results show that (ⅰ) the σ of films deposited at Ts = 25 ℃ is always lower than that of films deposited at higher temperature even after a heat treatment, and (ⅱ) there is a hierarchy in the conductivity such as σad < σarg < σox, where σad, σarg and σox are the conductivities of films as-deposited, annealed in Ar, and annealed in O2, respectively. The resistivity of an as-grown film at Ts = 25 ℃ is ρ = 5 × 104 Ω cm with an activation energy Ea = 0.43 eV, whereas that of a film heat-treated in O2 atmosphere is ρ = 0.62 Ω cm with Ea = 0.15 eV. We expect that C is a function of Ts and Ta due to the preferable desorption of oxygen during the growth. The increase in σ is attributed to the primary consequence of the structural modification from amorphous to polycrystalline state, and from oxygen deficiency of films grown with smaller O/V ratio at high substrate temperature. Similar effect was observed on the activation energy of the electrical conductivity which decreases with the increase of Ts. Sanchez et al. [167] reported a conductivity of 1.4 × 10–5 S cm–1 at 25 ℃ with an activation energy of W = 0.41 eV and a disordered activation energy Wd = 0.16 eV for flash-evaporated films. Kumar et al. [82] reported that thermally deposited films at Ts = 300 ℃ exhibit an electrical resistivity of 3.6 Ω cm. Luo et al. [142] studied the impact of the substrate temperature on the electrical properties of sputtered nanostructured films having semiconducting character. Four-point probe measurements show that the square resistance of film decreases exponentially from 46 to 33 kΩ/□ with the increase of Ts from 230 to 320 ℃.

    Figure 14. (a) Plots of σT vs. 1/T for flash-evaporated V2O5 films obtained in various conditions. (1) as-deposited at Ts = 25 ℃, (2) as-deposited at Ts = 250 ℃, (3) grown at Ts = 25 ℃ and annealed in O2, (4) grown at Ts = 25 ℃ and annealed in Ar, (5) grown at Ts = 250 ℃ and annealed in O2 and (6) grown at Ts = 250 ℃ and annealed in Ar. The annealing treatment was carried out at 300 ℃ for 48 h. (b) Plots of σRT and Wh vs. Ts.

    Rosaiah et al. [83] reported the effect of Ts on the electrical conductivity of e-beam deposited V2O5 films onto ITO-coated flexible Kapton substrate. They showed that σ increases from 2 × 10–6 to 3 × 10–2 S cm–1 by the increase of Ts from 30 to 300 ℃. Crystalline V2O5 thin films (26 nm grain size) deposited by thermal evaporation and annealed at 500 ℃ exhibited thermoelectric properties with a maximum Seebeck coefficient of –218 µV K–1 and electrical conductivity of 0.055 S cm–1 [28]. The electrical conductivity was ~10–7 S cm–1 with a high activation energy Ea = 0.9 eV for sputtered V2O5 films (85 nm thick) obtained in O2/Ar ratio of 20% [169]. Wang et al. [170] demonstrated that a rather thick (200 nm) crystalline V2O5 films can work properly as hole-extraction slab for organic photovoltaic devices. V2O5 thin films grown via r.f. reactive sputtering on fused silica substrates showed good electrical response for use as sensors methane and propane (50–3000 ppm) with an activation energy of 0.153 eV and a CPE exponent of n = 1.046, which approximates a CPE due to Debye-like capacitor [171].


    3.6. Intercalation of V2O5 films

    The Li+ intercalation into V2O5 films has been widely studied on specimens grown on various substrates using all deposition techniques described above. The insertion/de-insertion process occurs in a large potential window 3.5–1.5 V vs. Li+/Li according the reaction:

    V2O5(orange)+xLi++xeLixV2O5(green) (12)

    where x is the molar fraction of ions and electrons inserted in the host lattice. About 3e per mole of V2O5 could be reached at the end of the discharge (1.5 V), providing a theoretical specific capacity of 442 mAh g–1. The insertion of 3Li+ implies the formation of several lithiated phases during discharge with the transitions α→ε (0 < x < 0.5), ε→δ (0 < x < 0.5), δ→γ (0 < x < 0.5) that are characterized by voltage plateaus at 3.4, 3.2 and 2.3 V vs. Li+/Li, respectively. This is illustrated in Figure 15a for film formed by 50-nm grain size cycled at C/20 rate (20 mA g–1 current density). The corresponding cyclic voltammogram is shown in Figure 15b. The irreversible ω-Li3V2O5 is formed at the end of the discharge (1.85 V). Subsequent charge appears as a smooth S-shaped voltage profile with a mid-voltage at 2.67 V. As displayed in Figure 15b, four distinctive peaks are observed at 3.30, 3.05, 2.15 and 1.85 V vs. Li+/Li during a cathodic scan, which is a multi-step Li+-ion intercalation process. The corresponding phase transforms from α-V2O5 to ε-Li0.5V2O5 (3.30 V), δ-Li1V2O5 (3.05 V), γ-Li2V2O5 (2.15 V) and ω-Li3V2O5 (1.85 V) represented by the relations:

    V2O5+0.5Li+0.5eLi0.5V2O5 (13)
    Li0.5V2O5+0.5Li++0.5eLiV2O5 (14)
    LiV2O5+1Li++1eLi2V2O5 (15)
    Li2V2O5+1Li++1eLi3V2O5 (16)
    Figure 15. (a) Discharge-charge of Li//V2O5 thin-film cell carried out in the potential range 3.6–1.5 V vs. Li+/Li at C/20 rate (20 mA g–1 current density). (b) Cyclic voltammogram showing the formation of the LixV2O5 phases.

    In the following anodic scan, one peak is recorded at 2.67 V vs. Li+/Li corresponding to the delithiation of the ω-Li3V2O5 phase.

    The early work of Park et al. [172] showed that spin-coated V2O5 xerogel films can accommodate up to 3.3 moles Li/mole V2O5 in the potential range 3.8–1.9 V vs. Li+/Li. The films are highly reversible hosts and deliver a specific energy density 1137 Wh kg–1 and specific capacity 1682 C g–1. The Li+ insertion into V2O5 films deposited by CVD, rf-sputtered and ALD techniques onto various substrates such as Pt, Ti, stainless steel, glass and F-doped SnO2 was studied by Groult et al. [34,35,126,106,108]. Films grown on stainless steel and heat treated at 350 ℃ show a discharge capacity of 115 mAh g–1 at C/23 rate in the potential range 3.8–2.8 V vs. Li+/Li. The best performances (200 mAh g–1 at C/23 rate) were obtained in the electrochemical window 3.8–2.2 V for V2O5 deposited onto Ti foil. Jourdani et al. [173] reported an increase of the optical absorption edge from 2.3 to 2.8 eV and the (110) Bragg reflection shift from 2θ = 24.8° to 25.2° for spin coated V2O5 intercalated at –0.4 V vs. SCE.

    The structural evolution of Li-intercalated V2O5 films has been investigated by in situ and ex situ XRD, FTIR and Raman spectroscopy [174,175,176]. Meulenkamp et al. [177] investigated the different phases of lithiated sol-gel V2O5 films in the range 0 ≤ x ≤ 1.4 using in situ X-ray diffraction measurements and compared the structural changes with those of V2O5 films prepared by other techniques. Spectroscopic studies of Li-intercalated V2O5 polycrystalline films have been reported by several groups [71,174,178,179,180]. Cazzanelli et al. [99] analyzed the Raman spectra of xerogel films intercalated with lithium that showed a significant effect of the water content in the interlayer space inducing new Raman bands at 800 and 950 cm–1. Julien et al. [174] investigated the vibrational modifications of Li intercalated V2O5 films deposited by flash evaporation. They proposed that the tensile stress in the LixV2O5 film affects the Raman skeleton mode which shifts from 145 to 154 cm–1 due to an increase in the restoring force (Figure 16). Jung et al. [175] confirmed this hypothesis by the correlation of the spectral shift of the Raman mode at 145 cm–1 with the stress change as a function of the degree of intercalation in the LixV2O5 lattice.

    Figure 16. Frequency shift of the Ag vanadyl mode as a function of the lithium insertion in V2O5 films grown on Si (1000) wafer. The vibrational spectra reveal that the LixV2O5 phases possess considerably different local symmetry (from Ref. [71]).

    Micro-Raman spectroscopy was used to study of electrochemically intercalation LixV2O5 (0 ≤ x ≤ 1.8) samples obtained from V2O5 films deposited by various techniques, i.e., ALD and r.f. sputtering. Results showed the phase transformation from α-LixV2O5 to γ-LixV2O5 via ε and δ phases. A linear correlation exists between the V–O1 stretching (vanadyl) mode frequency and the V–O1 bond length in V2O5 ALD film [179,180]. Careful examination of the high-frequency region of RS spectra of LixV2O5 shows that, for the Li-rich phase (0 ≤ x ≤ 0.5), the vanadyl stretching mode at 984 cm–1 splits into two components, the first one at 975 cm–1, the second at 957 cm–1 for x ≥ 0.7. This is due to the evolution of the local environment of oxygen atoms, i.e., the continuous increase of the interlayer distance with x and it reflects the existence of two different lithium sites [159]. Ramana et al. [71] analyzed the Raman features of the LixV2O5 films prepared by PLD technique in the range 0.00 ≤ x ≤ 1.25. Crystalline V2O5 thin films synthesized by means of r.f. sputtering without post-heating treatment were obtained with different crystallographic preferential orientations by changing the growth parameters (Ts and deposition rate) as follows: (ⅰ) h00-oriented films (ab planes perpendicular to the substrate) were prepared at Ts = 300 ℃ for short time td, (ⅱ) longer deposition time at low Ts leads to 110-oriented films and (ⅲ) the 001-oriented films were obtained at high Ts irrespective on td [45]. Electrochemical tests of the h00-oriented films in Li/EC-DEC–LiPF6/V2O5 cells showed specific discharge capacities > 300 mAh g–1 in the potential range 3.8–1.5 V, with a capacity fading of 3% at C rate over 70 cycles.

    Numerous studies of the electrochemical insertion reactions of Li+ ions in the V2O5 film lattice were tested in Li cells with aprotic electrolyte such as 1 mol L-1 LiClO4 in propylene carbonate (PC) solution or 1 mol L–1 LiPF6 in ethylene carbonate (EC):dimethyl carbonate (DMC) (1:1) solution [58,181,182,183,184,185,186]. Park et al. [187] investigated the discharge capacity of V2O5 film cathode as a function of the thickness. Films (230 nm thick) obtained by sputter deposition at power of 100 W in working pressure of 1.3 Pa delivered a specific capacity of ~105 µAh cm–2 s–1 at current density 500 µA cm–2 s–1. Lithium microcells with optimized V2O5 films sputtered in O2/Ar (50/50) atmosphere were investigated by Yoon et al. [58]. When cycled in the range 1.5–3.7 V vs. Li+/Li, the cell delivered a specific capacity of 125 µAh cm–2 µm–1 after a first cycle of formation. Nano-porous crystallized V2O5 films (~35 nm crystallite size) fabricated by gel electrodeposition with the addition of block copolymer Pluoronic P123 delivered a specific discharge capacity of 240 mAh g–1 at 300 mA g–1 current density after 40 cycles [188]. Electrochemical studies of V2O5 films deposited by spin coating of vanadium-tri(isopropoxide)oxide gel show that the specific capacity and Li+ diffusion coefficient increased from 4.2 × 10–12 to 6.8 × 10–10 cm2 s–1 in the non-stoichiometric films [189]. Nanostructured V2O5 thin films made by anodic deposition from V2O5/H2O2 solution displayed a large discharge capacity of 596 mAh g–1 at a current density of 1.08 A g–1 and a fading rate of 1% per cycle in a cell using 1 mol L–1 LiClO4 in PC as the electrolyte and a Pt plate as the counter electrode [190]. The V2O5 thin film cathode with thickness of ~30 nm fabricated using biological templates delivered specific capacities of ~12 µAh cm–2 which is 7–8 times higher than that of planar V2O5 cathodes [191]. The effects of ageing on Li intercalation in V2O5 films, i.e., the evolution of the SEI, were investigated by a combination of techniques such as cyclic voltammetry, XPS, RBS and AFM for more than 3000 cycles [192].

    Kinetics of Li+ ions in the lattice of V2O5 films has been widely studied. Li-ion transport, i.e., ionic conduction (σion) and diffusion coefficient (DLi) have been investigated using cyclic voltammetry, galvanostatic titration and electrochemical impedance spectroscopy (EIS) [106,193,194]. DLi varies in the range 6 × 10–10–2 × 10–11 cm2 s–1 in sol-gel crystalline V2O5 thin films (1–3 µm thick) obtained by impedance spectroscopy [195]. Miyazaki et al. [196] investigated DLi of sputtered films oriented along the a-and b-axis using the chronoamperometry method. Kinetics along the a-axis DLi ≈ 10–11 cm2 s–1 was larger than along the b-axis 10–13DLi ≤ 10–14 cm2 s–1. McGraw et al. [197] obtained DLi vs. x(Li) in amorphous V2O5 films by PITT measurements on an h00 and 110 oriented films made by PLD method. DLi was initially 5 × 10–13 cm2 s–1 and decreased steadily from 1.2 × 10–13 cm2 s–1 at x = 0.4 to 5.52 × 10–14 cm2 s–1 at x = 1.5 in LixV2O5. Lantelme et al. [106] analyzed the phase transition process controlled by diffusion in V2O5 films deposited by CVD method. The component diffusion was estimated as a function of the film thickness: DLi = 0.48 × 10–12 cm2 s–1 for a semi-infinite configuration. DLi was also estimated from both PITT and EIS measurements using a semi-infinite diffusion model [198]. Evolution of the diffusion coefficient DLi in V2O5 film (160 nm thick) as a function of the electrode potential is shown in Figure 17. This figure indicates a good agreement between the two techniques.

    Figure 17. Evolution of the diffusion coefficient of Li+ ions in the V2O5 thin film lattice determined by EIS and PITT assuming a semi-infinite diffusion model (from Ref. [198]).

    Pyun and Bae reported a decrease of the diffusion coefficient of xerogel films from 10–10 to 10–12 cm2 s–1 as the electrode potential fell from 3.0 to 2.2 V vs. Li+/Li [193]. The transport mechanism was described as a transition from the diffusion-controlled insertion to an interfacial-controlled reaction due to a change of the lattice versus water in V2O5·nH2O xerogel. Ultrathin (17 nm) V2O5 thin were fabricated by spin coating and annealed at at 250 ℃ for 3 h in a muffle furnace. In the potential range, these films deliver a specific capacity of 262 mAh g–1 corresponding to Li1.78V2O5 when discharge in the potential range from +1 V to –1 V vs. SCE. The capacity loss is 0.03% per cycle over 600 cycles [199]. The effect of porosity on anomalous behavior of the Li intercalation into V2O5 film prepared by the polymer (Tergitol from Sigma-Aldrich Co.) surfactant templating method has been studied by AFM and EIS [200]. The diffusion impedance Zd, which displays a constant phase element (CPE) response can be expressed by:

    Zd=LnFAD1/2(Ec)(jω)α (17)

    where L is the film thickness, n the valence of inserted ion, A the active surface, F the Faraday constant, ∂E/∂C the slope of the coulometric titration and αd the CPE exponent. Analysis showed that αd is constant in the frequency range 0.63–0.10 Hz is significantly affected by the surface roughness and the formation of pores with a large size. The chemical diffusion coefficient of 1 × 10–11 cm2 s–1 was suggested. Thin films fabricated by electrostatic spray deposition (ESD) followed by heat treatment at 350 ℃ in air demonstrated a high energy efficiency. Ultra-high rate capability is obtained at 56C rate with a specific capacity of 86.7 mAh g–1 in the potential range 2.5–4.0 V vs. Li+/Li corresponding to the transitions of three phases: α ↔ ε ↔ δ. When cycled below 2.5 V, these ESD films have poor electrochemical performance due to irreversible structure degradation of the γ-phase [201,202]. Mui et al. [203] established a relationship between microstructure and kinetics in as-deposited and heat-treated sputtered V2O5 thin-film cathodes. DLi was determined by the GITT method for two submicrometric films: finer-grained 80 nm size and coarsed-grained 250 nm size. Both samples display similar values of DLi of 1 × 10–11 for the α-Li0.03V2O5 phase and 5 × 10–12 cm2 s–1 for the ε-Li0.35V2O5 and δ-Li0.9V2O5 phases, whereas DLi in finer-grain film decreases to 1 × 10–13 and 1 × 10–15 cm2 s–1 for the α–ε and ε–δ two-phase systems, respectively. Li+ diffusion coefficients of electrochemically inserted V2O5 films (0.6–3.6 µm thick) deposited by r.f. sputtering with h00 or 110 preferred orientation was studied using EIS measurements [194,204]. The easy diffusivity of Li+ ions (2 × 10–11DLi ≤ 3 × 10–10 cm2 s–1) along the b-direction is due to the lattice expansion along the c-axis. The plot DLi vs. x(Li) displays two minima at x = 0.14 and 0.6 that confirm the results of Lu et al. [198]. An energy barrier of 0.98 eV along the b-direction was found from the Arrhenius plot of DLi vs.1/T.


    4. Doped V2O5 films

    The poor electrochemical performance of V2O5 films during long term cycling attributed to low electronic conductivity can be improved by replacing a dopant for one of the V atoms forming V4+ cations [205,206]. Grégoire et al. [207] showed that octahedral chains are formed with metal dopant atoms making stable the structure during electrochemical process. The EXAFS analysis of Cu-and Zn-doped V2O5 showed that the environment of vanadium atoms is not altered by the doping atoms (M), while new V–O–M interactions are triplets [208]. The enhanced stability of M-doped vanadium pentoxide films used as electrode for lithium microbatteries and electrochromic devices comes from the distorted V2O5 layered structure. For example, the introduction of Ti dopant in spin-coated V2O5 films results in a disturbance of the diffraction peaks of (001) and (002) planes. The formation of V–O–Ti bond opens the interlayer space, distorts the lamellar structure and reduces the oxidation state of V [209]. Numerous metallic ions have been investigated as dopants, including Ag, Fe, Mo, Ta, Mg, Mn, Ti [17,66,93,202,209,210,211,212,213,214].

    Dip-coated Ag-doped V2O5 xerogel films with a molar ratio Ag/V in the range 0.005–0.5 showed an increase of the electronic conductivity by 2–3 orders of magnitude that allowed an insertion capacity of 4 Li per mole of AgxV2O5 with an excellent reversibility [215]. AgyV2O5 thin films were prepared by rf magnetron co-sputtering in 14% partial oxygen pressure pO2 = 14% from metallic Ag and V2O5 targets. Ag0.26V2O5 films were amorphous and showed improved intercalation properties due to the porosity and the additional influence of Ag+ ions to the redox reaction [205]. Nam et al. [216] investigated the electrochemical properties of copper-doped V2O5 thin film prepared by d.c. reactive magnetron co-sputtering within atmosphere O2:Ar ratio of 10:90. These films were tested as cathode materials in CuxV2O5/Lipon/Li cell and showed stable cycleability beyond 500 cycles with average capacity of 50 µAh cm–2 µm–1. Mo-doped V2O5 films were grown by thermal evaporation technique onto Ni substrates maintained at Ts = 250 ℃ and applied as electrode in supercapacitor with 1 mol L–1 KCl solution electrolyte. A maximum specific capacity of 175 mF cm–2 was obtained for 4 at.% Mo-doped V2O5 films [17]. The effect of the annealing temperature on the structure of Mo-doped V2O5 films electrochemically deposited resulted in the increase of the interlayer distance from 1.16 nm for as-deposited to 1.38 nm for film annealed at 250 ℃ that enhanced the Li+ ion motion in the electrochromic matrix [217].

    Tantalum-doped V2O5 thin films were deposited by sol-gel dip-coating method using anhydrous isopropyl alcohol as solvent [211]. 5 mol% Ta doped films as counter-electrode for electrochromic device exhibited a charge density of 70 mC cm–2 with high stability up to 1500 cycles. Mg-doped V2O5 films deposited by RF magnetron reactive sputtering onto SnO2-coated glass substrates were studied for various amount of dopant. Low Mg content (2 at.%) enhanced the specific discharge capacity and the diffusion coefficient of Li+ inserted ions, while high Mg content (15 at.%) favored the electrochromic window and the switching response time [212].

    Molybdenum acts as a n-type dopant in V2O5 that results in an increase of the carrier concentration and a shift of the absorption edge towards shorter wavelengths [93]. Several techniques were used to prepare V2O5 doped with Mn such as electrodeposition [218], e-beam evaporation [219], and dip-coating [213].

    Mn-doped V2O5 films were fabricated by dip-coating route with post-annealing at 250 ℃ in air for 3 h using ~1.67% of Mn(Ac)2·4H2O as Mn2+ source. Structural and elemental analyses showed that: (ⅰ) Mn doping suppresses the orthorhombic symmetry, (ⅱ) V4+ and V5+ coexist, (ⅲ) the formation of oxygen vacancies is due to the increase in the Mn2+ cation size. The suggested formula was (Mn, V)2O4.74(VÖ)0.26. The intercalation/deintercalation process is enhanced by the oxygen vacancies that eliminate the phase transition and reduce the irreversible capacity [182,213].

    Sn-doped V2O5 films were obtained by drop-casting and further heat treatment at 450 ℃ in air for 2 h. As measured by XPS experiments, the 10% Sn4+ ions accommodation is compensated by the reduction of V+5 to V4+ ions. The introduction of Sn atoms causes a slight expansion of the unit volume due to the small lattice expansion along a-and c-direction, which indicates the presence of Sn4+ ions between the VO5 layers, forming SnO6 octahedra [220]. Ce-doped V2O5 films prepared by sol-gel route and annealed at 400 ℃ showed a phase transition from α-V2O5 orthorhombic to β-V2O5 tetragonal structure. Film doped with 1 mol.% Ce exhibited the best ion storage capacity of ~207 mC cm–2 [221]. Titanium doped V2O5 films have been widely investigated as intercalation hosts for electrochromic devices [209,214,222,223,224,225]. Sol-gel dip-coated (100 – x)V2O5·xTiO2 films as electrochromic electrodes were formed with Ti content up to 20 mol.% by reacting vanadium and titanium tetra-isopropoxide as precursors [214]. The best results were obtained for the proper amount of Ti dopant (5 mol.%) showing that 550 nm-thick films switched with a coloration from pale yellow to a greenish state when reduced with 5.4 mC cm–2 of lithium. Wei et al. [209] reported a remarkable improved stability of the WO3-based electrochromic devices using Ti-doped V2O5 films. The device utilizing the optimized V2O5:Ti electrode (V:Ti = 2:1) lasted 200,000 cycles in the transmittance range 2–62% without degradation. An increase of the Li intercalation capacity from 14 to 27 mC cm–2 for 5% Ti-doped V2O5 films prepared by spin coating of inorganic sol-gel precursors was attributed to the non-stoichiometry and the smaller grain size [222]. Lee and Cao [223] demonstrated an Li+ intercalation performance improved by 100% in 20 mol.% Ti-doped V2O5 polycrystalline films that results from interaction force between adjacent V2O5 slabs. Highly stable V2O5–TiO2 films fabricated by sol-gel electrodeposition [224,225] were proposed as electrochromic windows using PProDOT-Me2 that is poly (3, 3-dimethyl-3, 4-dihydro-2H thieno[3, 4b][1,4]-di-oxepine). Tungsten-doped V2O5 films were deposited by rf magnetron sputtering on non-alkali glass substrate. Due to their good thermal insulation function, these films were applied as self-cleaning windows or anti-fogging glasses. The optimal thermal insulation temperature is 19.3 ℃ for 3 wt.% W-doped V2O5 thin films [226].


    5. Nanocomposite films

    V2O5-based nanocomposite films have been fabricated for several applications, such as electrochromic devices and gas sensors. Such a nanocomposite was obtained with V2O5 obtained from an aqueous solutions of equimolar vanadium chloride and ammonium tungstate and F-doped SnO2-coated glass substrates maintained at Ts = 400 ℃. Coloration efficiency (CE) of ~49 cm2 C–1 was obtained for a film mixed with 15% WO3 which is stable up to 1000 cycles [227]. Thin (240 nm) and thick (485 nm) V2O5–MoO3 films were deposited on conducting (SnO2:F) glass by e-beam (3 kV) and thermal vacuum deposition (EBD and TVD, respectively) for opto-electronic applications [228]. A specific capacity of 80 µAh cm–2 s–1 after 70 cycles in the potential range 1.5–3.9 V vs. Li+/Li was delivered by MoO3–V2O5 nanocomposite thin film electrodes deposited by r.f. sputtering from dual targets [229]. Mixed V2O5–WO3 thin films were deposited by pulsed spray pyrolysis technique onto glass [227]. Sol-gel deposited (1 – x)WO3·xV2O5 films at x = 0.035 showed the best electrochromic performance with a brownish blue [230]. Thin films of (V2O5)1x·(MoO3)x (0 ≤ x ≤ 1) were fabricated by e-beam evaporation in p(O2) = 2 × 10–2 mPa and Ts = 150 ℃. The optical bandgap increases and the electrical conductivity decreased with the increase of composition x [231]. V2O5–28 at.% MoO3 TVD thick films deposited onto borosilicate glass were amorphous, while thin films were crystalline (crystallite size 38 nm) with strong reflections along (131), (160) and (270) direction. The bandgap varies linearly with the atomic wt.% of MoO3 in V2O5–MoO3 EBD thin films, i.e., from 2.35 eV for 10 wt.% to 2.70 eV for 90 wt.%. Composite V2O5 xerogel films assembled with polypyrrole (PPy) showed high electrical conductivity and fast electrochemical response that makes them candidates for redox sensor in flow injection analysis [232]. Graphene/V2O5/MoO3 nanocomposite film was synthesized by dip-coating via sol-gel preparation of the precursors. This composite shows fast switching electrochromic performances with bleaching time of 1.25 s and the coloration time of 1.40 s [20]. V2O5/graphene nanocomposite films prepared by direct intercalation method using V2O5 sol and graphene were developed for their use in fast switching electrochromic devices. The intercalation of graphene improves the stability and the optical reversibility of the films, i.e., 1.5 times larger than that of V2O5 xerogel, due to the fast electron mobility of 1.5 × 104 cm2 V–1 s–1 in graphene [233]. Thin film electrodes for supercapacitors were made of V2O5/carbon nanotubes composites. A symmetric capacitor formed by identical V2O5/CNT electrodes operates in a wide potential range of 1.6 V in aqueous electrolyte delivering a volumetric energy density of 41 Wh dm–3 [234]. V2O5 xerogel films were modified by poly(ethylene-oxide) (PEO). The inserted structure is modified the interaction of hydrogens with oxygens of the V–O bonds that shields the interaction between Li+ ions and V2O5 slabs resulting in an improved anodic electrochromic property [235].


    6. Applications

    The early potential applications of nanostructured V2O5 were found in the field of lithium-ion batteries [236,237], thin-film microbatteries [12,108,238], actuators [239], catalysis [240], humidity sensors [241], electrochromic display devices [139,242], optical switching memories [243], antirelection coating [81].


    6.1. Li microbatteries

    Since the work of Bates et al. [12], V2O5 thin films have been extensively investigated as positive electrodes (so-called cathodes) in rechargeable lithium microbatteries constituted by a lithium metal anode, an amorphous inorganic electrolyte and amorphous or crystalline V2O5 [244]. The most popular solid-electrolyte thin film Li2.9PO3.3N0.46 (named Lipon) was deposited by r.f. magnetron sputtering of Li3PO4 in N2 ambient. Its ionic conductivity is 2 × 10–6 S cm–1 at 25 ℃ [245,246]. A typical solid-state battery (SSB) is built by 1 µm thick amorphous V2O5 film deposited at 25 ℃, a 1 µm thick solid electrolyte and 5 µm thick lithium film (5 times in excess) [247,248]. Currently, thin-film cells are evaluated at different current densities and their delivered capacity is expressed by surface and thickness in mC cm–2 µm–1 or µAh cm–2 µm–1. Considering the theoretical specific capacity of 440 mAh g–1 for LixV2O5 with uptake of 3 Li per formula and the density d = 3.36 g cm–3 of V2O5, the theoretical specific capacity is Qth = 147.8 µAh cm–2 µm–1 for a film as dense as the crystalline material. The schematic cross-section of a thin-film lithium microbattery is presented in Figure 18.

    Figure 18. Schematic cross-section of a thin-film lithium microbattery.

    Prior Li//V2O5 cell cycled in the potential window 3.6–1.5 V delivered a capacity ~100 µAh cm–2 µm–1 at 10 µA cm–2 [12]. Vanadium oxide films (VOx) prepared by several techniques such as PECVD, PLD, rf-sputtering at the National Renewable Energy Lab. (USA) exhibit high discharge capacity and are highly stable during electrochemical cycling. Both crystalline and amorphous PLD V2O5 films (Ts = 200 ℃) deliver a discharge specific capacity of 380 mAh g–1, while for the most stable films prepared by PECVD and exhibiting an O/V ratio close to that of V6O13 (0.5 μm thick) the discharge capacity exceeds 408 mAh g–1 over 4400 cycles. These results indicate that the PECVD technology is very attractive for manufacturing rechargeable lithium microbatteries [62,67]. Typical experimental conditions for the deposition of a thin-film battery made by r.f. sputtering are summarized in Table 2 [249,250]. A solid-state film cell including a crystalline V2O5 cathode (2.4 µm thick), a Lipon electrolyte film (1.4 µm thick) and a lithium metal film (3.5 µm thick) operated in the potential 3.8–2.15 V vs. Li+/Li. Figure 19 displays the discharge-charge profiles of the SSB at 10 µA cm–2 current density. After the first cycle of formation (SEI formation), a stable specific capacity of 30 µAh cm–2 was obtained with a cathode film 2.4 µm thick [250].

    Table 2. Experimental conditions for the deposition of thin-film battery active materials [250].
    Component Target Power density (W cm–2) Gas (sccm) Deposition pressure (Pa)
    V2O5 Vanadium 4.5 Ar:O2 (50:5) 0.55
    Lipon Li3PO4 2.2 N2 (80) 2
    Lithium Li Thermal evaporation
     | Show Table
    DownLoad: CSV
    Figure 19. Discharge-charge profiles of a c-V2O5/Lipon/Li SSB cycled between 3.8 and 2.15 V at 10 µA cm–2 current density (from Ref. [250]).

    Optimized conditions for rf-sputtering deposition (Ts = 250 ℃, Pw = 50 W) provide discharge capacity of 30.2 µAh cm–2 µm–1 with two potential plateaus at 3.5–3.6 V that corresponds to the Li uptake 0 < x < 1 in LixV2O5. Assuming a gravimetric density of 2.5 g cm–3 for rf-sputtered films, the discharge capacity is 121 mAh g–1. A capacity loss of 0.0076% per cycle was reported over 2000 cycles [251]. V2O5 thin films obtained by electrostatic spray-deposition (ESD) onto Pt substrates from triisopropoxy-vanadium oxide in 0.05 mol L–1 ethyl alcohol solution were tested in Li cells with 1 mol L–1 LiClO4/propylene carbonate (PC) as electrolyte. The crystallized films annealed at 275 ℃ exhibit a good cycleability and high capacity of 270 mAh g–1 at a current rate of 0.2C. After 25 cyles at 1C rate the capacity remains stable at 260 mAh g–1 [252]. Navone et al. [253] reported the fabrication of all-solid-state lithium microbatteries using crystalline sputtered V2O5 thin film (1.2 µm thick) as cathode and lithium phosphorus oxynitride (Lipon) (1 µm thick) as solid electrolyte. The electrochemical performances are strongly affected by the cathode morphology; its porosity depends on the oxygen flow during the film deposition; thus, a 0.6 µm thick porous film was optimized using an oxygen flow of 12.5 sccm (p(O2) = 0.05 Pa). Such a microcell film showed a stable discharge capacity of 35 µAh cm–2 at a current density of 10 µA cm–2 between 3.8 and 2.15 V vs. Li+/Li. Electrochemical performance of 50 µAh cm–2 was obtained with the use of a 1 µm thick dense film. All-solid-state thin-film lithium cells with the configuration Li/Lipon/LixV2O5/Cu were fabricated on stainless steel (SS) substrate using sequential deposition techniques. The amorphous V2O5 cathode film was lithiated by vacuum evaporation of pure Li metal to form the Li1.3V2O5 cathode film (500 nm thick). The microbattery was formed during the first charging cycle, where lithium anode was electroplated in between the Lipon electrolyte layer and the stainless-steel substrate. These "Li-free" cells (1 cm2 active area) delivered a specific capacity of 43 mAh cm–2 µm–1 at a current density of 0.1 mA cm–2 in the potential range 3.8–1.8 V vs. Li+/Li and showed a good stable cycleability up to 770 cycles at 0.1 mA cm–2 current density [254]. All-solid-state microbatteries Li1.5V2O5/Lipon/Li were fabricated by the successive deposition of active elements on Si/Ti substrate as follows: (ⅰ) the substrate was first coated with 0.2 m thick gold film to form a current collector; (ⅱ) crystalline LixV2O5 films with x = 0.8 and 1.5 (1 µm thick) were prepared by thermal evaporation of metallic lithium (0.2 µm thick) deposited on a magnetron sputtered pristine material. This cathode element was keep at rest for 5 days to ensure the diffusion of Li ions into the V2O5 matrix; (ⅲ) the in situ solid electrolyte film was obtained by reactive sputtering of Li3PO4 in N2 atmosphere (p(N2) = 2 Pa) at a power of 350 W to ensure a deposition rate of 034 µm h–1; (iv) the Li anode (3.5 µm thick) was obtained by thermal evaporation at growth rate of 150 nm min–1. The all-solid-state microbattery delivered a typical specific capacity of 50 µAh cm–2 in the potential range 2.15–3.80 V vs. Li+/Li [255].

    V2O5 thin films prepared by means of cathodic deposition from V2O5 and H2O2 in aqueous show excellent electrochemical properties as cathodes. The films had a specific discharge capacity of 240 mA h g–1 retained after 200 cycles with a 1.3C rate (200 mA g–1) that corresponds to an energy density of 723 Wh kg–1. Even at high 70C rate, V2O5 film electrodes retained a good discharge capacity of 120 mAh g–1 [104]. Ostreng et al. [113] reported remarkable electrochemical properties of strongly structured ALD V2O5 films as cathodes in Li microbatteries. The best performance was obtained a discharge rate of 120C for more than 1500 cycles for 10-nm nanostructured film. McGraw et al. [197] investigated the intercalation mechanism in LixV2O5 thin films deposited by by pulsed laser deposition and estimated the chemical diffusion coefficients of Li+ ions from potentiostatic intermittent titration technique (PITT) measurements. varies in the range from 1.7 × 10–12 cm2 s–1 to 5.8 × 10–15 cm2 s–1 in crystalline PLD films, while = 5 × 10–13 cm2 s–1 in amorphous V2O5 films. The general trend is a decrease of upon Li intercalation in the host lattice reaching the final value of 5.5 × 10–14 cm2 s–1 at Li1.5V2O5 (δ phase). The specific discharge capacity 334 mAh g–1 of 10% Sn-doped V2O5 film is twice that of pure V2O5 film [220]. This superior electrochemical performance is attributed the stabilization of VO5 slabs by Sn4+ ions that produces higher diffusion coefficient of Li+ ions. High-rate V2O5-based Li-ion thin film polymer cells show outstanding long-term cyclability. Cathode thin film growth was performed by r.f. sputtering (power 100 W) in high vacuum (1 mPa) at the 1.3 Pa partial pressure of the Ar/O2 gas mixture (ratio 2:1) [256]. The V2O5/solid polymer/Li cells were electrochemically tested in the potential range 2.5–3.8 V at current regimes ranging from 1.5C to 10C. Stable specific capacity of 100 mAh g–1 was maintained up to 5C after 300 cycles.

    Recent developments of V2O5 thin films as cathode elements of rechargeable microbatteries shows the general trends of amorphous films. Zeng et al. [257] reported that films prepared via sol-gel and liquid deposition method show a gradual activation upon cycling that yields higher capacity (~200 mAh g–1 after 200 cycles) and faster kinetics. This good performance was attributed to the novel nanostructure obtained from mild synthesis conditions. Mattelaer et al. [258,259] prepared amorphous films by ALD technique based on tetrakis[ethylmethylamino]vanadium in combination with water and ozone. Galvanostatic charging and discharging performed at 1C in the potential range 3.5–2.9 V (LiV2O5) and 4.0–1.5 V (Li3V2O5), the thin films cells deliver volumetric capacities of 488 and 810 mAh cm–3, respectively. Chae et al. [260] compared the electrochemical features of amorphous (a-V2O5) and crystalline (c-V2O5) films. The high performance of a-V2O5 with a reversible specific capacity > 600 mAh g–1 at 13C rate was attributed to the absence of irreversible phase transitions and lithium ions trapping. This remarkable rate property for a-V2O5 films is due to the vacant sites that open the Li+ diffusion pathways [261]. Zhang et al. [238] made a all-solid-state battery V2O5/Lipon/LiCoO2 prepared by r.f. magnetron sputtering in which the V2O5 (200 nm thick) acts as anode film. This microbattery (1.03 µm thick) delivered a volumetric capacity of 9.5 µAh cm–2 µm–1 after 40 cycles.


    6.2. Supercapacitors

    Electrochemical capacitors (ECs) known as "supercapacitors" employ the electrical double-layer phenomenon to store energy. There are two subclasses of ECs: the electrochemical double-layer capacitors (EDLCs) having a non-faradaic charge character and the pseudocapacitors based on faradaic electrochemical redox reaction. Among transition-metal oxides used as ECs electrode, V2O5 has attracted attention because of its pseudocapacitive features, broad oxidation states and high specific capacitance 294 F g–1. Reports on the properties of V2O5 thin films as electrodes for supercapacitors are rather sparse [17,234,262,263]. Films deposited onto Ni substrates by dc-magnetron sputtering in O2:Ar (1:8) atmosphere and grown with predominant (001) orientation are composed of 32-nm grain size and surface roughness of 14 nm. These films exhibit high specific capacitance of 238 F g–1 at current density of 1 mA cm–2 [263]. 4 at.% MoO3-doped V2O5 films obtained by thermal evaporation on Ni substrates, in which Mo atoms substitute for V, exhibit a specific capacitance of 175 F g–1 [17]. Wu et al. [234] fabricated V2O5/CNT films in three steps: (ⅰ) mixture of carbon nanotubes (CNTs) and V2O5 with ethyl cellulose in ethanol and terpilenol as uniform and viscous slurry, (ⅱ) evaporation of ethanol and "doctor blade" coating onto alumina substrate and (ⅲ) annealing at 350 ℃. Reversible redox reaction was evidenced by cyclic voltammetry providing a specific capacitance of 216 F g–1 at 5 mV s–1 that is 5.2 times of that of CNTs (41 F g–1) and a volumetric capacitance of ≈460 F cm–3. This electrode maintains a power density of 22.2 kW L–1.


    6.3. Electrochromic devices

    A typical electrochromic device (ECD) includes an electrochromic layer, an electrolyte layer and an ion storage layer. Both electrochromic and storage layers are deposited on transparent substrates coated by transparent conductive film. ECDs operate by a change in optical properties between different colors. V2O5 is utilized as a counter electrode material in ECDs [22,65,96,264,265,266,267,268,269]. During the optical switching, it provides electrochemical redox reactions that balance charge transfer at the electrochromic working electrode. The typical electrochromic reaction of V2O5 can be expressed by the double injection of ions and electrons (Eq 12). The transmittance variance (ΔT) is defined by [270]:

    ΔT(%)=TbTc (18)

    where Tb and Tc are the transmittances (in %) of bleached and colored states, respectively. The coloration efficiency (CE) is defined as the change in the optical density (ΔOD) per unit of electrode area (A) at a given wavelength, as follows:

    CE=ΔOD/(q/A)=log(Tb/Tc)/(q/A) (19)

    where q is the inserted charge capacity and the color efficiency (η) is defined by:

    η=ΔOD/q (20)

    Following the work of Sarminio et al. [271] there is a close correlation between the stress change in the crystalline film and the electrochromic properties. Tensile stress varies with the x(Li) composition, the largest obtains with the δ-LiV2O5 phase. V2O5 thin film electrochemically deposited onto ITO glass from a poly-vanadic acid sol were used in ECD cell [272]. The cell was tested more than 8 × 104 times and its response time was 2–20 s, depending upon the cell voltage. Shimizu et al. [273] developed an electrochromic display device using V2O5 thin film deposited by spin coating of V2O5 powder dissolved into a mixed solution of benzyl alcohol and iso-butanol. The films coated onto ITO glass followed by annealing at Ta > 300 ℃ are crystalline and exhibit excellent reversibility of the electrochromism. Mesoporous V2O5 films prepared by hydrolysis of vanadium tri-isopropoxide in the presence of polyethylene glycol (PEG) have shown that the surfactant promotes the formation of micrometric crystallites highly suitable for the development of advanced ECDs [274]. V2O5 thin films deposited by means of microwave plasma MOCVD have a transmittance of 70% at 400 nm absorption edge. The reversible Li intercalation reaction occurs with a change in the film color from yellow (V2O5) to blue (LiV2O5) [37]. The addition of the organic–inorganic template 3-isocyanatopropyltriethoxysilane (ICS) and poly(propylene glycol) bis-2(ammino-propyl-ether) (2-APPG) in sol-gel deposited V2O5 electrochromic thin film implements the storage capacity and increases the diffusivity of Li+ ions by modification of the surface structure of the film [275]. Magnetron sputtered V2O5z thin films onto flexible polyethylene terephthalate/indium tin oxide (PET/ITO) substrates at oxygen flow rate of 2 sccm were oxygen deficient with z = 0.169 and offered a light modulation of 43.3%. After 200 cycles, at wavelength of 400 nm, an optical density change and color efficiency of 0.38 and 102.5 cm2 C–1 were obtained, respectively [276]. Similar ECD devices were fabricated at 23 ℃ by Lin and Tsai [19], showing an oxygen deficiency z = 0.13 and a transmittance modulation of 36.5% at wavelength of 400 nm.

    Recent studies show that high switching speed and coloration efficiency are obtained for titanium-doped V2O5 thin film electrochromic devices [277] and spray-deposited lithium-doped V2O5 [278]. The ammonium intercalated V2O5 xerogel thin film prepared at 50 and 85 ℃ show two-step color changes: yellow/green and green/blue related to the redox process of V4+/V5+ states [96].


    6.4. Gas sensors

    Because of their excellent catalytic properties, V2O5 thin films have attracted great interest as gas sensors. The adsorbtion of gaseous molecules and catalytic reactions are due to the active sites formed by V5+ ions with d0 electronic configuration [279,280]. Especially in thin film form, V2O5 is used in sensor of volatile organic compounds (VOCs) emitted by industries such as benzene, toluene, acetone, methanol, xylene, etc. that are hazardous substances to human health [70,281,282,283,284,285,286,287]. For a short review of V2O5-based resistive gas sensors, see Ref. [171]. Nanocrystalline V2O5 thin films deposited by a spray pyrolysis technique at Ts = 300 ℃ had a preferential orientation along the (001) direction and revealed the highest sensing response in the presence of ethanol vapor [280]. Nanocrystalline V2O5 thin films synthesized by the ethanol suspension were developed as solid-state sensors. A reversible reaction occurs with ammonia with the formation of ammonium metavanadate (bleaching), which is turn back to the starting state after annealing in air at 350 ℃ [279]. The polycrystalline V2O5 films prepared by spray pyrolysis were found to be highly sensitive towards VOCs. The selective nature of the films deposited at Ts = 300 ℃ are shown in Figure 20 [92].

    Figure 20. Response of V2O5 films prepared bybspray pyrolysis at Ts = 300 ℃ for different VOCs vapors at concentration of 100 ppm (from [92]).

    V2O5 films fabricated by PLD technique were characterized as ammonia sensors even in the presence of NO used in diesel engine exhaust [70]. V2O5 thin films 200–600 nm thick were fabricated at 290 ℃ on fused silica and alumina by means of rf reactive sputtering from a metallic vanadium target for use as gas sensor. Thin films exhibit good response towards hydrogen (5–300 ppm), methane and propane (50–3000 ppm) [171]. Ultrathin crystalline epitaxial V2O5 films (10–90 nm thick) were deposited onto c-Al2O3 by ALD method with oxygen plasma. The films exhibiting a (001) preferred orientation are used as humidity sensors [119]. Nanostructured V2O5 films deposited by dc reactive magnetron sputtering have also been proposed as 2-propanol vapor sensors with an excellent response towards 5–200 ppm under ambient atmosphere [287].


    6.5. Healthy and safety issues

    A very detailed review (210 pages) on the toxicity of V2O5 can be found on the web site [287]. In short, V2O5 is toxic, and it is specified on the U.S. National Library of Medecine [288] that probable oral lethal dose for humans is between 5 and 50 mg/kg or between 7 drops and 1 teaspoonful for a 70 kg (150 lb.) person. In practice, however, there is little occasion to eat such an amount of vanadium pentoxide and absorption can occur in humans following inhalation exposure. In this case, V2O5 is readily absorbed through the lungs. The effects from exposure may include burns to the skin and eyes, tracheitis, bronchitis, emphysema, pulmonary edema, or bronchial pneumonia. The American Conference of Governmental Industrial Hygienists (ACGIH) has established a threshold limit value time-weighted average (TLV TWA) of 0.05 mg/cu m for respirable dust. When absorbed, V2O5 is eliminated, mainly by urine. In practice, the protection against this product thus concerns people who have to manipulate or prepare it in factories since they are placed in situation where over-exposure may occur. The hygiene and safety policy of laboratories and institutions that are concerned is given in [288].

    While hygiene and safety policy practice when handling or dealing with V2O5, this material plays an increasing role to make progress in environmental pollution and industrial safety issue, owing to its use for gas sensing application, reviewed in [289]. The list of targeted gas is constantly increasing [286].


    7. Concluding remarks

    V2O5 is one of the most important material for many applications, including optical switching devices, gas and humidity sensors, smart widows, window for solar cell, photocatalysis, for electrochromic devices [82], color filters [146], reflectance mirrors, gas sensors, surfaces with tunable emittance for temperature control of space vehicles [81,147], thermochromic thin films for optoelectronic applications [93]. Nanocrystal V2O5 thin film used as hole-extraction layer in normal architecture organic solar cells demonstrated the potential of this material for photovoltaic applications [170], and W-doped V2O5 film can be used for thermal insulation [226].

    So many applications explain the huge amount of works devoted to the preparation of V2O5 thin films. The difficulty is that their properties are strongly dependent of the mode of preparation, the film thickness, the orientation of the film, the crystallinity, the purity, the substrate [84], the nature and amount of dopants, the packing density of the film [85], temperature of the substrate [142], deposition rate, in short, deposition conditions [96], post-annealing treatment [217], the choice of precursors [100], porosity [188,201,262], doping.

    As electrochromic device, Ti-doped V2O5 electrode has lasted 200,000 cyclic switching times between the lowest (2%) and highest (62%) transmittance with no significant degradation of performance [209]. Nevertheless, V2O5 alone cannot be commercialized as an electrochromic device, because of its low electrical conductivity, poor coloration efficiency and narrow color variation. However, these drawbacks can be overcome by combining V2O5 with other compounds. A recent example is the graphene/V2O5/MoO3 film, which combines the richer colors of MoO3, the high conductivity of graphene and the remarkable electrochromic properties of V2O5. Such films present multi-electrochromic behavior (yellow → green → blue → gray-brown), excellent electrochromic performance, with tranmittence variation of 25.35%, bleaching time of 1.25 seconds, coloration time 1.40 seconds. In addition, it demonstrates good cycling [20]. Such films are thus very good fast switching electrochromic materials. Recently, the growth of large-scale electrodes by atmospheric pressure chemical vapor deposition has been made possible, with good performance for electrochromic devices with a change in optical density per unit inserted charge density reaches 336 cm2 C–1 at 630 nm, with a diffusion coefficient for the lithium as large as 9.19 × 10–11 cm2 s–1 [33].

    For use as a supercapacitors, nano-structured V2O5 is needed to increase the surface area. In particular, a grain size of 148 nm of the deposited film exhibited a high rate pseudo capacitance of 730 mF cm–2 at 1 mA cm–2 of current density [27]. Note the result was also obtained under such synthesis conditions that led to a c-axis oriented structure, giving better results than a (201) orientation. A film electrode composed of intertwined V2O5 nanowires and carbon nanotubes (CNTs) demonstrated a volumetric capacitance of ≈460 F cm–3 [234]. A capacity of 397 F g–1 was obtained after ultrasonic weltering [262]. V2O5 thin films with 4 at.% of Mo doping exhibited a maximum specific capacitance of 175 mF/cm2 at current density of 1 mA/cm2 [17].

    Regarding the electrochemical properties, high energy density (900 W h kg–1 at 200 mA g–1), power density (28 kW kg–1 at 10.5 A g–1), good cyclic stability over 200 cycles have been obtained with nanostructured V2O5 thin-film electrodes prepared by cathodic deposition [104,190]. These results, as well as results obtained by atomic layer deposition [111,113,114] show so that V2O5 nano-electrodes can now be used with high power and energy densities for thin film Li-ion batteries.

    V2O5 is also increasingly used as sensor. Fit has been tested for longer time periods up to 100 h as ammonia sensors. Very stable behavior was found with detection limit of 80 ppb for NH3 both in dry and humid air. The detection limits for NO and CO were 20 ppm and 50 ppm, respectively. The results suggest that the vanadium oxide thin-film sensing layers are good candidates for a Selective Catalytic Reduction (SCR) process control application [70]. The room temperature gas sensing response of V2O5 films was found to be highly selective towards xylene [92]. They also exhibited good response towards hydrogen (5–300 ppm), methane and propane (50–3000 ppm) [171]. Recently, intrinsic thermochromism of V2O5 films was reported as a perceptible thermally induced color change from bright yellow to deep orange [290].

    All these recent results show the continuous improvement in the synthesis of the films and the performance for tehes applications. In most of the works that have been published, however, the optimization of the film is made within one deposition technique, for one application. Therefore, further works should address two questions. The first one is to determine which deposition technique and synthesis parameters are best suited for each application. For instance, electrochemical properties of vanadium oxide are reported to be optimized for highly crystallized V2O5 [21], but better kinetics and higher volumetric capacities were observed for the amorphous vanadium oxides compared to their crystalline counterparts in Ref. [259]. Therefore, further works are needed to clarify the conditions that optimize the films for electrochemical applications. The second question is which kind of film should be used to optimize the film as a function of the application that is targeted. For instance, low Mg-doping (2%) is preferred for lithium ion battery applications, but 15% Mg-doping is preferred for electrochromic applications [212]. Only very few such reports have been made to relate the synthesis to applications, and further works along these lines would be very useful in the near future.


    Conflict of interest

    The authors declare that there is no conflict of interest regarding the publication of this manuscript.


    [1] Roscoe HE (1868) XXXVI.—Researches on vanadium. J Chem Soc 21: 322–350. doi: 10.1039/JS8682100322
    [2] Ditte A (1885) Recherches sur le vanadium, propriétés de l'acide vanadique. C R Acad Sci Paris 101: 698–702.
    [3] Hugues JM, Finger LW (1983) The crystal chemistry of shcherbinaite, naturally occurring V2O5. Am Mineral 68: 1220–1222.
    [4] Ketelaar JAA (1936) Crystal structure and shape of colloidal particles of vanadium pentoxide. Nature 137: 316.
    [5] Byström A, Wilhelmi KA, Brotzen O (1950) Vanadium pentoxide—a compound with five-coordinated vanadium atoms. Acta Chem Scand 4: 1119–1130. doi: 10.3891/acta.chem.scand.04-1119
    [6] Bachmann HG, Ahmed FR, Barnes WH (1961) The crystal structure of vanadium pentoxide. Z Krist -Cryst Mater 115: 110–131.
    [7] Enjalbert R, Galy J (1986) A refinement of the structure of V2O5. Acta Crystallogr C 42: 1467–1469. doi: 10.1107/S0108270186091825
    [8] Ramana CV, Hussain OM, Srinivasulu-Naidu B, et al. (1997) Spectroscopic characterization of electron-beam evaporated V2O5 thin films. Thin Solid Films 305: 219–226. doi: 10.1016/S0040-6090(97)00141-7
    [9] Julien C, Guesdon JP, Gorenstein A, et al. (1995) The influence of the substrate material on the growth of V2O5 flash-evaporated films. Appl Surf Sci 90: 389–391.
    [10] Balog P, Orosel D, Cancarevic Z, et al. (2007) V2O5 phase diagram revisited at high pressures and high temperatures. J Alloy Compd 429: 87–98. doi: 10.1016/j.jallcom.2006.04.042
    [11] Parija A, Prendergast D, Banerjee S (2017) Evaluation of multivalent cation insertion in single- and double-layered polymorphs of V2O5. ACS Appl Mater Inter 9: 23756–23765. doi: 10.1021/acsami.7b05556
    [12] Bates JB, Dudney NJ, Lubben DC, et al. (1995) Thin-film rechargeable lithium batteries. J Power Sources 54: 58–62. doi: 10.1016/0378-7753(94)02040-A
    [13] Julien C, Gorenstein A (1995) Materials design and optimization for thin-film microbatteries. Ionics 1: 193–210. doi: 10.1007/BF02426018
    [14] Julien C (1996) Technology of microbatteries and its applications, In: Radhakrishna S, Trends in Materials Science, New Delhi: Narosa Publ. House, 24–43.
    [15] Julien C (2000) Lithium microbatteries, In: Julien C, Stoynov Z, Materials for lithium-ion batteries, Amsterdam: Kluwer Acad Publ., 381–400.
    [16] Lampert CM (1998) Smart switchable glazing for solar energy and daylight control. Sol Energ Mat Sol C 52: 207–221. doi: 10.1016/S0927-0248(97)00279-1
    [17] Prakash NG, Dhananjaya M, Reddy BP, et al. (2016) Molybdenum doped V2O5 thin films electrodes for supercapacitors. Mater Today Proc 3: 4076–4081. doi: 10.1016/j.matpr.2016.11.076
    [18] Vernardou D, Sapountzis A, Spanakis E, et al. (2013) Electrochemical activity of electrodeposited V2O5 coatings. J Electrochem Soc 160: D6–D9. doi: 10.1149/2.043304jes
    [19] Lin YS, Tsai CW (2008) Reactive sputtering deposition of V2O5−z on flexible PET/ITO substrates for electrochromic devices. Surf Coat Tech 202: 5641–5645. doi: 10.1016/j.surfcoat.2008.06.037
    [20] Ma X, Lu S, Wan F, et al. (2016) Synthesis and electrochromic characterization of graphene/V2O5/MoO3 nanocomposite films. ECS J Solid State Sc 5: P572–P577. doi: 10.1149/2.0031610jss
    [21] Vernardou D, Loudoudakis D, Spanakis E, et al. (2014) Electrochemical properties of vanadium oxide coatings grown by hydrothermal synthesis on FTO substrates. New J Chem 38: 1959–1964. doi: 10.1039/C3NJ00931A
    [22] Mjejri I, Manceriu LM, Gaudon M, et al. (2016) Nano-vanadium pentoxide films for electrochromic displays. Solid State Ionics 292: 8–14. doi: 10.1016/j.ssi.2016.04.023
    [23] Beke S (2011) A review of the growth of V2O5 films from 1885 to 2010. Thin Solid Films 519: 1761–1771. doi: 10.1016/j.tsf.2010.11.001
    [24] Rao MC, Rao RK (2014) Thermal evaporated V2O5 thin films: thermodynamic properties. Int J Chem Tech Res 6: 3931–3934.
    [25] Chakraborty S, Sakata H, Yokoyama E, et al. (2007) Laser induced forward transfer technique for maskless patterning of amorphous V2O5 thin films. Appl Surf Sci 254: 638–643. doi: 10.1016/j.apsusc.2007.06.066
    [26] Kumar RTR, Karunagaran B, Venkatachalam S, et al. (2003) Influence of deposition temperature on the growth of vacuum evaporated V2O5 thin films. Mater Lett 57: 3820–3825. doi: 10.1016/S0167-577X(03)00185-X
    [27] Dhananjaya M, Prakash NG, Sandhya GL, et al. (2017) Microstructure and supercapacitor properties of V2O5 thin film prepared by thermal evaporation method. Mech Mater Sci Eng.
    [28] Santos R, Loureiro J, Nogueira A, et al. (2013) Thermoelectric properties of V2O5 thin films deposited by thermal evaporation. Appl Surf Sci 282: 590–594. doi: 10.1016/j.apsusc.2013.06.016
    [29] Julien C, Khelfa A, Benramdane N, et al. (1994) Lithium insertion in indium selenide films: application to microbatteries. Mater Sci Eng B-Adv 23: 105–115. doi: 10.1016/0921-5107(94)90342-5
    [30] Julien C, Guesdon JP, Gorenstein A, et al. (1995) The growth of V2O5 flash-evaporated films. J Mater Sci Lett 14: 934–936. doi: 10.1007/BF02427469
    [31] Murawski L, Gledel C, Sanchez C, et al. (1987) Electrical conductivity of V2O5 and LixV2O5 amorphous thin films. J Non-Cryst Solids 89: 98–106. doi: 10.1016/S0022-3093(87)80324-1
    [32] Ramana CV, Hussain OM, Naidu BS, et al. (1998) Physical investigations on electron-beam evaporated vanadium pentoxide films. Mater Sci Eng B-Adv 52: 32–39. doi: 10.1016/S0921-5107(97)00273-0
    [33] Vernardou D (2017) Using an atmospheric pressure chemical vapor deposition process for the development of V2O5 as an electrochromic material. Coatings 7: 24. doi: 10.3390/coatings7020024
    [34] Mantoux A, Groult H, Balnois E, et al. (2004) Vanadium oxide films synthesized by CVD and used as positive electrodes in secondary lithium batteries. J Electrochem Soc 151: A368–A373. doi: 10.1149/1.1641037
    [35] Groult H, Le Van K, Mantoux A, et al. (2007) Study of the Li+ insertion into V2O5 films deposited by CVD onto various substrates. J Power Sources 174: 312–320. doi: 10.1016/j.jpowsour.2007.09.045
    [36] Barreca D, Armelao L, Caccavale F, et al. (2000) Highly oriented V2O5 nanocrystalline thin films by plasma-enhanced chemical vapor deposition. Chem Mater 12: 98–103. doi: 10.1021/cm991095a
    [37] Watanabe H, Itoh K, Matsumoto O (2001) Properties of V2O5 thin films deposited by means of plasma MOCVD. Thin Solid Films 386: 281–285. doi: 10.1016/S0040-6090(00)01674-6
    [38] Barreca D, Battiston GA, Caccavalec F, et al. (1999) A PE-MOCVD route to V2O5 nanostructured thin films. J Phys IV France 9: 529–536.
    [39] Crociani L, Carta G, Natali M, et al. (2011) MOCVD of vanadium oxide films with a novel vanadium(III) precursor. Chem Vapor Depos 17: 6–8. doi: 10.1002/cvde.201004291
    [40] Sahana MB, Shivashankar SA (2004) Metalorganic chemical vapor deposition of highly oriented thin film composites of V2O5 and V6O13: suppression of the metal-semiconductor transition in V6O13. J Mater Res 19: 2859–2870. doi: 10.1557/JMR.2004.0394
    [41] Robbins J, Seman M (2005) Electrochromic devices deposited on low-temperature plastics by plasma-enhanced chemical vapor deposition. United States: USDOE Office of Energy Efficiency and Renewable Energy (EERE), Report No.: FG36-04GO14328.
    [42] Nandakumar NK, Seebauer EG (2011) Low temperature chemical vapor deposition of nanocrystalline V2O5 thin films. Thin Solid Films 519: 3663–3668. doi: 10.1016/j.tsf.2011.02.002
    [43] Koike S, Fujieda T, Sakai T, et al. (1999) Characterization of sputtered vanadium oxide films for lithium batteries. J Power Sources 81–82: 581–584.
    [44] Batista C, Teixeira V, Carneiro J (2008) Structural and morphological characterization of magnetron sputtered nanocrystalline vanadium oxide films for thermochromic smart surfaces. J Nano Res 2: 21–30. doi: 10.4028/www.scientific.net/JNanoR.2.21
    [45] Quinzeni I, Ferrari S, Quartarone E, et al. (2011) Structural, morphological and electrochemical properties of nanocrystalline V2O5 thin films deposited by means of radiofrequency magnetron sputtering. J Power Sources 196: 10228–10233. doi: 10.1016/j.jpowsour.2011.08.073
    [46] Giannetta HMR, Calaza C, Lamas DG, et al. (2015) Electrical transport properties of V2O5 thin films obtained by thermal annealing of layers grown by RF magnetron sputtering at room temperature. Thin Solid Films 589: 730–734. doi: 10.1016/j.tsf.2015.06.048
    [47] Poelman H, Tomaszewski H, Poelman D, et al. (2002) V2O5 thin films deposited by means of dc magnetron sputtering from ceramic V2O3 targets. Surf Interface Anal 34: 724–727. doi: 10.1002/sia.1397
    [48] Su Q, Lan W, Wang YY, et al. (2009) Structural characterization of β-V2O5 films prepared by dc reactive magnetron sputtering. Appl Surf Sci 255: 4177–4179. doi: 10.1016/j.apsusc.2008.11.002
    [49] Fateh N, Fontalvo GA, Mitterer C (2007) Structural and mechanical properties of dc and pulsed dc reactive magnetron sputtered V2O5 films. J Phys D Appl Phys 40: 7716–7719. doi: 10.1088/0022-3727/40/24/019
    [50] Gallasch T, Stockhoff T, Baither D, et al. (2011) Ion beam sputter deposition of V2O5 thin films. J Power Sources 196: 428–435. doi: 10.1016/j.jpowsour.2010.06.099
    [51] Silversmit G, Poelman H, DeGryse R (2004) Influence of magnetron deposition parameters on the stoichiometry of sputtered V2O5 films. Surf Interface Anal 36: 1163–1166. doi: 10.1002/sia.1866
    [52] De Castro MSB, Ferreira CL, de Avillez RR (2013) Vanadium oxide thin films produced by magnetron sputtering from a V2O5 target at room temperature. Infrared Phys Techn 60: 103–107. doi: 10.1016/j.infrared.2013.03.001
    [53] Raj PD, Gupta S, Sridharan M (2015) Nanostructured V2O5 thin films deposited at low sputtering power. Mat Sci Semicon Proc 39: 426–432. doi: 10.1016/j.mssp.2015.04.054
    [54] Fateh N, Fontalvo GA, Cha L, et al. (2008) Synthesis-structure relations for reactive magnetron sputtered V2O5 films. Surf Coat Tech 202: 1551–1555. doi: 10.1016/j.surfcoat.2007.07.010
    [55] Lourenço A, Gorenstein A, Passerini S, et al. (1998) Radio-frequency reactively sputtered VOx thin films deposited at different oxygen flows. J Electrochem Soc 145: 706–711. doi: 10.1149/1.1838327
    [56] Benmoussa M, Ibnouelghazi E, Bennouna A, et al. (1995) Structural, electrical and optical properties of sputtered vanadium pentoxide thin films. Thin Solid Films 265: 22–28. doi: 10.1016/0040-6090(95)06617-9
    [57] Benmoussa M, Outzourhit A, Bennouna A, et al. (2002) Electrochromism in sputtered V2O5 thin films: structural and optical studies. Thin Solid Films 405: 11–16. doi: 10.1016/S0040-6090(01)01734-5
    [58] Yoon YS, Kim JS, Choi SH (2004) Structural and electrochemical properties of vanadium oxide thin films grown by d.c. and r.f. reactive sputtering at room temperature. Thin Solid Films 460: 41–47.
    [59] Ottaviano L, Pennisi A, Simone F, et al. (2004) RF sputtered electrochromic V2O5 films. Opt Mater 27: 307–313. doi: 10.1016/j.optmat.2004.04.001
    [60] Lin YS, Tsai CW, Chen PW (2008) Electrochromic properties of V2O5−z thin films sputtered onto flexible PE/ITO substrates. Solid State Ionics 179: 290–297. doi: 10.1016/j.ssi.2008.01.054
    [61] Kang M, Chu M, Kim SW, et al. (2013) Optical and electrical properties of V2O5 nanorod films grown using an electron beam. Thin Solid Films 547: 198–201. doi: 10.1016/j.tsf.2013.03.060
    [62] Zhang JG, McGraw JM, Turner J, et al. (1997) Charging capacity and cycling stability of VOx films prepared by pulsed laser deposition. J Electrochem Soc 144: 1630–1634. doi: 10.1149/1.1837652
    [63] Julien C, Haro-Poniatowski E, Camacho-Lopez MA, et al. (1999) Growth of V2O5 thin films by pulsed laser deposition and their applications in lithium microbatteries. Mater Sci Eng B-Adv 65: 170–176. doi: 10.1016/S0921-5107(99)00187-7
    [64] Ramana CV, Smith RJ, Hussain OM, et al. (2004) On the growth mechanism of pulsed-laser deposited vanadium oxide thin films. Mater Sci Eng B-Adv 111: 218–225. doi: 10.1016/j.mseb.2004.04.017
    [65] Iida Y, Kaneko Y, Kanno Y (2008) Fabrication of pulsed-laser deposited V2O5 thin films for electrochromic devices. J Mater Process Tech 197: 261–267. doi: 10.1016/j.jmatprotec.2007.06.032
    [66] Iida Y, Kanno Y (2009) Doping effect of M (M = Nb, Ce, Nd, Dy, Sm, Ag, and/or Na) on the growth of pulsed-laser deposited V2O5 thin films. J Mater Process Tech 209: 2421–2427. doi: 10.1016/j.jmatprotec.2008.05.033
    [67] McGraw JM, Perkins JD, Hasoon F, et al. (2000) Pulsed laser deposition of oriented V2O5 thin films. J Mater Res 15: 2249–2265. doi: 10.1557/JMR.2000.0323
    [68] Ashrafi MA, Ranjbar M, Kalhori H, et al. (2017) Pulsed laser deposition of Mo–V–O thin films for chromogenic applications. Thin Solid Films 621: 220–228. doi: 10.1016/j.tsf.2016.11.041
    [69] Madhuri KV, Rao KS, Naidu BS, et al. (2002) Characterization of laser ablated V2O5 thin films. J Mater Sci-Mater El 13: 426–432.
    [70] Huotari J, Bjorklund R, Lappalainen J, et al. (2015) Pulsed laser deposited nanostructured vanadium oxide thin films characterized as ammonia sensors. Sensor Actuat B-Chem 217: 22–29. doi: 10.1016/j.snb.2015.02.089
    [71] Ramana CV, Smith RJ, Hussain OM, et al. (2005) Surface analysis of pulsed laser-deposited V2O5 thin films and their lithium intercalated products studied by Raman spectroscopy. Surf Interface Anal 37: 406–411. doi: 10.1002/sia.2018
    [72] Ramana CV, Smith RJ, Hussain OM (2003) Grain size effects on the optical characteristics of pulsed-laser deposited vanadium oxide thin films. Phys Status Solidi A 199: R4–R6. doi: 10.1002/pssa.200309009
    [73] Beke S, Giorgio S, Korosi L, et al. (2008) Structural and optical properties of pulsed laser deposited V2O5 thin films. Thin Solid Films 516: 4659–4664. doi: 10.1016/j.tsf.2007.08.113
    [74] Fang GJ, Liu ZL, Wang YQ, et al. (2000) Orientated growth of V2O5 electrochromic thin films on transparent conductive glass by pulsed excimer laser ablation technique. J Phys D Appl Phys 33: 3018–3021. doi: 10.1088/0022-3727/33/23/302
    [75] Ramana CV, Hussain OM, Pinto R, et al. (2003) Microstructural features of pulsed-laser deposited V2O5 thin films. Appl Surf Sci 207: 135–138. doi: 10.1016/S0169-4332(02)01411-3
    [76] Ramana CV, Smith RJ, Hussain OM, et al. (2004) Growth and surface characterization of V2O5 thin films made by pulsed laser deposition. J Vac Sci Technol A 22: 2453–2458. doi: 10.1116/1.1809123
    [77] Ramana CV, Hussain OM, Naidu BS, et al. (1997) Influence of substrate temperature on the composition and structural properties of electron beam evaporated V2O5 thin films. Vacuum 48: 431–434. doi: 10.1016/S0042-207X(96)00304-1
    [78] Ramana CV, Hussain OM (1997) Optical absorption behaviour of vanadium pentoxide thin films. Adv Funct Mater 7: 225–231.
    [79] Ramana CV, Hussain OM, Naidu BS, et al. (1998) Physical investigations on electron-beam evaporated vanadium pentoxide films. Mater Sci Eng B-Adv 52: 32–39. doi: 10.1016/S0921-5107(97)00273-0
    [80] Ramana CV, Hussain OM, Naidu BS (1997) Growth and structure of electron beam evaporated V2O5 thin films. Mater Chem Phys 50: 195–199. doi: 10.1016/S0254-0584(97)01914-7
    [81] Ramana CV, Hussain OM, Uthanna S, et al. (1998) Influence of oxygen partial pressure on the optical properties of electron beam evaporated vanadium pentoxide thin films. Opt Mater 10: 101–107. doi: 10.1016/S0925-3467(97)00168-7
    [82] Kumar A, Singh P, Kulkarni N, et al. (2008) Structural and optical studies of nanocrystalline V2O5 thin films. Thin Solid Films 516: 912–918. doi: 10.1016/j.tsf.2007.04.165
    [83] Rosaiah P, Sandhya GL, Suresh S, et al. (2016) Effect of substrate temperature on micro-structural, optical and electrical properties of V2O5 thin films. Adv Mater Proc 1: 215–219. doi: 10.5185/amp.2016/219
    [84] Thiagarajan S, Thaiyan M, Ganesan R (2015) Physical property exploration of highly oriented V2O5 thin films prepared by electron beam evaporation. New J Chem 39: 9471–9479. doi: 10.1039/C5NJ01582K
    [85] Ali HM, Abdel-Kakeem AM (2010) Structural and optical properties of electron-beam evaporated Al2O3-doped V2O5 thin films for various applications. Phys Status Solidi A 207: 132–138. doi: 10.1002/pssa.200925121
    [86] Bouzidi A, Benramdane N, Nakrela A, et al. (2002) First synthesis of vanadium oxide thin films by spray pyrolysis technique. Mater Sci Eng B-Adv 95: 141–147. doi: 10.1016/S0921-5107(02)00224-6
    [87] Boudaoud L, Benramdane N, Desfeux R, et al. (2006) Structural and optical properties of MoO3 and V2O5 thin films prepared by spray pyrolysis. Catal Today 113: 230–234. doi: 10.1016/j.cattod.2005.11.072
    [88] Varadaraajana V, Satishkumara BC, Nandab J, et al. (2011) Direct synthesis of nanostructured V2O5 films using solution plasma spray approach for lithium battery applications. J Power Sources 196: 10704–10711. doi: 10.1016/j.jpowsour.2011.09.016
    [89] Irani R, Rozati SM, Beke S (2013) Structural and optical properties of nanostructural V2O5 thin films deposited by spray pyrolysis technique: effect of the substrate temperature. Mater Chem Phys 139: 489–493. doi: 10.1016/j.matchemphys.2013.01.046
    [90] Wei Y, Li M, Zheng J, et al. (2013) Structural characterization and electrical and optical properties of V2O5 films prepared via ultrasonic spraying. Thin Solid Films 534: 446–451. doi: 10.1016/j.tsf.2013.01.093
    [91] Abd-Alghafour NM, Ahmed NM, Hassan Z, et al. (2016) Influence of solution deposition rate on properties of V2O5 thin films deposited by spray pyrolysis technique. AIP Conf Proc 1756: 090010. doi: 10.1063/1.4958791
    [92] Vijayakumar Y, Kumar-Mani G, Ramana-Reddy MV, et al. (2015) Nanostructured flower like V2O5 thin films and its room temperature sensing characteristics. Ceram Int 41: 2221–2227. doi: 10.1016/j.ceramint.2014.10.023
    [93] Nazemiyan M, Jalili S (2013) Record low temperature Mo doped V2O5 thermochromic thin films for optoelectronic applications. AIP Adv 3: 112103. doi: 10.1063/1.4829663
    [94] Akl AA (2006) Effect of solution molarity on the characteristics of vanadium pentoxide thin films. Appl Surf Sci 252: 8745–8750. doi: 10.1016/j.apsusc.2005.12.076
    [95] Wang J, Curtis CJ, Schulz DL, et al. (2004) Influences of treatment temperature and water content on capacity and rechargeability of V2O5 xerogel films. J Electrochem Soc 151: A1–A7. doi: 10.1149/1.1627342
    [96] Najdoski M, Koleva V, Samet A (2014) Effect of deposition conditions on the electrochromic properties of nanostructured thin films of ammonium intercalated vanadium pentoxide xerogel. J Phys Chem C 118: 9636–9646. doi: 10.1021/jp4127122
    [97] Chen Y, Xie K, Liu Z (1998) XPS studies of V2O5 thin film at different temperatures and oxygen partial pressures. Appl Surf Sci 126: 347–351. doi: 10.1016/S0169-4332(97)00791-5
    [98] Wang Y, Shang H, Chou T, et al. (2005) Effects of thermal annealing on the Li+ intercalation properties of V2O5·nH2O xerogel films. J Phys Chem B 109: 11361–11366.
    [99] Cazzanelli E, Mariotto G, Passerini S, et al. (1996) Raman spectroscopic investigations of Li-intercalated V2O5 xerogel. J Non-Cryst Solids 208: 89–98. doi: 10.1016/S0022-3093(96)00507-8
    [100] Gokdemir FP, Ozdemir O, Kutlu K (2014) Comparison of structural and electrochemical properties of V2O5 thin films prepared by organic/inorganic precursors. Electrochim Acta 121: 240–244. doi: 10.1016/j.electacta.2013.12.164
    [101] Raj DV, Ponpandian N, Mangalaraj M, et al. (2013) Effect of annealing and electrochemical properties of sol-gel dip coated nanocrystalline V2O5 thin films. Mat Sci Semicon Proc 16: 256–262. doi: 10.1016/j.mssp.2012.11.001
    [102] Wang Y, Cao G (2006) Li+-intercalation electrochemical/electrochromic properties of vanadium pentoxide films by sol electrophoretic deposition. Electrochim Acta 51: 4865–4872. doi: 10.1016/j.electacta.2006.01.026
    [103] Kim KH, Roh DK, Song IK, et al. (2010) Enhanced performance as a lithium ion battery cathode of electrodeposited V2O5 thin films by e-beam irradiation. J Solid State Electr 14: 1801–1805. doi: 10.1007/s10008-010-1114-9
    [104] Liu Y, Clark M, Zhang Q, et al. (2011) V2O5 nano-electrodes with high power and energy densities for thin film Li-ion batteries. Adv Energy Mater 1: 194–202. doi: 10.1002/aenm.201000037
    [105] Groult H, Balnois E, Mantoux A, et al. (2006) Two-dimensional recrystallisation processes of nanometric vanadium oxide thin films grown by atomic layer chemical vapor deposition (ALCVD) evidenced by AFM. Appl Surf Sci 252: 5917–5925. doi: 10.1016/j.apsusc.2005.08.014
    [106] Lantelme F, Mantoux A, Groult H, et al. (2002) Analysis of a phase transition process controlled by diffusion, application to lithium insertion into V2O5. Electrochim Acta 47: 3927–3938. doi: 10.1016/S0013-4686(02)00364-X
    [107] Keränen J, Auroux A, Ek S, et al. (2002) Preparation, characterization and activity testing of vanadia catalysts deposited onto silica and alumina supports by atomic layer deposition. Appl Catal A-Gen 228: 213–225. doi: 10.1016/S0926-860X(01)00975-9
    [108] Le Van K, Groult H, Mantoux A, et al. (2006) Amorphous vanadium oxide films synthesized by ALCVD for lithium rechargeable batteries. J Power Sources 160: 592–601. doi: 10.1016/j.jpowsour.2006.01.049
    [109] Musschoot J, Deduytsche D, Poelman H, et al. (2009) Comparison of thermal and plasma-enhanced ALD/CVD of vanadium pentoxide. J Electrochem Soc 156: P122–P126. doi: 10.1149/1.3133169
    [110] Badot J, Ribes S, Yousfi E, et al. (2000) Atomic layer epitaxy of vanadium oxide thin films and electrochemical behavior in presence of lithium ions. Electrochem Solid St 3: 485–488.
    [111] Chen X, Pomerantseva E, Banerjee P, et al. (2012) Ozone-based atomic layer deposition of crystalline V2O5 films for high performance electrochemical energy storage. Chem Mater 24: 1255–1261. doi: 10.1021/cm202901z
    [112] Badot JC, Mantoux A, Baffier N, et al. (2004) Electrical properties of V2O5 thin films obtained by atomic layer deposition (ALD). J Mater Chem 14: 3411–3415. doi: 10.1039/b410324f
    [113] Ostreng E, Gandrud KB, Hu Y, et al. (2014) High power nano-structured V2O5 thin film cathodes by atomic layer deposition. J Mater Chem A 2: 15044–15051. doi: 10.1039/C4TA00694A
    [114] Chen X, Pomerantseva E, Gregorczyk K, et al. (2013) Cathodic ALD V2O5 thin films for high-rate electrochemical energy storage. RSC Adv 3: 4294–4302. doi: 10.1039/c3ra23031g
    [115] Blanquart T, Niinisto J, Gavagnin M, et al. (2013) Atomic layer deposition and characterization of vanadium oxide thin films. RSC Adv 3: 1179–1185. doi: 10.1039/C2RA22820C
    [116] Daubert JS, Lewis NP, Gotsch HN, et al. (2015) Effect of meso- and micro-porosity in carbon electrodes on atomic layer deposition of pseudocapacitive V2O5 for high performance supercapacitors. Chem Mater 27: 6524–6534. doi: 10.1021/acs.chemmater.5b01602
    [117] Santos L, Swiatowska J, Lair V, et al. (2017) Mechanisms of enhanced lithium intercalation into thin film V2O5 in ionic liquids investigated by X-ray photoelectron spectroscopy and time-of-flight secondary ion mass spectroscopy. J Power Sources 364: 61–71. doi: 10.1016/j.jpowsour.2017.08.003
    [118] Wang X, Guo Z, Gao Y, et al. (2017) Atomic layer deposition of vanadium oxide thin films from tetrakis(dimethylamino)vanadium precursor. J Mater Res 32: 37–44. doi: 10.1557/jmr.2016.303
    [119] Sreedhara MB, Ghatak J, Bharath B, et al. (2017) Atomic layer deposition of ultrathin crystalline epitaxial films of V2O5. ACS Appl Mater Inter 9: 3178–3185.
    [120] Chain EE (1991) Optical properties of vanadium dioxide and vanadium pentoxide thin films. Appl Optics 30: 2782–2787. doi: 10.1364/AO.30.002782
    [121] Cezar AB, Graff IL, Varalda J, et al. (2014) Oxygen-vacancy-induced room-temperature magnetization in lamellar V2O5 thin films. J Appl Phys 116: 163904. doi: 10.1063/1.4899249
    [122] Ranea VA, Dammig-Quina PL (2016) The structure of the bulk and the (001) surface of V2O5. A DFT + U study. Mater Res Express 3: 085005. doi: 10.1088/2053-1591/3/8/085005
    [123] Chakrabarti A, Hermann K, Druzinic R, et al. (1999) Geometric and electronic structure of vanadium pentoxide: a density functional bulk and surface study. Phys Rev B 59: 10583–10590. doi: 10.1103/PhysRevB.59.10583
    [124] Eyert V, Hck KH (1998) Electronic structure of V2O5: role of octahedral deformations. Phys Rev B 57: 12727–12737. doi: 10.1103/PhysRevB.57.12727
    [125] Audière JP, Madi A, Grenet JC (1982) Electrical and thermal properties of highly quenched amorphous V2O5 thin films. J Mater Sci 17: 2973–2978. doi: 10.1007/BF00644678
    [126] Kumagai N, Kitamoto H, Baba M, et al. (1998) Intercalation of lithium in rf-sputtered vanadium oxide film as an electrode material for lithium-ion batteries. J Appl Electrochem 28: 41–48.
    [127] Benmoussa M, Outzourhit A, Jourdani R, et al. (2003) Structural, optical and electrochromic properties of sol-gel V2O5 thin films. Active Passive Electron Comp 26: 245–256. doi: 10.1080/0882751031000116223
    [128] Losurdo M, Barreca D, Bruno G, et al. (2001) Spectroscopic ellipsometry investigation of V2O5 nanocrystalline thin films. Thin Solid Films 384: 58–64. doi: 10.1016/S0040-6090(00)01820-4
    [129] Vernardou D, Paterakis P, Drosos H, et al. (2011) A study of the electrochemical performance of vanadium oxide thin films grown by atmospheric pressure chemical vapour deposition. Sol Energ Mat Sol C 95: 2842–2847. doi: 10.1016/j.solmat.2011.05.046
    [130] Guimond S, Sturm JM, Göbke D, et al. (2008) Well-ordered V2O5(001) thin films on Au(111): growth and thermal stability. J Phys Chem C 112: 11835–11846. doi: 10.1021/jp8011156
    [131] Szörényi T, Bali K, Hevesi I (1980) Structural characterization of amorphous vanadium pentoxide thin films prepared by chemical vapour deposition. J Non-Cryst Solids 35–36: 1245–1248.
    [132] Özer N (1997) Electrochemical properties of sol-gel deposited vanadium pentoxide films. Thin Solid Films 305: 80–87. doi: 10.1016/S0040-6090(97)00086-2
    [133] Kumar RTR, Karunagaran B, Kumar VS, et al. (2003) Structural properties of V2O5 thin films prepared by vacuum evaporation. Mat Sci Semicon Proc 6: 543–546. doi: 10.1016/j.mssp.2003.08.017
    [134] Gilles E, Boesman E (1966) EPR studies of V2O5 single crystals. I. Defect centres in pure, non-stoichiometric vanadium pentoxide. Phys Stat Solidi B 14: 337–347.
    [135] Pozarnsky G, Wright L, McCormijk A (1994) Effects of aging time on V2O5 sol-gel coatings. J Sol-Gel Sci Techn 3: 57–62. doi: 10.1007/BF00490149
    [136] Senapati S, Panda S (2016) Effect of aging of V2O5 sol on properties of nanoscale films. Thin Solid Films 599: 42–48. doi: 10.1016/j.tsf.2015.12.045
    [137] Anaissi FJ, Demets GJF, Alvarez EB, et al. (2001) Long-term aging of vanadium(V) oxide xerogel precursor solutions: structural and electrochemical implications. Electrochim Acta 47: 441–450. doi: 10.1016/S0013-4686(01)00737-X
    [138] Hirashima H, Gengyo M, Kojima C, et al. (1995) Effects of aging and drying on structure of V2O5 gels. J Non-Cryst Solids 186: 54–58. doi: 10.1016/0022-3093(95)00033-X
    [139] Talledo A, Granqvist CG (1995) Electrochromic vanadium-pentoxide-based films: structural, electrochemical and optical properties. J Appl Phys 77: 4655–4666. doi: 10.1063/1.359433
    [140] Luksich J, Aita CH (1991) Annealing response of disordered sputter deposited vanadium pentoxide (V2O5). J Vac Sci Technol A 9: 542–546. doi: 10.1116/1.577405
    [141] Song GY, Oh C, Sinha S, et al. (2017) Facile phase control of multivalent vanadium oxide tin films (V2O5 and VO2) by atomic layer deposition and postdeposition annealing. ACS Appl Mater Inter 9: 23909–23917. doi: 10.1021/acsami.7b03398
    [142] Luo Z, Wu Z, Xu X, et al. (2010) Impact of substrate temperature on the microstructure, electrical and optical properties of sputtered nanoparticle V2O5 thin films. Vacuum 85: 145–150. doi: 10.1016/j.vacuum.2010.05.001
    [143] El Sound AMA, Mansour B, Soliman LI (1994) Optical and electrical properties of V2O5 thin films. Thin Solid Films 247: 140–143. doi: 10.1016/0040-6090(94)90487-1
    [144] Akl AA (2010) Thermal annealing effect on the crystallization and optical dispersion of sprayed V2O5 thin films. J Phys Chem Solids 71: 223–229. doi: 10.1016/j.jpcs.2009.11.009
    [145] Maki K, Fukuda T, Momose H, et al. (2012) Structural and optical properties of reactive-sputtered films of V2O5: measurement of optical bandgap and roughness correction. J Fac Sci Tech Seikei Univ 49: 41–44.
    [146] Atuchin VV, Ayupov BM, Kochubey VA, et al. (2008) Optical properties of textured V2O5/Si thin films deposited by reactive magnetron sputtering. Opt Mater 30: 1145–1148. doi: 10.1016/j.optmat.2007.05.040
    [147] Wu QH, Thissen A, Jaegermann W, et al. (2004) Photoelectron spectroscopy study of oxygen vacancy on vanadium oxides surface. Appl Surf Sci 236: 473–478. doi: 10.1016/j.apsusc.2004.05.112
    [148] Wu QH, Thißen A, Jaegerman W (2005) Photoelectron spectroscopic study of Li intercalation into V2O5 thin films. Surf Sci 578: 203–212. doi: 10.1016/j.susc.2005.01.042
    [149] Benayad A, Martinez H, Gies A, et al. (2006) Vanadium pentoxide thin films used as positive electrode in lithium microbatteries: an XPS study during cycling. J Phys Chem Solids 67: 1320–1324. doi: 10.1016/j.jpcs.2006.01.089
    [150] Swiatowska-Mrowiecka J, Martin F, Maurice V, et al. (2008) The distribution of lithium intercalated in V2O5 thin films studied by XPS and ToF-SIMS. Electrochim Acta 53: 4257–4266. doi: 10.1016/j.electacta.2007.12.083
    [151] Salvi AM, Guascito MR, DeBonis A, et al. (2003) Lithium intercalation on amorphous V2O5 thin film, obtained by r.f. deposition, using in situ sample transfer for XPS analysis. Surf Interface Anal 35: 897–905.
    [152] Ibris N, Salvi AM, Liberatore M, et al. (2005) XPS study of the Li intercalation process in sol-gel-produced V2O5 thin film: influence of substrate and film synthesis modification. Surf Interface Anal 37: 1092–1104. doi: 10.1002/sia.2084
    [153] Alamarguy D, Castle JE, Ibris N, et al. (2006) Characterization of sol-gel crystalline V2O5 thin films after Li intercalation cycling. Surf Interface Anal 38: 801–804. doi: 10.1002/sia.2138
    [154] Fleutot B, Martinez H, Pecquenard B, et al. (2008) Surface film morphology (AFM) and chemical features (XPS) of cycled V2O5 thin films in lithium microbatteries. J Power Sources 180: 836–844. doi: 10.1016/j.jpowsour.2008.02.080
    [155] Alamarguy D, Castle JE, Liberatore M, et al. (2006) Distribution of intercalated lithium in V2O5 thin films determined by SIMS depth profiling. Surf Interface Anal 38: 847–850. doi: 10.1002/sia.2139
    [156] Julien CM, Massot M (2003) Lattice vibrations of materials for lithium rechargeable batteries III. Lithium manganese oxides. Mater Sci Eng B-Adv 100: 69–78.
    [157] Julien CM, Massot M (2003) Lattice vibrations of materials for lithium rechargeable batteries I. Lithium manganese oxides. Mater Sci Eng B-Adv 97: 217–230.
    [158] Abello L, Husson E, Repelin Y, et al. (1983) Vibrational spectra and valence force field of crystalline V2O5. Spectrochim Acta A 39: 641–651. doi: 10.1016/0584-8539(83)80040-3
    [159] Baddour-Hadjean R, Pereira-Ramos JP, Navone C, et al. (2008) Raman microspectrometry study of electrochemical lithium intercalation into sputtered crystalline V2O5 thin films. Chem Mater 20: 1916–1923. doi: 10.1021/cm702979k
    [160] Surca A, Orel B, Drazic G, et al. (1999) Ex situ and in situ infrared spectroelectrochemical investigations of V2O5 crystalline films. J Electrochem Soc 146: 232–242. doi: 10.1149/1.1391592
    [161] Fang GJ, Liu ZL, Wang Y, et al. (2001) Synthesis and structural, electrochromic characterization of pulsed laser deposited vanadium oxide thin films. J Vac Sci Technol A 19: 887–892. doi: 10.1116/1.1359533
    [162] Su Q, Liu Q, Ma ML, et al. (2008) Raman spectroscopic characterization of the microstructure of V2O5 films. J Solid State Electr 12: 919–923. doi: 10.1007/s10008-008-0515-5
    [163] Lee SH, Cheong HM, Seong MJ, et al. (2002) Microstructure study of amorphous vanadium oxide thin films using Raman spectroscopy. J Appl Phys 92: 1893–1897. doi: 10.1063/1.1495074
    [164] Lee SH, Cheong HM, Seong MJ, et al. (2003) Raman spectroscopic studies of amorphous vanadium oxide thin films. Solid State Ionics 165: 111–116. doi: 10.1016/j.ssi.2003.08.022
    [165] Zhang Y, Liu YW, Cheng YS, et al. (2005) Electrochemical impedance spectra of V2O5 xerogel films with intercalation of lithium ion. J Cent South Univ T 12: 309–314. doi: 10.1007/s11771-005-0151-5
    [166] Julien C, Khelfa A, Guesdon JP, et al. (1996) Electrical properties of flash-evaporated V2O5 films. Ionics 2: 380–385. doi: 10.1007/BF02375816
    [167] Sanchez C, Livage J, Audiere JP, et al. (1984) Influence of the quenching rate on the properties of amorphous V2O5 thin films. J Non-Cryst Solids 65: 285–300. doi: 10.1016/0022-3093(84)90053-X
    [168] Mott NF (1968) Conduction in glasses containing transition metal ions. J Non-Crystal Solids 1: 1–17.
    [169] Akl AA (2007) Crystallization and electrical properties of V2O5 thin films prepared by RF sputtering. Appl Surf Sci 253: 7094–7099. doi: 10.1016/j.apsusc.2007.02.054
    [170] Wang HQ, Li N, Guldal NS, et al. (2012) Nanocrystal V2O5 thin film as hole-extraction layer in normal architecture organic solar cells. Org Electron 13: 3014–3021. doi: 10.1016/j.orgel.2012.08.007
    [171] Schneider K, Lubecka M, Czapla A (2016) V2O5 thin film for gas sensor applications. Sensor Actuat B-Chem 236: 970–977. doi: 10.1016/j.snb.2016.04.059
    [172] Park HK, Smyrl WH, Ward MD (1995) V2O5 xerogel films as intercalation hosts for lithium. J Electrochem Soc 142: 1068–1073. doi: 10.1149/1.2044133
    [173] Jourdani R, Jadoual L, Ait El Fqih M, et al. (2018) Effects of lithium insertion and deinsertion into V2O5 thin films: optical, structural and absorption properties. Surf Interface Anal 50: 52–58. doi: 10.1002/sia.6331
    [174] Julien C, Ivanov I, Gorenstein A (1995) Vibrational modifications on lithium intercalation in V2O5 films. Mater Sci Eng B-Adv 33: 168–172. doi: 10.1016/0921-5107(95)80032-8
    [175] Jung H, Gerasopoulos K, Talin AA, et al. (2017) A platform for in situ Raman and stress characterizations of V2O5 cathode using MEMS device. Electrochim Acta 242: 227–239. doi: 10.1016/j.electacta.2017.04.160
    [176] Surca A, Orel B (1999) IR spectroscopy of crystalline V2O5 films in different stages of lithiation. Electrochim Acta 44: 3051–3057. doi: 10.1016/S0013-4686(99)00019-5
    [177] Meulenkam EA, van Klinken W, Schlatmann AR (1999) In-situ x-ray diffraction of Li intercalation in sol-gel V2O5 films. Solid State Ionics 126: 235–244. doi: 10.1016/S0167-2738(99)00243-X
    [178] Cazzanelli E, Mariotto, Passerini S, et al. (1994) Spectroscopic investigations of Li-intercalated V2O5 polycrystalline films. Solid State Ionics 70–71: 412–416.
    [179] Baddour-Hadjean R, Golabkan V, Pereira-Ramos JP, et al. (2002) A Raman study of the lithium insertion process in vanadium pentoxide thin films deposited by atomic layer deposition. J Raman Spectrosc 33: 631–638. doi: 10.1002/jrs.893
    [180] Baddour-Hadjean R, Navone C, Pereira-Ramos JP (2009) In situ Raman microspectrometry investigation of electrochemical lithium intercalation into sputtered crystalline V2O5 thin films. Electrochim Acta 54: 6674–6679. doi: 10.1016/j.electacta.2009.06.052
    [181] Levi MD, Lu Z, Aurbach D (2001) Li-insertion into thin monolithic V2O5 films electrodes characterized by a variety of electroanalytical techniques. J Power Sources 97–98: 482–485.
    [182] Li YM, Hibino M, Tanaka Y, et al. (2001) Evaluation of Mo-doped amorphous V2O5 films as a positive electrode for lithium batteries. Solid State Ionics 143: 67–72. doi: 10.1016/S0167-2738(01)00834-7
    [183] Navone C, Baddour-Hadjean R, Pereira-Ramos JP, et al. (2005) High-performance oriented V2O5 thin films prepared by DC sputtering for rechargeable lithium microbatteries. J Electrochem Soc 152: A1790–A1796. doi: 10.1149/1.1990160
    [184] Navone C, Pereira-Ramos JP, Baddour-Hadjean R, et al. (2006) High-capacity crystalline V2O5 thick films prepared by RF sputtering as positive electrodes for rechargeable lithium microbatteries. J Electrochem Soc 153: A2287–A2293. doi: 10.1149/1.2358852
    [185] Donsanti F, Kostourou K, Decker F, et al. (2006) Alkali ion intercalation in V2O5: preparation and laboratory characterization of thin films produced by ADL. Surf Interface Anal 38: 815–818. doi: 10.1002/sia.2237
    [186] Decker F, Donsanti F, Salvi AM, et al. (2008) Li+ distribution into V2O5 films resulting from electrochemical intercalation reactions. J Brazil Chem Soc 19: 667–671.
    [187] Park YJ, Ryu KS, Kim KM, et al. (2002) Electrochemical properties of vanadium oxide thin films deposited by rf sputtering. Solid State Ionics 154: 229–235.
    [188] Liu Y, Li J, Zhang Q, et al. (2011) Porous nanostructured V2O5 film electrode with excellent Li-ion intercalation properties. Electrochem Commun 13: 1276–1279. doi: 10.1016/j.elecom.2011.08.024
    [189] Sahana MB, Sudakar C, Thapa C, et al. (2007) Electrochemical properties of V2O5 thin films deposited by spin coating. Mater Sci Eng B-Adv 143: 42–50. doi: 10.1016/j.mseb.2007.08.002
    [190] Yu D, Qiao Y, Zhou X, et al. (2014) Mica-like vanadium pentoxide-nanostructured thin film as high-performance cathode for lithium-ion batteries. J Power Sources 266: 1–6. doi: 10.1016/j.jpowsour.2014.04.099
    [191] Pomerantseva E, Gerasopoulos K, Chen X, et al. (2012) Electrochemical performance of the nanostructured biotemplated V2O5 cathode for lithium-ion batteries. J Power Sources 206: 282–287. doi: 10.1016/j.jpowsour.2012.01.127
    [192] Swiatowska-Mrowiecka J, Maurice V, Zanna S, et al. (2007) Ageing of V2O5 thin films induced by Li intercalation multi-cycling. J Power Sources 170: 160–172. doi: 10.1016/j.jpowsour.2007.04.014
    [193] Pyun SI, Bae JS (1997) Electrochemical lithium intercalation into vanadium pentoxide xerogel film electrode. J Power Sources 68: 669–673. doi: 10.1016/S0378-7753(96)02639-0
    [194] Navone C, Baddour-Hadjean R, Pereira-Ramos JP, et al. (2008) A kinetic study of electrochemical lithium insertion into oriented V2O5 thin films prepared by rf sputtering. Electrochim Acta 53: 3329–3336. doi: 10.1016/j.electacta.2007.11.061
    [195] Vivier V, Farcy J, Pereira-Ramos JP (1998) Electrochemical lithium insertion in sol-gel crystalline vanadium pentoxide thin films. Electrochim Acta 44: 831–839. doi: 10.1016/S0013-4686(98)00234-5
    [196] Miyazaki H, Sakamura H, Kamei M, et al. (1999) Electrochemical evaluation of oriented vanadium oxide films deposited by reactive rf magnetron sputtering. Solid State Ionics 122: 223–229. doi: 10.1016/S0167-2738(99)00022-3
    [197] McGraw JM, Bahn CS, Parilla PA, et al. (1999) Li ion diffusion measurements in V2O5 and Li(Co1−xAlx)O2 thin-film battery cathodes. Electrochim Acta 45: 187–196. doi: 10.1016/S0013-4686(99)00203-0
    [198] Lu Z, Levi MD, Salitra G, et al. (2000) Basic electroanalytical characterization of lithium insertion into thin, well-crystallized V2O5 films. J Electroanal Chem 491: 211–221. doi: 10.1016/S0022-0728(00)00184-4
    [199] Li Y, Kunitake T, Aoki Y (2007) Synthesis and Li+ intercalation/extraction in ultrathin V2O5 layer and freestanding V2O5/Pt/PVA multilayer films. Chem Mater 19: 575–580. doi: 10.1021/cm0626233
    [200] Jung KN, Pyun SI (2006) Effect of pore structure on anomalous behavior of the lithium intercalation into porous V2O5 film electrode using fractal geometry concept. Electrochim Acta 51: 2646–2655. doi: 10.1016/j.electacta.2005.08.008
    [201] Wang S, Li S, Sun Y, et al. (2011) Three-dimensional porous V2O5 cathode with ultra-high rate capability. Energ Environ Sci 4: 2854–2857. doi: 10.1039/c1ee01172c
    [202] Li SR, Ge SY, Qiao Y, et al. (2012) Three-dimensional porous Fe0.1V2O5.15 thin film as a cathode material for lithium ion batteries. Electrochim Acta 64: 81–86.
    [203] Mui SC, Jasinski J, Leppert VJ, et al. (2006) Microstructure effects on the electrochemical kinetics of vanadium pentoxide thin-film cathodes. J Electrochem Soc 153: A1372–A1377. doi: 10.1149/1.2199327
    [204] Navone C, Pereira-Ramos JP, Baddour-Hadjean R, et al. (2005) Electrochemical and structural properties of V2O5 thin films prepared by DC sputtering. J Power Sources 146: 327–330. doi: 10.1016/j.jpowsour.2005.03.024
    [205] Wei Y, Ryu CW, Kim KB (2008) Cu-doped V2O5 as a high-energy density cathode material for rechargeable lithium batteries. J Alloy Compd 459: L13–L17. doi: 10.1016/j.jallcom.2007.04.275
    [206] Park HK (2005) Manganese vanadium oxides as cathodes for lithium batteries. Solid State Ionics 176: 307–312. doi: 10.1016/j.ssi.2004.07.014
    [207] Grégoire G, Baffier N, Kahn-Harari A, et al. (1998) X-Ray powder diffraction study of a new vanadium oxide Cr0.11V2O5.16 synthesized by a sol-gel process. J Mater Chem 8: 2103–2108.
    [208] Giorgetti M, Berrettoni M, Smyrl WH (2007) Doped V2O5-based cathode materials: where does the doping metal go? An X-ray absorption spectroscopy study. Chem Mater 19: 5991–6000.
    [209] Wei Y, Zhou J, Zheng J, et al. (2015) Improved stability of electrochromic devices using Ti-doped V2O5 film. Electrochim Acta 166: 277–284. doi: 10.1016/j.electacta.2015.03.087
    [210] Gies A, Pecquenard B, Benayad A, et al. (2008) Effect of silver co-sputtering on V2O5 thin films for lithium microbatteries. Thin Solid Films 516: 7271–7281. doi: 10.1016/j.tsf.2007.12.165
    [211] Avellaneda CO, Bulhoes LOS (2006) Optical and electrochemical properties of V2O5:Ta sol-gel thin films. Sol Energ Mat Sol C 90: 444–451. doi: 10.1016/j.solmat.2005.04.031
    [212] Panagopoulou M, Vernardou D, Koudouma E, et al. (2017) Tunable properties of Mg-doped V2O5 thin films for energy applications: Li-ion batteries and electrochromics. J Phys Chem C 121: 70–79. doi: 10.1021/acs.jpcc.6b09018
    [213] Yu DM, Zhang ST, Liu DW, et al. (2010) Effect of manganese doping on Li-ion intercalation properties of V2O5 films. J Mater Chem 20: 10841–10846. doi: 10.1039/c0jm01252a
    [214] Ozer N, Sabuncu S, Cronin J (1999) Electrochromic properties of sol-gel deposited Ti-doped vanadium oxide film. Thin Solid Films 338: 201–206. doi: 10.1016/S0040-6090(98)00974-2
    [215] Coustier F, Passerini S, Smyrl WH (1997) Dip-coated silver-doped V2O5 xerogels as host materials for lithium intercalation. Solid State Ionics 100: 247–258. doi: 10.1016/S0167-2738(97)00354-8
    [216] Nam SC, Lim YC, Park HY, et al. (2001) The effects of Cu-doping in V2O5 thin film cathode for microbattery. Korean J Chem Eng 18: 673–678. doi: 10.1007/BF02706385
    [217] Yong W, Zhang HL, Cao HT, et al. (2016) Effect of post-annealing on structural and electrochromic properties of Mo-doped V2O5 thin films. J Sol-Gel Sci Techn 77: 604–609. doi: 10.1007/s10971-015-3889-8
    [218] Guan S, Wei Y, Zhou J, et al. (2016) A method for preparing manganese-doped V2O5 films with enhanced cycling stability. J Electrochem Soc 163: H541–H545. doi: 10.1149/2.0761607jes
    [219] Gouda GM, Nagendra CL (2013) Preparation and characterization of thin film thermistors of metal oxides of manganese and vanadium (Mn–V–O). Sensor Actuat A-Phys 190: 181–190. doi: 10.1016/j.sna.2012.11.019
    [220] Li Y, Yao J, Uchaker E, et al. (2013) Sn-doped V2O5 film with enhanced lithium-ion storage performance. J Phys Chem C 117: 23507–23514. doi: 10.1021/jp406927m
    [221] Etemadi B, Mazloom J, Ghodsi FE (2017) Phase transition and surface morphology effects on optical, electrical and lithiation/delithiation behavior of nanostructured Ce-doped V2O5 thin films. Mat Sci Semicon Proc 61: 99–106. doi: 10.1016/j.mssp.2016.12.035
    [222] Sahana MB, Sudakar C, Thapa C, et al. (2009) The effect of titanium on the lithium intercalation capacity of V2O5 thin films. Thin Solid Films 517: 6642–6651. doi: 10.1016/j.tsf.2009.04.063
    [223] Lee K, Cao GZ (2005) Enhancement of intercalation properties of V2O5 film by TiO2 addition. J Phys Chem B 109: 11880–11885. doi: 10.1021/jp044651j
    [224] Kim S, Taya M, Xu C (2009) Contrast, switching speed, and durability of V2O5–TiO2 film-based electrochromic windows. J Electrochem Soc 156: E40–E45. doi: 10.1149/1.3031978
    [225] Kim S, Taya M (2012) Electrochromic windows based on V2O5–TiO2 and poly(3,3-dimethyl-3,4-dihydro-2H-thieno[3,4-b][1,4]dioxepine) coatings. Sol Energ Mat Sol C 107: 225–229. doi: 10.1016/j.solmat.2012.06.032
    [226] Zhang YF, Hsu CY, Hu CC (2017) Deposition of tungsten-doped V2O5 thin films on non-alkali glass substrate by RF magnetron sputtering for thermal insulation. Key Eng Mater 732: 16–23. doi: 10.4028/www.scientific.net/KEM.732.16
    [227] Patil CE, Tarwal NL, Jadhav PR, et al. (2014) Electrochromic performance of the mixed V2O5–WO3 thin films synthesized by pulsed spray pyrolysis technique. Curr Appl Phys 14: 389–395. doi: 10.1016/j.cap.2013.12.014
    [228] Acharya BS, Nayak BB (2008) Microstructural studies of nanocrystalline thin films of V2O5–MoO3 using X-ray diffraction, optical absorption and laser micro Raman spectroscopy. Indian J Pure Appl Phys 46: 866–875.
    [229] Kim YS, Ahn HJ, Shim HS, et al. (2006) Electrochemical and structural properties of MoO3–V2O5 nanocomposite thin film electrodes for lithium rechargeable batteries. Solid State Ionics 177: 1323–1326. doi: 10.1016/j.ssi.2006.05.031
    [230] Ozer N, Lampert CM (1999) Electrochromic performance of sol-gel deposited WO3–V2O5 films. Thin Solid Films 349: 205–211. doi: 10.1016/S0040-6090(99)00144-3
    [231] Madhuri KV, Naidu BS, Hussain OM (2002) Optical absorption studies on (V2O5)1−x–(MoO3)x thin films. Mater Chem Phys 77: 22–26.
    [232] Demets GJF, Anaissi FJ, Toma HE (2000) Electrochemical properties of assembled polypyrrole: V2O5 xerogel films. Electrochim Acta 46: 547–554. doi: 10.1016/S0013-4686(00)00635-6
    [233] Zhang X, Sun H, Li Z, et al. (2013) Synthesis and electrochromic characterization of vanadium pentoxide/graphene nanocomposite films. J Electrochem Soc 160: H587–H590. doi: 10.1149/2.064309jes
    [234] Wu J, Gao X, Yu H, et al. (2016) A scalable free-standing V2O5/CNT film electrode for supercapacitors with a wide operation voltage (1.6 V) in an aqueous electrolyte. Adv Funct Mater 26: 6114–6120.
    [235] Chen W, Mai L, Xu Q, et al. (2003) Synthesis and Li-insertion properties of poly(ethylene-oxide)/V2O5 nanocomposite films. Solid State Phenom 90–91: 19–24.
    [236] Murphy DW, Christian PA, DiSalvo FJ, et al. (1979) Lithium incorporation by vanadium pentoxide. Inorg Chem 18: 2800–2803. doi: 10.1021/ic50200a034
    [237] Cocciantelli JM, Ménétrier M, Delmas C, et al. (1995) On the δ → γ irreversible transformation in Li//V2O5 secondary batteries. Solid State Ionics 78: 143–150. doi: 10.1016/0167-2738(95)00015-X
    [238] Zhang L, Song J, Dong Q, et al. (2009) Application of V2O5 in thin film microbatteries prepared by rf magnetron sputtering. Adv Mater Res 79–82: 931–934.
    [239] Gu G, Schmid M, Chiu PW, et al. (1998) V2O5 nanofibre sheet actuators. Nat Mater 2: 316–319.
    [240] Ponzi M, Duschatzky C, Carrascull A, et al. (1998) Obtaining benzaldehyde via promoted V2O5 catalysts. Appl Catal A-Gen 169: 373–379. doi: 10.1016/S0926-860X(98)00026-X
    [241] Livage J (1991) Vanadium pentoxide gels. Chem Mater 3: 578–593. doi: 10.1021/cm00016a006
    [242] Granqvist CG (2002) Handbook of inorganic electrochromic materials, Amsterdam: Elsevier, 295–337.
    [243] Hirashima H, Ide M, Yoshida T (1986) Memory switching of V2O5-TeO2 glasses. J Non-Cryst Solids 86: 327–335. doi: 10.1016/0022-3093(86)90021-9
    [244] Julien C (1994) Thin film technology and microbatteries, In: Pistoia G, Lithium Batteries-New Materials-Development and Perspectives, Amsterdam: Elsevier, 167–237.
    [245] Bates JB, Dudney NJ, Gruzalski GR, et al. (1993) Fabrication and characterization of amorphous lithium electrolyte thin films and rechargeable thin-film batteries. J Power Sources 43–44: 103–110.
    [246] Nakazawa H, Sano K, Baba M (2005) Fabrication by using a sputtering method and charge–discharge properties of large-sized and thin-film lithium ion rechargeable batteries. J Power Sources 146: 758–761. doi: 10.1016/j.jpowsour.2005.03.155
    [247] Lee SH, Liu P, Tracy CE, et al. (1999) All-solid-state rocking chair lithium battery on a flexible Al substrate. Electrochem Solid St 2: 425–427. doi: 10.1149/1.1390859
    [248] Dudney NJ, Neudecker BJ (1999) Soli state thin-film lithium battery systems. Curr Opin Solid State Mater Sci 4: 479–482. doi: 10.1016/S1359-0286(99)00052-2
    [249] Oukassi S, Salot R, Pereira-Ramos JP (2009) Elaboration and characterization of crystalline rf deposited V2O5 positive electrode for thin film battery. Appl Surf Sci 256: 149–155. doi: 10.1016/j.apsusc.2009.07.115
    [250] Navone C, Tintignac S, Pereira-Ramos JP, et al. (2011) Electrochemical behaviour of sputtered c-V2O5 and LiCoO2 thin films for solid state lithium microbatteries. Solid State Ionics 192: 343–346. doi: 10.1016/j.ssi.2010.04.023
    [251] Liu P, Zhang JG, Turner JA, et al. (1998) Fabrication of LiV2O5 thin-film electrodes for rechargeable lithium batteries. Solid State Ionics 111: 145–151. doi: 10.1016/S0167-2738(98)00184-2
    [252] Kim YT, Gopukumar S, Kim KB, et al. (2003) Performance of electrostatic spray-deposited vanadium pentoxide in lithium secondary cells. J Power Sources 117: 110–117. doi: 10.1016/S0378-7753(03)00006-5
    [253] Navone C, Baddour-Hadjean R, Pereira-Ramos JP, et al. (2009) Sputtered crystalline V2O5 thin films for all-solid-state lithium microbatteries. J Electrochem Soc 156: A763–A767. doi: 10.1149/1.3170922
    [254] Lee SH, Liu P, Tracy CE (2003) Lithium thin-film battery with a reversed structural configuration SS/Li/Lipon/Li/Cu. Electrochem Solid St 6: A275–A277. doi: 10.1149/1.1623171
    [255] Navone C, Pereira-Ramos JP, Baddour-Hadjean R, et al. (2010) Lithiated c-V2O5 thin-film as positive electrode for rocking-chair solid-state lithium microbattery. Ionics 16: 577–580. doi: 10.1007/s11581-010-0460-z
    [256] Gerbaldi C, Destro M, Nair JR, et al. (2013) High-rate V2O5-based Li-ion thin film polymer cell with outstanding long-term cyclability. Nano Energy 2: 1279–1286. doi: 10.1016/j.nanoen.2013.06.007
    [257] Zeng Y, Gao G, Wu G, et al. (2015) Nanosheet-structured vanadium pentoxide thin film as a carbon- and binder-free cathode for lithium-ion battery applications. J Solid State Electr 19: 3319–3328. doi: 10.1007/s10008-015-2941-5
    [258] Mattelaer F, Kobe G, Rampelberg G, et al. (2016) Atomic layer deposition of vanadium oxides for thin-film lithium-ion battery applications. RSC Adv 6: 114658–114665. doi: 10.1039/C6RA25742A
    [259] Mattelaer F, Geryl K, Rampelberg G, et al. (2017) Amorphous and crystalline vanadium oxides as high-energy and high-power cathodes for three-dimensional thin-film lithium ion batteries. ACS Appl Mater Inter 9: 13121–13131. doi: 10.1021/acsami.6b16473
    [260] Chae OB, Kim J, Park I, et al. (2014) Reversible lithium storage at highly populated vacant sites in an amorphous vanadium pentoxide electrode. Chem Mater 26: 5874–5881. doi: 10.1021/cm502268u
    [261] Ku JH, Ryu JH, Kim SH, et al. (2012) Reversible lithium storage with high mobility at structural defects in amorphous molybdenum dioxide electrode. Adv Funct Mater 22: 3658–3664. doi: 10.1002/adfm.201102669
    [262] Sutar MA, Pawar MS, Rendale MK, et al. (2015) Enhancing electrochemical performance of V2O5 thin film by using ultrasonic weltering. IORS J Appl Phys 7: 41–45.
    [263] Sandhya GL, Dhananjaya M, Prakash NG, et al. (2017) Structural and supercapacitive performance of V2O5 thin films prepared by dc magnetron sputtering. IOSR J Appl Chem 10: 64–69.
    [264] Wruck D, Ramamurthi S, Rubin M (1989) Sputtered electrochromic V2O5 films. Thin Solid Films 182: 79–85. doi: 10.1016/0040-6090(89)90245-9
    [265] Talledo A, Andersson AM, Granqvist CG (1990) Electrochemically lithiated V2O5 films: An optically passive ion storage for transparent electrochromic devices. J Mater Res 5: 1253–1256. doi: 10.1557/JMR.1990.1253
    [266] Guan ZS, Yao JN, Yang YA, et al. (1998) Electrochromism of annealed vacuum-evaporated V2O5 films. J Electroanal Chem 443: 175–179. doi: 10.1016/S0022-0728(97)00594-9
    [267] Kaid MA (2006) Characterization of electrochromic vanadium pentoxide thin films prepared by spray pyrolysis. Egypt J Solids 29: 273–291.
    [268] Benmoussa M, Outzourhit A, Bennouna A, et al. (2008) Li+ ions diffusion into sol-gel V2O5 thin films: electrochromic properties. Eur Phys J Appl Phys 48: 10502.
    [269] Rosaiah P, Hussain OM (2013) Electron beam evaporated nano-crystalline V2O5 thin films for electrochromic and electrochemical applications, In: Giri P, Goswami D, Perumal A, Advanced Nanomaterials and Nanotechnology. Springer Proceedings in Physics, Berlin, Heidelberg: Springer, 143: 485–496. doi: 10.1007/978-3-642-34216-5_48
    [270] Chen LC, Ho KC (2001) Design equations for complementary electrochromic devices: application to the tungsten oxide–Prussian blue system. Electrochim Acta 46: 2151–2158. doi: 10.1016/S0013-4686(01)00368-1
    [271] Sarminio J, Talledd A, Andersson AA, et al. (1993) Stress and electrochromism induced by Li insertion in crystalline and amorphous V2O5 thin film electrodes. Electrochim Acta 12: 1637–1642.
    [272] Yoshino T, Baba N, Kouda Y (1987) Electrochromic properties of V2O5 thin films colloid-chemically deposited onto ITO glasses. Jpn J Appl Phys 26: 782–783. doi: 10.1143/JJAP.26.782
    [273] Shimizu Y, Nagase K, Miura N, et al. (1990) New preparation process of V2O5 thin film based on spin-coating from organic vanadium solution. Jpn J Appl Phys 29: L1708–L1711. doi: 10.1143/JJAP.29.L1708
    [274] Cremonesi A, Bersani D, Lottici PP, et al. (2006) Synthesis and structural characterization of mesoporous V2O5 thin films for electrochromic applications. Thin Solid Films 515: 1500–1505. doi: 10.1016/j.tsf.2006.04.029
    [275] Liberatore M, Decker F, Vuk AS, et al. (2006) Effect of the organic–inorganic template ICS-PPG on sol-gel deposited V2O5 electrochromic thin film. Sol Energ Mat Sol C 90: 434–443. doi: 10.1016/j.solmat.2005.04.035
    [276] Lin YS, Tsai CW, Chen PW (2008) Electrochromic properties of V2O5z thin films sputtered onto flexible PET/ITO substrates. Solid State Ionics 179: 290–297. doi: 10.1016/j.ssi.2008.01.054
    [277] Lu Y, Liu L, Mandler D, et al. (2013) High switching speed and coloration efficiency of titanium-doped vanadium oxide thin film electrochromic devices. J Mater Chem C 1: 7380–7386. doi: 10.1039/c3tc31508h
    [278] Kovendhan M, Joseph DP, Manimuthu P, et al. (2015) Prototype electrochromic device and dye sensitized solar cell using spray deposited undoped and 'Li' doped V2O5 thin film electrodes. Curr Appl Phys 215: 622–631.
    [279] Rizzo G, Arena A, Bonavita A, et al. (2010) Gasochromic response of nanocrystalline vanadium pentoxide films deposited from ethanol dispersions. Thin Solid Films 518: 7124–7127. doi: 10.1016/j.tsf.2010.07.010
    [280] Abbasi M, Rozati SM, Irani R, et al. (2015) Synthesis and gas sensing behavior of nanostructured V2O5 thin films prepared by spray pyrolysis. Mat Sci Semicon Proc 29: 132–138. doi: 10.1016/j.mssp.2014.01.008
    [281] Manno D, Serra A, Giulio MD, et al. (1997) Structural and electrical properties of sputtered vanadium oxide thin films for applications as gas sensing material. J Appl Phys 81: 2709–2714. doi: 10.1063/1.363973
    [282] Micocci G, Serra A, Tepore A, et al. (1997) Properties of vanadium oxide thin films for ethanol sensor. J Vac Sci Technol A 15: 34–38. doi: 10.1116/1.580471
    [283] Zhuiykov S, Wlodarski W, Li Y (2001) Nanocrystalline V2O5-TiO2 thin-films for oxygen sensing prepared by sol-gel process. Sensor Actuat B-Chem 77: 484–490. doi: 10.1016/S0925-4005(01)00739-0
    [284] Raj AD, Mangalaraj D, Ponpandian N, et al. (2010) Gas sensing properties of chemically synthesized V2O5 thin films. Adv Mater Res 123–125: 683–686.
    [285] Huotari J, Lappalainen J, Puustinen J, et al. (2013) Gas sensing properties of pulsed laser deposited vanadium oxide thin films with various crystal structures. Sensor Actuat B-Chem 187: 386–394. doi: 10.1016/j.snb.2012.12.067
    [286] Karthikeyan PS, Dhivya P, Raj PD, et al. (2016) V2O5 thin films for 2-propanol vapor sensing. Mater Today Proc 3: 1510–1516. doi: 10.1016/j.matpr.2016.04.035
    [287] Toxicological review of vanadium pentoxide (V2O5), 2011. Available from: https://ofmpub.epa.gov/eims/eimscomm.getfile?p_download_id=504127.
    [288] U.S. National Libray of Medecine, 2005. Available from: https://pubchem.ncbi.nlm.nih.gov/compound/ vanadium_pentoxide#section=GHS-Classification.
    [289] Eranna G, Joshi BC, Runthala DP, et al. (2010) Oxide materials for development of integrated gas sensors—A comprehensive review. Crit Rev Solid State 29: 111–188.
    [290] Kumar S, Qadir A, Maury F, et al. (2017) Visible thermochromism in vanadium pentaoxide coatings. ACS Appl Mater Inter 9: 21447–21456. doi: 10.1021/acsami.7b04484
  • This article has been cited by:

    1. Joseph McGhee, Vihar P. Georgiev, Simulation Study of Surface Transfer Doping of Hydrogenated Diamond by MoO3 and V2O5 Metal Oxides, 2020, 11, 2072-666X, 433, 10.3390/mi11040433
    2. H. Khmissi, Safwat A. Mahmoud, Alaa Ahmed Akl, Investigation of thermal annealing effect on the microstructure, morphology, linear and non-linear optical properties of spray deposited nanosized V2O5 thin films, 2021, 227, 00304026, 165979, 10.1016/j.ijleo.2020.165979
    3. T. D. Nguen, A. I. Zanko, D. A. Golosov, S. M. Zavadski, S. N. Melnikov, V. V. Kolos, Electrophysical properties of vanadium oxide films deposited by reactive magnetron sputtering, 2020, 18, 2708-0382, 94, 10.35596/1729-7648-2020-18-6-94-102
    4. Yongkang Xu, Yifeng Hu, Song Sun, Tianshu Lai, Changes in Resistance and Bandgap of V2O5 and Ge2Sb2Te5 during Phase Transition, 2021, 50, 0361-5235, 491, 10.1007/s11664-020-08599-5
    5. Q. J. Wang, H. Wang, Z. H. Zhou, J. Zuo, C. L. Zhang, The split-off terahertz radiating dipoles on thermally reduced α-V2O5 (001) surface, 2020, 12, 2040-3364, 21368, 10.1039/D0NR03889J
    6. Heungsoo Kim, Alberto Piqué, 2020, 9780128035818, 10.1016/B978-0-12-803581-8.12086-7
    7. Ziyauddin Khan, Prem Singh, Sajid Ali Ansari, Sai Rashmi Manippady, Amit Jaiswal, Manav Saxena, VO 2 Nanostructures for Batteries and Supercapacitors: A Review , 2021, 17, 1613-6810, 2006651, 10.1002/smll.202006651
    8. Christian M. Julien, Alain Mauger, Pulsed Laser Deposited Films for Microbatteries, 2019, 9, 2079-6412, 386, 10.3390/coatings9060386
    9. Md Hasan Ali, Md Mahabub Alam Moon, Md Masum Rana, Nasheat Siddequa, Structural, electronic and optical properties of Nb-doped β-V2O5 by first principles calculation, 2019, 6, 2053-1591, 116309, 10.1088/2053-1591/ab46cb
    10. Marjan Abdi Jamali, Alireza Arvani, Mostafa M. Amini, Vanadium Containing Metal‐organic Frameworks as Highly Efficient Catalysts for the Oxidation of Refractory Aromatic Sulfur Compounds, 2021, 13, 1867-3880, 293, 10.1002/cctc.202001327
    11. Huizhen Qin, Shunfei Liang, Lingyun Chen, Yang Li, Ziyang Luo, Shaowei Chen, Recent advances in vanadium-based nanomaterials and their composites for supercapacitors, 2020, 4, 2398-4902, 4902, 10.1039/D0SE00897D
    12. G Ravinder, C J Sreelatha, V Ganesh, Mohd. Shkir, Mohd Anis, P Chandar Rao, Thickness-dependent structural, spectral, linear, nonlinear and z-scan optical studies of V2O5 thin films prepared by a low-cost sol-gel spin coating technique, 2019, 6, 2053-1591, 096403, 10.1088/2053-1591/ab2992
    13. Top Khac Le, Manil Kang, Sok Won Kim, A review on the optical characterization of V2O5 micro-nanostructures, 2019, 45, 02728842, 15781, 10.1016/j.ceramint.2019.05.339
    14. Dane Tadeu Cestarolli, Elidia Maria Guerra, 2021, 10.5772/intechopen.96860
    15. Sandeep Yadav, Sonia Kumari, Renu Bala, Rajesh Thakur, Devendra Mohan, Characterization of Sn doped vanadium oxide thin films for nonlinear optical applications, 2021, 22147853, 10.1016/j.matpr.2021.02.712
    16. Rishabh Raj, Himanshu Gupta, L.P. Purohit, Performance of RF sputtered V2O5 interface layer in p-type CdTe/Ag Schottky diode, 2022, 126, 09253467, 112176, 10.1016/j.optmat.2022.112176
    17. Piyaporn Thangdee, Ekachai Chongsereecharoen, Thanun Chunjaemsri, Chinawat Ekwongsa, Pinit Kidkhunthod, Narong Chanlek, Natthapong Wongdamnern, Prapan Manyum, Saroj Rujirawat, Rattikorn Yimnirun, Local structure and structural properties of vanadium oxide thin films prepared by radio-frequency reactive magnetron sputtering at various oxygen flow rate, 2022, 586, 0015-0193, 213, 10.1080/00150193.2021.2014272
    18. Nai Yun Chang, Chuan Li, Jang-Hsing Hsieh, Reactive Sputtering Process Study for Vanadium Oxynitride Films, 2023, 13, 2079-6412, 459, 10.3390/coatings13020459
    19. Alex Pennetier, Willinton Y. Hernandez, Bright T. Kusema, Stéphane Streiff, Efficient hydrogenation of aliphatic amides to amines over vanadium-modified rhodium supported catalyst, 2021, 624, 0926860X, 118301, 10.1016/j.apcata.2021.118301
    20. J. Zuo, H. Wang, Q. Wang, Q.J. Wang, M. Zhang, Q.H. Li, Z.H. Zhou, X.Y. Wang, C.L. Zhang, Terahertz property of thermally reduced vanadium pentoxide films in vacuum, 2022, 196, 0042207X, 110756, 10.1016/j.vacuum.2021.110756
    21. S Tipawan Khlayboonme, Amorn Thedsakhulwong, Influence of the rf power and oxygen content on structural, electrical, and optical properties of V2O5 thin films prepared via reactive radio frequency sputtering, 2022, 9, 2053-1591, 076401, 10.1088/2053-1591/ac827a
    22. Samuel A. Hevia, Joseba Orive, Fernando Guzmán, Eduardo Cisternas, Fabian Dietrich, Roberto Villarroel, Judit Lisoni, High performance of V2O5 thin film electrodes for lithium-ion intercalation, 2022, 576, 01694332, 151710, 10.1016/j.apsusc.2021.151710
    23. Çağrı Durmuş, Tamer Akan, Vanadium pentoxide thin films deposited by the thermionic vacuum arc plasma, 2023, 770, 00406090, 139764, 10.1016/j.tsf.2023.139764
    24. S. K. Suresh Babu, D. Jackuline Moni, D. Gracia, Amsalu Gosu Adigo, Samson Jerold Samuel Chelladurai, Investigation on V2O5 Thin Films for Field Effect Transistor Applications, 2021, 2021, 1687-8442, 1, 10.1155/2021/2414589
    25. Jincheng Mei, Yi Li, Junyi Yan, Jiaqing Zhuang, Xingping Wang, Xin Zhang, Yuda Wu, Mengdi Zou, Chuang Peng, Wenyan Dai, Zhen Yuan, Ke Lin, Fabrication and photoelectric properties of n-V2O5/p-GaAs heterojunction, 2022, 152, 13698001, 107069, 10.1016/j.mssp.2022.107069
    26. Tina Tasheva, Vesselin Dimitrov, Correlation between structure and optical basicity of glasses in the TeO2-V2O5-MoO3 system, 2021, 570, 00223093, 120981, 10.1016/j.jnoncrysol.2021.120981
    27. Meenal D. Patil, Suprimkumar D. Dhas, Amol A. Mane, Annasaheb V. Moholkar, Clinker-like V2O5 nanostructures anchored on 3D Ni-foam for supercapacitor application, 2021, 133, 13698001, 105978, 10.1016/j.mssp.2021.105978
    28. Devendra Mohan, Anil Kumar, Monika Barala, Kavita Yadav, Third-order optical nonlinearity in molybdenum trioxide (MoO3
    )-doped vanadium pentaoxide (V2O5
    ) thin films, 2021, 75, 1434-6060, 10.1140/epjd/s10053-021-00242-0
    29. Alex Grant, Colm O'Dwyer, Real-time nondestructive methods for examining battery electrode materials, 2023, 10, 1931-9401, 011312, 10.1063/5.0107386
    30. S.A. Adewinbi, R.A. Busari, L.O. Animasahun, E. Omotoso, B.A. Taleatu, Effective pseudocapacitive performance of binder free transparent α-V2O5 thin film electrode: Electrochemical and some surface probing, 2021, 621, 09214526, 413260, 10.1016/j.physb.2021.413260
    31. T. D. Nguen, A. I. Zanko, D. A. Golosov, S. M. Zavadski, S. N. Melnikov, V. V. Kolos, T. Q. To, Influence of annealing on structure, phase and electrophysical properties of vanadium oxide films, 2021, 19, 2708-0382, 22, 10.35596/1729-7648-2021-19-3-22-30
    32. T. D. Nguyen, A. I. Zanko, D. A. Golosov, S. M. Zavadski, S. N. Melnikov, V. V. Kolos, N. К. Тоlochko, Formation and investigation of characteristics of thermoresistive structures based on vanadium oxide films, 2021, 19, 2708-0382, 85, 10.35596/1729-7648-2021-19-4-85-93
    33. Peng Hu, Ping Hu, Tuan Duc Vu, Ming Li, Shancheng Wang, Yujie Ke, Xianting Zeng, Liqiang Mai, Yi Long, Vanadium Oxide: Phase Diagrams, Structures, Synthesis, and Applications, 2023, 0009-2665, 10.1021/acs.chemrev.2c00546
    34. Devendra Mohan, Suraj Bhan, Purnima Arya, Reetu Sangwan, Kavita Yadav, Optical switching in molybdenum trioxide (MoO3)-doped vanadium pentaoxide (V2O5) thin films using Fabry–Perot cavity, 2023, 77, 1434-6060, 10.1140/epjd/s10053-023-00635-3
    35. S.Tipawan Khlayboonme, Transition between Monoclinic and Tetragonal Β Phases Induced by Reactive Oxygen Gas in Rf-Sputtered V2o5 Thin Films, 2022, 1556-5068, 10.2139/ssrn.4118246
    36. R. Rohith, Anandhu Thejas Prasannakumar, V. Manju, Ranjini R. Mohan, Sreekanth J. Varma, Flexible, symmetric supercapacitor using self-stabilized dispersion-polymerised polyaniline/V2O5 hybrid electrodes, 2023, 467, 13858947, 143499, 10.1016/j.cej.2023.143499
    37. Hridya C. Prakash, M. Sathish Kumar, Tsung-Wu Lin, Sudip K. Batabyal, Photo-assisted capacitive performance of V2O5 supercapacitor, 2023, 469, 00134686, 143229, 10.1016/j.electacta.2023.143229
    38. Matías Picuntureo, José Antonio García-Merino, Roberto Villarroel, Samuel A. Hevia, The Synthesis of Sponge-like V2O5/CNT Hybrid Nanostructures Using Vertically Aligned CNTs as Templates, 2024, 14, 2079-4991, 211, 10.3390/nano14020211
    39. G.R. Khan, T.A. War, Silver-impregnated vanadium pentoxide nanofillers for reinforcement in titanium-based missile surfaces, 2024, 50, 24519049, 102581, 10.1016/j.tsep.2024.102581
    40. Melethil Sabna, K. Safna, Sona C P, Peediyekkal Jayaram, Surface‐Modified Mesoporous V2‐xSb2xO5‐δ Ceramics for Improving the Electrostatic Chemisorption of Methylene Blue, 2024, 9, 2365-6549, 10.1002/slct.202400862
    41. Christian Julien, Alain Mauger, 2025, Chapter 6, 978-3-031-67469-3, 515, 10.1007/978-3-031-67470-9_6
    42. Ningaraju Gejjiganahalli Ningappa, Anil Kumar Madikere Raghunatha Reddy, Karim Zaghib, Advanced Polymer Electrolytes in Solid-State Batteries, 2024, 10, 2313-0105, 454, 10.3390/batteries10120454
    43. Doaa Essam, Ashour M. Ahmed, Ahmed A. Abdel-Khaliek, Mohamed Shaban, Mohamed Rabia, One pot synthesis of poly m-toluidine incorporated silver and silver oxide nanocomposite as a promising electrode for supercapacitor devices, 2025, 15, 2045-2322, 10.1038/s41598-024-84848-5
    44. Henry Thomas, A. Ashok Kumar, S. Kaleemulla, V. Rajagopal Reddy, Influence of substrate temperature on the electrical and photovoltaic properties of V2O5 modified n-Ge heterojunction, 2025, 225, 00381101, 109074, 10.1016/j.sse.2025.109074
    45. I. Kartharinal Punithavthy, A. Sherin Fathima, B. Arunkumar, M. Jothibas, S. Suthakaran, Lalitha Gnanasekaran, Manikandan Ayyar, Effect of Ag Doped V2O5 Thin Films on Structural, Morphological, Electrical Properties and Their Photocatalytic Application, 2025, 59, 1063-7826, 149, 10.1134/S1063782624602462
  • Reader Comments
  • © 2018 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(10474) PDF downloads(2545) Cited by(45)

Article outline

Figures and Tables

Figures(20)  /  Tables(2)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog